Skip to main content
Log in

Additive representation of separable preferences over infinite products

  • Published:
Theory and Decision Aims and scope Submit manuscript

Abstract

Let \(\mathcal{X }\) be a set of outcomes, and let \(\mathcal{I }\) be an infinite indexing set. This paper shows that any separable, permutation-invariant preference order \((\succcurlyeq )\) on \(\mathcal{X }^\mathcal{I }\) admits an additive representation. That is: there exists a linearly ordered abelian group \(\mathcal{R }\) and a ‘utility function’ \(u:\mathcal{X }{{\longrightarrow }}\mathcal{R }\) such that, for any \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\) which differ in only finitely many coordinates, we have \(\mathbf{x}\succcurlyeq \mathbf{y}\) if and only if \(\sum _{i\in \mathcal{I }} \left[u(x_i)-u(y_i)\right]\ge 0\). Importantly, and unlike almost all previous work on additive representations, this result does not require any Archimedean or continuity condition. If \((\succcurlyeq )\) also satisfies a weak continuity condition, then the paper shows that, for any \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\), we have \(\mathbf{x}\succcurlyeq \mathbf{y}\) if and only if \({}^*\!\sum _{i\in \mathcal{I }} u(x_i)\ge {}^*\!\sum _{i\in \mathcal{I }}u(y_i)\). Here, \({}^*\!\sum _{i\in \mathcal{I }} u(x_i)\) represents a nonstandard sum, taking values in a linearly ordered abelian group \({}^*\!\mathcal{R }\), which is an ultrapower extension of \(\mathcal{R }\). The paper also discusses several applications of these results, including infinite-horizon intertemporal choice, choice under uncertainty, variable-population social choice and games with infinite strategy spaces.

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.

Institutional subscriptions

Fig. 1

Similar content being viewed by others

Notes

  1. See Sect. 5 for the formal definition of \({}^*\!\mathbb{R }\). See Anderson (1991) or Goldblatt (1998) for good introductions to nonstandard analysis.

  2. See, e.g. Krantz et al. (1971, §1.5.2, §6.5.1 and §9.1) and (1990, §21.7), Pfanzagl (1968, §6.6 and §9.5), Narens (1974a), and Fuhrken and Richter (1991).

  3. See e.g. Sidgwick (1884, Book IV, Chapt. I, §1, p. 411), Ramsey (1928, p. 543) and Rawls (1928, §44, p. 253). Note that ‘instrumental’ time preferences may be ethically defensible, even if ‘pure’ time preferences are not. See Sect. 6 for further discussion of this point.

  4. It would remain infinite for any exponential discount rate less than 2 % per year.

  5. Note that I do not assume a probability distribution on \(\mathcal{I }\).

  6. Thus, the elements of \(\mathcal{X }\) are ‘extended alternatives’, which encode both the specific identity of a person and any ethically relevant information about her physical and mental state.

  7. That is: for any distinct \(r,s\in \mathcal{R }\), either \(r>s\) or \(s>r\), but not both.

  8. In Savage’s theory, this property is called the sure thing principle or Axiom P2. In axiomatic measurement theory, it is variously called (joint) independence or single cancellation. In social choice, separability is a special case of the axiom of independence of (or elimination of) indifferent individuals, which in turn is a special case of the Extended Pareto axiom.

  9. But see Sect. 6.

  10. Formally, \({}^*\!\mathcal{R }\) is an ultrapower of \(\mathcal{R }\) with respect to an ultrafilter \(\mathfrak{UF }\) defined over the set of all finite subsets of \(\mathcal{I }\). The precise construction of \({}^*\!\mathcal{R }\) is somewhat technical, and will be provided in Sect. 5 below.

  11. This recalls preference relations which have been proposed by Lauwers and Vallentyne (2004) and Asheim et al. (2010); see Sect. 7.2 for discussion.

  12. This recalls intertemporal preference relations proposed by Atsumi (1965) and Weizsäcker (1965); again see Sect. 7.2 for discussion.

  13. But see Sect. 6.

  14. ‘Generalized’ because \(u\) might actually be a monotone increasing transformation of the ‘true’ cardinal utility function of the individuals.

  15. In different contexts, the Archimedean property has been called continuity or substitutability.

  16. Of course, if \((\succcurlyeq )\) itself is strictly finitary, then \(\big ( {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathrm fin}}}\big )\) is \((\succcurlyeq )\).

  17. In the case \(\mathcal{I }=\mathbb{N }\), Proposition 5(a) is equivalent to Theorem 1.1 of Wakker (1986), which in turn builds on earlier, similar results by Camacho (1979a, b, 1980, 1982). For a similar result, representing separable preferences using utility averages (rather than utility sums), see Kothiyal et al. (2012).

  18. Technical note (to be read after Sect. 5): The results in this section require a specific—but quite natural—choice of ultrafilter \(\mathfrak{UF }\). Let \(\Pi :=\Pi _{\scriptscriptstyle {\mathrm{fs}}}\) be the group of all fixed-step permutations of \(\mathbb{N }\), from Example 16(c). Then define \(\mathfrak{UF }\) as in Lemma 15. Then use \(\mathfrak{UF }\) to define \({}^*\!\mathcal{R }, {}^*\!\sum \) and \(\big ({}^{^*}\!{ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\).

  19. Technical note (to be read after Sect. 5): Let \({\varvec{\mathcal{F }}}\) be the set of finite subsets of \(\mathcal{I }, \mathfrak{UF }\) be an ultrafilter on \({\varvec{\mathcal{F }}}\), and define \(\mathcal{R }\!:=\!\mathbb{R }^{\varvec{\mathcal{F }}}/\mathfrak{UF }\). Thus, \(\mathcal{R }\) is a hyperreal number field. Now define \({}^*\!\mathcal{R }\!:=\!\mathcal{R }^{\varvec{\mathcal{F }}}/\mathfrak{UF }\). Technically, \({}^*\!\mathcal{R }\) is another a hyperreal number field, but it is distinct from \(\mathcal{R }\). I could have referred to it as \(^{**}\mathbb{R }\), but this seemed like a notation too far.

  20. The anonymous referee has pointed out that this property was originally introduced by Finetti (1931).

  21. Strictly speaking, we would need to divide by the normalization factor \({}^*\!\sum _{i\in \mathcal{I }} 1\) for this to be true.

  22. Here I adopt the usual convention that \(0\cdot \log _2(0):=0\).

  23. To be somewhat more precise: if \((x_1,x_2,\ldots ,x_N)\) was a long sequence of independent, \(\mu \)-random variables, then, using an optimal encoding scheme, it would take approximately \(N\cdot H(\mu )\) bits to transmit complete information about the sequence \((x_1,x_2,\ldots ,x_N)\). This approximation becomes exact as \(N{\rightarrow }{\infty }\). For more information on entropy and information, see Cover and Thomas (2006).

  24. It is possible to define, e.g. the entropy of a Borel probability measure on a compact subset of \(\mathbb{R }^N\), but only relative to some ‘baseline’ measure—typically the Lebesgue measure. The baseline measure is then the measure of maximal entropy. But this raises the question: what is the right baseline measure?

  25. This follows from the construction of \(\mu \) in the proof of Proposition 10.

  26. Formally, an ultrafilter \(\mathfrak{UF }\) is equivalent to a finitely additive, \(\{0,1\}\)-valued probability measure defined on all subsets of \({\varvec{\mathcal{F }}}\). The elements of \(\mathfrak{UF }\) are the ‘sets of measure 1’, and the elements of \(\mathfrak{P }\setminus \mathfrak{UF }\) are the ‘sets of measure 0’. I mean ‘almost all’ with respect to this probability measure. Just as in classical probability theory, this use of ‘almost all’ is only meaningful with respect to a particular ultrafilter. If \(\mathfrak{UF }\) and \(\mathfrak{UF }'\) are two different ultrafilters, then there will exist a subset \({\varvec{\mathcal{G }}}\subset {\varvec{\mathcal{F }}}\) such that \(\mathfrak{UF }\) judges \({\varvec{\mathcal{G }}}\) to contain ‘almost all’ elements of \({\varvec{\mathcal{F }}}\), while \(\mathfrak{UF }'\) judges \({\varvec{\mathcal{F }}}\setminus {\varvec{\mathcal{G }}}\) to contain ‘almost all’ elements of \({\varvec{\mathcal{F }}}\).

  27. Proof: \(\big ({ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathfrak{UF }}}}}\big )\) is complete by Axiom (UF). Next, \(\big ({ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathfrak{UF }}}}}\big )\) is reflexive, because \({\varvec{\mathcal{F }}}\in \mathfrak{UF }\). Finally, \(\big ({ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathfrak{UF }}}}}\big )\) is transitive, by Axioms (F1) and (F2).

  28. Formally, \({}^*\!\mathcal{R }\) is called the ultrapower of \(\mathcal{R }\) modulo the ultrafilter \(\mathfrak{UF }\). It is conventional to denote ultrapower-related objects with the leading star \(*\).

  29. In fact, Lemma 13 is a special case of Łoś’s theorem, which roughly states that any first-order properties of any system of algebraic structures and/or \(N\)-ary relations on \(\mathcal{R }\) are ‘inherited’ by \({}^*\!\mathcal{R }\). For example, if \(\mathcal{R }\) is a linearly ordered field, then \({}^*\!\mathcal{R }\) will also be a linearly ordered field.

  30. A positive element of \({}^*\!\mathcal{R }\) is ‘infinitesimal’ if it is smaller than every positive element of \(\mathcal{R }'\). It is ‘infinite’ if it is bigger than every element of \(\mathcal{R }'\).

  31. For any infinite indexing set \(\mathcal{N }\) and any ultrafilter \(\varvec{\mathcal{U }\!\mathcal{F }}\) on \(\mathcal{N }\), one can obtain a hyperreal number field by taking the ultrapower \(\mathbb{R }^\mathcal{N }/\varvec{\mathcal{U }\!\mathcal{F }}\). Technically, different choices of \(\mathcal{N }\) and/or \(\varvec{\mathcal{U }\!\mathcal{F }}\) will lead to different fields, so it is somewhat inaccurate to speak of ‘the’ hyperreal number field \({}^*\!\mathbb{R }\) as if it was a single object. However, all of these fields satisfy the axioms of nonstandard analysis (e.g. they are non-Archimedean linearly ordered field extensions of the field \(\mathbb{R }\) of real numbers), hence they can be treated as ‘the same’ object for most purposes.

  32. That is: \({\left\langle \Delta \right\rangle }:= \{\delta _1^{n_1}\cdot \delta _2^{n_2}\cdots \delta _k^{n_k}\); \(k\in \mathbb{N }\), \(\delta _1,\ldots ,\delta _k\in \Delta \), and \(n_1,\ldots ,n_k\in \mathbb{Z }\}\).

  33. Basu and Mitra (2007a) show that a permutation group \(\Gamma \subset \Pi \) can be the symmetry group of some Paretian social welfare relation on \(\mathbb{R }^\mathbb{N }\) if and only if each single element of \(\Gamma \) has finite orbits. The condition of locally finite orbits is similar, but somewhat more restrictive.

  34. \(\Pi _{\scriptscriptstyle {\mathrm{fs}}}\)-invariant social welfare relations on \(\mathbb{R }^\mathbb{N }\) have been considered by Lauwers (1997), Fleurbaey and Michel (2003; §4.2) and Basu and Mitra (2007a; §5).

  35. Lauwers (1998); Basu and Mitra (2003, 2007a, b), Fleurbaey and Michel (2003; Thm. 1) and Sakai (2010; Thm.2) have analyzed this Pareto/anonymity conflict in greater detail. Seidenfeld et al. (2009) have observed an analogous conflict between permutation-invariance and the statewise dominance principle, in the setting of risky decisions.

  36. Standardization maps all infinite hyperreals to \(\pm {\infty }\), and strips the infinitesimal part away from all finite hyperreals, leaving only the real part.

  37. Of course, individuals can still derive (dis)utility from memory of the past, anticipation of the future, altruism/envy towards other people or the contingency of fate, as long as the relevant cognitive states are explicitly encoded in \(\mathcal{X }\).

  38. Of particular note is Jaffray (1974b), which gives necessary and sufficient conditions for real-valued additive representations in full generality, without the use of a continuity or solvability condition.

  39. Wakker (1986) built on Camacho’s (1979a, b, 1980, 1982) earlier theory of additive preferences over the space of arbitrary-length finite sequences in \(\mathcal{X }\). Later, Wakker and Zank (1999) developed another additive representation for infinite Cartesian powers, using measure theory. Meanwhile, Streufert (1995) has analyzed the separability structure of partially separable (but not necessarily additive) preferences on infinite Cartesian products.

  40. See also Chipman (1971). In fact, this result had earlier been proved independently in papers by Sierpiński, Cuesta and Mendelson; see Fishburn (1974; §5) for details. Starting from Chipman’s work, Herden and Mehta (2004) have developed continuous lexicographical ordinal utility functions.

  41. Nonstandard analysis also has other applications in economics, unrelated to the topic of this paper; see e.g. see Anderson (1991) and Stigum (1990).

  42. See footnote 19.

  43. Condition (i) is because \(r\) is finite. Condition (ii) ensures uniqueness by excluding binary expansions ending in an infinite sequence of \(1\)’s.

References

  • Anderson, R. M. (1991). Nonstandard analysis with applications to economics. In Handbook of mathematical economics (Vol. IV, pp. 2145–2208). Amsterdam: North-Holland.

  • Arrow, K. J., Sen, A. K., & Suzumura, K. (Eds.). (2002). Handbook of social choice and welfare (Vol. I). Amsterdam: North-Holland.

    Google Scholar 

  • Asheim, G. B., d’Aspremont, C., & Banerjee, K. (2010). Generalized time-invariant overtaking. Journal of Mathematical Economics, 46(4), 519–533.

    Article  Google Scholar 

  • Asheim, G. B., & Tungodden, B. (2004). Resolving distributional conflicts between generations. Economic Theory, 24(1), 221–230.

    Article  Google Scholar 

  • Atsumi, H. (1965). Neoclassical growth and the efficient program of capital accumulation. Review of Economic Studies, 32, 127–136.

    Article  Google Scholar 

  • Banerjee, K. (2006). On the extension of the utilitarian and Suppes-Sen social welfare relations to infinite utility streams. Social Choice and Welfare, 27(2), 327–339.

    Article  Google Scholar 

  • Basu, K., & Mitra, T. (2003). Aggregating infinite utility streams with intergenerational equity: The impossibility of being Paretian. Econometrica, 71(5), 1557–1563.

    Article  Google Scholar 

  • Basu, K., & Mitra, T. (2007a). On the existence of Paretian social welfare relations for infinite utility streams with extended anonymity. In J. Roemer & J. Suzumura (Eds.), Intergenerational equity and sustainability: Conference proceedings of the IWEA roundtable meeting on intergenerational equity (Palgrave) (pp. 85–100).

  • Basu, K., & Mitra, T. (2007b). Possibility theorems for aggregating infinite utility streams equitably. In J. Roemer & K. Suzumura (Eds.), Intergenerational equity and sustainability: Conference proceedings of the IWEA roundtable meeting on intergenerational equity (Palgrave) (pp. 69–84).

  • Basu, K., & Mitra, T. (2007c). Utilitarianism for infinite utility streams: A new welfare criterion and its axiomatic characterization. The Journal of Economic Theory, 133(1), 350–373.

    Article  Google Scholar 

  • Benci, V., Horsten, L., & Wenmackers, S. (2012). Axioms for non-archimedean probability. In J. De Vuyst & L. Demey (Eds.), Future directions for logic; Proceedings of PhDs in Logic III (Vol. 2, pp. 127–135). London: IfColog Proceedings London, College Publications.

    Google Scholar 

  • Benci, V., Horsten, L., & Wenmackers, S. (2013, forthcoming). Non-archimedean probability. Milan Journal of Mathematics.

  • Bernstein, A. R. (1974). A non-standard integration theory for unbounded functions. Zeitschrift für Mathematische Logik und Grundlagen der Mathematik, 20, 97–108.

    Article  Google Scholar 

  • Bernstein, A. R., & Wattenberg, F. (1969). Holt. New York: Rinehart and Winston.

    Google Scholar 

  • Blackorby, C., Bossert, W., & Donaldson, D. (1998). Uncertainty and critical-level population principles. Journal of Population Economics, 11(1), 1–20.

    Article  Google Scholar 

  • Blackorby, C., Bossert, W., & Donaldson, D. (2002). Utilitarianism and the theory of justice. In Arrow, et al. (Eds.), (pp. 543–596).

  • Blackorby, C., Bossert, W., & Donaldson, D. (2005). Population issues in social choice theory, welfare economics, and ethics. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Blume, L., Brandenburger, A., & Dekel, E. (1989). An overview of lexicographic choice under uncertainty. Annals of Operations Research, 19(1–4), 231–246. Choice under uncertainty.

    Google Scholar 

  • Blume, L., Brandenburger, A., & Dekel, E. (1991a). Lexicographic probabilities and choice under uncertainty. Econometrica, 59(1), 61–79.

    Article  Google Scholar 

  • Blume, L., Brandenburger, A., & Dekel, E. (1991b). Lexicographic probabilities and equilibrium refinements. Econometrica, 59(1), 81–98.

    Article  Google Scholar 

  • Camacho, A. (1979a). Maximizing expected utility and the rule of long run success. In M. Allais & O. Hagen (Eds.), Expected utility and the allais paradox. Dordrecht: D. Reidel.

  • Camacho, A. (1979b). On cardinal utility. Theory and Decision, 10(1–4), 131–145. Methodological developments in social sciences.

    Article  Google Scholar 

  • Camacho, A. (1980). Approaches to cardinal utility. Theory and Decision, 12(4), 359–379.

    Article  Google Scholar 

  • Camacho, A. (1982). Societies and social decision functions. Dordrecht: D. Reidel.

    Book  Google Scholar 

  • Chipman, J. S. (1960). The foundations of utility. Econometrica, 28, 193–224.

    Article  Google Scholar 

  • Chipman, J. S. (1971). On the lexicographic representation of preference orderings. In Preferences, utility and demand (a Minnesota Symposium) (pp. 276–288). New York: Harcourt Brace Jovanovich.

  • Cover, T. M., & Thomas, J. A. (2006). Elements of information theory (2nd ed.). Hoboken, NJ: Wiley.

    Google Scholar 

  • de Finetti, B. (1931). Sul significato soggettivo della probabilitá. Fundamenta Mathematicae, 17, 298–329. (trans. into English as On the subjective meaning of probability, in de Finetti, 1993).

  • de Finetti, B. (1993). Probabilità e induzione. In P. Monari & D. Cocch (Eds.), (2nd ed., Vol. 5, supplement of Statistica). Bologna: Cooperativa Libraria Universitaria Editrice Bologna (trans. by M. Khale).

  • Debreu, G. (1960). Topological methods in cardinal utility theory. In Mathematical methods in the social sciences (pp. 16–26). Stanford, CA: Stanford University Press.

  • Fishburn, P. (1999). Preference structures and their numerical representations. Theoretical Computer Science, 217(2), 359–383. oRDAL ’96 (Ottawa, ON).

    Google Scholar 

  • Fishburn, P. C. (1974). Lexicographic orders, utilities and decision rules: A survey. Management Science, 20, 1442–1471.

    Article  Google Scholar 

  • Fishburn, P. C. (1982). The foundations of expected utility. In Theory and decision library (Vol. 31). Dordrech: D. Reide.

  • Fishburn, P. C., & LaValle, I. H. (1998). Subjective expected lexicographic utility: Axioms and assessment. Annals of Operations Research, 80, 183–206.

    Article  Google Scholar 

  • Fleming, M. (1952). A cardinal concept of welfare. Quarterly Journal of Economics, 66(3), 284–366.

    Article  Google Scholar 

  • Fleurbaey, M., & Michel, P. (2003). Intertemporal equity and the extension of the Ramsey criterion. Journal of Mathematical Economics, 39(7), 777–802.

    Article  Google Scholar 

  • Fuhrken, G., & Richter, M. K. (1991). Additive utility. Economic Theory, 1(1), 83–105.

    Article  Google Scholar 

  • Goldblatt, R. (1998). Lectures on the hyperreals. In Graduate texts in mathematics (Vol. 188). New York: Springer.

  • Gorman, W. (1968). The structure of utility functions. Review of Economic Studies, 35(390), 367.

    Article  Google Scholar 

  • Gravett, K. A. H. (1956). Ordered abelian groups. Quarterly Journal of Mathematics: Oxford Series, 7(2), 57–63.

    Article  Google Scholar 

  • Halpern, J. Y. (2009). A nonstandard characterization of sequential equilibrium, perfect equilibrium, and proper equilibrium. International Journal of Game Theory, 38(1), 37–49.

    Article  Google Scholar 

  • Halpern, J. Y. (2010). Lexicographic probability, conditional probability, and nonstandard probability. Games and Economic Behavior, 68(1), 155–179.

    Article  Google Scholar 

  • Hammond, P. (1994). Elementary non-archimedean representations of probability for decision theory and games. In P. Humphreys (Ed.), Patrick Suppes: scientific philosoher (Vol. 1, pp. 25–49). Dordrecht: Kluwer.

    Chapter  Google Scholar 

  • Hammond, P. (1999). Consequentialism, non-archimedean probabilities, and lexicographic expected utility. In C. Bicchieri, R. Jeffrey, & B. Skyrms (Eds.), The logic of strategy. Oxford: Oxford University Press.

    Google Scholar 

  • Harsanyi, J. (1955). Cardinal welfare, individualistic ethics and interpersonal comparisons of utility. Journal of Political Economy, 63, 309–321.

    Article  Google Scholar 

  • Hausner, M. (1954). Multidimensional utilities. In R. Thrall & C. Coombs (Eds.), Decision processes (pp. 167–180). New York: Wiley.

    Google Scholar 

  • Hausner, M., & Wendel, J. G. (1952). Ordered vector spaces. Proceedings of the American Mathematical Society, 3, 977–982.

    Article  Google Scholar 

  • Henson, C. W. (1972). On the nonstandard representation of measures. Transactions of the American Mathematical Society, 172, 437–446.

    Article  Google Scholar 

  • Herden, G., & Mehta, G. B. (2004). The Debreu gap lemma and some generalizations. Journal of Mathematical Economics, 40(7), 747–769.

    Article  Google Scholar 

  • Herzberg, F. (2007). Internal laws of probability, generalized likelihoods and Lewis’ infinitesimal chances. British Journal of Philosophy of Science, 58(1), 25–43.

    Article  Google Scholar 

  • Herzberg, F. (2009). Elementary non-Archimedean utility theory. Mathematical Social Sciences, 58(1), 8–14.

    Article  Google Scholar 

  • Herzberg, F. (2011). Hyperreal expected utilities and Pascal’s wager. Logique et Anal. (N.S.), 54(213), 69–108.

    Google Scholar 

  • Holder, R. (2002). Fine-tuning, multiple universes and theism. Nous, 36(2), 295–312.

    Article  Google Scholar 

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

  • Jaffray, J.-Y., (1974a). Existence, propriétés de continuité, additivité de fonctions d’utilité sur un espace partiellement ou totalement ordonné. Ph.D. thesis, Université de Paris VI, Paris.

  • Jaffray, J.-Y. (1974b). On the extension of additive utilities to infinite sets. Journal of Mathematical Psychology, 11, 431–452.

    Article  Google Scholar 

  • Jaynes, E. (1968). Prior probabilities. IEEE Transactions on Systems Science and Cybernetics, 4(3), 227–241.

    Google Scholar 

  • Kannai, Y. I. (1992). Nonstandard concave utility functions. Journal of Mathematical Economics, 21(1), 51–58.

    Article  Google Scholar 

  • Kirman, A. P., & Sondermann, D. (1972). Arrow’s theorem, many agents, and invisible dictators. Journal of Economic Theory, 5(2), 267–277.

    Article  Google Scholar 

  • Kothiyal, A. A., Spinu, V., & Wakker, P. (April, 2013). Average utility: A preference foundation (preprint).

  • Krantz, D. H., Luce, R. D., Suppes, P., & Tversky, A. (1971). Foundations of measurement I: Additive and polynomial representations. New York: Academic Press.

    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 

  • Lauwers, L. (1997). Topological aggregation, the case of an infinite population. Social Choice and Welfare, 14(2), 319–332.

    Article  Google Scholar 

  • Lauwers, L. (1998). Intertemporal objective functions: Strong Pareto versus anonymity. Mathematical Social Sciences, 35(1), 37–55.

    Article  Google Scholar 

  • Lauwers, L. (2010). Ordering infinite utility streams comes at the cost of a non-Ramsey set. Journal of Mathematical Economics, 46(1), 32–37.

    Article  Google Scholar 

  • Lauwers, L., & Vallentyne, P. (2004). Infinite utilitarianism: More is always better. Economics and Philosophy, 20, 307–330.

    Article  Google Scholar 

  • Lauwers, L., & Van Liedekerke, L. (1995). Ultraproducts and aggregation. Journal of Mathematical Economics, 24(3), 217–237.

    Article  Google Scholar 

  • Lehmann, D. (2001). Expected qualitative utility maximization. Games and Economic Behavior, 35(1–2), 54–79.

    Article  Google Scholar 

  • Lehmann, D., & Magidor, M. (1992). What does a conditional knowledge base entail? Artificial Intelligence, 55, 1–60.

    Article  Google Scholar 

  • Luce, R. D., Krantz, D. H., Suppes, P., & Tversky, A. (1990). Foundations of measurement III: Representation, axiomatization, and invariance. San Diego, CA: Academic Press.

    Google Scholar 

  • McGee, V. (1994). Learning the impossible. In E. Eells & B. Skyrms (Eds.), Probability and conditionals (pp. 177–199). Cambridge: Cambridge University Press.

    Google Scholar 

  • Narens, L. (1974a). Measurement without Archimedean axioms. Philosophy of Science, 41, 374–393.

    Article  Google Scholar 

  • Narens, L. (1974b). Minimal conditions for additive conjoint measurement and qualitative probability. Journal of Mathematical Psychology, 11, 404–430.

    Article  Google Scholar 

  • Narens, L. (1985). Abstract measurement theory. Cambridge, MA: MIT Press.

    Google Scholar 

  • Norton, J. (2007). Probability disassembled. British Journal of Philosophy of Science, 58(2), 141–171.

    Article  Google Scholar 

  • Parfit, D. (1984). Reasons and persons. Oxford: Clarendon Press.

    Google Scholar 

  • Pedersen, A. P. (2013). Archimedes wake. Ph.D. thesis, Carnegie Mellon University, Department of Philosophy.

  • Pfanzagl, J (In cooperation with V. Baumann and H. Huber). (1968). Theory of measurement. Physica-Verlag, Würzburg: .

  • Ramsey, F. P. (1928). A mathematical theory of saving. Economic Journal, 38, 543–559.

    Article  Google Scholar 

  • Rawls, J. (1999). A Theory of Justice, revised ed. Oxford, UK: Oxford University Press.

  • Richter, M. K. (1971). Rational choice. In Preferences, utility and demand (a Minnesota symposium) (pp. 29–58). New York: Harcourt Brace Jovanovich.

  • Roemer, J., & Suzumura, K. (Ed.). (2007). In Intergenerational equity and sustainability. Basingstoke: Palgrave Macmillan.

  • Ryberg, J., & Tännsjö, T. (Eds.). (2004). In The Repugnant conclusion. Essays on population. Dordrecht: Kluwer.

  • Sakai, T. (2010a). A characterization and an impossibility of finite length anonymity for infinite generations. Journal of Mathematical Economics, 46(5), 877–883.

    Article  Google Scholar 

  • Sakai, T. (2010b). Intergenerational equity and an explicit construction of welfare criteria. Social Choice and Welfare, 35(3), 393–414.

    Article  Google Scholar 

  • Savage, L. J. (1954). The foundations of statistics. New York: Wiley.

    Google Scholar 

  • Seidenfeld, T., Schervish, M. J., & Kadane, J. B. (2009). Preference for equivalent random variables: A price for unbounded utilities. Journal of Mathematical Economics, 45(5–6), 329–340.

    Google Scholar 

  • Shorb, A. M. (1975). Completely additive measure and integration. Proceedings of the American Mathematical Society, 53(2), 453–459.

    Article  Google Scholar 

  • Sidgwick, H. (1884). The methods of ethics. London: Macmillan.

    Google Scholar 

  • Skala, H. J. (1974). Nonstandard utilities and the foundation of game theory. International Journal of Game Theory, 3, 67–81.

    Article  Google Scholar 

  • Skala, H. J. (1975). Non-Archimedean utility theory. Berlin: Springer.

    Book  Google Scholar 

  • Stigum, B. P. (1990). Toward a formal science of economics. Cambridge, MA: MIT Press.

    Google Scholar 

  • Streufert, P. A. (1995). A general theory of separability for preferences defined on a countably infinite product space. Journal of Mathematical Economics, 24(5), 407–434.

    Article  Google Scholar 

  • von Weizsäcker, C. (1965). Existence of optimal programs of accumulation for an infinite time horizon. Review of Economic Studies, 32, 85–104.

    Article  Google Scholar 

  • Wakker, P. (1986). The repetitions approach to characterize cardinal utility. Theory and Decision, 20(1), 33–40.

    Article  Google Scholar 

  • Wakker, P. (1989). Additive representations of preferences. Dordrecht: Kluwer.

    Book  Google Scholar 

  • Wakker, P., & Zank, H. (1999). State dependent expected utility for Savage’s state space. Mathematics of Operations Research, 24(1), 8–34.

    Article  Google Scholar 

  • Wattenberg, F. (1977). Nonstandard measure theory-Hausdorff measure. Proceedings of the American Mathematical Society, 65(2), 326–331.

    Google Scholar 

  • Wattenberg, F. (1979). Nonstandard measure theory: Avoiding pathological sets. Transactions of the American Mathematical Society, 250, 357–368.

    Article  Google Scholar 

  • Wenmackers, S. (2012, forthcoming). Ultralarge lotteries: Analyzing the lottery paradox using non-standard analysis. Journal of Applied Logic.

  • Wilson, N. (1995). An order of magnitude calculus. In Proceedings of Eleventh Conference on Uncertainty in Artificial Intelligence (UAI ’95). Morgan Kaufmann, San Mateo, CA, pp. 548–555.

  • Zame, W. R. (2007). Can intergenerational equity be operationalized? Theoretical Economics, 2, 187–202.

    Google Scholar 

Download references

Acknowledgments

Part of this paper was written while visiting the Department of Economics at the Université de Montréal. I would like to thank the UdM and CIREQ for their hospitality. I would also like to thank Klaus Nehring, Clemens Puppe and especially Peter Wakker for their detailed comments on earlier drafts of this paper. Finally, I am grateful to the anonymous referee for a very helpful and perceptive report, which contained many good suggestions. This research was supported by NSERC grant \(\#262620-2008\).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marcus Pivato.

Appendix: Proofs

Appendix: Proofs

Proof of Theorem 1

\({\Longleftarrow }\)’ It is easy to check that the additive preorder defined by (2) is \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariant and separable.

\(\Longrightarrow \)’ Fix some \(o\in \mathcal{X }\), and let \((\mathcal{A },+)\) be the free abelian group generated by \(\mathcal{X }\setminus \{o\}\). That is, \(\mathcal{A }\) consists of all formal \(\mathbb{Z }\)-linear combinations of the form ‘\(J_1 x_1 + J_2 x_2 + \cdots + J_N x_N\)’ where \(N\in \mathbb{N }, J_1,\ldots ,J_N\in \mathbb{Z }\setminus \{0\}\) and \(x_1,\ldots ,x_N\in \mathcal{X }\setminus \{o\}\) are distinct.

Let \(\mathcal{B }^{\prime }\subset \mathcal{A }\) be the set of all elements where \(J_1,\ldots ,J_N>0\). For any such \(b\in \mathcal{B }'\), let us define \(\mathbf{w}^b\in \mathcal{X }^\mathcal{I }\) as follows. Let \(\mathcal{J }_1,\mathcal{J }_2,\ldots ,\mathcal{J }_N\) be disjoint subsets of \(\mathcal{I }\), with \(|\mathcal{J }_n|=J_n\) for all \(n\in {\left[ 1\ldots N \right]}\). For all \(n\in {\left[ 1\ldots N \right]}\) and all \(j\in \mathcal{J }_n\), let \(w^b_j := x_n\). Meanwhile, for all \(i\in \mathcal{I }\setminus \mathcal{J }_1\sqcup \cdots \sqcup \mathcal{J }_N\), define \(w^b_i:=o\). (Heuristic: if we regard the elements of \(\mathcal{X }\) as ‘goods’ and ‘bads’, then \(\mathbf{w}^b\) represents a ‘bundle’ containing \(J_n\) units of \(x_n\) for each \(n\in {\left[ 1\ldots N \right]}\).) Since \((\succcurlyeq )\) is \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariant, it does not matter how we choose the sets \(\mathcal{J }_1,\ldots ,\mathcal{J }_N\); if \(\mathbf{w}^b\) and \({\widetilde{\mathbf{w}}}^b\) are two different elements of \(\mathcal{X }^\mathcal{I }\) built using the above recipe, then we automatically have \(\mathbf{w}^b\approx {\widetilde{\mathbf{w}}}^b\).

Finally, if \(0\) denotes the identity element of \(\mathcal{A }\), then let \(\mathcal{B }:=\{0\}\sqcup \mathcal{B }'\), and define \(\mathbf{w}^0\) by setting \(w^0_i:=o\) for all \(i\in \mathcal{I }\). For any \(b,b^{\prime }\in \mathcal{B }\), if \(b= J_1 x_1 + \cdots + J_N x_N\) and \(b'= J'_1 x'_1 + \cdots + J'_M x'_M\), then say \(b\) and \(b'\) are disjoint if the sets \(\{x_1,\ldots ,x_N\}\) and \(\{x'_1,\ldots ,x'_M\}\) are disjoint.

Claim 1 Let \(a\in \mathcal{A }\).

  1. (a)

    There exist unique disjoint \(a_+,a_-\in \mathcal{B }\) such that \(a=a_+-a_-\).

  2. (b)

    Let \(b,c\in \mathcal{B }\) be any other elements such that \(a=b-c\). Then \((\mathbf{w}^b\succcurlyeq \mathbf{w}^c) \iff (\mathbf{w}^{a_+} \succcurlyeq \mathbf{w}^{a_-})\).

Proof

  1. (a)

    If \(a=0\), then let \(0_+:=0\) and \(0_-:=0\). Then \(0_+,0_-\in \mathcal{B }\) are disjoint, and \(0=0_+-0_-\). Now suppose that \(a\ne 0\). Let \(\displaystyle a= \sum _{w\in \mathcal{W }} A_w w\), where \(\mathcal{W }\subseteq \mathcal{X }\) is a finite subset and \(A_w\in \mathbb{Z }\setminus \{0\}\) for all \(w\in \mathcal{W }\). Then \(\mathcal{W }:=\mathcal{W }_-\sqcup \mathcal{W }_+\), where \(\mathcal{W }_-:=\{w\in \mathcal{W }\); \(A_w<0\}\) and \(\mathcal{W }_+:=\{w\in \mathcal{W }\); \(A_w>0\}\). Let \(\displaystyle a_+:= \sum _{w\in \mathcal{W }_+} A_w w\) and \(\displaystyle a_-:= \sum _{w\in \mathcal{W }_-} (-A_w) w\). (If \(\mathcal{W }_+=\emptyset \), then \(a_+:=0\). If \(\mathcal{W }_-=\emptyset \), then \(a_-:=0\).) Then \(a_+,a_-\in \mathcal{B }\) are disjoint and \(a=a_+-a_-\).

  2. (b)

    Suppose that \(\displaystyle b:= \sum _{y\in \mathcal{Y }} B_y y\) and \(\displaystyle c:= \sum _{z\in \mathcal{Z }} C_z z\), for some finite subsets \(\mathcal{Y },\mathcal{Z }\subseteq \mathcal{X }\) and coefficients \(B_y\in \mathbb{N }\) for all \(y\in \mathcal{Y }\) and \(C_z\in \mathbb{N }\) for all \(z\in \mathcal{Z }\). If \(b-c=a\), then we must have \(\mathcal{W }\subseteq \mathcal{Y }\cup \mathcal{Z }\), \(\mathcal{Y }\setminus \mathcal{Z }\subseteq \mathcal{W }_+\) and \(\mathcal{Z }\setminus \mathcal{Y }\subseteq \mathcal{W }_-\). Furthermore:

    • \(B_y = A_y\) for all \(y\in \mathcal{Y }\setminus \mathcal{Z }\).

    • \(C_z = -A_z\) for all \(z\in \mathcal{Z }\setminus \mathcal{Y }\).

    • \(B_w - C_w = A_w\) for all \(w\in \mathcal{Y }\cap \mathcal{Z }\cap \mathcal{W }\).

    • \(B_x = C_x\) for all \(x\in (\mathcal{Y }\cap \mathcal{Z })\setminus \mathcal{W }\).

    Let \(J:=\max \left(\{B_y; y\in \mathcal{Y }\}\cup \{C_z; z\in \mathcal{Z }\}\right)\). For all \(x\in \mathcal{Y }\cup \mathcal{Z }\), let \(\mathcal{J }_x\subset \mathcal{I }\) be a subset of cardinality \(J\), and suppose that all these sets are disjoint. Since \((\succcurlyeq )\) is \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariant, we can permute the coordinates of \(\mathbf{w}^b\) and/or \(\mathbf{w}^c\) each in any desired finitary way without changing their \((\succcurlyeq )\)-ordering. Likewise, we can finitarily permute \(\mathbf{w}^{a_+}\) and/or \(\mathbf{w}^{a_-}\) without changing their \((\succcurlyeq )\)-ordering. Thus, without loss of generality, we can suppose:

    • \(\mathbf{w}^{a_+}\) assigns the value \(w\) to exactly \(A_w\) coordinates in \(\mathcal{J }_w\), for each \(w\in \mathcal{W }_+\).

    • \(\mathbf{w}^{a_-}\) assigns the value \(w\) to exactly \(-A_w\) coordinates in \(\mathcal{J }_w\), for each \(w\in \mathcal{W }_-\).

    • \(\mathbf{w}^{b}\) assigns the value \(y\) to exactly \(B_y\) coordinates in \(\mathcal{J }_y\), for each \(y\in \mathcal{Y }\).

    • \(\mathbf{w}^{c}\) assigns the value \(z\) to exactly \(C_z\) coordinates in \(\mathcal{J }_z\), for each \(z\in \mathcal{Z }\).

    Let us describe these as the ‘active’ coordinates. In all four cases, let us assign all other (‘inactive’) coordinates the value \(o\). We can further assume that:

    1. 1.

      For all \(y\in \mathcal{Y }\setminus \mathcal{Z }\), \(\mathbf{w}^{a_+}\) and \(\mathbf{w}^b\) involve the same \(B_y=A_y\) active coordinates.

    2. 2.

      For all \(z\in \mathcal{Z }\setminus \mathcal{Y }, \mathbf{w}^{a_-}\) and \(\mathbf{w}^c\) involve the same \(C_z=-A_z\) active coordinates.

    3. 3.

      For all \(x\in (\mathcal{Y }\cap \mathcal{Z })\setminus \mathcal{W }\), \(\mathbf{w}^b\) and \(\mathbf{w}^c\) involve the same set \(\mathcal{I }_x\subseteq \mathcal{J }_x\) of \(B_x=C_x\) active coordinates.

    4. 4.

      For all \(w\in \mathcal{Y }\cap \mathcal{Z }\cap \mathcal{W }_+\), the set of active coordinates in \(\mathbf{w}^b\) is the disjoint union of the active coordinates of \(\mathbf{w}^{a_+}\) and \(\mathbf{w}^c\). In this case, let \(\mathcal{I }_w\subseteq \mathcal{J }_w\) be the active coordinates of \(\mathbf{w}^c\).

    5. 5.

      For all \(w\in \mathcal{Y }\cap \mathcal{Z }\cap \mathcal{W }_-\), the set of active coordinates in \(\mathbf{w}^c\) is the disjoint union of the active coordinates of \(\mathbf{w}^{a_-}\) and \(\mathbf{w}^b\). In this case, let \(\mathcal{I }_w\subseteq \mathcal{J }_w\) be the active coordinates of \(\mathbf{w}^b\).

    Now suppose that \(\mathbf{w}^b\succcurlyeq \mathbf{w}^c\). Let us define

    $$\begin{aligned} \mathcal{J }:= \bigsqcup _{x\in (\mathcal{Y }\cap \mathcal{Z })\setminus \mathcal{W }} \! \mathcal{I }_x \ \sqcup \ \bigsqcup _{w\in \mathcal{Y }\cap \mathcal{Z }\cap \mathcal{W }_+} \! \mathcal{I }_w \ \sqcup \ \bigsqcup _{w\in \mathcal{Y }\cap \mathcal{Z }\cap \mathcal{W }_-} \!\mathcal{I }_w. \end{aligned}$$

    Then \(\mathbf{w}^b_{\mathcal{J }} = \mathbf{w}^c_{\mathcal{J }}\), by assumptions 3, 4 and 5. Meanwhile \(\mathbf{w}^{a_+}_{\mathcal{J }} = \mathbf{o}_\mathcal{J }= \mathbf{w}^{a_-}_{\mathcal{J }}\), by assumptions 4 and 5. On the other hand, if \(\mathcal{K }:=\mathcal{I }\setminus \mathcal{J }\), then we have \(\mathbf{w}^b_\mathcal{K }= \mathbf{w}^{a_+}_\mathcal{K }\) (by assumptions 1, 4 and 5) and \(\mathbf{w}^c_\mathcal{K }= \mathbf{w}^{a_-}_\mathcal{K }\) (by assumptions 2, 4 and 5). Thus, the separability of \((\succcurlyeq )\) implies that \((\mathbf{w}^b\succcurlyeq \mathbf{w}^c) \iff (\mathbf{w}^{a_+} \succcurlyeq \mathbf{w}^{a_-})\), as desired.\(\diamondsuit \text { Claim } 1\)

Let \(\mathcal{C }_+:=\{a\in \mathcal{A }\); \(\mathbf{w}^{a_+} \succ \mathbf{w}^{a_-}\}\), \(\mathcal{C }_-:=\{a\in \mathcal{A }\); \(\mathbf{w}^{a_+} \prec \mathbf{w}^{a_-}\}\), and \(\mathcal{C }_0:=\{a\in \mathcal{A }\); \(\mathbf{w}^{a_+} \approx \mathbf{w}^{a_-}\}\). Then \(\mathcal{A }=\mathcal{C }_-\sqcup \mathcal{C }_0\sqcup \mathcal{C }_+\). Let \(\mathcal{C }_{0+}:=\mathcal{C }_0\sqcup \mathcal{C }_+\) and \(\mathcal{C }_{0-}:=\mathcal{C }_0\sqcup \mathcal{C }_-\).

Claim 2 Let \(b,c\in \mathcal{A }\).

  1. (a)

    If \(b,c\in \mathcal{C }_{0+}\), then \(b+c\in \mathcal{C }_{0+}\). If also \(b\in \mathcal{C }_+\) or \(c\in \mathcal{C }_+\) then \(b+c\in \mathcal{C }_+\).

  2. (b)

    If \(b,c\in \mathcal{C }_{0-}\), then \(b+c\in \mathcal{C }_{0-}\). If also \(b\in \mathcal{C }_-\) or \(c\in \mathcal{C }_-\) then \(b+c\in \mathcal{C }_-\).

  3. (c)

    \(b\in \mathcal{C }_{0+}\) if and only if \(-b\in \mathcal{C }_{0-}\).

  4. (d)

    \(\mathcal{C }_0\) is a subgroup of \(\mathcal{A }\).

Proof

  1. (a)

    Claim 1(a) says that \(b=b_+-b_-\) and \(c=c_+-c_-\) for some disjoint \(b_+,b_-,c^+,c^-\in \mathcal{B }\). Clearly, \(b+c = (b_+ + c_+) - (b_- + c_-)\), and \((b_+ + c_+)\) and \((b_- + c_-)\) are also elements of \(\mathcal{B }\). Define \(\mathcal{J }_b^+:=\mathcal{I }(\mathbf{w}^{b_+},\mathbf{o})\), \(\mathcal{J }_b^-:=\mathcal{I }(\mathbf{w}^{b_-},\mathbf{o})\), \(\mathcal{J }_c^+:=\mathcal{I }(\mathbf{w}^{c_+},\mathbf{o})\) and \(\mathcal{J }_c^-:=\mathcal{I }(\mathbf{w}^{c_-},\mathbf{o})\). By \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariance, we can assume without loss of generality that:

  • \(\mathcal{J }_b^+, \mathcal{J }_b^-, \mathcal{J }_c^+\), and \(\mathcal{J }_c^-\) are disjoint.

  • \(\mathcal{I }(\mathbf{w}^{b_+ + c_+},\mathbf{o})=\mathcal{J }_b^+\sqcup \mathcal{J }_c^+\). Furthermore, \(\mathbf{w}^{b_+ + c_+}_{\mathcal{J }_b^+}=\mathbf{w}^{b_+}_{\mathcal{J }_b^+}\) and \(\mathbf{w}^{b_+ + c_+}_{\mathcal{J }_c^+}=\mathbf{w}^{c_+}_{\mathcal{J }_c^+}\).

  • \(\mathcal{I }(\mathbf{w}^{b_- + c_-},\mathbf{o})=\mathcal{J }_b^-\sqcup \mathcal{J }_c^-\). Furthermore, \(\mathbf{w}^{b_- + c_-}_{\mathcal{J }_b^-}=\mathbf{w}^{b_-}_{\mathcal{J }_b^-}\) and \(\mathbf{w}^{b_- + c_-}_{\mathcal{J }_c^-}=\mathbf{w}^{c_-}_{\mathcal{J }_c^-}\). Let \(\mathcal{J }:=\mathcal{I }\setminus (\mathcal{J }_b^+\sqcup \mathcal{J }_b^-\sqcup \mathcal{J }_c^+\sqcup \mathcal{J }_c^-)\). Then we can indicate the three assumptions above with the following equations:

    $$\begin{aligned} \begin{array}{rcccccc} \mathbf{w}^{b_+} &{}=&{} (\mathbf{w}^{b_+}_{\mathcal{J }_b^+} &{} \mathbf{o}_{\mathcal{J }_b^-} &{} \mathbf{o}_{\mathcal{J }_c^+} &{} \mathbf{o}_{\mathcal{J }_c^-} &{}\mathbf{o}_\mathcal{J });\\ \mathbf{w}^{b_-} &{}=&{} (\mathbf{o}_{\mathcal{J }_b^+} &{} \mathbf{w}^{b_-}_{\mathcal{J }_b^-} &{} \mathbf{o}_{\mathcal{J }_c^+} &{} \mathbf{o}_{\mathcal{J }_c^-} &{}\mathbf{o}_\mathcal{J });\\ \mathbf{w}^{c_+} &{}=&{} (\mathbf{o}_{\mathcal{J }_b^+} &{} \mathbf{o}_{\mathcal{J }_b^-} &{} \mathbf{w}^{c_+}_{\mathcal{J }_c^+} &{} \mathbf{o}_{\mathcal{J }_c^-} &{}\mathbf{o}_\mathcal{J });\\ \mathbf{w}^{c_-} &{}=&{} (\mathbf{o}_{\mathcal{J }_b^+} &{} \mathbf{o}_{\mathcal{J }_b^-} &{} \mathbf{o}_{\mathcal{J }_c^+} &{} \mathbf{w}^{c_-}_{\mathcal{J }_c^-} &{}\mathbf{o}_\mathcal{J });\\ \mathbf{w}^{b_+ + c_+} &{}=&{} (\mathbf{w}^{b_+}_{\mathcal{J }_b^+} &{}\mathbf{o}_{\mathcal{J }_b^-} &{} \mathbf{w}^{c_+}_{\mathcal{J }_c^+} &{} \mathbf{o}_{\mathcal{J }_c^-} &{} \mathbf{o}_\mathcal{J });\\ \text { and }\mathbf{w}^{b_- + c_-} &{}=&{} (\mathbf{o}_{\mathcal{J }_b^+} &{} \mathbf{w}^{b_-}_{\mathcal{J }_b^-}&{} \mathbf{o}_{\mathcal{J }_c^+} &{} \mathbf{w}^{c_-}_{\mathcal{J }_c^-} &{}\mathbf{o}_\mathcal{J }). \end{array} \end{aligned}$$
    (13)

    Now, \(b\in \mathcal{C }_{0+}\), so \(\mathbf{w}^{b_+} \succcurlyeq \mathbf{w}^{b_-}\). Applying separability in the \(\mathcal{J }_c^+\) coordinates to the first two equations of (13), we get

    $$\begin{aligned} (\mathbf{w}^{b_+}_{\mathcal{J }_b^+}, \mathbf{o}_{\mathcal{J }_b^-}, \mathbf{w}^{c_+}_{\mathcal{J }_c^+}, \mathbf{o}_{\mathcal{J }_c^-},\mathbf{o}_\mathcal{J }) \succcurlyeq (\mathbf{o}_{\mathcal{J }_b^+}, \mathbf{w}^{b_-}_{\mathcal{J }_b^-}, \mathbf{w}^{c_+}_{\mathcal{J }_c^+}, \mathbf{o}_{\mathcal{J }_c^-},\mathbf{o}_\mathcal{J }). \end{aligned}$$
    (14)

    Also, \(c\in \mathcal{C }_{0+}\), so \(\mathbf{w}^{c_+} \succcurlyeq \mathbf{w}^{c_-}\). Applying separability in the \(\mathcal{J }_b^-\) coordinates to the third and fourth equations of (13), we get

    $$\begin{aligned} (\mathbf{o}_{\mathcal{J }_b^+},\mathbf{w}^{b_-}_{\mathcal{J }_b^-}, \mathbf{w}^{c_+}_{\mathcal{J }_c^+}, \mathbf{o}_{\mathcal{J }_c^+-},\mathbf{o}_\mathcal{J }) \succcurlyeq (\mathbf{o}_{\mathcal{J }_b^+}, \mathbf{w}^{b_-}_{\mathcal{J }_b^-}, \mathbf{o}_{\mathcal{J }_c^+}, \mathbf{w}^{c_-}_{\mathcal{J }_c^+-},\mathbf{o}_\mathcal{J }). \end{aligned}$$
    (15)

    Combining Eqs. (14) and (15) via transitivity, we get

    $$\begin{aligned} (\mathbf{w}^{b_+}_{\mathcal{J }_b^+}, \mathbf{o}_{\mathcal{J }_b^-}, \mathbf{w}^{c_+}_{\mathcal{J }_c^+}, \mathbf{o}_{\mathcal{J }_c^-},\mathbf{o}_\mathcal{J }) \succcurlyeq (\mathbf{o}_{\mathcal{J }_b^+}, \mathbf{w}^{b_-}_{\mathcal{J }_b^-}, \mathbf{o}_{\mathcal{J }_c^+}, \mathbf{w}^{c_-}_{\mathcal{J }_c^+-},\mathbf{o}_\mathcal{J }). \end{aligned}$$
    (16)

    Now, matching the two sides of Eq. (16) with the last two equations in (13), we get \(\mathbf{w}^{b_+ + c_+} \succcurlyeq \mathbf{w}^{b_- + c_-}\). But \(b+c = (b_+ + c_+) - (b_- + c_-)\). Thus, Claim 1(b) implies that \(b+c\in \mathcal{C }_{0+}\). If \(b\in \mathcal{C }_{+}\), then \(\mathbf{w}^{b_+} \succ \mathbf{w}^{b_-}\), which makes Eq. (14) strict, which makes Eq. (16) strict, which means that \(\mathbf{w}^{b_+ + c_+} \succ \mathbf{w}^{b_- + c_-}\), and hence \(b+c\in \mathcal{C }_{+}\). Likewise, if \(c\in \mathcal{C }_{+}\), then Eqs. (15) and (16) become strict, so that \(b+c\in \mathcal{C }_{+}\).

    1. (b)

      Similar to (a).

    2. (c)

      Let \(b\in \mathcal{C }_{0+}\), and write \(b=b_+-b_-\). If \(c=-b\), then \(c = b_--b_+\), and these elements are disjoint, so the uniqueness part of Claim 1(a) implies that \(c_+=b_-\) and \(c_-=b_+\). We have \(\mathbf{w}^{b_+}\succcurlyeq \mathbf{w}^{b_-}\) (because \(b\in \mathcal{C }_{0+}\)); thus, \(\mathbf{w}^{c_-}\succcurlyeq \mathbf{w}^{c_+}\), so \(c\in \mathcal{C }_{0-}\). By identical argument, if \(-b\in \mathcal{C }_{0-}\), then \(b\in \mathcal{C }_{0+}\).

    3. (d)

      First note that (a) and (b) together imply that \(\mathcal{C }_0\) is closed under addition, and (c) implies that \(\mathcal{C }_0\) is closed under inverses. Finally, we have \(0\in \mathcal{C }_0\), because \(0_+=0_-=0\), so that \(\mathbf{w}^{0_-}=\mathbf{w}^{0_+}=\mathbf{w}^0=\mathbf{o}\). \(\diamondsuit \text { Claim } 2\)

Define \(\mathcal{R }:=\mathcal{A }/\mathcal{C }_0\); then \(\mathcal{R }\) is an abelian group. Let \(\phi :\mathcal{A }{{\longrightarrow }}\mathcal{R }\) be the quotient map. Then define \(\mathcal{R }_+:= \phi (\mathcal{C }_+)\) and \(\mathcal{R }_-:= \phi (\mathcal{C }_-)\).

Claim 3 (a) For all nonzero \(r\in \mathcal{R }\), either \(r\in \mathcal{R }_+\) or \(-r\in \mathcal{R }_+\), but not both.

(b) For all \(r,s\in \mathcal{R }_+\), we have \((r+s)\in \mathcal{R }_+\).

Proof

Mapping Claim 2(a) through \(\phi \) immediately yields (b).

  • To check (a), note that \(\mathcal{A }= \mathcal{C }_-\sqcup \mathcal{C }_0 \sqcup \mathcal{C }_+\); thus, \(\mathcal{R }= \phi (\mathcal{C }_-)\cup \phi (\mathcal{C }_0) \cup \phi (\mathcal{C }_+) = \mathcal{R }_- \cup \{0\} \cup \mathcal{R }_+\). Thus, any nonzero element of \(\mathcal{R }\) is either in \(\mathcal{R }_+\) or \(\mathcal{R }_-\).

  • It remains only to show that \(\mathcal{R }_+\) and \(\mathcal{R }_-\) are disjoint. By contradiction, suppose that \(r\in \mathcal{R }_+\cap \mathcal{R }_-\). Find \(b\in \mathcal{C }_+\) and \(c\in \mathcal{C }_-\) such that \(\phi (b)=\phi (c)=r\). Claim 2(c) implies that \(-c\in \mathcal{C }_+\). Thus, Claim 2(a) yields \(b-c\in \mathcal{C }_+\). But \(\phi (b-c)=\phi (b)-\phi (c)=r-r=0\), so \(b-c\in \mathcal{C }_0\). Contradiction. \(\diamondsuit \text { Claim } 3\)

Now define a binary relation \((>)\) on \(\mathcal{R }\) by setting \(r> s\) if and only if \(r-s\in \mathcal{R }_+\). Claim 3(a) implies that \((>)\) is complete and antisymmetric. Claim 3(b) implies that \((>)\) is transitive. Thus \((>)\) is a total order relation.

Finally, define \(u:\mathcal{X }{{\longrightarrow }}\mathcal{R }\) by \(u(x) := \phi (1\cdot x)\) (where \(1\cdot x\) denotes an element of \(\mathcal{A }\)). It remains to show that \((\succcurlyeq )\) satisfies statement (2). Let \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\), with \(\hbox {d}(\mathbf{x},\mathbf{y})<{\infty }\). Let \(\mathcal{K }:=\mathcal{I }(\mathbf{x},\mathbf{y})\) and \(\mathcal{J }:=\mathcal{I }\setminus \mathcal{K }\), and define \(\mathbf{x}',\mathbf{y}^{\prime }\in \mathcal{X }^\mathcal{I }\) by setting \(\mathbf{x}'_\mathcal{K }:=\mathbf{x}_\mathcal{K }\) and \(\mathbf{y}'_\mathcal{K }:=\mathbf{y}_\mathcal{K }\), while \(x'_j=y'_j=o\) for all \(j\in \mathcal{J }\).

Let \(x_1,\ldots ,x_N\) be the distinct elements of \(\mathcal{X }\setminus \{o\}\) which occur in \(\mathbf{x}'\), and for each \(n\in {\left[ 1\ldots N \right]}\), let \(J_n\in \mathbb{N }\) be the number of times we see \(x_n\). If \(a:=J_1 x_1+\cdots J_N x_N\in \mathcal{B }\), then there exists some \(\pi \in \Pi _{\scriptscriptstyle {\mathrm{fin}}}\) such that \(\pi (\mathbf{x}')=\mathbf{w}^a\); thus \(\mathbf{x}^{\prime }\approx \mathbf{w}^a\) by \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariance.

Likewise, let \(y_1,\ldots ,y_M\) be the distinct elements of \(\mathcal{X }\setminus \{o\}\) which occur in \(\mathbf{y}'\), and for each \(m\in {\left[ 1\ldots M \right]}\), let \(K_m\in \mathbb{N }\) be the number of times we see \(y_m\). If \(b:=K_1 y_1+\cdots K_M y_M\in \mathcal{B }\), then there exists some \(\tau \in \Pi _{\scriptscriptstyle {\mathrm{fin}}}\) such that \(\tau (\mathbf{y}')=\mathbf{w}^b\); thus \(\mathbf{y}^{\prime }\approx \mathbf{w}^b\) by \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariance. Now we have:

$$\begin{aligned} \left( \mathbf{x}\, {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle }}}\, \mathbf{y} \right)&\mathop {\Leftarrow \!\!\Rightarrow }\limits _{(*)}&\left( \mathbf{x}^{\prime }\, {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle }}}\,\mathbf{y}' \right) \mathop {\Leftarrow \!\!\Rightarrow }\limits _{(\dagger )} \left( \mathbf{w}^a\, {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle }}}\,\mathbf{w}^b \right) \mathop {\Leftarrow \!\!\Rightarrow }\limits _{(\diamond )} \left( a-b\in \mathcal{C }_{0+} \right) \nonumber \\&\mathop {\Leftarrow \!\!\Rightarrow }\limits _{(@)}&\left( \phi (a-b)\in \mathcal{R }_+\sqcup \{0\} \right) \mathop {\Leftarrow \!\!\Rightarrow }\limits _{(\ddagger )} \ \left( \phi (a-b)\ge 0 \right). \end{aligned}$$
(17)

\((*)\) is by separability, because \(\mathbf{x},\mathbf{y},\mathbf{x}^{\prime },\mathbf{y}^{\prime }\) satisfy the separability conditions (3) by construction. Next, \((\dagger )\) is because \(\mathbf{x}^{\prime }\approx \mathbf{w}^a\) and \(\mathbf{y}^{\prime }\approx \mathbf{w}^b\). Finally, \((\diamond )\) is by Claim 1(b) and the definition of \(\mathcal{C }_{0+}\), (@) is by definition of \(\mathcal{R }_+\), and \((\ddagger )\) is by definition of \((>)\). Now,

$$\begin{aligned} \phi (a-b)&= \phi (J_1 x_1+\cdots J_N x_N) - \phi (K_1 y_1+\cdots K_M y_M)\nonumber \\&\mathop {=}\limits _{(*)} \sum _{n=1}^N J_n\,u(x_n) -\sum _{m=1}^M K_m\,u(y_m)\nonumber \\&= \sum _{i\in \mathcal{I }(\mathbf{x},\mathbf{y})} u(x_i) - \sum _{i\in \mathcal{I }(\mathbf{x},\mathbf{y})} u(y_i) = \sum _{i\in \mathcal{I }} \left( u(x_i)-u(y_i)\right). \end{aligned}$$
(18)

Here \((*)\) is because \(\phi (1\cdot x_n)=u(x_n)\) and \(\phi (1\cdot y_m)=u(y_m)\). Combining statements (17) and (18) yields (2). Thus, \((\succcurlyeq )\) is the additive preorder on \(\mathcal{X }^\mathcal{I }\) defined by \(u\).

Universal property. Let \((\mathcal{R }',+,>)\) be another linearly ordered abelian group, and let \(u':\mathcal{X }{{\longrightarrow }}\mathcal{R }'\) be some function such that \((\succcurlyeq )\) is also the additive preorder defined by \(u'\). Let \(r':=u'(o)\), and define \(u'':\mathcal{X }{{\longrightarrow }}\mathcal{R }'\) by \(u''(x):=u'(x)-r'\). Thus \(u''(o)=0\), and \((\succcurlyeq )\) is also the additive preorder defined by \(u''\).

Define \(\gamma :\mathcal{A }{{\longrightarrow }}\mathcal{R }'\) by setting \(\gamma (\sum _{w\in \mathcal{W }} A_w\, w) = \sum _{w\in \mathcal{W }} A_w\,u''(w)\) for any finite \(\mathcal{W }\subseteq \mathcal{X }\setminus \{o\}\) and coefficients \(\{A_w\}_{w\in \mathcal{W }}\subseteq \mathbb{Z }\). Equivalently, \(\gamma (a):=\sum _\mathcal{I }\, \left( u''(\mathbf{w}^{a+})- u''(\mathbf{w}^{a-}) \right)\) for all \(a\in \mathcal{A }\). This is automatically a group homomorphism, because \(\mathcal{A }\) is the free abelian group generated by \(\mathcal{X }\setminus \{o\}\).

Claim 4 \(\mathcal{C }_0\subseteq \ker (\gamma )\).

Proof

If \(a\in \mathcal{C }_0\), then \(\mathbf{w}^{a+}\approx \mathbf{w}^{a-}\). Thus,

$$\begin{aligned} \gamma (a) := \sum _\mathcal{I }\, \left( u''(\mathbf{w}^{a+})- u''(\mathbf{w}^{a-}) \right) \mathop {=}\limits _{(*)} ~ 0. \end{aligned}$$

Thus, \(a\in \ker (\gamma )\), as desired. Here \((*)\) is by Eq. (2), because \(\mathbf{w}^{a+}\approx \mathbf{w}^{a-}\), and \((\succcurlyeq )\) is the additive preorder defined by \(u''\). \(\diamondsuit \text { Claim } 4\)

Claim 4 means that \(\gamma \) factors through \(\phi \) to yield a homomorphism \(\psi :\mathcal{R }{{\longrightarrow }}\mathcal{R }'\) such that \(\psi \circ \phi =\gamma \). (See Fig. 2.)

Claim 5 \(u'' = \psi \circ u\).

Fig. 2
figure 2

A commuting diagram illustrating the proof of the universal property in Theorem 1. Here \(i\) represents an inclusion map (regard \(\mathcal{X }\) as a subset of \(\mathcal{A }\) in the obvious way) and \(0\) represents a zero homomorphism

Proof

Let \(x\in \mathcal{X }\). Let \(a = 1\, x\) (an element of \(\mathcal{A }\)). Then \(u''(x) = \gamma (a) = \psi \circ \phi (a) = \psi (u(x))\). \(\diamondsuit \text { Claim } 5\)

It remains only to show that \(\psi \) is order-preserving. Let \(r\in \mathcal{R }\); then \(r=\phi (a)\) for some \(a\in \mathcal{A }\).

$$\begin{aligned} \left( r\!\ge \! 0 \right) \!\iff \!\left( a\!\in \!\mathcal{C }_{0+} \right)&\iff \left( \mathbf{w}^{a+}\succcurlyeq \mathbf{w}^{a-} \right) \mathop {\Leftarrow \!\!\Rightarrow }\limits _{(*)} \left( \displaystyle \sum _\mathcal{I }\, \left( u''(\mathbf{w}^{a+})\!-\! u''(\mathbf{w}^{a-})\right)\!\ge \! 0 \right)\\&\mathop {\Leftarrow \!\!\Rightarrow }\limits _{(\dagger )} \left( \gamma (a)\ge 0 \right) \mathop {\Leftarrow \!\!\Rightarrow }\limits _{(\diamond )} \left( \psi (r)\ge 0 \right), \end{aligned}$$

as desired. Here, \((*)\) is by Eq. (2), because \((\succcurlyeq )\) is the additive preorder defined by \(u''\). Next, \((\dagger )\) is by definition of \(\gamma \) and \((\diamond )\) is because \(\psi \circ \phi =\gamma \) and \(\phi (a)=r\). \(\square \)

Remark

The proof of the ‘existence’ part of Theorem 1 is somewhat analogous to the proof strategy used by Wakker (1986). Let \(\mathcal{Y }:=\mathcal{X }\setminus \{o\}\). Wakker begins with a separable, anonymous preference order \((\succcurlyeq )\) defined on all sequences in \(\mathcal{X }^\mathbb{N }\) which have only finitely many nonzero entries. He introduces a set \([\mathcal{Y }]\), which is essentially the free abelian monoid generated by \(\mathcal{Y }\) (denoted by \(\mathcal{B }'\) in my proof). The order \((\succcurlyeq )\) induces a complete preorder on \([\mathcal{Y }]\), which is compatible with the addition operation. Wakker also assumes \((\succcurlyeq )\) is Archimedean; he then invokes Theorem 3.2.1 of Krantz et al. (1971) to obtain a real-valued, additive utility representation for \((\succcurlyeq )\), equivalent to my Proposition 5(a).

Proof of Proposition 5(a)

Let \((\mathcal{R },+,>)\) be a linearly ordered abelian group, let \(u:\mathcal{X }{{\longrightarrow }}\mathcal{R }\), and let \(\big ({ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\) be the additive preorder defined by (2). Assume without loss of generality that the image set \(u(\mathcal{X })\) generates \(\mathcal{R }\) (otherwise replace \(\mathcal{R }\) with the subgroup generated by \(u(\mathcal{X })\).)

Claim 1 \(\big ({ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\) is Archimedean if and only if \(\mathcal{R }\) is Archimedean.

Proof

Let \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\), with \(\hbox {d}(\mathbf{x},\mathbf{o})<{\infty }\) and \(\hbox {d}(\mathbf{y},\mathbf{o})<{\infty }\). Suppose that

$$\begin{aligned} r = \sum _{i\in \mathcal{I }} \left( u(x_i) - u(o)\right)\,\text { and }\, s = \sum _{i\in \mathcal{I }} \left( u(y_i) - u(o)\right). \end{aligned}$$
(19)

For any \(N\in \mathbb{N }\), define \(\mathbf{x}^N\in \mathcal{X }^\mathcal{I }\) as in Sect. 3. Then

$$\begin{aligned} \sum _{i\in \mathcal{I }} \left( u(x_i^N) - u(y_i)\right) \!=\! N\cdot \sum _{i\in \mathcal{I }} \left( u(x_i) \!-\! u(o)\right) \!-\! \sum _{i\in \mathcal{I }} \left( u(y_i) \!-\! u(o)\right) \!=\! N\cdot r \!-\! s. \end{aligned}$$

Thus, \(\mathbf{x}^N\succ \mathbf{y}\) if and only if \(N\cdot r -s > 0\).

\({\Longleftarrow }\)’ If \(\mathbf{x}\succ \mathbf{o}\), then \(r>0\). Since \(\mathcal{R }\) is Archimedean, there exists \(N\in \mathbb{N }\) such that \(N\cdot r > s\), and thus, \(\mathbf{x}^N\succ \mathbf{y}\).

\(\Longrightarrow \)’ For any \(r,s\in \mathcal{R }\), we can construct some \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\) satisfying (19) (because \(u(\mathcal{X })\) generates \(\mathcal{R }\) by hypothesis). Since \((\succcurlyeq )\) is Archimedean, there exists \(N\in \mathbb{N }\) such that \(\mathbf{x}^N\succ \mathbf{y}\) and thus, \(N\cdot r > s\). \(\diamondsuit \text { Claim } 1\)

Now combining Theorem 1, Hölder’s Theorem, and Claim 1 yields the result. \(\square \)

Part (b) of Proposition 5 follows immediately from part (a) and Theorem 2, which will be proved below. The proofs of Theorem 2 and the results from Sect. 4 all depend on the results from Sect. 5, so I will prove those results first.

Proofs from Sect. 5, and the proof of Theorem 2

I will now prove the results of Sect. 5, culminating in the proof of Theorem 2. First, let us define some convenient notation. For any \(\mathbf{x}\in \mathcal{X }^\mathcal{I }\) and any function \(u:\mathcal{X }{{\longrightarrow }}\mathcal{R }\), let us define \({}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{x}):= \displaystyle {}^*\!\sum _{i\in \mathcal{I }} u(x_i)\). Also, for any \(\mathcal{F }\in {\varvec{\mathcal{F }}}\), define \(\sum _{\mathcal{F }}\,u(\mathbf{x}):=\displaystyle \sum _{f\in \mathcal{F }} u(x_f)\).

Proof of Lemma 13

Let \(0^{\varvec{\mathcal{F }}}\) be the constant \(0\) element of \(\mathcal{R }^{\varvec{\mathcal{F }}}\), and define \(\mathcal{Z }:=\{r\in \mathcal{R }^{\varvec{\mathcal{F }}}\); \(r {\stackrel{\displaystyle \approx }{{\scriptscriptstyle \mathfrak{UF }}}} 0^{\varvec{\mathcal{F }}}\}\);

Claim 1 \(\mathcal{Z }\) is a subgroup in \(\mathcal{R }^{\varvec{\mathcal{F }}}\).

Proof

For any \(r\in \mathcal{R }^{\varvec{\mathcal{F }}}\), let \({\varvec{\mathcal{O }}}(r):=\{\mathcal{F }\in {\varvec{\mathcal{F }}}\); \(r(\mathcal{F })=0\}\). Then \(r\in \mathcal{Z }\) if and only if \({\varvec{\mathcal{O }}}(r)\in \mathfrak{UF }\). But, for any \(r,s\in \mathcal{R }^{\varvec{\mathcal{F }}}\), we have \({\varvec{\mathcal{O }}}(r- s) \supseteq {\varvec{\mathcal{O }}}(r)\cap {\varvec{\mathcal{O }}}(s)\); thus, if \(r\in \mathcal{Z }\) and \(s\in \mathcal{Z }\), then axioms (F1) and (F2) imply \(r- s\in \mathcal{Z }\). \(\diamondsuit \text { Claim } 1\)

For any \(r,s\in \mathcal{R }^{\varvec{\mathcal{F }}}\), it is easy to check that \(r \, {\stackrel{\displaystyle \approx }{{\scriptscriptstyle \mathfrak{UF }}}}\, s\) if and only if \((r-s)\in \mathcal{Z }\). Thus, \({}^*\!\mathcal{R }\) is just the quotient group \(\mathcal{R }^{\varvec{\mathcal{F }}}/\mathcal{Z }\). The relation \((>)\) defines a linear order on \({}^*\!\mathcal{R }\). \(\square \)

Proof of Lemma 14

Definition (4) says that the function \(\mathcal{X }^\mathcal{I }\ni \mathbf{x}\mapsto {}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{x})\in {}^*\!\mathcal{R }\) is a utility function for the preorder \(\big ({}^{^*}\!{ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\). The value of \({}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{x})\) is well-defined for all \(\mathbf{x}\in \mathcal{X }^\mathcal{I }\), and \({}^*\!\mathcal{R }\) is totally ordered, so \(\big ({}^{^*}\!{ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\) is a complete preorder on \(\mathcal{X }^\mathcal{I }\).

Separable. Let \(\mathcal{J }\subset \mathcal{I }\) and let \(\mathcal{K }:=\mathcal{I }\setminus \mathcal{J }\). Suppose that \(\mathbf{x},\mathbf{y},\mathbf{x}',\mathbf{y}^{\prime }\in \mathcal{X }^\mathcal{I }\) satisfy the separability conditions (3). I must show that \((\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\mathbf{y})\Leftrightarrow (\mathbf{x}' \,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\mathbf{y}')\).

Define \({\varvec{\mathcal{F }}}(\mathbf{x},\mathbf{y}):=\{\mathcal{F }\in {\varvec{\mathcal{F }}}\); \(\sum _{\mathcal{F }}\,u(\mathbf{x})\ge \sum _{\mathcal{F }}\,u(\mathbf{y})\}\). Then \(({}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{x}) \ge {}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{y}))\Leftrightarrow ({\varvec{\mathcal{F }}}(\mathbf{x},\mathbf{y})\in \mathfrak{UF })\) and \(({}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{x}') \ge {}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{y}'))\Leftrightarrow ({\varvec{\mathcal{F }}}(\mathbf{x}',\mathbf{y}')\in \mathfrak{UF })\). Thus, it suffices to show that \(({\varvec{\mathcal{F }}}(\mathbf{x},\mathbf{y})\in \mathfrak{UF })\Leftrightarrow ({\varvec{\mathcal{F }}}(\mathbf{x}',\mathbf{y}')\in \mathfrak{UF })\). In fact, I will show that \({\varvec{\mathcal{F }}}(\mathbf{x}',\mathbf{y}')= {\varvec{\mathcal{F }}}(\mathbf{x},\mathbf{y})\).

For any \(\mathcal{F }\in {\varvec{\mathcal{F }}}\), the Eqs. (3) imply:

$$\begin{aligned} \begin{array}{rclcrcl} \mathrm{(a)} \quad \sum \nolimits _{\mathcal{F }\cap \mathcal{J }}\,u(\mathbf{x})&{}=&{}\sum \nolimits _{\mathcal{F }\cap \mathcal{J }}\,u(\mathbf{y}), &{}\quad &{} \mathrm{(b)} \quad \sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{x})&{}=&{}\sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{x}'),\\ \mathrm{(c)} \quad \sum \nolimits _{\mathcal{F }\cap \mathcal{J }}\,u(\mathbf{x}')&{}=&{}\sum \nolimits _{\mathcal{F }\cap \mathcal{J }}\,u(\mathbf{y}'), &{}\text { and }&{} \mathrm{(d)} \quad \sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{y})&{}=&{}\sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{y}'). \end{array} \end{aligned}$$

Furthermore, \(\mathcal{F }=(\mathcal{F }\cap \mathcal{J })\sqcup (\mathcal{F }\cap \mathcal{K })\) (because \(\mathcal{I }:=\mathcal{J }\sqcup \mathcal{K }\)); thus

$$\begin{aligned} \begin{array}{rclrcl} \mathrm{(e)}\quad \sum \nolimits _{\mathcal{F }}\,u(\mathbf{x})&{}=&{}\sum \nolimits _{\mathcal{F }\cap \mathcal{J }}\,u(\mathbf{x})+\sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{x}), &{} \sum \nolimits _{\mathcal{F }}\,u(\mathbf{y})&{}=&{}\sum \nolimits _{\mathcal{F }\cap \mathcal{J }}\,u(\mathbf{y})\\ +\sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{y});\\ \mathrm{(f)}\quad \sum \nolimits _{\mathcal{F }}\,u(\mathbf{x}')&{}=&{}\sum \nolimits _{\mathcal{F }\cap \mathcal{J }}\,u(\mathbf{x}')+\sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{x}'), &{} \sum \nolimits _{\mathcal{F }}\,u(\mathbf{y}')&{}=&{}\sum \nolimits _{\mathcal{F }\cap \mathcal{J }}\,u(\mathbf{y}')\\ +\sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{y}'). \end{array} \end{aligned}$$

Thus, for all \(\mathcal{F }\in {\varvec{\mathcal{F }}}\), we have:

$$\begin{aligned} \left( \mathcal{F }\in {\varvec{\mathcal{F }}}(\mathbf{x},\mathbf{y}) \right)&\iff \left( \sum \nolimits _{\mathcal{F }}\,u(\mathbf{x})\ge \sum \nolimits _{\mathcal{F }}\,u(\mathbf{y}) \right) \mathop {\Leftarrow \!\!\Rightarrow }\limits _{(*)} \left( \sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{x})\ge \sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{y}) \right)\\&\mathop {\Leftarrow \!\!\Rightarrow }\limits _{(\dagger )} \left( \sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{x}')\!\ge \! \sum \nolimits _{\mathcal{F }\cap \mathcal{K }}\,u(\mathbf{y}') \right) \mathop {\Leftarrow \!\!\Rightarrow }\limits _{(\diamond )} \left( \sum \nolimits _{\mathcal{F }}\,u(\mathbf{x}')\!\ge \! \sum \nolimits _{\mathcal{F }}\,u(\mathbf{y}') \right)\\&\iff \left( \mathcal{F }\in {\varvec{\mathcal{F }}}(\mathbf{x}',\mathbf{y}') \right). \end{aligned}$$

Thus, \({\varvec{\mathcal{F }}}(\mathbf{x}',\mathbf{y}')= {\varvec{\mathcal{F }}}(\mathbf{x},\mathbf{y})\). Here, \((*)\) is by Eqs. (a) and (e); \((\dagger )\) is by Eqs. (b) and (d); and \((\diamond )\) is by Eqs. (c) and (f). \(\square \)

Proof of Lemma 15

  1. (a)

    If \(\epsilon \in \Pi \) is the identity permutation, then clearly \({\varvec{\mathcal{F }}}(\epsilon )={\varvec{\mathcal{F }}}\in \mathfrak{UF }\). Also, for any \(\pi \in \Pi \), it is clear that \({\varvec{\mathcal{F }}}(\pi ^{-1})={\varvec{\mathcal{F }}}(\pi )\), so \(\pi \in \Pi _\mathfrak{UF }\) if and only if \(\pi ^{-1}\in \Pi _\mathfrak{UF }\). Finally, let \(\pi _1,\pi _2\in \Pi _\mathfrak{UF }\). Then \({\varvec{\mathcal{F }}}(\pi _1\circ \pi _2) \supseteq {\varvec{\mathcal{F }}}(\pi _1)\cap {\varvec{\mathcal{F }}}(\pi _2)\). But \({\varvec{\mathcal{F }}}(\pi _1)\cap {\varvec{\mathcal{F }}}(\pi _2)\in \mathfrak{UF }\) by Axiom (F1); thus, \({\varvec{\mathcal{F }}}(\pi _1\circ \pi _2)\in \mathfrak{UF }\) by (F2); thus, \(\pi _1\circ \pi _2\in \Pi _\mathfrak{UF }\).

  2. (b)

    Let \(\mathbf{x}\in \mathcal{X }^\mathcal{I }\) and \(\pi \in \Pi _\mathfrak{UF }\). Then for every \(\mathcal{F }\in {\varvec{\mathcal{F }}}\), we have \(\sum _{\mathcal{F }}\,u(\mathbf{x}) = \sum _{\pi (\mathcal{F })}\,u[\pi (\mathbf{x})]\). But if \(\mathcal{F }\in {\varvec{\mathcal{F }}}(\pi )\), then \(\pi (\mathcal{F })=\mathcal{F }\), so we get \(\sum _{\mathcal{F }}\,u(\mathbf{x})=\sum _{\mathcal{F }}\,u[\pi (\mathbf{x})]\). If \({\varvec{\mathcal{F }}}(\pi )\in \mathfrak{UF }\), then this implies that \({}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{x})={}^* \sum _{_{\mathcal{I }}}\,u[\pi (\mathbf{x})]\); thus, \(\mathbf{x}\,{}^{^*}\!\!\! {\stackrel{\displaystyle \approx }{{\scriptscriptstyle u}}}\pi (\mathbf{x})\). \(\square \)

Proof of Lemma 17

  1. (a)

    Recall that \(\mathfrak{F }_\Gamma :=\{{\varvec{\mathcal{E }}}\subseteq {\varvec{\mathcal{F }}}\); \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\subseteq {\varvec{\mathcal{E }}}\) for some finite \(\mathcal{J }\subset \mathcal{I }\) and finite \(\Delta \subseteq \Gamma \}\).

Claim 1 \(\mathfrak{F }\) is a free filter.

Proof

We must check axioms (F0)–(F2).

  1. (F0)

    For any finite subset \(\Delta \subseteq \Gamma \), let \({\varvec{\mathcal{O }}}_\Delta \) be the orbit partition generated by \({\left\langle \Delta \right\rangle }\). Then \({\varvec{\mathcal{O }}}_\Delta \) has an infinite number of elements, because \(\mathcal{I }\) is infinite, whereas each element of \({\varvec{\mathcal{O }}}_\Delta \) is a finite subset (because \(\Gamma \) has locally finite orbits). For any finite subset \(\mathcal{J }\subseteq \mathcal{I }\) let \({\varvec{\mathcal{O }}}_\Delta (\mathcal{J }):=\{\mathcal{O }\); \(\mathcal{O }\in {\varvec{\mathcal{O }}}_\Delta \) and \(\mathcal{J }\cap \mathcal{O }\ne \emptyset \}\); then \({\varvec{\mathcal{O }}}_\Delta (\mathcal{J })\) is finite, and \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J }):=\{\bigsqcup _{\mathcal{P }\in {\varvec{\mathcal{P }}}} \mathcal{P }\); \({\varvec{\mathcal{P }}}\subseteq {\varvec{\mathcal{O }}}_\Delta \) any finite subset such that \({\varvec{\mathcal{O }}}_\Delta (\mathcal{J })\subseteq {\varvec{\mathcal{P }}}\}\). Thus \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\) is infinite, because \({\varvec{\mathcal{O }}}_\Delta \setminus {\varvec{\mathcal{O }}}_\Delta (\mathcal{J })\) is infinite. Thus, if \({\varvec{\mathcal{E }}}\subseteq {\varvec{\mathcal{F }}}\) is finite, then we cannot have \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\subseteq {\varvec{\mathcal{E }}}\) for any finite subsets \(\mathcal{J }\subseteq \mathcal{I }\) and \(\Delta \subseteq \Gamma \); thus, \({\varvec{\mathcal{E }}}\not \in \mathfrak{F }\).

  2. (F1)

    Let \({\varvec{\mathcal{D }}},{\varvec{\mathcal{E }}}\in \mathfrak{F }\). Then there exist finite subsets \(\mathcal{J },\mathcal{K }\subset \mathcal{I }\) and \(\Delta ,\mathsf{ E }\subseteq \Gamma \) such that \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\subseteq {\varvec{\mathcal{D }}}\) and \({\varvec{\mathcal{F }}}_\mathsf{ E }(\mathcal{K })\subseteq {\varvec{\mathcal{E }}}\). Thus, \(\mathcal{J }\cup \mathcal{K }\) and \(\Delta \cup \mathsf{ E }\) are also finite, and we have \({\varvec{\mathcal{F }}}_{\Delta \cup \mathsf{ E }}(\mathcal{J }\cup \mathcal{K })={\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\cap {\varvec{\mathcal{F }}}_\mathsf{ E }(\mathcal{K }) \subseteq {\varvec{\mathcal{D }}}\cap {\varvec{\mathcal{E }}}\). Thus, \({\varvec{\mathcal{D }}}\cap {\varvec{\mathcal{E }}}\in \mathfrak{F }\) also.

  3. (F2)

    Let \({\varvec{\mathcal{D }}}\in \mathfrak{F }\); then \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\subseteq {\varvec{\mathcal{D }}}\) for some finite subsets \(\mathcal{J }\subset \mathcal{I }\) and \(\Delta \subseteq \Gamma \). Thus, for any \({\varvec{\mathcal{E }}}\subseteq {\varvec{\mathcal{F }}}\), if \({\varvec{\mathcal{D }}}\subseteq {\varvec{\mathcal{E }}}\), then \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\subseteq {\varvec{\mathcal{E }}}\) also, so \({\varvec{\mathcal{E }}}\in \mathfrak{F }\). \(\diamondsuit \text { Claim } 1\)

Now apply the Ultrafilter Lemma to obtain some ultrafilter \(\mathfrak{UF }\) which contains \(\mathfrak{F }_\Gamma \).

  • (b) Let \(\pi \in \Pi _{\scriptscriptstyle {\mathrm{fin}}}\); then \(\mathcal{I }(\pi )\) is finite. Thus, for any finite subset \(\Delta \subseteq \Gamma \), we have \({\varvec{\mathcal{F }}}_\Delta [\mathcal{I }(\pi )]\in \mathfrak{F }_\Gamma \subseteq \mathfrak{UF }\). But \({\varvec{\mathcal{F }}}_\Delta [\mathcal{I }(\pi )]\subseteq {\varvec{\mathcal{F }}}(\pi )\); thus, axiom (F2) implies \({\varvec{\mathcal{F }}}(\pi )\in \mathfrak{UF }\), so \(\pi \in \Pi _\mathfrak{UF }\).

Now let \(\gamma \in \Gamma \), and fix a finite \(\mathcal{J }\subseteq \mathcal{I }\). If \(\mathcal{F }\in {\varvec{\mathcal{F }}}_{\{\gamma \}}(\mathcal{J })\), then \(\gamma (\mathcal{F })=\mathcal{F }\), so \(\mathcal{F }\in {\varvec{\mathcal{F }}}(\gamma )\). Thus, \({\varvec{\mathcal{F }}}_{\{\gamma \}}(\mathcal{J }) \subseteq {\varvec{\mathcal{F }}}(\gamma )\), so \({\varvec{\mathcal{F }}}(\gamma )\in \mathfrak{F }_\Gamma \), and thus \({\varvec{\mathcal{F }}}(\gamma )\in \mathfrak{UF }\). Thus, \(\gamma \in \Pi _\mathfrak{UF }\). \(\square \)

Proof of Lemma 18

If \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\) and \(\hbox {d}(\mathbf{x},\mathbf{y})<{\infty }\), then the set \(\mathcal{I }(\mathbf{x},\mathbf{y})\) is finite. Recall that \({\varvec{\mathcal{F }}}[\mathcal{I }(\mathbf{x},\mathbf{y})]:=\{\mathcal{J }\in {\varvec{\mathcal{F }}}\); \(\mathcal{I }(\mathbf{x},\mathbf{y})\subseteq \mathcal{J }\}\). For any \(\mathcal{J }\in {\varvec{\mathcal{F }}}[\mathcal{I }(\mathbf{x},\mathbf{y})]\), we have:

$$\begin{aligned} \sum _{\mathcal{J }}\,u(\mathbf{x}) - \sum _{\mathcal{J }}\,u(\mathbf{y}) = \sum _{j\in \mathcal{J }} u(x_j)- \sum _{j\in \mathcal{J }} u(y_j) = \sum _{i\in \mathcal{I }(\mathbf{x},\mathbf{y})} u(x_i) - \sum _{i\in \mathcal{I }(\mathbf{x},\mathbf{y})} u(y_i), \end{aligned}$$

because \(x_j=y_j\) for all \(j\in \mathcal{J }\setminus \mathcal{I }(\mathbf{x},\mathbf{y})\).

If \(\mathbf{x}\, {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}\), then \(\displaystyle \sum _{i\in \mathcal{I }(\mathbf{x},\mathbf{y})} u(x_i)\ge \sum _{i\in \mathcal{I }(\mathbf{x},\mathbf{y})} u(y_i)\); thus, \(\sum _{\mathcal{J }}\,u(\mathbf{x})\ge \sum _{\mathcal{J }}\,u(\mathbf{y})\) for all \(\mathcal{J }\in {\varvec{\mathcal{F }}}_\Delta [\mathcal{I }(\mathbf{x},\mathbf{y})]\). But \({\varvec{\mathcal{F }}}_\Delta [\mathcal{I }(\mathbf{x},\mathbf{y})]\in \mathfrak{UF }\) because \(\mathfrak{UF }\) is a suitable ultrafilter. Thus, \({}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{x})\ge {}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{y})\). Thus \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}\).

Likewise, if \(\mathbf{x}\, {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u}}}\,\mathbf{y}\), then \(\displaystyle \sum _{i\in \mathcal{I }(\mathbf{x},\mathbf{y})} u(x_i) > \sum _{i\in \mathcal{I }(\mathbf{x},\mathbf{y})} u(y_i)\); thus, \(\sum _{\mathcal{J }}\,u(\mathbf{x})> \sum _{\mathcal{J }}\,u(\mathbf{y})\) for all \(\mathcal{J }\in {\varvec{\mathcal{F }}}[\mathcal{I }(\mathbf{x},\mathbf{y})]\in \mathfrak{UF }\); thus, \({}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{x}) > {}^* \sum _{_{\mathcal{I }}}\,u(\mathbf{y})\). Thus \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u}}}\,\mathbf{y}\). \(\square \)

Proof of Lemma 19

Claim 1 Let \({\varvec{\mathcal{G }}}\in \mathfrak{UF }\).

  1. (a)

    \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u,{\varvec{\mathcal{G }}}}}}\, \mathbf{y}\) if and only if \(\sum _{\mathcal{J }}\,u(\mathbf{x})\ge \sum _{\mathcal{J }}\,u(\mathbf{y})\) for all \(\mathcal{J }\in {\varvec{\mathcal{G }}}\).

  2. (b)

    \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u,{\varvec{\mathcal{G }}}}}}\, \mathbf{y}\) if and only if \(\sum _{\mathcal{J }}\,u(\mathbf{x})> \sum _{\mathcal{J }}\,u(\mathbf{y})\) for all \(\mathcal{J }\in {\varvec{\mathcal{G }}}\).

Proof

Fix \(\mathcal{J }\in {\varvec{\mathcal{G }}}\) and \(\mathbf{z}\in \mathcal{X }^\mathcal{I }\), and let \(\mathbf{x}':=\mathbf{x}_{\mathcal{J }}\mathbf{z}_{{\mathcal{I }\setminus \mathcal{J }}}\) and \(\mathbf{y}':=\mathbf{y}_{\mathcal{J }}\mathbf{z}_{{\mathcal{I }\setminus \mathcal{J }}}\). Then \(\mathcal{I }(\mathbf{x}',\mathbf{y}')\subseteq \mathcal{J }\), so \(\hbox {d}(\mathbf{x}',\mathbf{y}')<{\infty }\) (because \(\mathcal{J }\) is finite). Thus, the \(\big ({ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\)-order of \(\mathbf{x}'\) and \(\mathbf{y}'\) is well-defined, and Lemma 18 says that

$$\begin{aligned}&\qquad \big ( \mathbf{x}^{\prime }\, {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}' \big )\ \iff \ \big ( \mathbf{x}^{\prime }\,\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}' \big ) \,\,\text { and }\left( \mathbf{x}^{\prime }\, {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u}}}\,\mathbf{y}' \right) \ \iff \ \left( \mathbf{x}^{\prime }\,\,{}^{^*}\! {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u}}}\,\mathbf{y}' \right).\end{aligned}$$
(20)
$$\begin{aligned}&\text {Furthermore,}\quad \sum _{i\in \mathcal{I }}\left( u(x'_i)-u(y'_i)\right) = \sum _{\mathcal{J }} u(\mathbf{x})- \sum _{\mathcal{J }} u(\mathbf{y}). \end{aligned}$$
(21)
  1. (a)

    \(\Longrightarrow \)’ Suppose that \(\mathbf{x}\,{}^{^*}\!\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u,{\varvec{\mathcal{G }}}}}}\, \mathbf{y}\). Then for any \(\mathcal{J }\in \mathcal{G }\) and \(\mathbf{z}\in \mathcal{X }^\mathcal{I }\), if \(\mathbf{x}'\) and \(\mathbf{y}'\) are defined as above, then \(\mathbf{x}^{\prime }\,\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}'\). Thus, statement (20) says that \(\mathbf{x}^{\prime }\, {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}'\), so \(\sum _{i\in \mathcal{I }}\left( u(x'_i)-u(y'_i)\right)\ge 0\). But then statement (21) implies that \(\sum _{\mathcal{J }}\,u(\mathbf{x})\ge \sum _{\mathcal{J }}\,u(\mathbf{y})\). ‘\({\Longleftarrow }\)’ Fix \(\mathbf{z}\in \mathcal{X }^\mathcal{I }\) and \(\mathcal{J }\in {\varvec{\mathcal{G }}}\). If \(\mathbf{x}'\) and \(\mathbf{y}'\) are defined as above, then statement (21) implies that \(\displaystyle \sum _{i\in \mathcal{I }}\left( u(x'_i)-u(y'_i)\right)\ge 0\), so \(\mathbf{x}^{\prime }\, {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}'\), so statement (20) says \(\mathbf{x}^{\prime }\,\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}'\). This holds for all \(\mathcal{J }\in {\varvec{\mathcal{G }}}\) and \(\mathbf{z}\in \mathcal{X }^\mathcal{I }\); thus, \(\mathbf{x}\,{}^{^*}\!\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u,{\varvec{\mathcal{G }}}}}}\, \mathbf{y}\).

  2. (b)

    The proof is similar to (a); change ‘\(\ge \)’ to ‘\(>\)’ and ‘\(\succcurlyeq \)’ to ‘\(\succ \)’ everywhere. \(\diamondsuit \) Claim 1

  1. (a)

    \(\Longrightarrow \)’ Suppose that \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\, \mathbf{y}\). Thus, if \({\varvec{\mathcal{G }}}:=\{\mathcal{F }\in {\varvec{\mathcal{F }}}\); \(\sum _{\mathcal{F }}\,u(\mathbf{x})\ge \sum _{\mathcal{F }}\,u(\mathbf{y})\}\), then \({\varvec{\mathcal{G }}}\in \mathfrak{UF }\). By definition, \(\sum _{\mathcal{J }}\,u(\mathbf{x})\ge \sum _{\mathcal{J }}\,u(\mathbf{y})\) for all \(\mathcal{J }\in {\varvec{\mathcal{G }}}\). Thus, Claim 1(a) says that \(\mathbf{x}\,{}^{^*}\!\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u,{\varvec{\mathcal{G }}}}}}\, \mathbf{y}\). ‘\({\Longleftarrow }\)’ If \(\mathbf{x}\,{}^{^*}\!\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u,{\varvec{\mathcal{G }}}}}}\, \mathbf{y}\); then Claim 1(a) says that \(\sum _{\mathcal{J }}\,u(\mathbf{x})\ge \sum _{\mathcal{J }}\,u(\mathbf{y})\) for all \(\mathcal{J }\in {\varvec{\mathcal{G }}}\). But \({\varvec{\mathcal{G }}}\in \mathfrak{UF }\), so this means that \({}^*\!\sum _{\mathcal{I }}\,u(\mathbf{x})\ge {}^*\!\sum _{\mathcal{I }}\,u(\mathbf{y})\), which means \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\,\mathbf{y}\).

  2. (b)

    The proof is the same as (a), but using Claim 1(b) instead of Claim 1(a). \(\square \)

Proof of Lemma 20

\({\Longleftarrow }\)’ follows from Lemmas 18 and 19.

\(\Longrightarrow \)’ Let \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\), and suppose that \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\, \mathbf{y}\). Then Lemma 19(a) yields some \({\varvec{\mathcal{G }}}\in \mathfrak{UF }\) such that, for all \(\mathbf{z}\in \mathcal{X }^\mathcal{I }\) and \(\mathcal{J }\in {\varvec{\mathcal{G }}}\), we have \(\mathbf{x}_{\mathcal{J }}\mathbf{z}_{{\mathcal{I }\setminus \mathcal{J }}}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\mathbf{y}_{\mathcal{J }}\mathbf{z}_{{\mathcal{I }\setminus \mathcal{J }}}\), and thus, \(\mathbf{x}_{\mathcal{J }}\mathbf{z}_{{\mathcal{I }\setminus \mathcal{J }}} {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}} \mathbf{y}_{\mathcal{J }}\mathbf{z}_{{\mathcal{I }\setminus \mathcal{J }}}\), by Lemma 18. But then \(\mathbf{x}_{\mathcal{J }}\mathbf{z}_{{\mathcal{I }\setminus \mathcal{J }}} {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathrm fin}}}\mathbf{y}_{\mathcal{J }}\mathbf{z}_{{\mathcal{I }\setminus \mathcal{J }}}\), because \(\big ( {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathrm fin}}}\big )=\big ({ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\) by hypothesis.

This holds for all \(\mathbf{z}\in \mathcal{X }^\mathcal{I }\) and \(\mathcal{J }\in {\varvec{\mathcal{G }}}\); thus, (C1) forces \(\mathbf{x}\succcurlyeq \, \mathbf{y}\), because \((\succcurlyeq )\) is \(\mathfrak{UF }\)-continuous. Thus, we have \((\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\mathbf{y})\Longrightarrow (\mathbf{x}\succcurlyeq \mathbf{y})\).

Likewise, if \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u}}}\, \mathbf{y}\), then Lemma 19(b) and (C2) imply that \(\mathbf{x}\succ \, \mathbf{y}\).

Lemma 14 says \(\big ({}^{^*}\!{ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\) is a complete preorder on \(\mathcal{X }^\mathcal{I }\). It follows that \((\succcurlyeq )=\big ({}^{^*}\!{ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\). \(\square \)

Proof of Theorem 2

\({\Longleftarrow }\)’ follows from Lemmas 14, 15(b), 17(b) and 19.

\(\Longrightarrow \)’ Let \(\big ( {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathrm fin}}}\big )\) be the finitary part of \((\succcurlyeq )\). Then \(\big ( {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathrm fin}}}\big )\) is \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariant and separable, so Theorem 1 yields a linearly ordered abelian group \((\mathcal{R },+,>)\) and \(u:\mathcal{X }{{\longrightarrow }}\mathcal{R }\) such that \(\big ( {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle \mathrm fin}}}\big )=\big ({ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\). Then Lemma 20 implies that \((\succcurlyeq )=\big ({}^{^*}\!{ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\).

The existence of an \(\mathcal{R }\) and \(u\) with the universal property follows from the construction in Theorem 1. \(\square \)

Proofs from Sects. 4 and 7.3

Proposition 6 is actually a special case of a more general result, which elucidates the behaviour of the hyperadditive preorder \(\big ({}^{^*}\!{ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\) when \(\mathfrak{UF }\) is constructed as in Lemma 15.

Lemma 22

Let \(\Pi \) be a permutation group with locally finite orbits on \(\mathcal{I }\), and let \(\mathfrak{UF }\) be as in Lemma 15. Let \(\mathcal{R }\) be a linearly ordered abelian group, let \(u:\mathcal{X }{{\longrightarrow }}\mathcal{R }\), and define \({}^*\!\mathcal{R }\) and \(\big ({}^{^*}\!{ {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}}\big )\) using \(\mathfrak{UF }\). Let \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\). Suppose that there is some finite \(\mathcal{J }\subset \mathcal{I }\) and some finite \(\Delta \subset \Pi \) such that, for all finite \(\mathcal{K }\subseteq \mathcal{I }\) we have:

$$\begin{aligned} \left( \mathcal{J }\subseteq \mathcal{K } \,and\, \delta (\mathcal{K })=\mathcal{K }\, for \,all \,\delta \in \Delta \right) \Longrightarrow \left( \displaystyle \sum _{k\in \mathcal{K }} u(x_k) \ge \sum _{k\in \mathcal{K }} u(y_k) \right).\qquad \quad \end{aligned}$$
(22)

Then \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\, \mathbf{y}\). If the inequalities on the right side of statement (22) are strict, then \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u}}}\, \mathbf{y}\)

Proof

By hypothesis, \(\sum _{k\in \mathcal{K }} u(x_k) \ge \sum _{k\in \mathcal{K }} u(y_k)\) for all \(\mathcal{K }\in {\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\). But \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\in \mathfrak{F }_\Gamma \) by the definition of \(\mathfrak{F }_\Gamma \),  and  \(\mathfrak{F }_\Gamma \subseteq \mathfrak{UF }\) by Lemma 15. Thus, \(\displaystyle {}^*\!\sum _{i\in \mathcal{I }} u(x_i) \ge {}^*\!\sum _{i\in \mathcal{I }} u(y_i)\). Thus, \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\, \mathbf{y}\). \(\square \)

Proof of Proposition 6

Let \(\Pi :=\Pi _{\scriptscriptstyle {\mathrm{fs}}}\) be the group of fixed-step permutations from Example 16(c), and then define \(\mathfrak{UF }\) as in Lemma 15. Let \(\pi :\mathbb{N }{{\longrightarrow }}\mathbb{N }\) be a permutation which cyclically permutes the elements of the interval \([m S+1\, \ldots \, (m+1)S]\), for every \(m\in \mathbb{N }\). Then \(\pi \) is a fixed-step permutation with \(T_\pi =S\). Note that the only \(\pi \)-invariant subsets of \(\mathbb{N }\) are intervals of the form \([n S\!+\!1 \,\ldots \, m S]\) for some \(n\le m\in \mathbb{N }\). Let \(\mathcal{J }:=[1\ldots \,M_0 S]\). Then for all finite \(\mathcal{K }\subseteq \mathbb{N }\) we have:

$$\begin{aligned} \left( \mathcal{J }\subseteq \mathcal{K } \text { and } \pi (\mathcal{K })=\mathcal{K } \right)&\Longrightarrow&\left( \mathcal{K }=[1\ldots \, M S], \,for \,some \,M\ge M_0 \right)\\&\mathop {\Longrightarrow }\limits _{(*)} \left( \displaystyle \sum _{k\in \mathcal{K }} u(x_k) > \sum _{k\in \mathcal{K }} u(y_k) \right), \end{aligned}$$

where \((*)\) is by hypothesis. Thus, setting \(\Delta :=\{\pi \}\) in Lemma 22, we deduce that \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u}}}\, \mathbf{y}\). \(\square \)

Proof of Corollary 7

If \(\mathcal{R }\subseteq \mathbb{R }\), then \(\mathcal{R }\) is Archimedean. Thus, if condition (b) holds, then there is some \(M_0\in \mathbb{N }\) such that

$$\begin{aligned} \sum _{t=1}^{t=M\,P} u(x_t) \ > \ \sum _{t=1}^{t=M\,P} u(y_t), \quad \text{ for } \text{ all } M\ge M_0\text{. } \end{aligned}$$
(23)

On the other hand, if condition (a) holds, then, set \(M_0:=\lceil P/T\rceil \). Then statement (23) is true. Now set \(S:=P\) in Proposition 6 to get \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succ }{{\scriptscriptstyle u}}}\, \mathbf{y}\). \(\square \)

Proof of Proposition 8

Recall that \(\mathcal{R }:={}^*\!\mathbb{R }\), and the expected utilities of mixed strategy pairs take values in \({}^*\!\mathcal{R }\).Footnote 42 If \(\rho ^*\) is a uniformly distributed mixed strategy for Row, then there is some \(r\in \mathcal{R }\) such that \(EU_{\scriptscriptstyle \mathrm{col}}(\rho ^*,c)=r\) for all \(c\in \mathbb{Z }\). Now define \(R:={}^*\!\sum _{i\in \mathcal{I }} r\) (an element of \({}^*\!\mathcal{R }\)). Then for any mixed strategy \(\kappa :\mathcal{I }{{\longrightarrow }}\mathbb{Z }\) for Column, we have

$$\begin{aligned} EU_{\scriptscriptstyle \mathrm{col}}(\rho ^*,\kappa ) = {}^*\!\sum _{i\in \mathcal{I }} EU_{\scriptscriptstyle \mathrm{col}}\left(\rho , \kappa (i)\right) = {}^*\!\sum _{i\in \mathcal{I }} r = R. \end{aligned}$$

In particular, \(EU_{\scriptscriptstyle \mathrm{col}}(\rho ^*,\kappa ^*)=R\). Thus, \(EU_{\scriptscriptstyle \mathrm{col}}(\rho ^*,\kappa ^*)\ge EU_{\scriptscriptstyle \mathrm{col}}(\rho ^*,\kappa )\) for all mixed strategies \(\kappa \), so \(\kappa ^*\) is a best response to \(\rho ^*\).

By an exactly symmetric argument, if \(\kappa ^*\) is a uniformly distributed mixed strategy for Column, then there is some (negative) \(R^{\prime }\in {}^*\!\mathcal{R }\) such that \(EU_{\scriptscriptstyle \mathrm{col}}(\rho ,\kappa ^*)\!=\!R'\) for any mixed strategy \(\rho \). In particular, \(EU_{\scriptscriptstyle \mathrm{col}}(\rho ^*,\kappa ^*)\!=\!R'\), so \(\rho ^*\) is a best response to \(\kappa ^*\).

Proof of Example 9(b)

This construction is well-known, but I include it for completeness. Recall that \({\varvec{\mathcal{P }}}(\mathcal{I })\) is the set of all subsets of \(\mathcal{I }\). A free ultrafilter on \(\mathcal{I }\) is a subset \(\varvec{\mathcal{U }\!\mathcal{F }}\subset {\varvec{\mathcal{P }}}(\mathcal{I })\) such that:

  1. (a)

    No finite subset of \(\mathcal{I }\) is in \(\varvec{\mathcal{U }\!\mathcal{F }}\).

  2. (b)

    If \(\mathcal{J },\mathcal{K }\in \varvec{\mathcal{U }\!\mathcal{F }}\), then \(\mathcal{J }\cap \mathcal{K }\in \varvec{\mathcal{U }\!\mathcal{F }}\).

  3. (c)

    For any \(\mathcal{U }\in \varvec{\mathcal{U }\!\mathcal{F }}\) and \(\mathcal{J }\subseteq \mathcal{I }\), if \(\mathcal{U }\subset \mathcal{J }\), then \(\mathcal{J }\in \varvec{\mathcal{U }\!\mathcal{F }}\) also.

  4. (d)

    For any \(\mathcal{J }\subseteq \mathcal{I }\), exactly one of \(\mathcal{J }\) or \(\mathcal{I }\setminus \mathcal{J }\) is in \(\varvec{\mathcal{U }\!\mathcal{F }}\).

As in Sect. 5, the Ultrafilter Lemma guarantees the existence of a free ultrafilter on \(\mathcal{I }\). Note that properties (a) and (d) together imply that every cofinite subset of \(\mathcal{I }\) is in \(\varvec{\mathcal{U }\!\mathcal{F }}\). Now define \(\nu :{\varvec{\mathcal{P }}}(\mathcal{I }){{\longrightarrow }}\{0,1\}\) by setting \(\nu [\mathcal{J }]:=1\) for all \(\mathcal{J }\in \varvec{\mathcal{U }\!\mathcal{F }}\), while \(\nu [\mathcal{J }]:=0\) for any \(\mathcal{J }\in {\varvec{\mathcal{P }}}(\mathcal{I })\setminus \varvec{\mathcal{U }\!\mathcal{F }}\) [in which case \(\mathcal{I }\setminus \mathcal{J }\in \varvec{\mathcal{U }\!\mathcal{F }}\), by property (d)]. Using the four defining properties of \(\varvec{\mathcal{U }\!\mathcal{F }}\), it is easy to verify that \(\nu \) is a probability measure.

It remains to show that \(\nu \) is uniform. Let \(\pi \in \Pi _{\scriptscriptstyle {\mathrm{fin}}}\), and let \(\mathcal{J }\subseteq \mathcal{I }\). Let \(\mathcal{J }_0:=\{j\in \mathcal{J }\); \(\pi (j)\ne j\}\), and let \(\mathcal{J }_1:=\pi [\mathcal{J }_0]\). Then \(\mathcal{J }_1\) is disjoint from \(\mathcal{J }\setminus \mathcal{J }_0\) (because \(\pi \) is a bijection). Also \(\mathcal{J }_0\) and \(\mathcal{J }_1\) are finite sets, because \(\pi \in \Pi _{\scriptscriptstyle {\mathrm{fin}}}\); thus, \(\nu [\mathcal{J }_0]=\nu [\mathcal{J }_1]=0\). But \(\pi (\mathcal{J }) = (\mathcal{J }\setminus \mathcal{J }_0)\sqcup \mathcal{J }_1\). Thus, \(\nu [\pi (\mathcal{J })] = \nu [\mathcal{J }]-\nu [\mathcal{J }_0]+\nu [\mathcal{J }_1]=\nu [\mathcal{J }]-0+0=\nu [\mathcal{J }]\), as desired. \(\square \)

Proof of Proposition 10

If \(\mathcal{I }\) is finite, then all rankings and all probability measures on \(\mathcal{I }\) are \(\mathfrak{UF }\)-continuous, so these conditions are vacuous. Meanwhile, \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariance just means invariance under all permutations of \(\mathcal{I }\). It is well-known that the only permutation-invariant probability measure on \(\mathcal{I }\) is the (real-valued) probability measure with \(\mu [\mathcal{J }]:=|\mathcal{J }|/|\mathcal{I }|\) for all \(\mathcal{J }\subseteq \mathcal{I }\), and the only \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariant likelihood ranking is the one defined by \(\mu \). Thus, for the rest of the proof, I will suppose that \(\mathcal{I }\) is infinite.

Existence of \((\succcurlyeq ):\) Let \({\varvec{\mathcal{F }}}\) be the set of all finite subsets of \(\mathcal{I }\). Fix a suitable ultrafilter \(\mathfrak{UF }\) on \({\varvec{\mathcal{F }}}\). Then define the preorder \((\succcurlyeq )\) on \({\varvec{\mathcal{P }}}(\mathcal{I })\) as follows: for all \(\mathcal{J },\mathcal{K }\subseteq \mathcal{I }\),

$$\begin{aligned} \left( \mathcal{J }\succcurlyeq \mathcal{K } \right) \iff \left( {\left\{ \mathcal{F }\in {\varvec{\mathcal{F }}} \; ; \; |\mathcal{J }\cap \mathcal{F }| \ge |\mathcal{K }\cap \mathcal{F }| \right\} }\in \mathfrak{UF } \right). \end{aligned}$$

In particular, if \(\mathcal{J },\mathcal{K }\subset \mathcal{I }\) are finite, then we have \(\mathcal{J }\succcurlyeq \mathcal{K }\) if and only if \(|\mathcal{J }|\ge |\mathcal{K }|\). It is easy to verify that this is a likelihood ranking. It is \(\mathfrak{UF }\)-continuous by construction. Finally, it is uniform because \(\mathfrak{UF }\) is a suitable ultrafilter (hence \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariant).

Existence of \(\mu \), and representation of \((\succcurlyeq )\) by \(\mu \): Let \(\mathcal{X }:=\{0,1\}\). Then there is an obvious correspondence between \({\varvec{\mathcal{P }}}(\mathcal{I })\) and \(\mathcal{X }^\mathcal{I }\), obtained by identifying each \(\mathcal{J }\subseteq \mathcal{I }\) with its indicator function \(\varvec{1}_\mathcal{J }\in \mathcal{X }^\mathcal{I }\). The likelihood ranking \((\succcurlyeq )\) can thus be regarded as a complete preorder on \(\mathcal{X }^\mathcal{I }\). The second condition in the definition of ‘likelihood ranking’ implies this preorder is separable. If \((\succcurlyeq )\) is uniform, then it is \(\Pi _{\scriptscriptstyle {\mathrm{fin}}}\)-invariant. Thus, Theorem 1 yields a linearly ordered abelian group \(\mathcal{R }\) and a utility function \(u:\mathcal{X }{{\longrightarrow }}\mathcal{R }\) such that the finitary part of \((\succcurlyeq )\) is the additive preorder defined by \(u\) via formula (2).

However, an inspection of the proof of Theorem 1 reveals that \(\mathcal{R }\) is a quotient of the free abelian group on \(|\mathcal{X }|-1\) generators. Since \(|\mathcal{X }|=2\), this is the free abelian group on one generator—namely \(\mathbb{Z }\). Thus, \(\mathcal{R }\) is a quotient of \(\mathbb{Z }\). But \(\mathcal{R }\) is linearly ordered, so it is torsion-free. Thus, \(\mathcal{R }\) is in fact isomorphic to \(\mathbb{Z }\) (with the natural order). It will be convenient to embed \(\mathbb{Z }\) into \(\mathbb{R }\) in the natural way. Thus, without loss of generality, we can suppose that \(\mathcal{R }=\mathbb{R }\). Thus, \({}^*\!\mathcal{R }={}^*\!\mathbb{R }\).

The first condition in the definition of ‘likelihood ranking’ implies that \(u(0)< u(1)\). Without loss of generality, suppose that \(u(0)=0\) and \(u(1)=1\). Thus, for any \(\mathbf{x}\in \mathcal{X }^\mathcal{I }\), we have

$$\begin{aligned} {}^*\!\sum _{i\in \mathcal{I }} u(x_i) = {}^*\!\sum _{i\in \mathcal{I }} x_i. \end{aligned}$$
(24)

Next, the \(\mathfrak{UF }\)-continuity of \((\succcurlyeq )\) as a likelihood ranking is obviously equivalent to the \(\mathfrak{UF }\)-continuity of \((\succcurlyeq )\) as a preorder on \(\mathcal{X }^\mathcal{I }\). Thus, Lemma 19 implies that \((\succcurlyeq )\) is the hyperadditive preorder defined by \(u\) via Eq. (4). Combining this with Eq. (24), and the identification of each subset of \(\mathcal{I }\) with its indicator function, we get:

$$\begin{aligned} \left( \mathcal{J }\succcurlyeq \mathcal{K } \right) \iff \left( \displaystyle {}^*\!\sum _{i\in \mathcal{I }}\varvec{1}_\mathcal{J }(i) \ge {}^*\!\sum _{i\in \mathcal{I }}\varvec{1}_\mathcal{K }(i) \right), \quad \text{ for } \text{ all } \mathcal{J },\mathcal{K }\subseteq \mathcal{I }\text{. } \end{aligned}$$
(25)

Define \(I:={}^*\!\sum _{i\in \mathcal{I }} 1\); this is a positive (infinite) element of \({}^*\!\mathbb{R }\). Define \(\epsilon :=1/I\). Then \(\epsilon \) is a positive infinitesimal hyperreal. For any \(\mathcal{J }\subseteq \mathcal{I }\), define

$$\begin{aligned} \mu [\mathcal{J }] := \epsilon \cdot {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_\mathcal{J }(i). \end{aligned}$$

In particular, \(\mu [\mathcal{I }] = \epsilon \cdot I = 1\). We must check that \(\mu \) is an \({}^*\!\mathbb{R }\)-valued probability measure on \(\mathcal{I }\). Properties (a) and (c) are satisfied by construction. To see (b), let \(\mathcal{J },\mathcal{K }\subset \mathcal{I }\) be disjoint. Then \(\varvec{1}_{\mathcal{J }\sqcup \mathcal{K }}=\varvec{1}_\mathcal{J }+ \varvec{1}_\mathcal{K }\). Thus,

$$\begin{aligned} {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_{\mathcal{J }\sqcup \mathcal{K }}(i)&= {}^*\!\sum _{i\in \mathcal{I }} \left( \varvec{1}_\mathcal{J }(i) +\varvec{1}_\mathcal{K }(i) \right) \mathop {=}\limits _{(*)} {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_\mathcal{J }(i) + {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_\mathcal{K }(i) , \end{aligned}$$

where \((*)\) follows from the definition of the summation operator \({}^*\!\sum \). Thus,

$$\begin{aligned} \mu [\mathcal{J }\sqcup \mathcal{K }]&= \epsilon \cdot {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_{\mathcal{J }\sqcup \mathcal{K }}(i) = \epsilon \cdot {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_\mathcal{J }(i) + \epsilon \cdot {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_\mathcal{K }(i) = \mu [\mathcal{J }]+\mu [\mathcal{K }], \end{aligned}$$

as desired. Thus, \(\mu \) is an \({}^*\!\mathbb{R }\)-valued probability measure on \(\mathcal{I }\). To see that \(\mu \) has full support, note that \(\mu \{i\}=\epsilon \) for all \(i\in \mathcal{I }\). Finally,

$$\begin{aligned} \left( \mu [\mathcal{J }]\ge \mu [\mathcal{K }] \right) \iff \left( \displaystyle {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_\mathcal{J }(i)\ge {}^*\!\sum _{i\in \mathcal{I }} \varvec{1}_\mathcal{J }(i) \right) \mathop {\Leftarrow \!\!\Rightarrow }\limits _{(*)} \left( \mathcal{J }\succcurlyeq \mathcal{K } \right), \end{aligned}$$

where \((*)\) is by statement (25). In other words, \((\succcurlyeq )\) is the likelihood ranking defined by \(\mu \). Thus, \(\mu \) is uniform because \((\succcurlyeq )\) is uniform. Likewise, \(\mu \) is \(\mathfrak{UF }\)-continuous, because \((\succcurlyeq )\) is \(\mathfrak{UF }\)-continuous.

Uniqueness of \(\mu \): Now let \(\mu \) be any uniform, \(\mathfrak{UF }\)-continuous \({}^*\!\mathbb{R }\)-valued probability measure on \(\mathcal{I }\). Then for all \(\mathcal{J },\mathcal{K }\subseteq \mathcal{I }\), we have

$$\begin{aligned} \left( \mu [\mathcal{J }]\ge \mu [\mathcal{K }] \right)&\mathop {\Leftarrow \!\!\Rightarrow }\limits _{(*)}&\left( {\left\{ \mathcal{F }\in {\varvec{\mathcal{F }}} \; ; \; \mu [\mathcal{J }\cap \mathcal{F }] \ge \mu [\mathcal{K }\cap \mathcal{F }] \right\} }\in \mathfrak{UF } \right) \\&\mathop {\Leftarrow \!\!\Rightarrow }\limits _{(\dagger )} \left( {\left\{ \mathcal{F }\in {\varvec{\mathcal{F }}} \; ; \; |\mathcal{J }\cap \mathcal{F }| \ge |\mathcal{K }\cap \mathcal{F }| \right\} }\in \mathfrak{UF } \right). \end{aligned}$$

Thus, \(\mu \) is uniquely determined by \(\mathfrak{UF }\). Here, \((*)\) is by the \(\mathfrak{UF }\)-continuity of \(\mu \), combined with ultrafilter axiom (UF). Meanwhile, \((\dagger )\) is because the uniformity of \(\mu \) implies that, for any finite subsets \(\mathcal{G },\mathcal{H }\subset \mathcal{I }\), we have \(\mu [\mathcal{G }]\ge \mu [\mathcal{H }]\) if and only if \(|\mathcal{G }|\ge |\mathcal{H }|\).

Uniqueness of \((\succcurlyeq )\): Finally, let \((\succcurlyeq )\) be any \(\mathfrak{UF }\)-continuous, uniform likelihood ranking. Then we have shown that \((\succcurlyeq )\) can be represented by an \(\mathfrak{UF }\)-continuous, uniform \({}^*\!\mathbb{R }\)-valued probability measure \(\mu \). But we have also shown that there is only one such measure. Thus, there is only one such likelihood ranking. \(\square \)

Recall that \({}^*\!\log _2:{}^*\!\mathbb{R }_+{{\longrightarrow }}{}^*\!\mathbb{R }\) is the unique hyperreal extension of the base-2 logarithm function. The next result is needed in the proof of Proposition 11.

Lemma 23

Let \(\ell :=\ln (2)\). For any positive \(x<z\in {}^*\!\mathbb{R }\), we have

$$\begin{aligned} {}^*\!\log _2(z)-{}^*\!\log _2(x) < \frac{z-x}{\ell \,x }. \end{aligned}$$

Proof

First suppose that \(x\) and \(z\) are positive real numbers, and consider the ordinary real logarithm. The Mean Value Theorem yields some \(y\in {\left( x,z \right)}\) such that

$$\begin{aligned} \frac{\log _2(z)-\log _2(x)}{z-x} = \log _2'(y) = \frac{1}{\ell \,y }. \end{aligned}$$

Thus, \(\displaystyle \log _2(z)-\log _2(x)=\frac{z-x}{\ell \,y } < \frac{z-x}{\ell \,x }\), because \(y>x\).

This proves the claim for any for any real \(z>x>0\). Now the general statement follow by applying the Transfer Principle. \(\square \)

Proof of Proposition 11

Let \({\varvec{\mathcal{F }}}\) be the set of all finite subsets of \(\mathcal{I }\), and let \(\mathfrak{UF }\) be a suitable ultrafilter on \({\varvec{\mathcal{F }}}\) (as defined in Sect. 5). For any \(\mathcal{J }\in {\varvec{\mathcal{F }}}\), we define

$$\begin{aligned} H_\mathcal{J }(\nu ):= -\sum _{j\in \mathcal{J }} \nu \{j\}\,{}^*\!\log _2[\nu \{j\}] \quad \text { and }\quad H_\mathcal{J }(\mu ):= -\sum _{j\in \mathcal{J }} \mu \{j\}\,{}^*\!\log _2[\mu \{j\}]. \end{aligned}$$

Thus, \(H(\nu )\) (an element of \(^{**}\mathbb{R }\)) is in fact the \(\mathfrak{UF }\)-equivalence class of the function \([{\varvec{\mathcal{F }}}\ni \mathcal{J }\mapsto H_\mathcal{J }(\nu )\in {}^*\!\mathbb{R }]\) (and likewise for \(H(\mu )\)). Thus, if \({\varvec{\mathcal{G }}}:=\{\mathcal{J }\in {\varvec{\mathcal{F }}}\); \(H_\mathcal{J }(\nu ) < H_\mathcal{J }(\mu )\}\), then \(H(\nu )< H(\mu )\) if and only if \({\varvec{\mathcal{G }}}\in \mathfrak{UF }\). So this is what we must show.

Recall that there is some positive infinitesimal \(\epsilon \in {}^*\!\mathbb{R }\) such that \(\mu \{i\}=\epsilon \) for all \(i\in \mathcal{I }\).

Claim 1 For any finite subset \(\mathcal{J }\subset \mathcal{I }\), we have \(H_\mathcal{J }(\mu )=-\mu [\mathcal{J }]\cdot {}^*\!\log _2(\epsilon )\).

Proof

\(\displaystyle H_\mathcal{J }(\mu ) = -\sum _{j\in \mathcal{J }} \mu \{j\}\,{}^*\!\log _2[\mu \{j\}] = -\sum _{j\in \mathcal{J }} \epsilon \,{}^*\!\log _2(\epsilon ) = -|\mathcal{J }|\epsilon \,{}^*\!\log _2(\epsilon ) = -\mu [\mathcal{J }]\,{}^*\!\log _2(\epsilon )\). \(\diamondsuit \text { Claim } 1\)

For any finite subset \(\mathcal{J }\subset \mathcal{I }\), let \(\nu _\mathcal{J }\) and \(\mu _\mathcal{J }\) be the conditional probability measures induced on \(\mathcal{J }\) by \(\nu \) and \(\mu \). That is, let us define \(\nu _\mathcal{J }[\mathcal{K }]:=\nu [\mathcal{K }]/\nu [\mathcal{J }]\) and \(\mu _\mathcal{J }[\mathcal{K }]:=\mu [\mathcal{K }]/\mu [\mathcal{J }]\) for all \(\mathcal{K }\subseteq \mathcal{J }\). It follows that \(\mu _\mathcal{J }\) is the uniform measure on \(\mathcal{J }\), because \(\mu \) is the uniform measure on \(\mathcal{I }\). Let us define \(H(\nu _\mathcal{J }):=-\sum _{j\in \mathcal{J }} \nu _\mathcal{J }\{j\}\,{}^*\!\log _2(\nu _\mathcal{J }\{j\})\). (If \(\nu \) was an ordinary real-valued probability measure, then we could remove the ‘\(*\)’ and this would just be the defining formula (10) for the entropy of a probability measure on a finite set.) Let us likewise define \(H(\mu _\mathcal{J })\).

Claim 2 For any finite subset \(\mathcal{J }\subset \mathcal{I }\), we have

  1. (a)

    \(\displaystyle H(\nu _\mathcal{J }) = \frac{1}{\nu [\mathcal{J }]} H_\mathcal{J }(\nu ) + {}^*\!\log _2(\nu [\mathcal{J }])\), and

  2. (b)

    \(\displaystyle H(\mu _\mathcal{J }) = \frac{1}{\mu [\mathcal{J }]} H_\mathcal{J }(\mu ) + {}^*\!\log _2(\mu [\mathcal{J }])\).

Proof

$$\begin{aligned} H(\nu _\mathcal{J })&= -\sum _{j\in \mathcal{J }}\nu _\mathcal{J }\{j\}\,{}^*\!\log _2(\nu _\mathcal{J }\{j\}) = -\sum _{j\in \mathcal{J }}\frac{\nu \{j\}}{\nu [\mathcal{J }]}\,{}^*\!\log _2\left(\frac{\nu \{j\}}{\nu [\mathcal{J }]}\right)\\&= -\frac{1}{\nu [\mathcal{J }]}\sum _{j\in \mathcal{J }} \nu \{j\}\,\left({}^*\!\log _2\left(\nu \{j\}\right)-{}^*\!\log _2\left(\nu [\mathcal{J }]\right)\right) \\&= \frac{1}{\nu [\mathcal{J }]} H_\mathcal{J }(\nu ) +\frac{1}{\nu [\mathcal{J }]}\sum _{j\in \mathcal{J }} \nu \{j\}\,{}^*\!\log _2\left(\nu [\mathcal{J }]\right) \\&= \frac{1}{\nu [\mathcal{J }]} H_\mathcal{J }(\nu ) +\frac{\nu [\mathcal{J }]}{\nu [\mathcal{J }]}\,{}^*\!\log _2\left(\nu [\mathcal{J }]\right) = \frac{1}{\nu [\mathcal{J }]} H_\mathcal{J }(\nu ) + {}^*\!\log _2(\nu [\mathcal{J }]), \end{aligned}$$

which proves (a). A similar computation proves (b). \(\diamondsuit \text { Claim } 2\)

Claim 3 Let \({\varvec{\mathcal{E }}}:=\{\mathcal{F }\in {\varvec{\mathcal{F }}}; \nu [\mathcal{F }]\le \mu [\mathcal{F }]\}\). Then \({\varvec{\mathcal{E }}}\in \mathfrak{UF }\).

Proof

Since \(\nu \) is normal and \(\mu \) is a \({}^*\!\sum \)-additive probability measure, we have

$$\begin{aligned} {}^*\!\sum _{i\in \mathcal{I }} \nu \{i\} \le 1 = \mu [\mathcal{I }] = {}^*\!\sum _{i\in \mathcal{I }} \mu \{i\}. \end{aligned}$$

Thus, if \({\varvec{\mathcal{E }}}':=\left\rbrace \mathcal{J }\in {\varvec{\mathcal{F }}}; \ \sum _{j\in \mathcal{J }} \nu \{j\} \le \sum _{j\in \mathcal{J }} \mu \{j\}\right\lbrace \), then we must have \({\varvec{\mathcal{E }}}^{\prime }\in \mathfrak{UF }\), by the definition of the summation operator \({}^*\!\sum \). But for any finite subset \(\mathcal{J }\subset \mathcal{I }\), we have \(\nu [\mathcal{J }]=\sum _{j\in \mathcal{J }} \nu \{j\} \) and \(\mu [\mathcal{J }]=\sum _{j\in \mathcal{J }} \mu \{j\}\), by finite additivity. Thus, in fact \({\varvec{\mathcal{E }}}'={\varvec{\mathcal{E }}}\). \(\diamondsuit \text { Claim } 3\)

Now, if \(\nu \ne \mu \), then there exists some finite subset \(\mathcal{K }\subset \mathcal{I }\) such that the measure \(\nu _\mathcal{K }\) is not uniform on \(\mathcal{K }\). Thus, for any finite subset \(\mathcal{E }\subset \mathcal{I }\), if \(\mathcal{K }\subseteq \mathcal{E }\), then the measure \(\nu _\mathcal{E }\) is not uniform either.

Claim 4 Let \(\varvec{\mathcal{J }}:=\{\mathcal{E }\in {\varvec{\mathcal{E }}}; \mathcal{K }\subseteq \mathcal{E }\}\). Then \(\varvec{\mathcal{J }}\in \mathfrak{UF }\).

Proof

Let \({\varvec{\mathcal{F }}}(\mathcal{K }):=\{\mathcal{F }\in {\varvec{\mathcal{F }}}\); \(\mathcal{K }\subseteq \mathcal{F }\}\). Then \({\varvec{\mathcal{F }}}(\mathcal{K })\in \mathfrak{UF }\) because \(\mathfrak{UF }\) is a suitable ultrafilter. Meanwhile, \({\varvec{\mathcal{E }}}\in \mathfrak{UF }\) by Claim 3. But \(\varvec{\mathcal{J }}:={\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}(\mathcal{K })\); thus, \(\varvec{\mathcal{J }}\in \mathfrak{UF }\) by ultrafilter axiom (F1). \(\diamondsuit \text { Claim } 4\)

Claim 5 If \(\mathcal{J }\in \varvec{\mathcal{J }}\), then \(H(\nu _\mathcal{J })< H(\mu _\mathcal{J })\).

Proof

\(\mathcal{J }\) is finite, and \(\mu _\mathcal{J }\) is the uniform measure on \(\mathcal{J }\); thus, it is the (unique) measure of maximal entropy on \(\mathcal{J }\) (Cover and Thomas 2006). By the definition of \(\varvec{\mathcal{J }}\), the measure \(\nu _\mathcal{J }\) is non-uniform; thus, \(H(\nu _\mathcal{J })< H(\mu _\mathcal{J })\). \(\diamondsuit \text { Claim } 5\)

Claim 6 If \(\mathcal{J }\in \varvec{\mathcal{J }}\), then \(H_\mathcal{J }(\nu ) < H_\mathcal{J }(\mu )\).

Proof

First note that

$$\begin{aligned} H_\mathcal{J }(\nu )&\mathop {=}\limits _{(*)} \nu [\mathcal{J }] \,\left( H(\nu _\mathcal{J }) - {}^*\!\log _2(\nu [\mathcal{J }])\right) \ \ \mathop {<}\limits _{(\dagger )} \ \ \nu [\mathcal{J }] \,\left( H(\mu _\mathcal{J }) - {}^*\!\log _2(\nu [\mathcal{J }])\right) \\&\mathop {=}\limits _{(\diamond )} \nu [\mathcal{J }] \,\left(\frac{1}{\mu [\mathcal{J }]} H_\mathcal{J }(\mu ) + {}^*\!\log _2(\mu [\mathcal{J }]) - {}^*\!\log _2(\nu [\mathcal{J }])\right) \\&= \frac{\nu [\mathcal{J }]}{\mu [\mathcal{J }]} H_\mathcal{J }(\mu ) +\nu [\mathcal{J }] \,\left( {}^*\!\log _2(\mu [\mathcal{J }]) - {}^*\!\log _2(\nu [\mathcal{J }])\right), \end{aligned}$$

where \((*)\) is by Claim 2(a), \((\dagger )\) is by Claim 5 and \((\diamond )\) is by Claim 2(b). It follows that

$$\begin{aligned} H_\mathcal{J }(\nu ) \!-\! H_\mathcal{J }(\mu ) \ < \ \left(\frac{\nu [\mathcal{J }]}{\mu [\mathcal{J }]} \!-\!1\right) \, H_\mathcal{J }(\mu ) \!+\!\nu [\mathcal{J }] \,\left( {}^*\!\log _2(\mu [\mathcal{J }]) - {}^*\!\log _2(\nu [\mathcal{J }])\right).\nonumber \\ \end{aligned}$$
(26)

Now,

$$\begin{aligned} \left(\frac{\nu [\mathcal{J }]}{\mu [\mathcal{J }]} -1\right) \, H_\mathcal{J }(\mu )&= \frac{\nu [\mathcal{J }]-\mu [\mathcal{J }]}{\mu [\mathcal{J }]} \, H_\mathcal{J }(\mu ) \nonumber \\&= \frac{H_\mathcal{J }(\mu )}{\mu [\mathcal{J }]}\left(\nu [\mathcal{J }]-\mu [\mathcal{J }]\right)\nonumber \\&\mathop {=}\limits _{(*)} -{}^*\!\log _2(\epsilon )\,\left(\nu [\mathcal{J }]-\mu [\mathcal{J }]\right)\nonumber \\&= {}^*\!\log _2(\epsilon )\,\left(\mu [\mathcal{J }]-\nu [\mathcal{J }]\right), \end{aligned}$$
(27)

where \((*)\) is by Claim 1. Meanwhile,

$$\begin{aligned}&\nu [\mathcal{J }] \,\left( {}^*\!\log _2(\mu [\mathcal{J }]) - {}^*\!\log _2(\nu [\mathcal{J }])\right) \nonumber \\&\mathop {<}\limits _{(*)} \nu [\mathcal{J }] \, \frac{\mu [\mathcal{J }]-\nu [\mathcal{J }]}{\ell \,\nu [\mathcal{J }]} =\frac{\mu [\mathcal{J }]-\nu [\mathcal{J }]}{\ell }, \end{aligned}$$
(28)

where \((*)\) is by Lemma 23, because \(\nu [\mathcal{J }]<\mu [\mathcal{J }]\) because \(\mathcal{J }\in \varvec{\mathcal{J }}\subseteq {\varvec{\mathcal{E }}}\). Substituting Eq. (27) and inequality (28) into inequality (26), we get

$$\begin{aligned} H_\mathcal{J }(\nu ) - H_\mathcal{J }(\mu )&< {}^*\!\log _2(\epsilon )\,\left(\mu [\mathcal{J }]-\nu [\mathcal{J }]\right) +\frac{\mu [\mathcal{J }]-\nu [\mathcal{J }]}{\ell } \\&= \left(\frac{1}{\ell }+{}^*\!\log _2(\epsilon )\right)\,\left(\mu [\mathcal{J }]-\nu [\mathcal{J }]\right) \, \mathop {<}\limits _{(*)} \, 0, \end{aligned}$$

and hence, \(H_\mathcal{J }(\nu ) < H_\mathcal{J }(\mu )\), as desired. Here \((*)\) is because \(\mu [\mathcal{J }]>\nu [\mathcal{J }]\), whereas \({}^*\!\log _2(\epsilon )<-\frac{1}{\ell }\), because \(\epsilon <1/2^{1/\ell }\) (indeed, \(\epsilon \) is infinitesimal). \(\diamondsuit \text { Claim } 6\)

Now Claims 4 and 6 implies that \(H(\nu )<H(\mu )\), as desired. \(\square \)

The next result shows that the ‘normality’ hypothesis of Proposition 11 is not redundant. Let \(\mu \) be the uniform measure on \(\mathcal{I }\), from Proposition 10.

Proposition 24

If \(\mathcal{I }\) is infinite, then there exists an \({}^*\!\mathbb{R }\)-valued probability measure \(\lambda \) on \(\mathcal{I }\) which is not normal, such that \(H(\lambda )>H(\mu )\).

Proof

Let \(r\in {}^*\!\mathbb{R }\) be a positive value such that \(\mu [\mathcal{J }]>r\) for all infinite \(\mathcal{J }\subset \mathcal{I }\). (For example, we could set \(r:=\mu [\mathcal{F }]\) for some nonempty but finite subset \(\mathcal{F }\subset \mathcal{I }\).) Let \(\nu \) be a probability measure such that \(\nu [\mathcal{J }]=0\) for all finite \(\mathcal{J }\subset \mathcal{I }\) [e.g. the measure from Example 9(b)]. Define the function \(\lambda :{\varvec{\mathcal{P }}}(\mathcal{I }){{\longrightarrow }}{}^*\!\mathbb{R }\) by

$$\begin{aligned} \lambda (\mathcal{J }) := \frac{2\mu [\mathcal{J }] - r\,\nu [\mathcal{J }]}{2-r} , \quad \text{ for } \text{ all } \mathcal{J }\subseteq \mathcal{I }\text{. } \end{aligned}$$

Then \(\lambda \) is finitely additive, because \(\mu \) and \(\nu \) are finitely additive. Also,

$$\begin{aligned} \lambda [\emptyset ]&= \frac{2\mu [\emptyset ]-r\,\nu [\emptyset ]}{2-r} = \frac{0}{2-r} = 0,\\ \text { and }\lambda [\mathcal{I }]&= \frac{2\mu [\mathcal{I }]-r\,\nu [\mathcal{I }]}{2-r} = \frac{2-r}{2-r} = 1. \end{aligned}$$

It remains to show that \(\lambda [\mathcal{J }]\ge 0\) for all nonempty \(\mathcal{J }\subseteq \mathcal{I }\). If \(\mathcal{J }\) is finite, then \(\lambda [\mathcal{J }]=2\mu [\mathcal{J }]/(2-r)>0\), because \(\mu [\mathcal{J }]>0\) because \(\mu \) has full support. If \(\mathcal{J }\) is infinite, then

$$\begin{aligned} \lambda [\mathcal{J }] = \frac{2\mu [\mathcal{J }] - r\,\nu [\mathcal{J }]}{2-r} \ge \frac{2\mu [\mathcal{J }] - r }{2-r} > \frac{2r- r }{2-r} > 0, \end{aligned}$$

as desired. Thus, \(\lambda \) is a probability measure. To see that \(\lambda \) is not normal, note that \(\lambda \{i\}=2\mu \{i\}/(2-r)\), for all \(i\in \mathcal{I }\). Thus,

$$\begin{aligned} {}^*\!\sum _{i\in \mathcal{I }}\lambda \{i\} ={}^*\!\sum _{i\in \mathcal{I }} \frac{2\mu \{i\}}{2-r} =\frac{2}{2-r} {}^*\!\sum _{i\in \mathcal{I }} \mu \{i\} \ \ \mathop {=}\limits _{(*)} \ \ \frac{2}{2-r} \, \mu [\mathcal{I }] =\frac{2}{2-r} \, > \ , 1. \end{aligned}$$

Here, \((*)\) is because \(\mu \) is \({}^*\!\sigma \)-additive.

Finally, see that \(H(\lambda )>H(\mu )\), note that

$$\begin{aligned} H(\lambda )&\mathop {=}\limits _{(*)} -{}^*\!\sum _{i\in \mathcal{I }}\lambda \{i\}\cdot {}^*\!\log _2[\lambda \{i\}] = -{}^*\!\sum _{i\in \mathcal{I }}\frac{2\,\mu \{i\} }{2-r}\cdot {}^*\!\log _2\left(\frac{2\,\mu \{i\} }{2-r}\right)\\&= -\frac{2}{2-r}{}^*\!\sum _{i\in \mathcal{I }}\mu \{i\}\cdot \left( {}^*\!\log _2[\mu \{i\}] + {}^*\!\log _2(2)\!-\!{}^*\!\log _2(2-r)\right) \\&= -\frac{2}{2-r}{}^*\!\sum _{i\in \mathcal{I }}\mu \{i\}\cdot {}^*\!\log _2[\mu \{i\}] -\frac{2}{2-r}\left( {}^*\!\log _2(2)-{}^*\!\log _2(2\!-\!r)\right){}^*\!\sum _{i\in \mathcal{I }}\mu \{i\}\\&\mathop {=}\limits _{(\diamond )} \frac{2}{2-r} H(\mu ) - \frac{2}{2-r}\left( {}^*\!\log _2(2)-{}^*\!\log _2(2-r)\right)\,\mu [\mathcal{I }]\\&\mathop {>}\limits _{(\dagger )} \left(1+\frac{r}{2-r}\right) H(\mu ) - \frac{2r}{(2-r)^2\,\ell } \qquad \text{(with } \ell :=\ln (2)\text{). } \end{aligned}$$

Here, both \((*)\) and \((\diamond )\) invoke defining formula (11), and \((\diamond )\) also uses the fact that \(\mu \) is \({}^*\!\sum \)-additive. Finally, \((\dagger )\) is because \(\mu [\mathcal{I }]=1\) and Lemma 23 says

$$\begin{aligned} {}^*\!\log _2(2)-{}^*\!\log _2(2-r) < \frac{r}{(2-r)\,\ell }, \end{aligned}$$

Thus,

$$\begin{aligned} H(\lambda )-H(\mu )&> \frac{r}{2-r} H(\mu ) - \frac{2r}{(2-r)^2\,\ell } = \frac{r}{2-r}\left(H(\mu )-\frac{2}{(2-r)\,\ell }\right) > 0, \end{aligned}$$

because \(H(\mu )>\frac{2}{(2-r)\,\ell }\) (indeed, \(H(\mu )\) is infinite). Thus, \(H(\lambda )>H(\mu )\), as claimed.

\(\square \)

Proof of Proposition 12

Any element of \(r\in \mathbb{R }\) has a unique binary expansion \(r=\sum _{z\in \mathbb{Z }} r_z 2^z\) for some sequence \(\{r_z\}_{z\in \mathbb{Z }}\) taking values in \(\{0,1\}\), such that: (i) there exists some \(N\in \mathbb{N }\) such that \(r_z=0\) for all \(z>N\); and (ii) for any \(M\in \mathbb{N }\), there exists some \(z<-M\) with \(r_z=0\).Footnote 43 For all \(n\in \mathbb{Z }\), define \(\pi _n:\mathbb{R }{{\longrightarrow }}\mathbb{R }\) by \(\pi _n(r):=(1-r_n) 2^n + \sum _{z\in \mathbb{Z }\setminus \{n\}} r_z 2^z\) (i.e. \(\pi _n\) toggles the \(n\)th binary digit of \(r\)). Let \(\Gamma \) be the group generated by \(\{\pi _n\}_{n\in \mathbb{Z }}\); then \(\Gamma \) has locally finite orbits (because any finite subset of \(\Gamma \) can only act upon a finite set of digits).

Let \(\mathcal{I }:=\mathbb{R }\), and define \(\mathfrak{UF }\) be as in Lemma 17; then \(\mathfrak{UF }\) is a suitable ultrafilter, and \( \Gamma \subseteq \Pi _\mathfrak{UF }\). Let \(\mu \) be the full uniform measure on \(\mathbb{R }\) defined by \(\mathfrak{UF }\) in Proposition 10. Then Lemma 15(b) implies that

$$\begin{aligned} \mu [\pi _n(\mathcal{J })]=\mu [\mathcal{J }], \, \text{ for } \text{ any } \mathcal{J }\subseteq \mathbb{R } \text{ and } n\in \mathbb{Z }\text{. } \end{aligned}$$
(29)

Now let \(M:=\mu {\left[ 0,1 \right)}\) and let \(\lambda \) be the Lebesgue measure on \(\mathbb{R }\). Statement (29) and the finite additivity of \(\mu \) imply that

$$\begin{aligned} \mu {\left[ \frac{n}{2^k},\frac{m}{2^k} \right)} = \frac{(m-n)}{2^k}M = M\cdot \lambda {\left[ \frac{n}{2^k},\frac{m}{2^k} \right)}, \end{aligned}$$
(30)

for any \(n,m,k\in \mathbb{Z }\) with \(n<m\). An interval of this kind is called a dyadic interval. A  dyadic subset is a finite disjoint union of dyadic intervals. In particular, if \(a<b\in \mathbb{R }\) have finite binary expansions, then the interval \([a,b)\) is a dyadic subset. (For example, if \(a=\frac{1}{8}\) and \(b=\frac{3}{4}\), then \([a,b)=[\frac{1}{8},\frac{1}{4}) \cup [\frac{1}{4},\frac{1}{2}) \cup [\frac{1}{2},\frac{3}{4})\).) If \(\mathcal{D }\subset \mathbb{R }\) is any dyadic subset, then Eq. (30) and the finite additivity of \(\mu \) and \(\lambda \) imply that

$$\begin{aligned} \mu [\mathcal{D }] = M\cdot \lambda [\mathcal{D }]. \end{aligned}$$
(31)

Now let \(a<b\in \mathbb{R }\) be arbitrary, and let \(\mathcal{J }\) be either the interval \((a,b)\), or \([a,b)\), or \((a,b]\), or \([a,b]\).

Claim 1 For any (real) \(\epsilon >0\), there exist dyadic subsets \(\mathcal{D }_1,\mathcal{D }_2\subset \mathbb{R }\) with \(\mathcal{D }_1\subseteq \mathcal{J }\subseteq \mathcal{D }_2\) and \(\lambda [\mathcal{J }]-\epsilon \le \lambda [\mathcal{D }_1]\le \lambda [\mathcal{J }] \le \lambda [\mathcal{D }_2]\le \lambda [\mathcal{J }]+\epsilon \) .

Proof

Suppose that \(a\) and \(b\) have dyadic expansions \(a=\sum _{z\in \mathbb{Z }} a_z 2^z\) and \(b=\sum _{z\in \mathbb{Z }} b_z 2^z\), where \(a_z,b_z\in \{0,1\}\) for all \(z\in \mathbb{Z }\). Find \(n\in \mathbb{N }\) such that \(1/2^n<\epsilon /2\). Define

$$\begin{aligned} \begin{array}{rclcrcl} {\underline{a}}&{}:=&{}\displaystyle \sum _{z=-n}^{\infty }a_z 2^z,&{}&{} {\overline{a}}&{}:=&{} \displaystyle {\underline{a}}+ \frac{1}{2^n},\\ {\underline{b}}&{}:=&{}\displaystyle \sum _{z=-n}^{\infty }a_z 2^z,&{}\text { and }&{} {\overline{b}}&{}:=&{}\displaystyle {\underline{b}}+ \frac{1}{2^n}. \end{array} \end{aligned}$$

Then \({\underline{a}}\le a< {\overline{a}}\) and \({\underline{b}}\le b< {\overline{b}}\). Thus, if \(\mathcal{D }_1:=[{\overline{a}},{\underline{b}})\) and \(\mathcal{D }_2:=[{\underline{a}},{\overline{b}})\), then \(\mathcal{D }_1\subseteq \mathcal{J }\subseteq \mathcal{D }_2\), and thus \(\lambda [\mathcal{D }_1]\le \lambda [\mathcal{J }]\le \lambda [\mathcal{D }_2]\). By construction, \({\underline{a}}, {\overline{a}}, {\underline{b}}\) and \({\overline{b}}\) all have finite binary expansions; thus \(\mathcal{D }_1\) and \(\mathcal{D }_2\) are dyadic subsets. Finally, note that

$$\begin{aligned} \lambda [\mathcal{D }_2]-\lambda [\mathcal{D }_1] \!=\! ({\overline{b}}-{\underline{a}}) - ({\underline{b}}-{\overline{a}}) \!=\! ({\overline{b}}-{\underline{b}}) \!+\! ({\overline{a}}-{\underline{a}}) \!=\! \frac{1}{2^n}\!+\!\frac{1}{2^n} \!=\!\frac{2}{2^n} < \epsilon . \end{aligned}$$

The claim follows. \(\diamondsuit \text { Claim } 1\)

Now, if \(\mathcal{D }_1,\mathcal{D }_2\) are in Claim 1, then we also have

$$\begin{aligned} \lambda [\mathcal{D }_1] ~\mathop {=}\limits _{(*)}~ \frac{\mu [\mathcal{D }_1]}{M} \underset{\dagger }{(\le )} \frac{\mu [\mathcal{J }]}{M} \underset{\dagger }{(\le )} \frac{\mu [\mathcal{D }_2]}{M} \mathop {=}\limits _{(*)} \lambda [\mathcal{D }_2], \end{aligned}$$

where both \((*)\) are invocations of statement (31), while the \((\dagger )\) inequalities are because \(\mathcal{D }_1\subseteq \mathcal{J }\subseteq \mathcal{D }_2\). Since \(\lambda [\mathcal{J }]-\epsilon \le \lambda [\mathcal{D }_1]\) and \(\lambda [\mathcal{D }_2]\le \lambda [\mathcal{J }]+\epsilon \), we conclude that

$$\begin{aligned} \lambda [\mathcal{J }]-\epsilon \le \frac{\mu [\mathcal{J }]}{M} \le \lambda [\mathcal{J }]+\epsilon . \end{aligned}$$

Since this holds for any \(\epsilon >0\), we conclude that \(\mu [\mathcal{J }]/M-\lambda [\mathcal{J }]\) is either zero or infinitesimal.

This proves the theorem statement for a single interval. Any simple set is a finite union of intervals; thus, the general statement follows because both \(\mu \) and \(\lambda \) are finitely additive, and a finite sum of infinitesimals is still infinitesimal. \(\square \)

Proof of Proposition 21

Fix \(\mathbf{x},\mathbf{y}\in \mathcal{X }^\mathcal{I }\). Let \({\varvec{\mathcal{F }}}_\mathbf{x}:=\{\mathcal{F }\in {\varvec{\mathcal{F }}}\); \(\sum _\mathcal{F }\, u(\mathbf{x}) \ge \sum _\mathcal{F }\, u(\mathbf{y})\}\), while \({\varvec{\mathcal{F }}}_\mathbf{y}:={\varvec{\mathcal{F }}}_\mathbf{x}^\complement = \{\mathcal{F }\in {\varvec{\mathcal{F }}}\); \(\sum _\mathcal{F }\, u(\mathbf{x}) < \sum _\mathcal{F }\, u(\mathbf{y})\}\).

\(\Longrightarrow \)’ If \(\mathbf{x}\, {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}\!^{^{\Gamma }}\,\,\mathbf{y}\), then there exists some \(\mathcal{J }\in {\varvec{\mathcal{F }}}\) and finite \(\Delta \subset \Gamma \) such that \(\displaystyle \sum _{\mathcal{F }} u(\mathbf{x}) \ \ge \ \sum _{\mathcal{F }} u(\mathbf{y})\) for all \(\mathcal{F }\in {\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\). Thus, \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\subseteq {\varvec{\mathcal{F }}}_\mathbf{x}\). But \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\in \mathfrak{UF }\) for any \(\Gamma \)-admissible free ultrafilter \(\mathfrak{UF }\); thus, \({\varvec{\mathcal{F }}}_\mathbf{x}\in \mathfrak{UF }\) by (F2). Thus, \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}^{_{\mathfrak{UF }}}\, \mathbf{y}\).

\({\Longleftarrow }\)’ (by contrapositive) Suppose that . Then there is no \(\mathcal{J }\in {\varvec{\mathcal{F }}}\) and finite \(\Delta \subseteq \Gamma \) with \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\subseteq {\varvec{\mathcal{F }}}_\mathbf{x}\). Thus, \({\varvec{\mathcal{F }}}_\Delta (\mathcal{J })\cap {\varvec{\mathcal{F }}}_\mathbf{y}\ne \emptyset \) for every \(\mathcal{J }\in {\varvec{\mathcal{F }}}\) and finite \(\Delta \subseteq \Gamma \). Thus,

$$\begin{aligned} {\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\ne \emptyset , \qquad \text{ for } \text{ every } {\varvec{\mathcal{E }}}\in \mathfrak{F }_\Gamma \text{. } \end{aligned}$$
(32)

Now define \(\mathfrak{F }_\Gamma ^\mathbf{y}:=\{{\varvec{\mathcal{D }}}\subseteq {\varvec{\mathcal{F }}}; {\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\subseteq {\varvec{\mathcal{D }}}\) for some \({\varvec{\mathcal{E }}}\in \mathfrak{F }_\Gamma \}\).

Claim 1 (a) \(\mathfrak{F }_\Gamma ^\mathbf{y}\) is a free filter, and (b) \(\mathfrak{F }_\Gamma \subseteq \mathfrak{F }_\Gamma ^\mathbf{y}\).

Proof

(a) \(\mathfrak{F }_\Gamma ^\mathbf{y}\) satisfies (F1) because \(\mathfrak{F }_\Gamma \) satisfies (F1) and (32). It satisfies (F2) by construction. To verify (F0), it suffices to show that \({\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\) is infinite for any \({\varvec{\mathcal{E }}}\in \mathfrak{F }_\Gamma \).

By contradiction, suppose that \({\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\) is finite. Each element of \({\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\) is a finite subset of \(\mathcal{I }\). Let \(\mathcal{F }:= \bigcup \{\mathcal{E }\); \(\mathcal{E }\in {\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\}\); then \(\mathcal{F }\) is also finite, hence, a proper subset of \(\mathcal{I }\). Let \(\mathcal{D }\subset \mathcal{I }\) be a nonempty finite subset disjoint from \(\mathcal{F }\); then no element of \({\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\) contains \(\mathcal{D }\). Thus, for any finite \(\Delta \subseteq \Gamma \), we have \({\varvec{\mathcal{F }}}_\Delta (\mathcal{D })\cap {\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}=\emptyset \). But \({\varvec{\mathcal{F }}}_\Delta (\mathcal{D })\cap {\varvec{\mathcal{E }}}\in \mathfrak{F }_\Gamma \) by (F2), because \({\varvec{\mathcal{F }}}_\Delta (\mathcal{D })\in \mathfrak{F }_\Gamma \) by definition. This contradicts (32). By contradiction, \({\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\) must be infinite.

(b) Let \({\varvec{\mathcal{E }}}\in \mathfrak{F }_\Gamma \). Then \({\varvec{\mathcal{E }}}\supseteq {\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\); hence \({\varvec{\mathcal{E }}}\in \mathfrak{F }_\Gamma ^\mathbf{y}\) by definition. \(\diamondsuit \text { Claim } 1\)

Claim 1(a) and the Ultrafilter Lemma says there exists a free ultrafilter \(\mathfrak{UF }^\mathbf{y}\) containing \(\mathfrak{F }^\mathbf{y}_\Gamma \). Claim 1(b) implies that \(\mathfrak{UF }^\mathbf{y}\) also contains \(\mathfrak{F }_\Gamma \), so it is \(\Gamma \)-admissible.

Claim 2 \(\mathbf{x}\, \,{}^{^*}\!\!\! {\stackrel{\displaystyle \prec }{{\scriptscriptstyle u}}}^{\!\mathfrak{UF }^\mathbf{y}}\, \mathbf{y}\).

Proof

Let \({\varvec{\mathcal{E }}}\in \mathfrak{F }_\Gamma \) be arbitrary. Then \({\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\in \mathfrak{F }_\Gamma ^\mathbf{y}\) by definition, and \({\varvec{\mathcal{E }}}\cap {\varvec{\mathcal{F }}}_\mathbf{y}\subseteq {\varvec{\mathcal{F }}}_\mathbf{y}\), so \({\varvec{\mathcal{F }}}_\mathbf{y}\in \mathfrak{F }_\Gamma ^\mathbf{y}\) by (F2), so \({\varvec{\mathcal{F }}}_\mathbf{y}\in \mathfrak{UF }^\mathbf{y}\), which means \(\mathbf{x}\, \,{}^{^*}\!\!\! {\stackrel{\displaystyle \prec }{{\scriptscriptstyle u}}}^{\!\!\mathfrak{UF }^\mathbf{y}}\, \mathbf{y}\).\(\diamondsuit \text { Claim } 2\)

Thus, it is false that \(\mathbf{x}\,{}^{^*}\! {\stackrel{\displaystyle \succcurlyeq }{{\scriptscriptstyle u}}}^{\!_{\mathfrak{UF }}}\,\mathbf{y}\) for every \(\Gamma \)-admissible ultrafilter \(\mathfrak{UF }\) on \({\varvec{\mathcal{F }}}\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Pivato, M. Additive representation of separable preferences over infinite products. Theory Decis 77, 31–83 (2014). https://doi.org/10.1007/s11238-013-9391-2

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11238-013-9391-2

Keywords

JEL Classification

Navigation