Skip to main content
Log in

Anticipated stochastic choice

  • Research Article
  • Published:
Economic Theory Aims and scope Submit manuscript

Abstract

The purpose of this study is to characterize stochastic choice influenced by the objective or subjective positions of alternatives in a menu. The main theorem axiomatizes the anticipated stochastic choice (ASC) representation, wherein the decision maker maximizes the expected utility by the cognitive control of a probability measure over mental states that trigger the ex post choice of alternatives. A key prerequisite for this axiomatization is that the randomization between menus is identified with their perfectly correlated mixture, which includes only mixtures of specific alternative pairs. The essential uniqueness of an ASC representation defines an index of rationality that is relevant to a preference for commitment. Special cases of ASC include exact utility maximization, uncontrolled stochastic choice, trembling hands, and choice with limited attention. Furthermore, ASC accommodates potentially stochastic choice anomalies such as the attraction effect, cyclical choice, and position effects.

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.

Fig. 1

Similar content being viewed by others

Notes

  1. We use the term “position-dependence” rather than the more general term “menu-dependence” because the former provides a natural interpretation for the correlated choices of alternatives and potentially full support stochastic choice function, which are key characterizations of the following analysis.

  2. Anticipated mistakes (i.e., ex ante awareness of ex post mistakes) have been discussed both in psychology (Broadbent et al. 1982) and in economics (Kahneman et al. 1997; Manzini and Mariotti 2015; Piccione and Rubinstein 1997). Gross (1998) also argued that under the influence of emotion, the DM often applies the situation selection strategy, that is, she attempts to avoid a situation that possibly creates a negative psychological state leading to an undesirable future choice, which can be interpreted as a preference for commitment.

  3. As usual, menu x is naturally identified with a degenerated random menu that generates x with a probability of one, and an alternative \(\beta \) is identified with a singleton menu \(\{\beta \}\).

  4. By definition, we have \(\lambda P \oplus (1-\lambda ) Q = (1-\lambda ) Q \oplus \lambda P\).

  5. We restrict our attention to finite (random) menus because of a key characterization of the following analysis, namely a surjection or bijection from one menu to the other can be consistently defined for all menu pairs; this is not generally the case for (uncountably) infinite menus. Furthermore, many studies have examined a preference over (the convex hulls of) finite menus (e.g., Chatterjee and Krishna 2009; Dekel et al. 2009; Gul and Pesendorfer 2001).

  6. The timeline outlined here is an interpretation, rather than part of the model. The only item that is assumed to be observable is choice over random menus in period 1.

  7. Whether we can characterize an informative preference representation with imperfectly correlated mixtures of menus (i.e., a randomization between identical menus not being indifferent to the menu itself or the Minkowski sum) is an open question. However, this problem may be solved by imposing certain axioms. For example, the set betweenness proposed by Gul and Pesendorfer (2001) renders the preference representation dependent only on two alternatives in a menu, implying that a randomization between menus is indifferent to a doubleton set comprising the mixtures of two specific alternative pairs. Similar axioms may characterize more general cases.

  8. Peter Hammond suggests the alternative terminology of selection function for this type of choice function.

  9. Chandrasekher (2015) independently developed a technique to apply the arguments in the Anscombe–Aumann framework to menu preferences by providing additional structures to the menus. A major difference in our approach is the imposition of perfectly correlated mixtures of menus, whereas Chandrasekher employed uncorrelated mixtures. This difference eventually results in representations with a vN–M utility function (combined with the stochastic choice function) in ASC, and a Strotzian value function in Chandrasekher’s model.

  10. We discuss the implication of the preference representation obtained by reversing ARI’s implication in Sect. 6.3.

  11. Our approach is in line with that of Noor and Takeoka (2010, 2015) in that both consider a menu preference that violates the Independence of Irrelevant Alternatives axiom and the Weak Axiom of Revealed Preference (WARP). However, Noor and Takeoka focus on the deterministic choice of alternatives from a menu, whereas we allow for stochastic choice.

  12. Although we do not provide a formal axiomatization of the trembling-hand ASC representation, for simplicity, it is derived by exploiting an axiomatization of \(\epsilon \)-contamination (e.g., Nishimura and Ozaki 2006) and the similarity between ASC and MMEU.

  13. This representation is reminiscent of that of Chatterjee and Krishna (2009), who considered a DM that maximizes the normative utility with a probability of \(1-\epsilon \) and the “alter-ego” utility with a probability of \(\epsilon \). Specifically, if we interpret the composite functions of u and \(\phi \) as exactly maximized state-dependent utility functions, (3) implies that the DM maximizes the normative utility with a probability of \(1-\epsilon \) and the multiple “alter-ego” utilities with a probability of \(\epsilon \). Unlike their study, however, our model permits nonextreme points of the menu’s convex hull to be chosen with a positive probability in the “\(\epsilon \)” event, which is crucial to interpreting it as the trembling-hand case.

  14. A similar “implementation error” interpretation of limited attention in the context of menu preference is discussed by Manzini and Mariotti (2015).

  15. A similar argument indicates that the set betweenness axiom introduced by Gul and Pesendorfer (2001) (i.e., \(x \succsim y\) implying \(x \succsim x \cup y \succsim y\)) is equivalent to exact utility maximization for a regular ASC: the latter trivially implies the former. Conversely, under the lemma proved by Gul and Pesendorfer, set betweenness implies in ASC that only the mental states associated with the normatively best and worst alternatives in menu x are assigned a positive probability by the optimal probability measure over mental states chosen in (2). Moreover, the normatively worst alternative is assigned a zero probability; otherwise, regularity condition (c) can replace the alternative, so that the optimal probability measure over mental states assigns a positive probability to an alternative that is neither normatively best nor normatively worst in the menu, which contradicts the implication of Gul and Pesendorfer’s lemma. Thus, \({\mathcal {M}}_{|x|}\) includes a probability measure that assigns a probability of one to an arbitrary mental state. It follows from the closedness and convexity that \({\mathcal {M}}_{|x|}\) comprises all the probability measures over mental states.

  16. A condition similar to Definition 5a can also be found in the desire for commitment axiom proposed by Dekel et al. (2009).

  17. The reader may suspect that the above argument relies on the fact that x and y have different cardinalities and are thus evaluated by different sets of probability measures over mental states. However, ASC also accommodates similar choice anomalies for menus of identical cardinality. For example, Tyszka (1983) reported that, for \(\alpha , \beta , \gamma \), and \(\delta \in {\varDelta }(Z)\), the majority of the participants choose \(\alpha \) from \(x = \{\alpha , \beta , \gamma \}\), despite choosing \(\beta \) from \(y = \{\alpha , \beta , \delta \}\). Assuming that \(\alpha \succ \beta \succ \gamma \sim \delta , c(x) = (\alpha , \gamma , \beta ), c(y) = (\beta , \delta , \alpha )\), and \({\mathcal {M}}_3 = \{\mu \in {\varDelta }(S_3) : \mu (s_i) \le 1-\epsilon \ \text {for} \ i = 1, 2 \ \text {and} \ \mu (s_3) \le \epsilon ' \}\) for sufficiently small \(\epsilon \) and \(\epsilon '\), the above choice pattern is consistent with the ASC preference.

  18. This result does not imply that the attraction effect generally accompanies a preference for commitment. However, the attraction effect (and other choice anomalies) can be associated with a specific type of preference for commitment: in this example, the DM would naturally prefer menu x to y, because she is ex ante aware that decoy \(\delta _\beta \) attracts her attention to the suboptimal alternative \(\beta \), and is thus willing to exclude \(\delta _\beta \) from the menu. Another interpretation is that the menu choice is made by the social planner, who prefers to exclude suboptimal alternatives from the menu to prevent the DM from making an erroneous alternative choice. A discussion relevant to this interpretation can be found in Manzini and Mariotti (2015).

  19. Nehring’s (1999) argument parallels the relationship between the maximization and minimization approaches, as he used concave capacities to denote utility maximization. In contrast, Schmeidler (1989) used convex capacities to model ambiguity aversion.

References

  • Ahn, D.S.: Ambiguity without a state space. Rev. Econ. Stud. 75, 3–28 (2007)

    Article  Google Scholar 

  • Ahn, D.S., Sarver, T.: Preference for flexibility and random choice. Econometrica 81, 341–361 (2013)

    Article  Google Scholar 

  • Aliprantis, C.D., Border, K.C.: Infinite Dimensional Analysis: A Hitchhiker’s Guide, 3rd edn. Springer Verlag, Berlin, Germany (2006)

    Google Scholar 

  • Bandyopadhyay, T., Dasgupta, I., Pattanaik, P.K.: Stochastic revealed preference and the theory of demand. J. Econ. Theory 84, 95–110 (1999)

    Article  Google Scholar 

  • Broadbent, D.E., Cooper, P.F., FitzGerald, P., Parkes, K.R.: The cognitive failures questionnaire (CFQ) and its correlates. Br. J. Clin. Psychol. 19, 177–188 (1982)

    Google Scholar 

  • Bruine de Bruin, W.: Save the last dance for me: unwanted serial position effects in jury evaluations. Acta Psychol. 118, 245–260 (2005)

    Article  Google Scholar 

  • Chandrasekher, M.: Incomplete menu preferences and ambiguity. Mimeo, Arizona State University. http://www.public.asu.edu/~mchandr6/Working (2015). Accessed 15 March 2016

  • Chatterjee, K., Krishna, R.V.: A “dual self” representation for stochastic temptation. Am. Econ. J. Microecon. 1, 148–167 (2009)

    Article  Google Scholar 

  • Chatterjee, K., Krishna, R.V.: Uniquely representing “a preference for uniformity”. BE J. Theor. Econ. 12, 1–34 (2012)

    Google Scholar 

  • Christenfeld, N.: Choices from identical options. Psychol. Sci. 6, 50–55 (1995)

    Article  Google Scholar 

  • Dekel, E., Lipman, B.: Costly self-control and random self-indulgence. Econometrica 80, 1271–1302 (2012)

    Article  Google Scholar 

  • Dekel, E., Lipman, B., Rustichini, A.: Representing preferences with a unique subjective state space. Econometrica 69, 891–934 (2001)

    Article  Google Scholar 

  • Dekel, E., Lipman, B., Rustichini, A.: Temptation-driven preferences. Rev. Econ. Stud. 76, 937–971 (2009)

    Article  Google Scholar 

  • Epstein, L., Marinacci, M., Seo, K.: Coarse contingencies and ambiguity. Theor. Econ. 2, 355–394 (2007)

    Google Scholar 

  • Ergin, H., Sarver, T.: A unique costly contemplation representation. Econometrica 78, 1285–1339 (2010)

    Article  Google Scholar 

  • Ergin, H., Sarver, T.: Hidden actions and preferences for timing of resolution of uncertainty. Theor. Econ. 10, 489–541 (2015)

    Article  Google Scholar 

  • Fudenberg, D., Levine, D.: A dual-self model of impulse control. Am. Econ. Rev. 96, 1449–1476 (2006)

    Article  Google Scholar 

  • Ghirardato, P., Marinacci, M.: Ambiguity made precise: a comparative foundation. J. Econ. Theory 102, 251–289 (2002)

    Article  Google Scholar 

  • Gilboa, I., Schmeidler, D.: Maxmin expected utility with non-unique prior. J. Math. Econ. 18, 141–53 (1989)

    Article  Google Scholar 

  • Gross, J.J.: The emerging field of emotion regulation: an integrative view. Rev. Gen. Psychol. 2, 271–299 (1998)

    Article  Google Scholar 

  • Gul, F., Pesendorfer, W.: Temptation and self-control. Econometrica 69, 1403–1435 (2001)

    Article  Google Scholar 

  • Gul, F., Pesendorfer, W.: Random expected utility. Econometrica 74, 121–146 (2006)

    Article  Google Scholar 

  • Huber, J., Payne, J., Puto, C.: Adding asymmetrically dominated alternatives: violations of regularity and the similarity hypothesis. J. Consum. Res. 9, 90–98 (1982)

    Article  Google Scholar 

  • Kahneman, D., Wakker, P.P., Sarin, R.: Back to Bentham? Explorations of experienced utility. Q. J. Econ. 112, 375–405 (1997)

    Article  Google Scholar 

  • Kreps, D.M.: A representation theorem for “preference for flexibility”. Econometrica 47, 565–577 (1979)

    Article  Google Scholar 

  • Krueger, L.E., Salthouse, T.A.: Influence of cognitive abilities and age on word recall performance across trials and list segments. Am. J. Psychol. 124, 291–300 (2011)

    Article  Google Scholar 

  • Larsen, R.J.: Toward a science of mood regulation. Psychol. Inq. 11, 129–141 (2000)

    Article  Google Scholar 

  • Loewenstein, G., O’Donoghue, T.: Animal spirits: affective and deliberative processes in economic behavior. In: CAE Working Paper 04-14, Cornell University. https://cae.economics.cornell.edu/04-14 (2004). Accessed 15 March 2016

  • Luce, R.D.: Individual Choice Behavior. Wiley, New York (1959)

    Google Scholar 

  • Luce, R.D., Suppes, P.: Preference, utility and subjective probability. In: Luce, R.D., Bush, R.R., Galanter, E. (eds.) Handbook of Mathematical Psychology, vol. III, pp. 249–410. Wiley, New York (1965)

    Google Scholar 

  • Manzini, P., Mariotti, M.: Stochastic choice and consideration sets. Econometrica 82, 1153–1176 (2014)

    Article  Google Scholar 

  • Manzini, P., Mariotti, M.: Imperfect attention and menu evaluation. Mimeo, University of St. Andrews. http://www.st-andrews.ac.uk/~wwwecon/repecfiles/4/1319 (2015). Accessed 15 March 2016

  • Masatlioglu, Y., Nakajima, D., Ozbay, E.Y.: Revealed attention. Am. Econ. Rev. 102, 2183–2205 (2012)

    Article  Google Scholar 

  • Murdock, B.B.: The serial position effect of free recall. J. Exp. Psychol. 64, 482–488 (1962)

    Article  Google Scholar 

  • Nehring, K.: Preference for flexibility in a Savage framework. Econometrica 67, 101–119 (1999)

    Article  Google Scholar 

  • Nehring, K.: Self-control through second-order preferences. Mimeo, University of California, Davis. http://www.parisschoolofeconomics.eu/IMG/pdf/sop-9-06 (2006). Accessed 15 March 2016

  • Nishimura, K.G., Ozaki, H.: An axiomatic approach to epsilon-contamination. Econ. Theory 27, 333–340 (2006)

    Article  Google Scholar 

  • Noor, J., Takeoka, N.: Uphill self-control. Theor. Econ. 5, 127–158 (2010)

    Article  Google Scholar 

  • Noor, J., Takeoka, N.: Menu-dependent self-control. J. Math. Econ. 61, 1–20 (2015)

    Article  Google Scholar 

  • Olszewski, W.: Preferences over sets of lotteries. Rev. Econ. Stud. 74, 567–595 (2007)

    Article  Google Scholar 

  • Piccione, M., Rubinstein, A.: On the interpretation of decision problems with imperfect recall. Games Econ. Behav. 20, 3–24 (1997)

    Article  Google Scholar 

  • Posner, M.I., Snyder, C.R.R.: Attention and cognitive control. In: Solso, R.L. (ed.) Information Processing and Cognition: The Loyola Symposium, pp. 55–85. Lawrence Erlbaum Associates, Hillsdale (1975)

    Google Scholar 

  • Rubinstein, A., Salant, Y.: A model of choice from lists. Theor. Econ. 1, 3–17 (2006)

    Google Scholar 

  • Sarver, T.: Anticipating regret: why fewer options may be better. Econometrica 76, 263–305 (2008)

    Article  Google Scholar 

  • Schmeidler, D.: Subjective probability and expected utility without additivity. Econometrica 57, 571–587 (1989)

    Article  Google Scholar 

  • Selten, R.: Reexamination of the perfectness concept for equilibrium points in extensive games. Int. J. Game Theory 4, 25–55 (1975)

    Article  Google Scholar 

  • Sherman, J.W., Gawronski, B., Gonsalkorale, K., Hugenberg, K., Allen, T.J., Groom, C.J.: The self-regulation of automatic associations and behavioral impulses. Psychol. Rev. 115, 314–335 (2008)

    Article  Google Scholar 

  • Simonson, I.: Choice based on reasons: the case of attraction and compromise effects. J. Consum. Res. 16, 158–174 (1989)

    Article  Google Scholar 

  • Slovic, P.: Choice between equally-valued alternatives. J. Exp. Psychol. 1, 280–287 (1975)

    Google Scholar 

  • Strotz, R.H.: Myopia and inconsistency in dynamic utility maximization. Rev. Econ. Stud. 23, 165–180 (1955)

    Article  Google Scholar 

  • Tentori, K., Osherson, D., Hasher, L., May, C.: Wisdom and aging: irrational preferences in college students but not older adults. Cognition 81, B87–B96 (2001)

    Article  Google Scholar 

  • Thaler, R.H., Shefrin, H.M.: An economic theory of self-control. J. Polit. Econ. 89, 392–406 (1981)

    Article  Google Scholar 

  • Tyszka, T.: Contextual multiattribute decision rules. In: Sjoberg, I., Tyszka, T., Wise, J. (eds.) Human Decision Making, pp. 243–256. Doxa, Bodafors (1983)

    Google Scholar 

  • Wright, P., Barbour, F.: Phased decision strategies: sequels to an initial screening. In: Starr, M., Zeleny, M. (eds.) Studies in Management Sciences, Multiple Criteria Decision Making, pp. 91–109. North-Holland, Amsterdam (1977)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nobuo Koida.

Additional information

I am grateful to Itzhak Gilboa, Peter Hammond, Youichiro Higashi, Kazuya Hyogo, Atsushi Kajii, Paola Manzini, Marco Mariotti, Hitoshi Matsushima, Daisuke Nakajima, Yutaka Nakamura, Christopher Olivola, Daisuke Oyama, Hiroyuki Ozaki, Marciano Siniscalchi, Norio Takeoka, and the conference and seminar participants at SAET 2015, JEA 2011 Fall Meeting, Kyoto University, the University of Tokyo, Keio University, Okayama University, and Tohoku University for their helpful discussions and comments. I also thank two anonymous referees for their valuable comments that have significantly improved the paper. Financial support from the Japan Society for the Promotion of Science (JSPS) Grants-in-Aid for Scientific Research (Grant Numbers 24730168, 26380241) is gratefully appreciated. The usual disclaimer applies.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 80 KB)

Appendix

Appendix

1.1 Proof of Theorem 1

For the sufficiency part of Theorem 1 (a regular ASC representation implies the axioms), we only present proofs for Axioms 2, 7b, and 7c, because the sufficiency for the other axioms is straightforward.

To confirm that Axiom 2 holds, we note that the affineness of u implies that \(\int _{\mathcal {A}} u(\phi (x, s)) dP(x) = \lambda _1 u(\phi (x_1, s)) + \cdots + \lambda _m u(\phi (x_m, s))\) for all finite random menus \(P = (\lambda _1, x_1; \ldots ; \lambda _m, x_m)\) that generate menu \(x_i\) with probability \(\lambda _i\) for \(i = 1, \ldots , m\), and all \(s \in S\). Then, Axiom 2b follows from regularity conditions (a) and (b), and considering \(P = (\lambda , x; 1-\lambda , x)\) also implies Axiom 2a.

On the other hand, Axioms 7b and 7c are implied by the regularity conditions of \(\phi \). First, Axiom 7b is proved by induction. Let \(x^*_1 = \{\beta ^1_1\}\) for an arbitrary \(\beta ^1_1 \in {\varDelta }(Z)\). Next, fix \(n \ge 1\) and assume that \(x^*_i\) satisfies the implication of Axiom 7b for \(i = 1, \ldots , n\). The construction of \(\phi \) implies, for any given \(x^*_{n+1} \in {\mathcal {A}}\) such that \(|x^*_{n+1}| = n+1\), that \(\lambda x^*_n \oplus (1-\lambda ) x^*_{n+1} \sim \cup ^{n+1}_{i=1} \{\lambda \phi (x^*_n, s_i) +(1-\lambda )\phi (x^*_{n+1}, s_i)\}\) for all \(\lambda \in [0, 1]\). By defining \(\beta ^{n+1}_i \equiv \phi (x^*_{n+1}, s_i)\) for \(i = 1, \ldots , n+1\), regularity conditions (a) and (b) imply that \(x^*_{n+1} = \{\beta ^{n+1}_1, \ldots , \beta ^{n+1}_{n+1}\}\) and \(\lambda x^*_n \oplus (1-\lambda ) x^*_{n+1} \sim \cup ^{n+1}_{i=1} \{\lambda \beta ^n_{\min \{i, n\}} + (1-\lambda ) \beta ^{n+1}_i\}\) for all \(\lambda \in [0, 1]\). Because the same argument holds for all \(n \ge 1\), we obtain Axiom 7b. Second, regularity condition (c) implies that, for all \(x \in {\mathcal {A}}, 1 \le k \le |x|\), and \({\bar{\beta }} \in {\varDelta }(Z)\), there exists \(x' \in {\mathcal {A}}\) such that \(|x'| = |x|, \phi (x', s_i) \sim \phi (x, s_i)\) for all \(i \ne k\), and \(\phi (x', s_k) \sim {\bar{\beta }}\), which implies Axiom 7c.

Below, we prove the necessity part of the theorem. The first lemma indicates that Axiom 2 derives a bijection between menus of identical cardinality that describes the DM’s perception of random menus.

Lemma 1

Axiom 2 implies the following axiom:

Axiom \(2^{*}\) For all \(x, y \in {\mathcal {A}}\) such that \(|x| = |y|\), there exists a bijection \(\tau : x \rightarrow y\) such that \(\lambda x \oplus (1-\lambda ) y \sim \{\lambda \alpha + (1-\lambda ) \tau (\alpha ): \alpha \in x \}\) for all \(\lambda \in [0, 1]\).

Proof

Axiom 2b (specifically, Definition 1a) implies that, for given \(x, y \in {\mathcal {A}}\) such that \(|x| = |y|\), there exists \(C \subseteq x \times y\) such that \(\lambda x \oplus (1-\lambda ) y \sim \{\lambda \alpha + (1-\lambda ) \beta : (\alpha , \beta ) \in C \}\) for all \(\lambda \in [0, 1]\). By construction, we also have \(\lambda ' y \oplus (1-\lambda ') x \sim \{\lambda ' \beta + (1-\lambda ') \alpha : (\alpha , \beta ) \in C \}\) for all \(\lambda ' \in [0, 1]\). From Axiom 2b (Definition 1b), it follows that

$$\begin{aligned} \lambda '' x \oplus (1-\lambda '') x \sim \{\lambda '' \alpha + (1-\lambda '') \alpha ': (\alpha , \beta ) \in C \ \text {and} \ (\alpha ', \beta ) \in C\} \end{aligned}$$
(4)

for all \(\lambda '' \in [0, 1]\). Axiom 2a implies that there exists some C such that the right-hand side of (4) is equivalent to x.

Now, assume that for all C satisfying (4), there is no bijection \(\tau : x \rightarrow y\), such that \(\beta = \tau (\alpha )\) if and only if \((\alpha , \beta ) \in C\). This implies that (i) there exists \({\tilde{\alpha }} \in x\) such that \(({\tilde{\alpha }}, \beta ) \not \in C\) for all \(\beta \in y\), (ii) there exist distinct \({\tilde{\alpha }}, {\hat{\alpha }} \in x\) such that \(({\tilde{\alpha }}, \beta ), ({\hat{\alpha }}, \beta ) \in C\) for some \(\beta \in y\), or (iii) there exist distinct \({\tilde{\beta }}, {\hat{\beta }} \in y\) such that \((\alpha , {\tilde{\beta }}), (\alpha , {\hat{\beta }}) \in C\) for some \(\alpha \in x\). In case (i), we assume without loss of generality that there is no \(C' \subseteq x \times y\) such that \(C' \supsetneq C\) satisfies (4) in replacing C. Then, the right-hand side of (4) cannot be indifferent to x for all \(\lambda '' \in [0, 1]\) because it never includes \({\tilde{\alpha }}\). In case (ii), we assume without loss of generality that there is no \(C' \subseteq x \times y\) such that \(C' \subsetneq C\) satisfies (4) in replacing C. Then, the right-hand side of (4) cannot be indifferent to x for all \(\lambda '' \in [0, 1]\), because it includes \(\lambda {\tilde{\alpha }} + (1-\lambda ) {\hat{\alpha }}\). Finally, case (iii) is equivalent to case (ii) after exchanging the roles of x and y. Accordingly, all the cases contradict Axiom 2a.

The above discussion holds for all such menus x and y, and thus, we obtain the desired result. \(\square \)

Next, we construct the sets of mental states and a regular choice function.

Lemma 2

There exist \(S_n = \{s_1, \ldots , s_n\}, S = \cup ^\infty _{n=1} \{S_n\}\), and a regular \(\phi \) such that

$$\begin{aligned} \lambda x \oplus (1-\lambda ) y \sim \cup _{s \in S_{\max \{|x|, |y|\}}} \{\lambda \phi (x, s) + (1-\lambda ) \phi (y, s)\} \end{aligned}$$
(5)

for all \(x, y \in {\mathcal {A}}\) and \(\lambda \in [0, 1]\).

Proof

We prove the lemma by induction. First, define \(S_1 = \{s_1\}\) and \(\phi (x, s_1) = \beta \) for all \(x = \{\beta \} \in {\mathcal {A}}_1\). These definitions, combined with Lemma 1, imply regularity condition (a) and (5) for all singleton menus x and y.

Next, set \(n \ge 1\) and assume that \(S_n\) and \(\phi \) satisfy regularity condition (a) and (5) for all \(x, y \in \mathcal{A}_n\). Axiom 7b implies that, for menus \(x^*_n = \{\beta ^n_1, \ldots , \beta ^n_n\}\) and \(x^*_{n+1} = \{\beta ^{n+1}_1, \ldots , \beta ^{n+1}_{n+1}\}\) defined in the axiom, \(\lambda x^*_n \oplus (1-\lambda ) x^*_{n+1} \sim \cup ^{n+1}_{i=1}\{\lambda \beta ^n_{\min \{i, n\}} + (1-\lambda ) \beta ^{n+1}_i\}\) for all \(\lambda \in [0, 1]\). Accordingly, by defining \(S_{n+1} \equiv S_n \cup \{s_{n+1}\}, \phi (x^*_n, s_i) \equiv \beta ^n_{\min \{i, n\}}\), and \(\phi (x^*_{n+1}, s_i) \equiv \beta ^{n+1}_i\) for \(i = 1, \ldots , n+1\), we obtain \(\lambda x^*_n \oplus (1-\lambda ) x^*_{n+1} \sim \cup _{s \in S_{n+1}}\{\lambda \phi (x^*_n, s) + (1-\lambda ) \phi (x^*_{n+1}, s)\}\) for all \(\lambda \in [0, 1]\) and \(\phi (x^*_n, s_{n+1}) = \phi (x^*_n, s_n)\).

Now, iteratively applying Axiom 2b (i.e., \(\succsim \) having consistent perceptions of random menus) to the above argument implies that, for all \(m \le n+1, \lambda x^*_m \oplus (1-\lambda ) x^*_{n+1} \sim \cup _{s \in S_{n+1}} \{\lambda \phi (x^*_m, s) + (1-\lambda ) \phi (x^*_{n+1}, s)\}\) for all \(\lambda \in [0, 1]\) and \(\phi (x^*_m, s_i) = \phi (x^*_m, s_m)\) for \(i = m+1, \ldots , n+1\). Moreover, Lemma 1 implies that, for all \(x, y \in {\mathcal {A}}\) such that \(|x| = m\) and \(|y| = n+1\), there exist bijections \({\hat{\tau }}: x^*_m \rightarrow x\) and \({\tilde{\tau }}: x^*_{n+1} \rightarrow y\) such that \(\lambda ' x^*_m \oplus (1-\lambda ') x \sim \{\lambda ' \beta ^m_i + (1-\lambda ') {\hat{\tau }} (\beta ^m_i): i = 1, \ldots , m \}\) for all \(\lambda ' \in [0, 1]\) and \(\lambda '' x^*_{n+1} \oplus (1-\lambda '') y \sim \{\lambda '' \beta ^{n+1}_i + (1-\lambda '') {\tilde{\tau }} (\beta ^{n+1}_i): i = 1, \ldots , n+1 \}\) for all \(\lambda '' \in [0, 1]\). Accordingly, by defining \(\phi (x, s_i) = {\hat{\tau }}(\beta ^m_{\min \{i, m\}})\) and \(\phi (y, s_i) = {\tilde{\tau }}(\beta ^{n+1}_i)\) for \(i = 1, \ldots , n+1\), Axiom 2b implies that \(\lambda x \oplus (1-\lambda ) y \sim \cup _{s \in S_{n+1}} \{\lambda \phi (x, s) + (1-\lambda ) \phi (y, s)\}\) for all \(\lambda \in [0, 1]\), and \(\phi (x, s_i) = \phi (x, s_m)\) for \(i = m+1, \ldots , n+1\). This implies (5) for all \(x, y \in {\mathcal {A}}_{n+1}\). The construction of \(\phi (y, \cdot )\) also gives regularity condition (a) for all \(y \in {\mathcal {A}}_{n+1}\). Moreover, because the above argument implies that \(\phi (x, s_{n'}) = \phi (x, s_{|x|})\) for all \(x \in {\mathcal {A}}\) and \(n' \in {\mathbb {N}}\) such that \(n' \ge |x|\), regularity condition (b) is obtained.

Finally, by the construction of \(\phi \), Axiom 7c implies that, for all \(1 \le k \le n, {\bar{\beta }} \in {\varDelta }(Z)\), and \(x = \{\beta _1, \ldots , \beta _n\}\) such that \(\phi (x, s_i) = \beta _i\) for \(i = 1, \ldots , n\), there exists \(x' = \{\beta '_1, \ldots , \beta '_n\}\) such that \(\phi (x', s_i) = \beta '_i\) for \(i = 1, \ldots , n, \beta '_i \sim \beta _i\) for all \(i \ne k\), and \(\beta '_k \sim {\bar{\beta }}\). Regularity condition (c) is obtained by iteratively applying the argument for \(k = 1, \ldots , n\). \(\square \)

The iterative application of (5) in Lemma 2 implies that a finite random menu \(P = (\lambda _1, x_1; \ldots ; \lambda _m, x_m)\) is indifferent to a finite nonrandom menu \(\Psi (P) \equiv \cup _{s \in S_{{\bar{n}}(P)}} \{ \lambda _1 \phi (x_1, s)+ \cdots +\lambda _m \phi (x_m, s)\}\). Thus, we focus on the preference over finite menus, rather than that over random menus.

Next, we show that there is an affine function u (i.e., \(u(\lambda \beta + (1-\lambda ) \gamma ) = \lambda u(\beta ) + (1-\lambda ) u(\gamma )\) for all \(\lambda \in [0, 1]\) and \(\beta , \gamma \in {\varDelta }(Z)\)) representing the commitment ranking.

Lemma 3

There is an affine function u that represents \(\succsim \) on \({\varDelta }(Z)\); that is, for all \(\beta , \beta ' \in {\varDelta }(Z), u(\beta ) \ge u(\beta ')\) whenever \(\{\beta \} \succsim \{\beta '\}\). Furthermore, u is unique up to a positive affine transformation.

Proof

The conclusion is straightforward from Axioms 1, 3, and 4. Note that the independence axiom for alternatives follows from singleton independence. \(\square \)

The following lemma indicates that there is a functional J representing the restriction of \(\succsim \) on \({\mathcal {A}}\).

Lemma 4

Let u be an affine function derived from Lemma 3. Then, there is a functional \(J: {\mathcal {A}} \rightarrow \mathfrak {R}\) such that

  1. (a)

    for all \(x, y \in {\mathcal {A}}, x \succsim y\) if and only if \(J(x) \ge J(y)\), and

  2. (b)

    for all \(\beta \in {\varDelta }(Z)\) and \(x = \{\beta \}, J(x) = u(\beta )\).

Proof

For singleton menus, J is uniquely defined by (b). To define J for all menus, fix \({\bar{\beta }}\) and \({\underline{\beta }}\) such that \(\{{\bar{\beta }}\} \succsim x \succsim \{{\underline{\beta }}\}\) for all \(x \in {\mathcal {A}}\). It follows from Axiom 3 and Lemma 1 that there exists a unique \(\lambda \in [0, 1]\) such that \(x \sim \lambda \{{\bar{\beta }}\} \oplus (1-\lambda ) \{{\underline{\beta }}\} \sim \{\lambda {\bar{\beta }} + (1-\lambda ) {\underline{\beta }}\}\). Thus, by defining \(J(x) = J(\{\lambda {\bar{\beta }} + (1-\lambda ) {\underline{\beta }}\}), J\) also satisfies (a). \(\square \)

Now, fix the cardinality n of the menus and the function u such that \(u(\beta ) > 1\) and \(u(\beta ') < -1\) for some \(\beta , \beta ' \in {\varDelta }(Z)\) (which is allowed by Axiom 7a). We denote by B the linear space of all functions \(a: S_n \rightarrow \mathfrak {R}\) endowed with state-wise scalar multiplication and addition; that is, for all \(a, b \in B\) and \(\lambda \in \mathfrak {R}, \lambda a\) and \(a + b\) are defined by \((\lambda a)(s) = \lambda (a(s))\) and \((a + b)(s) = a (s) + b (s)\) for all \(s \in S_n\), respectively. Because \(S_n\) is finite, B is equivalent to the linear subspace of simple functions, which we denote by \(B_0\). We also define \(K = u({\varDelta }(Z))\) and denote by \(B_0(K)\) the set of simple functions that have range K. For all \(\xi \in \mathfrak {R}\), we denote by \(\xi ^* \in B\) a constant function such that \(\xi ^*(s) = \xi \) for all \(s \in S_n\). Now, define \(\phi ^x_n(\cdot ) \equiv \phi (x, \cdot )\) on \(S_n\). From regularity condition (c) of \(\phi \), which we proved in Lemma 2, it follows that the set of functions \(u \circ \phi ^x_n\) with x ranging over \(\mathcal{A}_n\) is equivalent to \(B_0(K)\), i.e., \(\{u \circ \phi ^x_n : x \in {\mathcal {A}}_n \} = B_0(K)\).

Let \({\tilde{u}}: {\mathcal {A}}_n \rightarrow K^{S_n}\) be such that \({\tilde{u}}(x) = u \circ \phi ^x_n\) for all \(x \in {\mathcal {A}}_n\). The following lemmas, which are counterparts of Gilboa and Schmeidler’s (1989) lemmas, characterize the functional I on B derived from the axioms. Note that, unlike Gilboa and Schmeidler’s argument, I is sublinear instead of superlinear and is denoted by the maximum of integrals rather than the minimum.

Lemma 5

There is a functional \(I: B \rightarrow \mathfrak {R}\) such that

  1. (a)

    for all \(x \in {\mathcal {A}}\) such that \(|x| = n, I({\tilde{u}}(x)) = J(x)\),

  2. (b)

    I is monotonic, that is, \(a \ge b\) implies \(I(a) \ge I(b)\) for all \(a, b \in B\),

  3. (c)

    I is sublinear (subadditive and homogeneous of degree one), and

  4. (d)

    I is C-independent, that is, \(I(a+\xi ^*) = I(a)+I(\xi ^*)\) for all \(a \in B\) and \(\xi \in \mathfrak {R}\).

Proof

We define I on \(B_0(K)\) by condition (a). This also implies that \(I({\tilde{u}}(\{\beta \})) = J(\{\beta \}) = u(\beta )\) for all \(\beta \in {\varDelta }(Z)\), and thus, \(I(1^*) = 1\). The monotonicity (b) of I follows from Lemma 2 and Axiom 6. We indicate that I satisfies (c) and (d).

First, we show \(I(\lambda a) = \lambda I(a)\) for all \(\lambda \in (0, 1]\) and \(a, \lambda a \in B_0(K)\). Let \(y \in {\mathcal {A}}\) be such that \(|y| = n\) and \(a = {\tilde{u}}(y)\), and \(\beta \in {\varDelta }(Z)\) be such that \({\tilde{u}}(\{\beta \}) = 0^*\). (The existence of such y is guaranteed by regularity condition (c) of \(\phi \).) Now, let \(x = \cup _{s \in S_n} \{\lambda \phi (y, s) + (1-\lambda ) \beta \}\). Lemma 2 implies \(x \sim \lambda y \oplus (1-\lambda ) \{\beta \}\), and thus, we have \(J(x) = I({\tilde{u}}(x)) = I(\lambda a + (1-\lambda ) {\tilde{u}}(\{\beta \})) = I(\lambda a)\). Next, let \(\beta '\) be such that \(\{\beta '\} \sim y\). Then, by Axiom 4 and Lemma 1, \(x \sim \lambda y \oplus (1-\lambda ) \{\beta \} \sim \lambda \{\beta '\} \oplus (1-\lambda ) \{\beta \} \sim \{\lambda \beta ' + (1-\lambda ) \beta \}\). That is, \(J(x) = J(\{\lambda \beta ' + (1-\lambda ) \beta \}) = \lambda J(\{\beta '\}) + (1-\lambda ) J(\{\beta \}) = \lambda I({\tilde{u}}(y)) + (1-\lambda ) I({\tilde{u}}(\{\beta \})) = \lambda I(a)\). Accordingly, we obtain \(I(\lambda a) = \lambda I(a)\). Now, define \(I(a) = \frac{1}{\lambda }I(\lambda a)\) for all \(\lambda > 0\) and \(\lambda a \in B_0(K)\). By the positive homogeneity of I on \(B_0(K)\) that we have shown, I(a) is homogeneous of degree one for all \(a \in B\).

Second, we show that I is C-independent. By homogeneity, it suffices to show that \(I(\frac{1}{2} a + \frac{1}{2} \xi ^*) = \frac{1}{2} I(a) + \frac{1}{2} I(\xi ^*)\) for all \(a, \xi ^* \in B_0(K)\). Let \(x \in {\mathcal {A}}\) be such that \(|x| = n\) and \(a = {\tilde{u}}(x), \beta ' \in {\varDelta }(Z)\) be such that \(\{\beta '\} \sim x\), and \(\beta \in {\varDelta }(Z)\) be such that \({\tilde{u}}(\{\beta \}) = \xi ^*\). By Lemma 2 and Axiom 4, \(\cup _{s \in S_n} \{\frac{1}{2} \phi (x, s) + \frac{1}{2} \beta \} \sim \frac{1}{2}x \oplus \frac{1}{2}\{\beta \} \sim \frac{1}{2}\{\beta '\} \oplus \frac{1}{2}\{\beta \} \sim \{\frac{1}{2}\beta ' + \frac{1}{2}\beta \}\), which implies that \(I(\frac{1}{2}a+\frac{1}{2}\xi ^*) = J(\{\frac{1}{2} \beta ' +\frac{1}{2} \beta \}) = \frac{1}{2}J(\{\beta '\}) + \frac{1}{2} J(\{\beta \}) = \frac{1}{2}J(x) + \frac{1}{2} J(\{\beta \}) = \frac{1}{2}I(a) + \frac{1}{2} I(\xi ^*)\).

Finally, we show that I is subadditive. By homogeneity, it suffices to show that \(I(\frac{1}{2}a + \frac{1}{2}b) \le \frac{1}{2}I(a) + \frac{1}{2}I(b)\) for all \(a, b \in B_0(K)\). Let \(x, y \in {\mathcal {A}}\) be such that \(|x| = |y| = n, a = {\tilde{u}}(x)\), and \(b = {\tilde{u}}(y)\). Suppose \(I(a) = I(b)\), that is, \(x \sim y\). Then, it follows from Axiom 5 and Lemma 2 that \(x \succsim \frac{1}{2}x \oplus \frac{1}{2}y \sim \cup _{s \in S_n} \{\frac{1}{2} \phi (x, s) + \frac{1}{2} \phi (y, s)\}\), implying that \(I(a) = \frac{1}{2}I(a)+\frac{1}{2}I(b) \ge I(\frac{1}{2}a+\frac{1}{2}b)\). Next, suppose that \(I(a) > I(b)\). Define \(\xi = I(a) - I(b)\) and \(c = b + \xi ^*\). Note that \(I(c) = I(b + \xi ^*) = I(b) + \xi = I(a)\) (the second equality follows from C-independence). Accordingly, we obtain

$$\begin{aligned} I\left( \frac{1}{2}a+\frac{1}{2}b\right) + \frac{1}{2} \xi = I\left( \frac{1}{2}a+\frac{1}{2}c\right) \le \frac{1}{2}I(a)+\frac{1}{2}I(c) = \frac{1}{2}I(a)+\frac{1}{2}I(b)+ \frac{1}{2}\xi , \end{aligned}$$

which completes the proof. The first and third equalities follow from C-independence, while the second inequality follows from \(I(a) = I(c)\). \(\square \)

Lemma 6

Let I be a monotonic, sublinear, and C-independent functional on B with \(I(1^*) = 1\). Then, there is a closed and convex set \(\mathcal{M}_n\) of finitely additive probability measures over \(S_n\) such that \(I(b) = \max _{\mu \in {\mathcal {M}}_n}\int _{S_n} b d\mu \) for all \(b \in B\). Furthermore, \({\mathcal {M}}_n\) is unique.

Proof

Fix \(b \in B\) such that \(I(b) > 0\). We first show that there is a (finitely additive) probability measure \(\mu _b\) over \(S_n\) such that \(I(b) = \int _{S_n} b d\mu _b\) and \(I(a) \ge \int _{S_n} a d\mu _b\) for all \(a \in B\). Define

$$\begin{aligned} D_1 = \text {co}(\{a \in B : a \ge 1^*\} \cup \{a \in B : a \ge b/I(b)\}) \end{aligned}$$

(where \(\text {co}(\cdot )\) denotes the convex hull of the set \((\cdot )\)) and

$$\begin{aligned} D_2 = \{a \in B : I(a) < 1\}. \end{aligned}$$

Let \(d_1 = \lambda a + (1-\lambda ) a'\), where \(a \ge 1^*, a' \ge b/I(b)\), and \(\lambda \in [0, 1]\). Then, it follows from monotonicity, homogeneity, and C-independence that \(I(d_1) \ge \lambda + (1-\lambda )I(a') \ge 1\), which implies that \(D_1 \cap D_2 = \emptyset \). Furthermore, both \(D_1\) and \(D_2\) have inner points and are convex (the convexity of \(D_2\) follows from the sublinearity of I). Thus, by a separating hyperplane theorem (e.g., Aliprantis and Border 2006), there is a linear functional \(F_b\) and \(\lambda \in \mathfrak {R}\) such that

$$\begin{aligned} F_b(d_1) \ge \lambda \ge F_b(d_2) \end{aligned}$$
(6)

for all \(d_1 \in D_1\) and \(d_2 \in D_2\). Because we clearly have \(\lambda > 0\) (otherwise \(F_b\) must be identically zero), we set \(\lambda = 1\) without loss of generality.

Then, (6) implies that \(F_b(1^*) \ge 1\). In addition, because \(1^*\) is a limit point of \(D_2\), the inverse inequality also holds, and we can conclude that \(F_b(1^*) = 1\). Furthermore, \(F_b\) is nonnegative because, for all nonempty \(E \subseteq S_n\) and the indicator function \(1_E\) of E, we have \(1^*-1_E \in D_2\) and \(F_b(1_E) + F_b(1^*-1_E) = F_b(1^*) = 1\), which implies that \(F_b(1_E) \ge 0\).

Accordingly, because \(F_b\) is a nonnegative linear functional, the Riesz representation theorem (e.g., Aliprantis and Border 2006) implies that there is a finitely additive probability measure \(\mu _b\) such that \(F_b(a) = \int _{S_n} a \hbox {d} \mu _b\) for all \(a \in B\). We show that

\(F_b(a) \le I(a)\) for all \(a \in B\) and \(F_b(b) = I(b)\). First, assume that \(I(a) > 0\). Because \(a/I(a)-(1/m)^* \in D_2\) for all \(m \in {\mathbb {N}}\) and \(F_b(a)\) is continuous with respect to a, we have \(F_b(a) \le I(a)\) from (6). A similar implication for \(I(a) \le 0\) follows from C-independence (set \(\xi \in \mathfrak {R}\) such that \(I(a+\xi ^*) > 0\)). Second, we focus on the special case \(a = b\). Note that \(b/I(b) \in D_1\), and so it follows from (6) that \(F_b(b) \ge I(b)\). Because the previous argument indicates that the inverse inequality also holds, we have \(F_b(b) = I(b)\).

Now, let \({\mathcal {M}}_n \equiv \overline{\text {co}(\{\mu _b : b \in B, I(b) > 0 \})}\), for \(\mu _b\) defined above. It follows from the previous paragraph that \(I(a) \ge \max _{\mu \in \mathcal {M}_n}\int _{S_n} a \hbox {d}\mu \) for all \(a \in B\). It has also been shown that, for all \(a \in B\) such that \(I(a) > 0\), there is a probability measure \(\mu _a \in {\mathcal {M}}_n\) such that \(I(a) =\int _ {S_n} a \hbox {d}\mu _{a}\), which implies that \(I(a) \le \max _ {\mu \in {\mathcal {M}}_n} \int _ {S_n} a \text{ d } \mu \). Applying C-independence, a similar argument also holds for \(I(a) \le 0\).

Finally, we show the uniqueness of \({\mathcal {M}}_n\). Suppose that there are distinct sets \({\mathcal {M}}_n\) and \({\mathcal {M}}'_n\) satisfying the statements of this lemma; that is, \(I({\tilde{u}}(x)) = \max _{\mu \in {\mathcal {M}} _n} \int _{S_n} u(\phi (x, s)) \text{ d }\mu (s)\) and \(I'({\tilde{u}}(x)) = \max _{\mu \in {\mathcal {M}}'_n} \int _{S_n} u (\phi (x, s)) \hbox {d}\mu (s)\) both represent \(\succsim \) for all \(x \in {\mathcal {A}}\) such that \(|x| = n\). Choose \({\tilde{\mu }} \in {\mathcal {M}}_n\setminus {\mathcal {M}}'_n\) (if such \({\tilde{\mu }}\) does not exist, choose \({\tilde{\mu }} \in \mathcal{M}'_n\setminus {\mathcal {M}}_n\) instead and proceed accordingly). Then, because \({\mathcal {M}}'_n\) is convex, the separating hyperplane theorem indicates that there exists \(a \in B\) such that \(\int _{S_n} a \hbox {d}{\tilde{\mu }} > \max _{\mu \in {\mathcal {M}}'_n} \int _{S_n} a \hbox {d} \mu \). It follows from regularity condition (c) that there exists \(y \in \mathcal{A}\) such that \(|y| = n\) and \(a(\cdot ) = u \circ \phi (y, \cdot )\), implying \(I({\tilde{u}}(y)) > I'({\tilde{u}}(y))\), which is a contradiction. \(\square \)

We now conclude the proof of Theorem 1. Because Lemmas 5 and 6 hold for an arbitrary \(n \in {\mathbb {N}}\), Lemmas 1– 6 imply that, by defining \(S = \{s_1, s_2, \ldots \}, {\mathcal {M}} = \cup _{i=n}^{\infty } {\mathcal {M}}_n\), and a regular \(\phi \), we obtain an ASC representation for all \(x \in {\mathcal {A}}\). The essential uniqueness follows from the construction of \(S, {\mathcal {M}}\), and \(\phi \).

In particular, the uniqueness of \(\phi \) with respect to the relevant mental states is shown as follows. Suppose that there exist regular \(\phi \) and \(\phi '\) representing the preference, and \(x, y \in \mathcal{A}\) such that \(\phi (x, {\tilde{s}}) = \phi '(x, {\tilde{s}})\) and \(\phi (y, {\tilde{s}}) \ne \phi '(y, {\tilde{s}})\) for some relevant \({\tilde{s}} \in S\). Without loss of generality, we assume \(|x| = |y|\). By construction, \(\lambda y \oplus (1-\lambda ) x \sim \cup _{s \in S_{|y|}} \{\lambda \phi (y, s) + (1-\lambda ) \phi (x, s)\}\) and \(\lambda ' x \oplus (1-\lambda ') y \sim \cup _{s' \in S_{|y|}} \{\lambda ' \phi '(x, s') + (1-\lambda ') \phi '(y, s')\}\) for all \(\lambda , \lambda ' \in [0, 1]\), and so Axiom 2b implies that \(\lambda '' y \oplus (1-\lambda '') y \sim \{\lambda '' \phi (y, s) + (1-\lambda '') \phi '(y, s'): \phi (x, s) = \phi '(x, s') \ \text {for some} \ s, s' \in S_{|y|}\} \equiv z_{\lambda ''}\) for all \(\lambda '' \in [0, 1]\). However, because \(z_{\lambda ''}\) includes \(\lambda '' \phi (y, {\tilde{s}}) + (1-\lambda '') \phi '(y, {\tilde{s}}), \phi (y, {\tilde{s}}) \ne \phi '(y, {\tilde{s}})\), and \({\tilde{s}}\) is relevant, we conclude without loss of generality that \(z_{\lambda ''}\) is not indifferent to y for all \(\lambda '' \in [0, 1]\), contradicting Axiom 2a.

Finally, the desired representation for all random menus P follows from the affineness of u. \(\square \)

1.2 Proof of Theorem 2

The sufficiency (\({\mathcal {M}}_n = {\varDelta }(S_n)\) implies monotonicity) is straightforward. We show the necessity. First, we prove the following lemma.

Lemma 7

Suppose that \(\succsim \) admits a regular ASC representation \((u, \phi , S, {\mathcal {M}})\) and satisfies monotonicity. Let \(x \in {\mathcal {A}}, {\hat{\beta }} \in x\), and \({\hat{s}} \in S_{|x|}\) be such that \({\hat{\beta }} \succ \beta \) for all \(\beta \in x, \beta \ne {\hat{\beta }}\), and \(\phi (x, {\hat{s}}) = {\hat{\beta }}\). Then, there exists \(\mu \in {\mathcal {M}}_{|x|}\) such that \(\mu ({\hat{s}}) = 1\).

Proof

Let \(x \in {\mathcal {A}}, {\hat{\beta }} \in x\), and \({\hat{s}} \in S_{|x|}\) be such that \({\hat{\beta }} \succ \beta \) for all \(\beta \in x, \beta \ne {\hat{\beta }}\), and \(\phi (x, {\hat{s}}) = {\hat{\beta }}\). Suppose that \(\mu ({\hat{s}}) < 1\) for all \(\mu \in {\mathcal {M}}_{|x|}\). Then, we have \(u({\hat{\beta }}) > V(x) = \max _{\mu \in \mathcal{M}_{|x|}}\int _{S_{|x|}} u(\phi (x, s))d\mu (s)\), or \(\{{\hat{\beta }}\} \succ x\), which contradicts the monotonicity. \(\square \)

Because Lemma 7 applies to all such x and \({\hat{\beta }}\), there exists \(\mu \in {\mathcal {M}}_n\) such that \(\mu (s) = 1\) for any given \(n \in {\mathbb {N}}\) and \(s \in S_n\). Furthermore, because \({\mathcal {M}}_n\) is closed and convex, \({\mathcal {M}}_n \supseteq \overline{\text {co}(\{\mu \in {\varDelta }(S_n): \mu (s) = 1 \text { for some} { \ } s \in S_n\})} = {\varDelta }(S_n)\). \({\mathcal {M}}_n \subseteq {\varDelta }(S_n)\) is straightforward, which concludes the proof. \(\square \)

1.3 Proof of Theorem 3

The sufficiency (\({\mathcal {M}}^1 \supseteq {\mathcal {M}}^2\) implies that \({\succsim }_2\) exhibits a greater preference for commitment to a singleton menu than \(\succsim _1\)) is straightforward.

Conversely, we show that \(\succsim _2\) with a greater preference for commitment to a singleton menu than \(\succsim _1\) implies \({\mathcal {M}}^1 \supseteq {\mathcal {M}}^2\). Assume that there exists \(n \in {\mathbb {N}}\) such that \(\mu ' \in {\mathcal {M}}^2_n \setminus {\mathcal {M}}^1_n\), where \({\mathcal {M}}^i_n \in {\mathcal {M}}^i\) for \(i = 1\) and 2 denote the sets of probability measures over \(S_n\) in each ASC representation. Because \({\mathcal {M}}^1_n\) is convex, the separating hyperplane theorem implies that there exists \(a: S_n \rightarrow \mathfrak {R}\) such that \(\max _{\mu \in {\mathcal {M}}^1_n} \int _{S_n} a \hbox {d}\mu < \int _{S_n} a \hbox {d}\mu '\), and so it follows from regularity condition (c) that there exists some \(x \in {\mathcal {A}}\) such that \(|x| = n\) and \(a(\cdot ) = u \circ \phi (x, \cdot )\). Without loss of generality, we also assume that \(\beta \in x\) exists such that \(u(\beta ) = \int _{S_n} u (\phi (x, s)) \hbox {d}\mu '(s) < \max _{\mu \in {\mathcal {M}}^2_n} \int _{S_n} u(\phi (x, s)) \hbox {d}\mu (s)\). Then,

$$\begin{aligned} \max _{\mu \in {\mathcal {M}}^1_n} \int _{S_n} u(\phi (x, s)) \hbox {d}\mu (s)< & {} \int _{S_n} u(\phi (x, s)) \hbox {d}\mu '(s) \\= & {} u(\beta ) < \max _{\mu \in \mathcal{M}^2_n} \int _{S_n} u(\phi (x, s)) \hbox {d}\mu (s), \end{aligned}$$

implying that \(\{\beta \} \succ _1 x\) and \(x \succ _2 \{\beta \}\), which is a contradiction. \(\square \)

1.4 Proof of Observation 2

Showing that Axiom 2’ implies Axiom 2” is straightforward. Conversely, let \(x, y \in {\mathcal {A}}\) be such that \(|x| \le |y|\). Axiom 2”b (specifically, Definition 1a) implies that there exist a nonempty \(C_0 \subseteq x \times y\) and \(({\tilde{\alpha }}, {\tilde{\beta }}) \in C_0\) such that \(\lambda _0 x \oplus (1-\lambda _0) y \sim \{\lambda _0 \alpha + (1-\lambda _0) \beta : (\alpha , \beta ) \in C_0\}\) for all \(\lambda _0 \in [0, 1]\). Next, Axiom 2”a implies that for \(C_1 \equiv x \times x\) and all \(\lambda _1 \in [0, 1], \lambda _1 x \oplus (1-\lambda _1) x \sim \{\lambda _1 \alpha + (1-\lambda _1) \alpha ': (\alpha , \alpha ') \in C_1\}\), which gives \((\alpha , {\tilde{\alpha }}) \in C_1\) for all \(\alpha \in x\). Axiom 2”a also implies that for \(C_2 \equiv y \times y\) and all \(\lambda _2 \in [0, 1], \lambda _2 y \oplus (1-\lambda _2) y \sim \{\lambda _2 \beta ' + (1-\lambda _2) \beta : (\beta ', \beta ) \in C_2\}\), which obtains \(({\tilde{\beta }}, \beta ) \in C_2\) for all \(\beta \in y\). The iterative application of Axiom 2”b (specifically, Definition 1b) implies that for \(C = \{(\alpha , \beta ): (\alpha , {\tilde{\alpha }}) \in C_1, ({\tilde{\beta }}, \beta ) \in C_2\} = x \times y\) and all \(\lambda \in [0, 1], \lambda x \oplus (1-\lambda ) y \sim \{\lambda \alpha + (1-\lambda ) \beta : (\alpha , \beta ) \in C\} = \lambda x + (1-\lambda ) y\), which concludes the proof. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Koida, N. Anticipated stochastic choice. Econ Theory 65, 545–574 (2018). https://doi.org/10.1007/s00199-016-1025-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-016-1025-9

Keywords

JEL Classification

Navigation