Skip to main content
Log in

Desirability foundations of robust rational decision making

  • S.I.: DecTheory&FutOfAI
  • Published:
Synthese Aims and scope Submit manuscript

Abstract

Recent work has formally linked the traditional axiomatisation of incomplete preferences à la Anscombe-Aumann with the theory of desirability developed in the context of imprecise probability, by showing in particular that they are the very same theory. The equivalence has been established under the constraint that the set of possible prizes is finite. In this paper, we relax such a constraint, thus de facto creating one of the most general theories of rationality and decision making available today. We provide the theory with a sound interpretation and with basic notions, and results, for the separation of beliefs and values, and for the case of complete preferences. Moreover, we discuss the role of conglomerability for the presented theory, arguing that it should be a rationality requirement under very broad conditions.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Such a correspondence follows from a separating hyperplane theorem.

  2. In the weak* topology, which is the smallest topology for which all the evaluation functionals given by \(f(P):=P(f)\), where \(f \in \mathcal {L}\), are continuous.

  3. A note of caution to prevent confusion in the reader: the adjective ‘strict’ denotes two unrelated things in desirability and in preferences. In preferences it characterises irreflexive relations, while in desirability it formalises an Archimedean condition as it will become clear in Sect. 3. We are keeping the same adjective in both cases for historical reasons and given that there should be no possibility to create ambiguity by doing so.

  4. In the following we shall just call it conglomerability for short.

  5. Our conditional horse lotteries should not be confused with Luce and Krantz’s (1971) conditional acts. In such a case acts are conditional on subsets of \(\varOmega \) and are defined with the aim to increase their theory’s expressiveness. Our horse lotteries p are conditional in the sense that \(p(\omega ,\cdot )\) intuitively represents a collection of conditional probabilities. This is just a technical point that makes our theory simpler and easier to extend to the case of infinitely many prizes compared to the case of horse lotteries.

  6. This is not the only way to reject the negative gambles: e.g., we could accept the zero gamble and directly require in an axiom that the negative gambles not be acceptable. However, we want to deal with strict preferences in this paper and for that we need irreflexivity, that is, D2.

  7. Strictly speaking, \(P_0\) should be defined on the subset of \(\mathcal {X}\)-measurable gambles of \(\mathcal {L}({\mathcal {W}}\times \mathcal {X})\). We neglect this detail to make things simpler.

  8. For a discussion of the rationality of this axiom and the connection with other, weaker, requirements, see (Zaffalon and Miranda 2013, Section 6.4).

  9. This is due to the involvement of conglomerability in the definition of independence; de Bock (2017) has recently shown that some other notions of independence continue to exist by dropping conglomerability.

  10. Given that \(\varOmega \) is possibly infinite, this entails an assumption of conglomerability; without such an assumption one could obtain uninformative preferences, see Miranda and Zaffalon (2015a, Section 5.1) for details.

  11. We refer to the work by Moral (2005, Section 2.4), de Cooman and Miranda (2012, Theorem 13) and de Bock and de Cooman (2015b, Theorem 8) for somewhat related results on finite referential spaces.

References

  • Allais, M. (1953). Le comportement de l’homme rationnel devant le risque: Critique des postulats et axiomes de l’école Américaine. Econometrica, 21(4), 503–546.

    Google Scholar 

  • Anscombe, F. J., & Aumann, R. J. (1963). A definition of subjective probability. The Annals of Mathematical Statistics, 34, 199–2005.

    Google Scholar 

  • Armstrong, T. (1990). Conglomerability of probability measures on Boolean algebras. Journal of Mathematical Analysis and Applications, 150(2), 335–358.

    Google Scholar 

  • Augustin, T., Coolen, F., de Cooman, G., & Troffaes, M. (Eds.). (2014). Introduction to imprecise probabilities. Hoboken: Wiley.

    Google Scholar 

  • Benavoli, A., Facchini, A., & Zaffalon, M. (2016a). Quantum mechanics: The Bayesian theory generalized to the space of Hermitian matrices. Physical Review A, 94, 042106.

    Google Scholar 

  • Benavoli, A., Facchini, A., & Zaffalon, M. (2016b). Quantum rational preferences and desirability. In Proceedings of The 1st International Workshop on “Imperfect Decision Makers: Admitting Real-World Rationality” (NIPS 2016), JMLR.org, PMLR (Vol. 58, pp. 87–96).

  • Berti, P., Miranda, E., & Rigo, P. (2017). Basic ideas underlying conglomerability and disintegrability. International Journal of Approximate Reasoning, 88, 387–400.

    Google Scholar 

  • Bewley, T. (2002). Knightian decision theory: Part I. Decisions in Economics and Finance, 25, 79–110. (reprint of Discussion Paper 807, Cowles Foundation, (1986)).

    Google Scholar 

  • Carnap, R. (1980). The problem of a more general concept of regularity (Vol. 2, pp. 145–155). Berkeley: University of Califormia Press.

    Google Scholar 

  • Cerreira-Vioglio, S., Maccheroni, F., & Marinacci, M. (2016). Stochastic dominance analysis without the independence axiom. Management Science, 63(4), 1097–1109.

    Google Scholar 

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

    Google Scholar 

  • Corani, G., & Zaffalon, M. (2008). Learning reliable classifiers from small or incomplete data sets: The naive credal classifier 2. Journal of Machine Learning Research, 9, 581–621.

    Google Scholar 

  • Couso, I., & Moral, S. (2011). Sets of desirable gambles: Conditioning, representation, and precise probabilities. International Journal of Approximate Reasoning, 52, 1034–1055.

    Google Scholar 

  • Cozman, F. G. (2012). Sets of probability distributions, independence, and convexity. Synthese, 186(2), 577–600.

    Google Scholar 

  • De Bock, J. (2017). Independent natural extension for sets of desirable gambles: Williams-coherence to the rescue. In A. Antonucci, G. Corani, I. Couso, & S. Destercke (Eds.), ISIPTA’17: Proceedings of the 10th International Symposium on Imprecise Probability: Theories and Applications (Vol. 62, pp. 121–132). PMLR.

  • De Bock, J., & de Cooman, G. (2014). An efficient algorithm for estimating state sequences in imprecise hidden Markov models. Journal of Artificial Intelligence Research, 50, 189–233.

    Google Scholar 

  • De Bock, J., & de Cooman, G. (2015a). Conditioning, updating and lower probability zero. International Journal of Approximate Reasoning, 67, 1–36.

    Google Scholar 

  • De Bock, J., & de Cooman, G. (2015b). Credal networks under epistemic irrelevance: The sets of desirable gambles approach. International Journal of Approximate Reasoning, 56, 178–207.

    Google Scholar 

  • De Cooman, G., & Hermans, F. (2008). Imprecise probability trees: Bridging two theories of imprecise probability. Artificial Intelligence, 172(11), 1400–1427.

    Google Scholar 

  • De Cooman, G., & Miranda, E. (2009). Forward irrelevance. Journal of Statistical Planning and Inference, 139(2), 256–276.

    Google Scholar 

  • De Cooman, G., & Miranda, E. (2012). Irrelevance and independence for sets of desirable gambles. Journal of Artificial Intelligence Research, 45, 601–640.

    Google Scholar 

  • De Cooman, G., & Quaeghebeur, E. (2012). Exchangeability and sets of desirable gambles. International Journal of Approximate Reasoning, 53, 363–395. (special issue in honour of Henry E. Kyburg, Jr.).

    Google Scholar 

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

    Google Scholar 

  • De Finetti, B. (1937). La prévision: ses lois logiques, ses sources subjectives. Annales de l’Institut Henri Poincaré 7:1–68, English translation in Kyburg Jr. and Smokler (1964).

  • De Finetti, B. (1970). Teoria delle Probabilità. Turin: Einaudi.

    Google Scholar 

  • De Finetti, B. (1974). Theory of probability: A critical introductory treatment (Vol. 1). Chichester: Wiley, English translation of de Finetti (1970).

  • Del Amo, A. G., & Ríos Insua, D. (2002). A note on an open problem in the foundations of statistics. RACSAM, 96(1), 55–61.

    Google Scholar 

  • DiBella, N. (2018). The qualitative paradox of non-conglomerability. Synthese, 195(3), 1181–1210.

    Google Scholar 

  • Dietterich, T. G. (2017). Steps toward robust artificial intelligence. AI Magazine, 38(3), 3–24.

    Google Scholar 

  • Domshlak, C., Hüllermeier, E., Kaci, S., & Prade, H. (2011). Preferences in AI: An overview. Artificial Intelligence, 175(7–8), 1037–1052.

    Google Scholar 

  • Dubins, L. E. (1975). Finitely additive conditional probabilities, conglomerability and disintegrations. The Annals of Probability, 3, 88–99.

    Google Scholar 

  • Dubois, D., Fargier, H., & Perny, P. (2003). Qualitative decision theory with preference relations and comparative uncertainty: An axiomatic approach. Artificial Intelligence, 148(1–2), 219–260.

    Google Scholar 

  • Dubra, J., Maccheroni, F., & Ok, E. (2004). Expected utility theory without the completeness axiom. Journal of Economic Theory, 115, 118–133.

    Google Scholar 

  • Galaabaatar, T., & Karni, E. (2013). Subjective expected utility with incomplete preferences. Econometrica, 81(1), 255–284.

    Google Scholar 

  • Ghirardato, P., & Siniscalchi, M. (2012). Ambiguity in the small and in the large. Econometrica, 80(6), 2827–2847.

    Google Scholar 

  • Hacking, I. (1967). Slightly more realistic personal probability. Philosophy of Science, 34(4), 311–325.

    Google Scholar 

  • Hájek, A., & Smithson, M. (2012). Rationality and indeterminate probabilities. Synthese, 187(1), 33–48.

    Google Scholar 

  • Hill, B., & Lane, D. (1985). Conglomerability and countable additivity. Sankhya, 47, 366–379.

    Google Scholar 

  • Howson, C. (2009). Can logic be combined with probability? Probably. Journal of Applied Logic, 7, 177–187.

    Google Scholar 

  • Howson, C. (2011). Bayesianism as a pure logic of inference. In P. S. Bandyopadhyay & M. R. Forster (Eds.), Handbook of the philosophy of science, vol 7: Philosophy of statistics (pp. 441–471). Amsterdam: Elsevier.

    Google Scholar 

  • Kikuti, D., Cozman, F. G., & Filho, R. S. (2011). Sequential decision making with partially ordered preferences. Artificial Intelligence, 175(7–8), 1346–1365.

    Google Scholar 

  • Kyburg, H.E. Jr., & Smokler, H.E. (Eds.) (1964) Studies in subjective probability (2 Ed.). New York: Wiley (with new material) (1980).

  • Luce, R. D., & Krantz, D. H. (1971). Conditional expected utility. Econometrica, 39(2), 253–271.

    Google Scholar 

  • Machina, M. (1982). “Expected utility” analysis without the independence axiom. Econometrica, 50(2), 277–323.

    Google Scholar 

  • Miranda, E. (2008). A survey of the theory of coherent lower previsions. International Journal of Approximate Reasoning, 48(2), 628–658.

    Google Scholar 

  • Miranda, E., & de Cooman, G. (2007). Marginal extension in the theory of coherent lower previsions. International Journal of Approximate Reasoning, 46(1), 188–225.

    Google Scholar 

  • Miranda, E., & Zaffalon, M. (2010). Notes on desirability and conditional lower previsions. Annals of Mathematics and Artificial Intelligence, 60(3–4), 251–309.

    Google Scholar 

  • Miranda, E., & Zaffalon, M. (2015a). Independent products in infinite spaces. Journal of Mathematical Analysis and Applications, 425(1), 460–488.

    Google Scholar 

  • Miranda, E., & Zaffalon, M. (2015b). On the problem of computing the conglomerable natual extension. International Journal of Approximate Reasoning, 56, 1–27.

    Google Scholar 

  • Miranda, E., & Zaffalon, M. (2016). Conformity and independence with coherent lower previsions. International Journal of Approximate Reasoning, 78, 125–137.

    Google Scholar 

  • Miranda, E., Zaffalon, M., & de Cooman, G. (2012). Conglomerable natural extension. International Journal of Approximate Reasoning, 53(8), 1200–1227.

    Google Scholar 

  • Moral, S. (2005). Epistemic irrelevance on sets of desirable gambles. Annals of Mathematics and Artificial Intelligence, 45, 197–214.

    Google Scholar 

  • Nau, R. (2006). The shape of incomplete preferences. The Annals of Statistics, 34(5), 2430–2448.

    Google Scholar 

  • Nau, R. (2011). Risk, ambiguity, and state-preference theory. Economic Theory, 48(2–3), 437–467.

    Google Scholar 

  • Ok, E., Ortoleva, P., & Riella, G. (2012). Incomplete preferences under uncertainty: Indecisiveness in beliefs versus tastes. Econometrica, 80(4), 1791–1808.

    Google Scholar 

  • Pedersen, A. P. (2014). Comparative expectations. Studia Logica, 102(4), 811–848.

    Google Scholar 

  • Pigozzi, G., Tsoukiàs, A., & Viappiani, P. (2016). Preferences in artificial intelligence. Annals of Mathematics and Artificial Intelligence, 77(3–4), 361–401.

    Google Scholar 

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

    Google Scholar 

  • Seidenfeld, T., Schervish, M. J., & Kadane, J. B. (1995). A representation of partially ordered preferences. The Annals of Statistics, 23, 2168–2217. (reprinted in Seidenfeld et al. (1999), (pp. 69–129)).

    Google Scholar 

  • Seidenfeld, T., Schervisch, M. J., & Kadane, J. B. (1998). Non-conglomerability for finite-valued finitely additive probability. Sankhya, 60(3), 476–491.

    Google Scholar 

  • Seidenfeld, T., Schervish, M. J., & Kadane, J. B. (1999). Rethinking the foundations of statistics. Cambridge: Cambridge University Press.

    Google Scholar 

  • Seidenfeld, T., Schervisch, M. J., & Kadane, J. B. (2010). Coherent choice functions under uncertainty. Synthese, 172(1), 157–176.

    Google Scholar 

  • Shafer, G., Gillett, P. R., & Scherl, R. B. (2003). A new understanding of subjective probability and its generalization to lower and upper prevision. International Journal of Approximate Reasoning, 33, 1–49.

    Google Scholar 

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

    Google Scholar 

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

    Google Scholar 

  • Tversky, A., & Kahneman, D. (1992). Advances in prospect theory: Cumulative representation of uncertainty. Journal of Risk and Uncertainty, 5(4), 297–323.

    Google Scholar 

  • Van Camp, A. (2018). Choice functions as a tool to model uncertainty. Ph.D. thesis, University of Ghent.

  • Van Camp, A., de Cooman, G., Miranda, E., & Quaeghebeur, E. (2018a). Coherent choice functions, desirability and indifference. Fuzzy Sets and Systems, 341(C), 1–36.

    Google Scholar 

  • Van Camp, A., Miranda, E., & de Cooman, G. (2018b). Lexicographic choice functions. International Journal of Approximate Reasoning, 92, 97–119.

    Google Scholar 

  • Von Neumann, J., & Morgestern, O. (1947). Theory of games and economic behaviour. Princeton: Princeton University Press.

    Google Scholar 

  • Walley, P. (1991). Statistical reasoning with imprecise probabilities. London: Chapman and Hall.

    Google Scholar 

  • Williams, P. M. (1975). Notes on conditional previsions. Tech. rep., School of Mathematical and Physical Science, University of Sussex, UK, reprinted in Williams (2007).

  • Williams, P. M. (2007). Notes on conditional previsions. International Journal of Approximate Reasoning, 44, 366–383. (revised journal version of Williams (1975)).

    Google Scholar 

  • Zaffalon, M., & Miranda, E. (2013). Probability and time. Artificial Intelligence, 198(1), 1–51.

    Google Scholar 

  • Zaffalon, M., & Miranda, E. (2015). Desirability and the birth of incomplete preferences. CoRR abs/1506.00529. arxiv:1506.00529.

  • Zaffalon, M., & Miranda, E. (2017). Axiomatising incomplete preferences through sets of desirable gambles. Journal of Artificial Intelligence Research, 60, 1057–1126.

    Google Scholar 

Download references

Acknowledgements

The authors are grateful to the anonymous referees for a careful reading of the paper and in particular to Referee 1 for having spotted some technical problems that are now fixed.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marco Zaffalon.

Additional information

Work partially supported by the Swiss NSF Grant No. 200021_146606 / 2 and by Project TIN2014-59543-P.

Appendices

A Decomposition and completeness of preferences

We now turn to the formalisms of desirability and coherent lower previsions to briefly show how they can represent and generalise well-known concepts in rational preferences. The following discussion is a summary of the one we presented in Zaffalon and Miranda (2017, Section 5), with some additional remarks and proofs for the case of infinitely many values.

1.1 A.1 In terms of sets of desirable gambles

We begin by considering the case where our assessments are modelled by sets of desirable gambles.

1.1.1 A.1.1 State independence

State independence is the condition that allows us to have separate models for beliefs and values.

We showed in Zaffalon and Miranda (2017, Section 5.3) that the traditional notion of state independence in the literature of preferences is equivalent to the notion of ‘strong independence’ in imprecise probability. In the case of totally ordered Archimedean preferences (i.e., precise probability and utility models), this means that probability and utility are stochastically independent models, when utility is mathematically represented through a probability function (like we do in this paper, see Sect. 3.4).

Such a notion of independence extends over multiple probability-utility pairs, and even to sets of desirable gambles. However, it ceases to exist when both \(\varOmega \) and \(\mathcal {X}\) are infinite (Miranda and Zaffalon 2015a, Example 1).Footnote 9 This holds also for most other notions of independence while, interestingly, that need not be the case for notions of ‘irrelevance’: these are asymmetric notions where mutual (symmetric) independence is replaced by independence in only one direction. In the case of this paper, for instance, this means assuming that values are independent of states but not vice versa; we say that states are irrelevant to values.

This makes of irrelevance a very natural candidate for a notion of state independence that exists no matter the cardinalities of the spaces involved. But there is more to it, as it is arguable that irrelevance qualifies itself as the right notion to use in the present context, unlike independence. The reason is that while it makes sense that states may be irrelevant to values, the opposite seems to be questionable: in fact, as we have already remarked, it appears meaningless to update a model on elements of \(\mathcal {X}\), which do not occur; as a consequence the irrelevance of values to states appears meaningless too.

Let us then consider the following definition of irrelevance for desirability:

Definition 29

(\(\varOmega \)-\(\mathcal {X}\) irrelevant product for gambles) A coherent set of gambles \(\mathcal {R}\) on \(\varOmega \times \mathcal {X}\) is called an \(\varOmega \)-\(\mathcal {X}\) irrelevant product of its marginal sets of gambles \(\mathcal {R}_{\varOmega }',\mathcal {R}_{\mathcal {X}}'\) if it includes the set

$$\begin{aligned} \mathcal {R}|\varOmega :=\{h\in \mathcal {L}(\varOmega \times \mathcal {X}): (\forall \omega \in \varOmega )\ h(\omega ,\cdot )\in \mathcal {R}_{\mathcal {X}}'\cup \{0\}\}{\setminus }\{0\}. \end{aligned}$$

We have already considered these products, in another context and for lower previsions, in Miranda and Zaffalon (2015a) (see also the work by de Cooman and Miranda 2009; de Bock and de Cooman 2014, 2015b).

The rationale of the definition is the following. First, the requirement that \(h(\omega ,\cdot )\in \mathcal {R}_\mathcal {X}'\cup \{0\}\) is there so that the inferences conditional on \(\{\omega \}\) encompassed by \(\mathcal {R}|\varOmega \) should yield the marginal set \(\mathcal {R}_\mathcal {X}'\). This is just a way to formally state that \(\omega \) is irrelevant to \(\mathcal {X}\). The same is repeated for every \(\omega \), so that \(\mathcal {R}|\varOmega \) can be regarded as being born out of aggregating all the irrelevant conditional sets.Footnote 10 Finally, that \(\mathcal {R}\) contains \(\mathcal {R}|\varOmega \) is imposed to make sure that \(\mathcal {R}\) is a model coherent with the irrelevance of beliefs on values.

If \(\mathcal {R}\) satisfies Definition 29 we say it models state-independent preferences. The least-committal among these models is given byFootnote 11

Proposition 5

Given two marginal coherent sets of desirable gambles \(\mathcal {R}_{\varOmega },\mathcal {R}_{\mathcal {X}}\) defined on the subsets of \(\mathcal {L}(\varOmega \times \mathcal {X})\) given by the \(\varOmega \)-measurable and \(\mathcal {X}\)-measurable gambles, respectively, their smallest \(\varOmega \)-\(\mathcal {X}\) irrelevant product is given by the marginal extension of \(\mathcal {R}_{\varOmega }\) and \(\mathcal {R}|{\{\omega \}}:=\mathcal {R}_{\mathcal {X}}\) for all \(\omega \in \varOmega \). Namely,

$$\begin{aligned} \hat{\mathcal {R}}:=\{g+h: g\in \mathcal {R}_{\varOmega }\cup \{0\}, h\in \mathcal {R}|\varOmega \cup \{0\}\}{\setminus }\{0\}. \end{aligned}$$
(12)

Definition 29 and Eq. (12) represent our main proposal to model state-independent preferences.

1.1.2 A.1.2 State dependence

We say that a set of desirable gambles \(\mathcal {R}\) models state-dependent preferences when it does not satisfy the conditions in Definition 29, i.e., when it does not include the set \(\mathcal {R}|\varOmega \) derived from its marginal \(\mathcal {R}_\mathcal {X}\) and the assumption of \(\varOmega \)-\(\mathcal {X}\) irrelevance.

1.1.3 A.1.3 Completeness

So far we have discussed the general case of coherent sets of gambles \(\mathcal {R}\) without discussing in particular the so-called complete or maximal coherent sets for which it holds that \(f\notin \mathcal {R}\Rightarrow -f\in \mathcal {R}\) for all \(f\in \mathcal {L},f\ne 0\). Given the equivalence between coherent sets of gambles in \(\mathcal {L}(\varOmega \times \mathcal {X})\) and coherent preferences in \(\mathcal {H}\times \mathcal {H}\), discussing the case of maximal sets amounts to discussing the case of complete preferences.

Definition 30

(Completeness of beliefs and values for gambles) A coherent set of desirable gambles \(\mathcal {R}\subseteq \mathcal {L}(\varOmega \times \mathcal {X})\) is said to represent complete beliefs if \(\mathcal {R}_{\varOmega }\) is maximal. It is said to represent complete values if \(\mathcal {R}_{\mathcal {X}}\) is maximal. Finally, if \(\mathcal {R}\) is maximal, then it is said to represent complete preferences.

The rationale of this definition is straightforward: when we say, for instance, that \(\mathcal {R}\) represents complete beliefs, we mean that there is never indecision between two options that are concerned only with \(\varOmega \). The situation is analogous in the case of complete values. Finally, in the case of complete preferences, the definition is just the direct application of the definition of maximality.

1.2 A.2 In terms of lower previsions

We now consider the case where preferences are modelled by means of a coherent lower prevision \({\underline{P}}\) on \(\mathcal {L}(\varOmega \times \mathcal {X})\).

1.2.1 A.2.1 State independent preferences

We proceed as in the case of desirability by focusing on irrelevance of states to values.

Definition 31

(\(\varOmega \)-\(\mathcal {X}\) irrelevant product for lower previsions) Given marginal coherent lower previsions \({\underline{P}}_{\varOmega }',{\underline{P}}_{\mathcal {X}}'\), let

$$\begin{aligned} {\underline{P}}(f|\{\omega \}):={\underline{P}}_\mathcal {X}'(f(\omega ,\cdot )), \end{aligned}$$

for all \(f\in \mathcal {L}(\varOmega \times \mathcal {X})\), and

$$\begin{aligned} {\underline{P}}(\cdot |\varOmega ):=\sum _{\omega \in \varOmega }I_{\{\omega \}}{\underline{P}}(\cdot |\{\omega \}). \end{aligned}$$

Then a coherent lower prevision \({\underline{P}}\) on \(\mathcal {L}(\varOmega \times \mathcal {X})\) is called an \(\varOmega \)-\(\mathcal {X}\) irrelevant product of \({\underline{P}}_{\varOmega }',{\underline{P}}_{\mathcal {X}}'\) if

$$\begin{aligned} {\underline{P}}\ge {\underline{P}}_\varOmega '({\underline{P}}(\cdot |\varOmega )). \end{aligned}$$

In this case we also say that \({\underline{P}}\) models state-independent preferences.

Here \({\underline{P}}(\cdot |\varOmega )\) plays the role that \(\mathcal {R}|\varOmega \) took in Section A.1, that is, conglomerating all the conditional information. The concatenation \({\underline{P}}_\varOmega '({\underline{P}}(\cdot |\varOmega ))\) is a marginal extension: in particular, it is the least-committal coherent \(\varOmega \)-conglomerable model built out of the given marginals and the assessment of irrelevance that defines \({\underline{P}}(\cdot |\varOmega )\) through \({\underline{P}}_\mathcal {X}'\). Every coherent lower prevision that dominates \({\underline{P}}_\varOmega '({\underline{P}}(\cdot |\varOmega ))\) is compatible with the irrelevance assessment but is also more informative than that:

Proposition 6

Given two marginal coherent lower previsions \({\underline{P}}_{\varOmega }',{\underline{P}}_{\mathcal {X}}'\), their smallest \(\varOmega \)-\(\mathcal {X}\) irrelevant product is given by the marginal extension

$$\begin{aligned} {\underline{P}}_\varOmega '({\underline{P}}(\cdot |\varOmega )). \end{aligned}$$

We refer to the work of Walley (1991, Section 6.7), Miranda and de Cooman (2007), de Cooman and Miranda (2009) and de Cooman and Hermans (2008) for additional information on the marginal extension and its use in a number of different contexts.

1.2.2 A.2.2 State dependent preferences

As in the case of desirable gambles, we define state dependence as the lack of state independence: this means that if we consider a coherent lower prevision \({\underline{P}}\) with marginals \({\underline{P}}_{\varOmega }',{\underline{P}}_{\mathcal {X}}'\), it is said to model state-dependent preferences when it does not dominate the concatenation \({\underline{P}}_{\varOmega }'({\underline{P}}(\cdot |\varOmega ))\), where \({\underline{P}}(\cdot |\varOmega )\) is derived from \({\underline{P}}_{\mathcal {X}}'\) by an assumption of epistemic irrelevance.

1.2.3 A.2.3 Completeness

The definition of completeness for lower previsions is a rephrasing of that for desirable gambles:

Definition 32

(Completeness of beliefs and values for lower previsions) A coherent lower prevision \({\underline{P}}\) on \(\mathcal {L}(\varOmega \times \mathcal {X})\) is said to represent complete beliefs if its marginal \({\underline{P}}_{\varOmega }\) is linear. It is said to represent complete values if its marginal \({\underline{P}}_{\mathcal {X}}\) is linear. Finally, if \({\underline{P}}\) is linear, then it is said to represent complete preferences.

There are a number of equivalent ways in which we can characterise the linearity of previsions in terms of the corresponding set of desirable gambles:

Proposition 7

Let \({\underline{P}}\) be a coherent lower prevision on \(\mathcal {L}\) and \(\mathcal {R}\) its corresponding coherent set of strictly desirable gambles. The following are equivalent:

  1. (i)

    \({\underline{P}}\) is a linear prevision.

  2. (ii)

    If \(f\notin \mathcal {R}\) then \(\varepsilon -f\in \mathcal {R}\) for all \(\varepsilon >0\).

  3. (iii)

    \(\mathcal {R}\) is negatively additive, meaning that \(f,g\notin \mathcal {R}\Rightarrow (\forall \varepsilon >0)\ f+g-\varepsilon \notin \mathcal {R}\).

We can then focus on any of those formulations to immediately deduce characterisations for all the cases of completeness in preferences:

Proposition 8

Consider a coherent preference relation \(\succ \) on gambles represented by a coherent lower prevision \({\underline{P}}\) on \(\mathcal {L}(\varOmega \times \mathcal {X})\), whose corresponding coherent set of gambles is denoted by \(\mathcal {R}\). Let \(\mathcal {R}_\varOmega ,\mathcal {R}_\mathcal {X}\) denote its marginals. Then:

  1. (i)

    \({\underline{P}}\) represents complete beliefs \(\Leftrightarrow \) \(\mathcal {R}_\varOmega \) is negatively additive.

  2. (ii)

    \({\underline{P}}\) represents complete values \(\Leftrightarrow \) \(\mathcal {R}_\mathcal {X}\) is negatively additive.

  3. (iii)

    \({\underline{P}}\) represents complete preferences \(\Leftrightarrow \) \(\mathcal {R}\) is negatively additive.

B Proofs

Proof of Proposition 1

Using the mixture-independence axiom A4 together with the assumption,

$$\begin{aligned} p\succ q\Leftrightarrow \alpha p+(1-\alpha )r\succ \alpha q+(1-\alpha )r=\alpha p+(1-\alpha )s\Leftrightarrow r\succ s. \end{aligned}$$

Proof of Proposition 2

By \(p\gneq q\) we deduce that \(0\lneq p-q\le 1\). Then \((p-q)\in \mathcal {H}\) and \((p-q)\succ 0\) by A1. And since

$$\begin{aligned} \frac{1}{2}(p-q)=\frac{1}{2}[(p-q)-0], \end{aligned}$$

we deduce by Proposition 1 that \(p\succ q\).

Proof of Lemma 1

Consider a set of desirable gambles \(\mathcal {D}\subseteq \mathcal {L}_1(\varOmega \times \mathcal {X})\) and define the linear relation \(\succ \subseteq \mathcal {H}\times \mathcal {H}\) by \(p\succ q\) if and only if \(p-q\in \mathcal {D}\).

Observe also that for any \(f\in \mathcal {D}\) we can always find \(p,q\in \mathcal {H}\) such that \(f=p-q\): it is sufficient to let \(p:=\max (f,0)\) and \(q:=\min (0,f)\).

Let us now prove injectivity. Assume by contradiction that there are sets \(\mathcal {D}_1\ne \mathcal {D}_2\) in \(\mathcal {L}_1\) such that the respective linear relations are equal: \(\succ _1=\succ _2\). Since the sets are different, then we can assume without loss of generality that there is \(f_1\in \mathcal {D}_1{\setminus }\mathcal {D}_2\). At the same time for all \(p,q\in \mathcal {H}\) with \(p-q=f_1\), we have that \(p\succ _1q\) and \(p\succ _2 q\). This implies that there is \(f_2\in \mathcal {D}_2\) such that \(f_2=p-q\). But then \(f_1=f_2\), a contradiction.

Let us focus on surjectivity. Consider a linear relation \(\succ \subseteq \mathcal {H}\times \mathcal {H}\) and let \(\mathcal {D}':=\{f\in \mathcal {L}_1(\varOmega \times \mathcal {X}):(\exists p,q\in \mathcal {H})\ f=p-q,p\succ q\}\). The relation it induces is given by \(p\succ 'q\Leftrightarrow p-q\in \mathcal {D}'\). Now, if \(p\succ q\), then \(p-q\in \mathcal {D}'\), whence \(p\succ 'q\). Conversely, if \(p\succ 'q\), then \(p-q\in \mathcal {D}'\); as a consequence, there are \(p',q'\in \mathcal {H}\), \(p'\succ q'\) such that \(p-q=p'-q'\). But relation \(\succ \) is linear, whence \(p\succ q\). It follows that \(\succ =\succ '\).

Proof of Theorem 1

Consider a set of desirable gambles \(\mathcal {D}\subseteq \mathcal {L}_1\) and its corresponding linear relation \(\succ \). They are related through

$$\begin{aligned}&\mathcal {D}=\{f\in \mathcal {L}_1(\varOmega \times \mathcal {X}):(\exists p,q\in \mathcal {H})\ f=p-q,p\succ q\}, \\&p\succ q\Leftrightarrow p-q\in \mathcal {D}. \end{aligned}$$
(\(\hbox {A1}\Leftrightarrow \hbox {D1}'\)):

If \(p\succ 0\) for all \(p\in \mathcal {H},p\ne 0\), then \(p\in \mathcal {D}\); and since \(\mathcal {L}_1^+=\mathcal {H}{\setminus }\{0\}\), then we have the thesis. Conversely, if \(\mathcal {L}_1^+\subseteq \mathcal {D}\), then \(p\in \mathcal {D}\) for all \(p\in \mathcal {H}{\setminus }\{0\}\); as a consequence \(p\succ 0\) for all \(p\in \mathcal {H},p\ne 0\).

(\(\hbox {A2}\Leftrightarrow \hbox {D2}'\)):

By contradiction, if \(p\succ p\) for some \(p\in \mathcal {H}\), then \(0\in \mathcal {D}\); and if \(0\in \mathcal {D}\), then for some \(p\in \mathcal {D}\), \(p\succ p\).

(\(\hbox {A3}\Leftrightarrow \hbox {D3}'\)):

Assume that the relation is transitive and take \(f,g\in \mathcal {D}\) such that \(f=p-q,g=q-r\) for some \(p,q,r\in \mathcal {H}\). Then we have that \(p\succ q\) and \(q\succ r\); by transitivity we conclude that \(p\succ r\), whence \(f+g\in \mathcal {D}\).

Conversely, assume that \(\mathcal {D}\) is additive and take \(p\succ q\), \(q\succ r\). Letting \(f:=p-q,g:=q-r\), we get by additivity that \(f+g=p-r\in \mathcal {D}\), so that \(p\succ r\).

(\(\hbox {A4}\Leftrightarrow \hbox {D4'}\)):

Assume that \(\succ \) satisfies mixture independence and take \(f\in \mathcal {D}\) and \(\lambda >0\) such that \(\lambda f\in \mathcal {L}_1\). Then there are \(p\succ q\) such that \(f=p-q\). Moreover, there must be \(r,s\in \mathcal {H}\) such that \(\lambda f=r-s\) given that \(\lambda f\le 1\) (it is enough as usual to take the positive and negative parts of \(\lambda f\), respectively). Using Proposition 1, we get that \(r\succ s\), whence \(\lambda f\in \mathcal {D}\).

Conversely, assume that \(f\in \mathcal {D}\) implies that \(\lambda f\in \mathcal {D}\) for all \(\lambda >0\) such that \(\lambda f\in \mathcal {L}_1\). Taking \(p\succ q\), we have then that \(\lambda p+(1-\lambda )r\succ \lambda q+(1-\lambda )r\) for all \(p,q,r\in \mathcal {H}\) and \(\lambda \in (0,1]\). On the other hand, if \(\lambda p+(1-\lambda )r\succ \lambda q+(1-\lambda )r\) for all \(p,q,r\in \mathcal {H}\) and \(\lambda \in (0,1]\), then trivially \(p\succ q\).

(\(\hbox {A0}\Leftrightarrow \hbox {D0'}\)):

Assume that for all \(p\succ q\), with \(p\gneq q\) not holding, there is \(\alpha \in (0,1)\) such that \(\alpha p\succ q\). We want to show that for all \(f\in \mathcal {D}{\setminus }\mathcal {L}_1^+\) and all \(g\in \mathcal {L}_1\) such that \(\max (0,-f)\le g\le \min (1,1-f)\), there is \(\alpha \in (0,1)\) such that \(\alpha f-(1-\alpha )g\in \mathcal {D}\).

Let \(q:=g\) and \(p:=f+q\), so that \(f=p-q\). Let us show that \(q\in \mathcal {H}\); it is enough to prove that \(\max (0,-f)\le \min (1,1-f)\). In case \(f(\omega ,x)\ge 0\), the inequality reduces itself to \(0\le 1-f(\omega ,x)\), which is true since \(f(\omega ,x)\le 1\). In case \(f(\omega ,x)<0\), the inequality becomes \(-f(\omega ,x)\le 1\), which is again true since \(f(\omega ,x)\ge -1\).

Let us similarly show that \(p=f+q\in \mathcal {H}\) too. We know that \(f+\max (0,-f)\le p\le f+\min (1,1-f)\). Again, if \(f(\omega ,x)\ge 0\), we see that \(0\le f(\omega ,x)\le p(\omega ,x)\le 1\); if \(f(\omega ,x)<0\), we have that \(0\le p(\omega ,x)\le f+1\le 1\).

At this point we can apply the hypothesis, recalling that \(f\notin \mathcal {L}_1^+\) and hence that \(p\gneq q\) does not hold, and so we deduce that there is \(\alpha \in (0,1)\) such that \(\alpha p-q=\alpha (p-q)-(1-\alpha )q=\alpha f-(1-\alpha )q=\alpha f-(1-\alpha )g\in \mathcal {D}\).

Conversely, suppose that for all \(f\in \mathcal {D}{\setminus }\mathcal {L}_1^+\) and all \(g\in \mathcal {L}_1\) such that \(\max (0,-f)\le g\le \min (1,1-f)\), there is \(\alpha \in (0,1)\) such that \(\alpha f-(1-\alpha )g\in \mathcal {D}\). Let us show that \(p\succ q\), with \(p\gneq q\) not holding, implies that there is \(\alpha \in (0,1)\) such that \(\alpha p\succ q\).

Let \(f:=p-q\), from which it follows that \(f\notin \mathcal {L}_1^+\), and \(g:=q\). Observe that if \(\max (0,-f)\le g\le \min (1,1-f)\) holds, then from the assumption the thesis follows immediately. In other words, we have to show that \(\max (0,q-p)\le q\le \min (1,q+(1-p))\). This is immediate since on the left-hand side we decrease q by some non-negative amount while on the right-hand side we are increasing it by a non-negative amount.

Proof of Lemma 2

That D3\(''\) implies D3\('\) is trivial. Therefore let us assume that \(\mathcal {D}\) satisfies D3\('\) and prove that D3\(''\) holds.

Consider \(f,g\in \mathcal {D}\) such that \(f+g\in \mathcal {L}_1\). By D4\('\) we have that \(\frac{1}{2}f,\frac{1}{2}g\in \mathcal {D}\). We can always write \(f=p-q,g=r-s\) for some \(p,q,r,s\in \mathcal {H}\). It follows that

$$\begin{aligned} \left( \frac{1}{2}p+\frac{1}{2}r\right) -\left( \frac{1}{2}q+\frac{1}{2}r\right) \in \mathcal {D},\\ \left( \frac{1}{2}q+\frac{1}{2}r\right) -\left( \frac{1}{2}q+\frac{1}{2}s\right) \in \mathcal {D}. \end{aligned}$$

Using D3\('\) we obtain that \((\frac{1}{2}p+\frac{1}{2}r)-(\frac{1}{2}q+\frac{1}{2}s)=\frac{1}{2}f+\frac{1}{2}g\in \mathcal {D}\). Using D4\('\) again, we have the thesis.

Proof of Theorem 2

Let us prove the five points of the theorem.

  1. 1.

    Assume that \(\mathcal {D}\) satisfies D1\('\)–D4\('\) and show that \(\mathcal {R}_\mathcal {D}\) is coherent.

    1. D1.

      Consider \(f\in \mathcal {L}^+\) and let \(g:=f/\sup |f|\), so that \(g\in \mathcal {L}_1^+\). As a consequence of D1\('\), \(g\in \mathcal {D}\), whence \(f\in \mathcal {R}_\mathcal {D}\).

    2. D2.

      Trivial.

    3. D3.

      Take \(f,g\in \mathcal {R}_\mathcal {D}\), so that there are \(f',g'\in \mathcal {D},\lambda ',\lambda ''>0\) with \(f=\lambda 'f',g=\lambda ''g'\). If we let \(\lambda :=\lambda '+\lambda ''\), we have that \(f/(2\lambda ),g/(2\lambda )\in \mathcal {D}\) by D4\('\) and at the same time \((f+g)/(2\lambda )\in \mathcal {L}_1\), so that \((f+g)/(2\lambda )\in \mathcal {D}\) by D3\(''\). As a consequence, \(f+g\in \mathcal {R}_\mathcal {D}\).

    4. D4.

      Trivial.

  2. 2.

    That \(\mathcal {D}_\mathcal {R}\) satisfies D1\('\)–D4\('\) once \(\mathcal {R}\) is coherent is straightforward.

  3. 3.

    We need to show that, under coherence of \(\mathcal {R},\mathcal {D}\), it holds that \(\mathcal {R}_{\mathcal {D}_{\mathcal {R}}}=\mathcal {R}\) and \(\mathcal {D}_{\mathcal {R}_{\mathcal {D}}}=\mathcal {D}\). For the first equality, that \(\mathcal {R}_{\mathcal {D}_\mathcal {R}}\supseteq \mathcal {R}\) follows from

    $$\begin{aligned} \mathcal {R}_{\mathcal {D}_\mathcal {R}}=\{\lambda g: \lambda>0,g\in \mathcal {D}_\mathcal {R}\}=\left\{ \lambda \frac{g}{\sup |g|}: \lambda >0,g\in \mathcal {R}\right\} \supseteq \mathcal {R}\end{aligned}$$

    by taking \(\lambda :=\sup |g|\).

    To show that \(\mathcal {R}_{\mathcal {D}_\mathcal {R}}\subseteq \mathcal {R}\), consider any h in \(\mathcal {R}\), implying that \(\frac{h}{\sup |h|}\in \mathcal {D}_{\mathcal {R}}\). Note that \(\sup |h| > 0\); indeed, if \(\sup |h|\) would be 0 then \(h=0\). For any given \(\lambda >0\) , this implies that \(\lambda \frac{h}{\sup |h|}=\lambda ' h\) for \(\lambda ':=\frac{\lambda }{\sup |h|}>0\) indeed belongs to \(\mathcal {R}\), because this set satisfies D4.

    Let us prove now the equality \(\mathcal {D}_{\mathcal {R}_{\mathcal {D}}}=\mathcal {D}\). Take \(\mathcal {D}\) satisfying D1\('\)–D4\('\). To see that \(\mathcal {D}_{\mathcal {R}_{\mathcal {D}}} \subseteq \mathcal {D}\), take \(f\in \mathcal {D}_{\mathcal {R}_{\mathcal {D}}}\). Then there is some \(g\in \mathcal {R}_\mathcal {D}\) such that \(f=\frac{g}{\sup |g|}\). Since \(g\in \mathcal {R}_\mathcal {D}\), there is some \(h\in \mathcal {D}\) and some \(\lambda >0\) such that \(g=\lambda h\), whence \(f=\frac{\lambda h}{\sup |\lambda h|}=\frac{h}{\sup |h|}\). Since \(f\in \mathcal {L}_1\), we deduce from D4\('\) that \(f\in \mathcal {D}\).

    Conversely, take \(f\in \mathcal {D}\). Then \(f\in \mathcal {R}_\mathcal {D}\) and as a consequence \(\frac{f}{\sup |f|}\in \mathcal {D}_{\mathcal {R}_\mathcal {D}}\). Since \(\mathcal {D}_{\mathcal {R}_\mathcal {D}}\) satisfies D4\('\) and \(f\in \mathcal {L}_1\), we deduce that \(f\in \mathcal {D}_{\mathcal {R}_\mathcal {D}}\).

  4. 4.

    Since \(\mathcal {D}\subseteq \mathcal {R}_\mathcal {D}\) and \(\mathcal {R}_\mathcal {D}\) is coherent (apply 1), then \(\mathcal {D}\) avoids partial loss by (Miranda and Zaffalon 2010, Proposition 3(e)). This implies that the minimal coherent extension of \(\mathcal {D}\) exists. Let us call it \(\mathcal {E}\). That \(\mathcal {R}_\mathcal {D}\subseteq \mathcal {E}\) is trivial. Since \(\mathcal {R}_\mathcal {D}\) is coherent, then by definition of \(\mathcal {E}\), it must be that \(\mathcal {R}_\mathcal {D}\supseteq \mathcal {E}\). As a consequence \(\mathcal {R}_\mathcal {D}\) is the natural extension.

    Let us show that \(\mathcal {R}_\mathcal {D}\cap \mathcal {L}_1\subseteq \mathcal {D}\) (the converse inclusion is trivial). Consider \(f\in \mathcal {R}_\mathcal {D}\cap \mathcal {L}_1\). Then \(f=\lambda g\) for some \(\lambda >0,g\in \mathcal {D}\). And since \(\lambda g\in \mathcal {L}_1\), we get by D4\('\) that \(f=\lambda g\in \mathcal {D}\). This shows that \(\mathcal {D}\) is coherent relative to \(\mathcal {L}_1\).

  5. 5.

    (\(\Rightarrow \)) Showing that \(\mathcal {D}_\mathcal {R}\) satisfies D0\('\) is equivalent to showing that the preference relation it induces is weakly Archimedean, thanks to Theorem 1. Therefore consider \(g\in \mathcal {D}_\mathcal {R}{\setminus }\mathcal {L}_1^+\) such that \(g=p-q\) for some \(p,q\in \mathcal {H}\). We want to show that there is \(\alpha \in (0,1)\) such that \(\alpha p-q\in \mathcal {D}_\mathcal {R}\).

    Thanks to point 4, we have that \(g\in \mathcal {R}{\setminus }\mathcal {L}^+\) and hence by strict desirability that there is \(\delta >0\) such that \(g-\delta \in \mathcal {R}\). Choose \(\alpha \in (0,1)\) so that \((1-\alpha )p\le \delta \). Then \(\alpha p-q\ge p-q-\delta =g-\delta \), whence \(\alpha p-q\in \mathcal {D}_\mathcal {R}\).

    (\(\Leftarrow \)) Finally, let us assume that \(\mathcal {D}_\mathcal {R}\) satisfies D0\('\)–D4\('\) and show that for all \(h\in \mathcal {R}{\setminus }\mathcal {L}^+\) and \(k\in \mathcal {L}\), there is \(\alpha \in (0,1)\) such that \(\alpha h+(1-\alpha )k\in \mathcal {R}\). If we prove this, then it is enough to take a constant \(k<0\) to get that \(h-\delta \in \mathcal {R}\), with \(\delta :=-\frac{1-\alpha }{\alpha }k>0\). This means that \(\mathcal {R}\) would be strictly desirable.

    Now then consider the positive and negative parts of k, namely, \(k^+,k^-\), such that \(k=k^+-k^-\). Note that for the thesis it is enough to show that there is \(\alpha \in (0,1)\) such that \(\alpha \frac{h}{2}-(1-\alpha )k^-\in \mathcal {R}\), given that \(\alpha \frac{h}{2}+(1-\alpha )k^+\) belongs to \(\mathcal {R}\) already, whence by additivity the result follows.

    To that end, choose \(\lambda >0\) such that \(f:=\frac{h}{2\lambda }\) and \(g:=\frac{k^-}{\lambda }\) satisfy the inequalities in D0\('\). Then there is \(\alpha \in (0,1)\) such that \(\alpha \frac{h}{2\lambda }-(1-\alpha )\frac{k^-}{\lambda }\in \mathcal {D}_\mathcal {R}\); and as a consequence, \(\alpha \frac{h}{2}-(1-\alpha )k^-\in \mathcal {R}\).

Proof of Theorem 3

Thanks to Theorems 1 and 2, we have an equivalence between coherent preference relations \(\succ \) and sets of desirable gambles coherent relative to \(\mathcal {L}_1\); the latter are in a one-to-one correspondence with their natural extensions, from which we deduce the first part of the theorem. The second part is granted again by Theorems 1 and 2.

Proof of Proposition 3

Assume first of all that \(\succ \) is negatively transitive, and let \(f,g\notin \mathcal {R}\). Then \(f\nsucc 0\) and \(g\nsucc 0\), from which it follows that \(0\nsucc -g\). Applying the negative transitivity of \(\succ \), we deduce that \(f\nsucc -g\), or, equivalently, that \(f+g \nsucc 0\), whence \(f+g\notin \mathcal {R}\).

Conversely, assume that Eq. (7) holds and let \(p,q,r\in \mathcal {H}\) satisfy \(p\nsucc q\) and \(q\nsucc r\). Then \(p-q\notin \mathcal {R}\) and \(q-r\notin \mathcal {R}\), whence \(p-r\notin \mathcal {R}\), and using the correspondence between \(\succ \) and \(\mathcal {R}\) we conclude that \(p\nsucc r\).

Proof of Proposition 4

From (10) and (11), we see that simply defining

$$\begin{aligned} p_2:=\psi (p_1) \end{aligned}$$
(13)

for all horse lotteries \(p_1\in \mathcal {H}_z\), we obtain a correspondence between \(\mathcal {H}\) and \(\mathcal {H}_z\) made of elements that are equally valuable for us: their respective probability currencies are just the same. Moreover, the correspondence is one-to-one as it follows from Zaffalon and Miranda (2017, Remark 5 in p. 1098) (note also that \(\psi \) is a linear operator).

It is now a simple step to prove the two inclusions.

\(\mathcal {R}_2\subseteq \mathcal {R}_1\):

If \(\lambda (p_2-q_2)\in \mathcal {R}_2\) for some \(p_2,q_2\in \mathcal {H}\) and \(\lambda >0\), then also \(\lambda \psi (p_1-q_1)\in \mathcal {R}_1\), where \(p_1,q_1\) are obtained through (13). This follows from the identity \(p_2-q_2=\psi (p_1-q_1)\), as obtained from linearity of \(\psi \), and the fact that the lotteries defined through (13) are equally valuable.

\(\mathcal {R}_1\subseteq \mathcal {R}_2\):

Similarly, consider any \(\lambda \psi (p_1-q_1)\in \mathcal {R}_1\) and let \(p_2:=\psi (p_1),q_2:=\psi (q_1)\), so that \(p_2,q_2\in \mathcal {H}\). It follows that \(\lambda (p_2-q_2)\in \mathcal {R}_2\) for the same reasons given in the proof of the converse inclusion.

Proof of Proposition 5

It follows from (Miranda et al. 2012, Proposition 29) that \(\hat{\mathcal {R}}\) is the smallest conglomerable and coherent set of gambles that includes \(\mathcal {R}_\varOmega \) and \(\mathcal {R}|\{\omega \}\) (for all \(\omega \in \varOmega \)). It includes \(\mathcal {R}|\varOmega \) by construction. We are left to show that \(\hat{\mathcal {R}}\) induces both \(\mathcal {R}_{\varOmega }\) and \(\mathcal {R}_{\mathcal {X}}\).

We begin by proving that the \(\varOmega \)-marginal of \(\hat{\mathcal {R}}\) is \(\mathcal {R}_{\varOmega }\). Consider an \(\varOmega \)-measurable gamble \(f\in \hat{\mathcal {R}}\). Then there are \(g\in \mathcal {R}_{\varOmega }\) and \(h\in \mathcal {R}|\varOmega \) such that \(f\ge g+h\). For any \(\omega \in \varOmega \) , it holds that

$$\begin{aligned} 0 \le \sup _{x\in \mathcal {X}} h(\omega ,x) \le \sup _{x\in \mathcal {X}} (f(\omega ,x)-g(\omega ,x))=f'(\omega )-g'(\omega ), \end{aligned}$$

where in last equality we are using that both fg are \(\varOmega \)-measurable, and are denoting by \(f',g'\) their equivalent representations as gambles on \(\varOmega \).

Thus, \(f\ge g\), and since \(\mathcal {R}_{\varOmega }\) is a coherent set of gambles we conclude that also \(f\in \mathcal {R}_{\varOmega }\). The converse inclusion follows from Eq. (12).

Next, consider an \(\mathcal {X}\)-measurable gamble \(f\in \hat{\mathcal {R}}\). Then there are \(g\in \mathcal {R}_{\varOmega }\) and \(h\in \mathcal {R}|\varOmega \) such that \(f\ge g+h\). If \(g\ne 0\), then there exists \(\omega \in \varOmega \) such that \(g'(\omega )> 0\). Then,

$$\begin{aligned} fI_{\{\omega \}}\ge g'(\omega )+hI_{\{\omega \}}\ge hI_{\{\omega \}}, \end{aligned}$$

and since \(h(\omega ,\cdot )\in \mathcal {R}_{\mathcal {X}}\cup \{0\}\), either \(h(\omega ,\cdot )\in \mathcal {R}_{\mathcal {X}}\), in which case \(f(\omega ,\cdot )\in \mathcal {R}_{\mathcal {X}}\), or \(h(\omega ,\cdot )=0\), whence \(f(\omega ,\cdot )\ge g'(\omega )>0\), meaning that also \(f(\omega ,\cdot )\in \mathcal {R}_{\mathcal {X}}\). Finally, the proof when \(g=0\) is similar (in that case we can pick any \(\omega \in \varOmega \) such that \(hI_{\{\omega \}}\ne 0\), since one is bound to exist). Again, the converse inclusion follows from Eq. (12).

Proof of Proposition 6

The result follows immediately from Definition 31, taking into account that \({\underline{P}}_\varOmega '({\underline{P}}(\cdot |\varOmega ))\) is a coherent lower prevision by Walley (1991, Section 6.7.2).

Proof of Proposition 7

We make a circular proof.

\((i)\Rightarrow (ii)\):

It \(f\notin \mathcal {R}\), it follows from Eq. (3) that \({\underline{P}}(f)={\overline{P}}(f)\le 0\), whence \({\underline{P}}(-f)=-{\overline{P}}(f)\ge 0\), and therefore \({\underline{P}}(\varepsilon -f)>0\) for every \(\varepsilon >0\). Thus, Eq. (3) implies that \(\varepsilon -f\in \mathcal {R}\).

\((ii)\Rightarrow (iii)\):

Assume ex-absurdo the existence of \(f,g\notin \mathcal {R}\) such that \(f+g-\varepsilon \in \mathcal {R}\). It follows from (ii) that \(\frac{\varepsilon }{4}-f,\frac{\varepsilon }{4}-g\in \mathcal {R}\), and applying D3 we conclude that \(-\frac{\varepsilon }{2}\in \mathcal {R}\), a contradiction with D2.

\((iii)\Rightarrow (i)\):

If \({\underline{P}}\) is not linear, we can find a gamble f such that \({\underline{P}}(f)<{\overline{P}}(f)\). The conjugacy of the lower and upper previsions and Eq. (1) implies that for every \(\varepsilon >0\) the gambles \(f-{\underline{P}}(f)-\varepsilon \) and \({\overline{P}}(f)-f-\varepsilon \) do not belong to \(\mathcal {R}\), and (iii) implies then that for every \(\delta >0\) the gamble \((f-{\underline{P}}(f)-\varepsilon )+({\overline{P}}(f)-f-\varepsilon )-\delta \) does not belong to \(\mathcal {R}\). But for \(\varepsilon ,\delta \) satisfying \(2\varepsilon +\delta <{\overline{P}}(f)-{\underline{P}}(f)\) this sum is a non-negative gamble. This is a contradiction with D1.

Proof of Proposition 8

This is an immediate consequence of Proposition 7.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zaffalon, M., Miranda, E. Desirability foundations of robust rational decision making. Synthese 198 (Suppl 27), 6529–6570 (2021). https://doi.org/10.1007/s11229-018-02010-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-018-02010-x

Keywords

Navigation