ambiguity aversion, robustness, and the variational representation of...

53
Ambiguity Aversion, Robustness, and the Variational Representation of Preferences Author(s): Fabio Maccheroni, Massimo Marinacci, Aldo Rustichini Source: Econometrica, Vol. 74, No. 6 (Nov., 2006), pp. 1447-1498 Published by: The Econometric Society Stable URL: http://www.jstor.org/stable/4123081 Accessed: 18/03/2009 06:50 Your use of the JSTOR archive indicates your acceptance of JSTOR's Terms and Conditions of Use, available at http://www.jstor.org/page/info/about/policies/terms.jsp. JSTOR's Terms and Conditions of Use provides, in part, that unless you have obtained prior permission, you may not download an entire issue of a journal or multiple copies of articles, and you may use content in the JSTOR archive only for your personal, non-commercial use. Please contact the publisher regarding any further use of this work. Publisher contact information may be obtained at http://www.jstor.org/action/showPublisher?publisherCode=econosoc. Each copy of any part of a JSTOR transmission must contain the same copyright notice that appears on the screen or printed page of such transmission. JSTOR is a not-for-profit organization founded in 1995 to build trusted digital archives for scholarship. We work with the scholarly community to preserve their work and the materials they rely upon, and to build a common research platform that promotes the discovery and use of these resources. For more information about JSTOR, please contact [email protected]. The Econometric Society is collaborating with JSTOR to digitize, preserve and extend access to Econometrica. http://www.jstor.org

Upload: nguyenquynh

Post on 06-May-2018

225 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

Ambiguity Aversion, Robustness, and the Variational Representation of PreferencesAuthor(s): Fabio Maccheroni, Massimo Marinacci, Aldo RustichiniSource: Econometrica, Vol. 74, No. 6 (Nov., 2006), pp. 1447-1498Published by: The Econometric SocietyStable URL: http://www.jstor.org/stable/4123081Accessed: 18/03/2009 06:50

Your use of the JSTOR archive indicates your acceptance of JSTOR's Terms and Conditions of Use, available athttp://www.jstor.org/page/info/about/policies/terms.jsp. JSTOR's Terms and Conditions of Use provides, in part, that unlessyou have obtained prior permission, you may not download an entire issue of a journal or multiple copies of articles, and youmay use content in the JSTOR archive only for your personal, non-commercial use.

Please contact the publisher regarding any further use of this work. Publisher contact information may be obtained athttp://www.jstor.org/action/showPublisher?publisherCode=econosoc.

Each copy of any part of a JSTOR transmission must contain the same copyright notice that appears on the screen or printedpage of such transmission.

JSTOR is a not-for-profit organization founded in 1995 to build trusted digital archives for scholarship. We work with thescholarly community to preserve their work and the materials they rely upon, and to build a common research platform thatpromotes the discovery and use of these resources. For more information about JSTOR, please contact [email protected].

The Econometric Society is collaborating with JSTOR to digitize, preserve and extend access to Econometrica.

http://www.jstor.org

Page 2: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

Econometrica, Vol. 74, No. 6 (November, 2006), 1447-1498

AMBIGUITY AVERSION, ROBUSTNESS, AND THE VARIATIONAL REPRESENTATION OF PREFERENCES

BY FABIO MACCHERONI, MASSIMO MARINACCI, AND ALDO RUSTICHINI1

We characterize, in the Anscombe-Aumann framework, the preferences for which there are a utility function u on outcomes and an ambiguity index c on the set of proba- bilities on the states of the world such that, for all acts f and g,

ftg min(J u(f)dp+ c(p) >min u(g)dp +c(p)

The function u represents the decision maker's risk attitudes, while the index c captures his ambiguity attitudes. These preferences include the multiple priors preferences of Gilboa and Schmeidler and the multiplier preferences of Hansen and Sargent. This provides a rigorous decision-theoretic foundation for the latter model, which has been widely used in macroeconomics and finance.

KEYWORDS: Ambiguity aversion, model uncertainty, robustness.

1. INTRODUCTION

AMBIGUITY HAS BEEN A CLASSIC ISSUE in decision theory since the seminal work of Ellsberg (1961). The fundamental feature of ambiguity pointed out by Ellsberg is that there may be no belief on the states of the world that the decision maker holds and that rationalizes his choices.

A widely used class of preferences that model ambiguity are the multiple priors preferences axiomatized by Gilboa and Schmeidler (1989) (also known as maxmin expected utility preferences). Agents with these preferences rank payoff profiles f according to the criterion

(1) V(f) =min u(f)dp, peCf where C is a given convex subset of the set A of all probabilities on states. The set C is interpreted as a set of priors held by agents, and ambiguity is reflected by the multiplicity of the priors.

1An extended version of this paper was previously circulated with the title "Variational Repre- sentation of Preferences under Ambiguity" ICER Working Paper 5/2004, March 2004. We thank Erio Castagnoli, Rose-Anne Dana, Larry Epstein, Peter Klibanoff, Bart Lipman, Mark Machina, Jianjun Miao, Sujoy Mukerji, Emre Ozdenoren, Ben Polak, and, especially, Andy Postlewaite and four anonymous referees for helpful discussions and suggestions. We also thank several seminar audiences. Part of this research was done while the first two authors were visiting the Depart- ment of Economics of Boston University, which is thanked for its hospitality. They also grate- fully acknowledge the financial support of the Ministero dell'Istruzione, dell'Universith e della Ricerca. Rustichini gratefully acknowledges the financial support of the National Science Foun- dation (Grant SES-04-52477).

1447

Page 3: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1448 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

In the past few years the possibility that agents may not hold a single be- lief on the states of the world has widely informed the research in macroeco- nomics and finance. In particular, there has been a growing dissatisfaction in macroeconomics toward the strong uniformity on the agents' beliefs that is im- posed by the rational expectations hypothesis. Under this hypothesis, all agents share the same probability distribution on some relevant economic phenom- enon and each agent has to be firmly convinced that the model he has adopted is the correct one. This is a strong requirement because agents can have dif- ferent models, each of them being only an approximation of the underlying true model, and they may be aware of the possibility that their model is mis- specified. A weakening of this requirement allows agents to entertain different priors on the economy.

A large part of the research has modeled ambiguity with multiple priors models (see, for instance, Epstein and Wang (1994) and Chen and Epstein (2002)). More recently, a different alternative has been explored, starting with the work of Hansen and Sargent (2000, 2001) which in turn builds on earlier work in the field of robust control in the engineering and optimal control litera- ture. In this robust preferences approach, the objective functions of the agents take into account the possibility that their model q may not be the correct one, but only an approximation. Specifically, agents rank payoff profiles f according to the choice criterion

(2) V(f)=-min

fu(f)dp + OR(p

Iq)},

where R(. II q):A -* [0, oo] is the relative entropy with respect to q (see Sec- tion 4.2 for the definition). Preferences represented by criterion (2) are called multiplier preferences.2

Agents who behave according to this choice criterion are considering the possibility that q may not be the appropriate law that governs the phenom- enon in which they are interested and for this reason they take into account other possible models p. The relative likelihood of these alternative models is measured by their relative entropy, while the positive parameter 0 reflects the weight that agents are giving to the possibility that q might not be the cor- rect model. As the parameter 0 becomes larger, agents focus more on q as the correct model, giving less importance to possible alternative models p (Propo- sition 22).

Hansen and Sargent (2000) have pointed out that this model uncertainty can be viewed as the outcome of ambiguity, possibly resulting from the poor quality of the information on which agents base the choice of their model. For this reason, the motivation behind multiplier preferences is closely connected

2Hansen and Sargent (2001) also considered a class of multiple priors preferences with C = {pe A: R(p II q) < r7}, which they called constraint preferences.

Page 4: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1449

to the motivation that underlies multiple priors preferences. So far, however, this connection has been stated simply as intuitively appealing rather than es- tablished on formal grounds. In particular, no behavioral foundation of the preferences in (2) has been provided.

Here we establish this connection by presenting a general class of prefer- ences with common behavioral features that includes both multiplier and mul- tiple prior preferences as special cases. The nature of the connection between the two main models that we have been discussing so far can be first established formally. The multiple priors criterion (1) can be written as

(3) V(f) = min u(f)dp+8 c(p)},

where Sc: A -- [0, oc] is the indicator function of C (in the sense of convex analysis; see Rockafellar (1970)) given by

c(P) O, if

pEj C,

oc, otherwise.

Like the relative entropy, the indicator function also is a convex function de- fined on the simplex A.

This reformulation clarifies the formal connection with the multiplier pref- erences in (2) and suggests the general representation

(4) V(f) = min u(f) dp + c(p) ,

where c: A -- [0, oo] is a convex function on the simplex. In this paper, we show that the connection is substantive and not formal by establishing that the two models have a common behavioral foundation.

We first axiomatize (Theorem 3) the representation (4) by showing how it rests on a simple set of axioms that generalizes the multiple priors axiomati- zation of Gilboa and Schmeidler (1989). We then show (Proposition 8) how to interpret in a rigorous way the function c as an index of ambiguity aver- sion: the lower is c, the higher is the ambiguity aversion exhibited by the agent. The relative entropy OR(p 11 q) and the indicator function 8c(p) can thus be viewed as special instances of ambiguity indices. The assumptions on behavior for this general representation are surprisingly close to those given by Gilboa and Schmeidler (1989), and result from a simple weakening of their certainty independence axiom, so the similarity between the different representations has a sound behavioral foundation.

Once we have established this common structure, we can analyze the re- lationship between ambiguity aversion and probabilistic sophistication. For example, the multiplier preferences used by Hansen and Sargent are proba- bilistically sophisticated (and the same is true for their constraint preferences).

Page 5: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1450 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

Is this a fundamental property of robust preferences that makes them pro- foundly different from multiple priors? In the common structure provided by the representation (4), the answer is simple and clear. The preferences in (4) are probabilistically sophisticated if and only if the ambiguity index c is symmetric. For multiple priors preferences, this property translates into a sym- metry property of the set C. For multiplier preferences (and for constraint preferences), the condition is reflected by the symmetry of the relative entropy.

The property of being probabilistically sophisticated is therefore a property of the specific ambiguity index, not of the model. Being a symmetry property, this property is fragile: small perturbations destroy it and produce Ellsberg- type behavior. Even if one adopts the view that such behavior is necessary for ambiguity aversion, the preferences in (4) are typically ambiguity averse (in the precise sense of Theorem 14).

Although our original motivation came from multiple priors and multiplier preferences, the class of preferences we axiomatize goes well beyond these two classes of preferences. In particular, multiplier preferences are a very special example of a new class of preferences, called divergence preferences, that we introduce and study in this paper. These preferences are able to accommodate Ellsberg-type behavior and, unlike multiple priors preferences, they are in gen- eral smooth (Proposition 23), an important feature for applications. This new class of preferences can provide a tractable alternative to multiple priors pref- erences in economic applications that deal with ambiguity, and the works of Hansen and Sargent can be viewed as an instance of this (see Theorem 16).

This claim is further substantiated by the observation that divergence preferences include as a special case a third classic class of preferences-the mean-variance preferences of Markowitz (1952) and Tobin (1958). Recall that mean-variance preferences are represented by the preference functional

(5) V(f) = f dq - Var(f).

In Theorem 24, we show that, on the domain of monotonicity of V, the equality

f dq- Var(f) = min f dp + OG(p II q) 20 ~ ~p(-A I

holds, where G(- I q): A -- [0, oo] is the relative Gini concentration index (see Section 4.2.2 for the definition). As a result, the mean-variance preference functional (5) is a special case of our representation (4). Interestingly, the asso- ciated index of ambiguity aversion is the relative version of the classic Gini in- dex. After Shannon's entropy, a second classic concentration index thus comes up in our analysis.

Summing up, in this paper we generalize a popular class of preferences that deal with ambiguity-the multiple priors preferences-and in this way we are

Page 6: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1451

able to introduce divergence preferences, a large new class of preferences un- der ambiguity that are in general smooth and that include, as special cases, two widely used classes of preferences, the multiplier preferences of Hansen and Sargent and the mean-variance preferences of Markowitz and Tobin. We are thus able to provide both a rigorous decision-theoretic foundation on two widely used classes of preferences and a setting in which the two most clas- sic concentration indices, Shannon's entropy and Gini's index, have a natural decision-theoretic interpretation. We finally characterize the preferences in this class that are probabilistically sophisticated and show that this property coincides with a symmetry property of the cost function. This property is frag- ile and so the preferences we characterize are typically not probabilistically sophisticated.

1.1. Ambiguity Aversion

In addition to Schmeidler's (1989) original notion based on preference for randomization (see Axiom A.5 in Section 3), there are two main notions of ambiguity aversion in the literature: those proposed by Epstein (1999) and by Ghirardato and Marinacci (2002). The key difference in the two approaches lies in the different notion of ambiguity neutrality they use: while Ghirardato and Marinacci (2002) identify ambiguity neutrality with subjective expected utility, Epstein (1999) more generally identifies ambiguity neutrality with prob- abilistic sophistication. In a nutshell, Ghirardato and Marinacci (2002) claim that, unless the setting is rich enough, probabilistically sophisticated prefer- ences may be compatible with behavior that intuitively can be viewed as gen- erated by ambiguity. For this reason, they consider only subjective expected utility preferences as ambiguity neutral preferences.3

For the general class of preferences we axiomatize, the relative merits of these two notions of ambiguity aversion are the same as for the special case represented by multiple priors preferences. Although in the paper we adopt the view and terminology of Ghirardato and Marinacci (2002), the appeal of our analysis does not depend on this choice; in particular, we expect that the ambiguity features of our preferences can be studied along the lines of Epstein (1999), in the same way as it has been done in Epstein (1999) for multiple priors preferences.

To clarify this issue further, in Section 3.5 we study the form that probabilis- tic sophistication takes in our setting. As we already mentioned, we show that our preferences are probabilistically sophisticated once their ambiguity indices satisfy a symmetry property. As a result, probabilistic sophistication is not pe- culiar to some particular specification of our preferences, but, to the contrary,

3We refer to Epstein (1999) and Ghirardato and Marinacci (2002) for a detailed presentation and motivation of their approaches. Notice that the notion of ambiguity aversion in Ghirardato and Marinacci (2002) is what provides a foundation for the standard comparative statics exercises in ambiguity for multiple priors preferences that are based on the size of the set of priors.

Page 7: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1452 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

all examples of our preferences will include special cases of probabilistically sophisticated preferences. For instance, both multiplier and mean-variance preferences are easily seen to be examples of probabilistically sophisticated di- vergence preferences, although Example 17 shows that this is not the case for general divergence preferences, even for those that are very close to multiplier and mean-variance preferences.

We close by briefly mentioning a possible alternative interpretation of our preferences. Consider an agent who has to make choices with only limited information and without a full understanding of what is going on. Some re- cent psychological literature (e.g., Keren and Gerritsen (1999), Kiihberger and Perner (2003)) suggests that in this case the agent may behave as if he were playing against an informed opponent who might take advantage of this uncer- tainty and turn it against him.

This is a psychological attitude that can be relevant in many choice sit- uations, including Ellsberg-type choice situations, and our preference func- tional (4) can be viewed as modeling such psychological treat. In fact, agents who rank payoff profiles according to (4) can be viewed as believing they are playing a zero-sum game against (a malevolent) Nature. This view, however, has not been firmly established in the psychological and neuroscience litera- tures, and it is the object of current research (see Hsu, Bhatt, Adolphs, Tranel, and Camerer (2005) and Rustichini (2005)); for this reason, we do not expati- ate further on this interpretation.

The paper is organized as follows. After introducing the setup in Section 2, we present the main representation result in Section 3. In the same section, we discuss the ambiguity attitudes featured by the preferences we axiomatize and we give conditions that make them probabilistically sophisticated. In Section 4, we study two important examples of our preferences, that is, the multiple priors preferences of Gilboa and Schmeidler (1989) and divergence preference, a new class of preferences that includes, as special cases, multiplier preferences and mean-variance preferences. Proofs and related material are collected in the Appendixes.

2. SETUP

Consider a set S of states of the world, an algebra X of subsets of S called events, and a set X of consequences. We denote by .T the set of all the (simple) acts: finite-valued functions f :S -- X, which are X-measurable. Moreover, we denote by Bo(2) the set of all real-valued X-measurable simple functions, so that u(f) E Bo(V) whenever u: X -- R, and we denote by B(s) the supnorm closure of BO(X).

Given any x E X, define x E T to be the constant act such that x(s) = x for all s E S. With the usual slight abuse of notation, we thus identify X with the subset of the constant acts in F. If f E F, x E X, and A E I, we denote by xAf E FT the act that yields x if s E A and f(s) if s A.

Page 8: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1453

We assume additionally that X is a convex subset of a vector space. For instance, this is the case if X is the set of all the lotteries on a set of prizes, as happens in the classic setting of Anscombe and Aumann (1963). Using the linear structure of X, we can define as usual for every f, g E T and a E [0, 1] the act af + (1 - a)g E F, which yields oaf(s) + (1 - a)g(s) E X for every s E S.

We model the decision maker's preferences on Tf by a binary relation >. As usual, >- and - denote, respectively, the asymmetric and symmetric parts of >-. If f E F, an element xf E X is a certainty equivalent for f if f - xf.

3. REPRESENTATION

3.1. Axioms

In the sequel we make use of the following properties of >:

AXIOM A.1-Weak Order: If f, g, h E ?, (a) either f > g or g >- f, and (b) f > g and g > h imply f > h.

AXIOM A.2-Weak Certainty Independence: If f, g E ?, x, y E X, and a E (0, 1),

af + (1 - a)x > Lag + (1 - a)x

= of + (1 - aa)y c ag + (1 - a)y.

AXIOM A.3 -Continuity: If f, g, h E Y, the sets {a E [0, 1]: af + (1 - a)g g> h} and {a e [0, 1]: h > af + (1 - a)g} are closed.

AXIOM A.4-Monotonicity: If f, g E ? and f(s) -

g(s) for all s E S, then f rg.

AXIOM A.5-Uncertainty Aversion: If f, g E ? and a E (0, 1),

f~g = af +(1-a)g rjf.

AXIOM A.6-Nondegeneracy: f >- g for some f, g E F.

Axioms A.1, A.3, A.4, and A.6 are standard assumptions. Axioms A.3 and A.6 are technical assumptions, while Axioms A.1 and A.4 require preferences to be complete, transitive, and monotone. The latter requirement is basically a state-independence condition, which says that decision makers always (weakly) prefer acts that deliver statewise (weakly) better payoffs, regardless of the state where the better payoffs occur. If a preference relation >- satisfies Axioms A.1, A.3, and A.4, then each act f E F admits a certainty equivalent xf E X. (See the proof of Lemma 28 in Appendix B.)

Page 9: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1454 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

Axiom A.5, due to Schmeidler (1989), is a smoothing axiom that can be in- terpreted as an ambiguity aversion axiom, as discussed at length in Gilboa and Schmeidler (1989), Schmeidler (1989), Epstein (1999), and Ghirardato and Marinacci (2002).

Axiom A.2 is a weak independence axiom-a variation of an axiom of Gilboa and Schmeidler (1989). It requires independence only with respect to mixing with constant acts, provided the mixing weights are kept constant. Axiom A.2 is weaker than the original axiom of Gilboa and Schmeidler (1989), and it is this weakening that makes it possible to go beyond the multiple priors model. Be- cause of its importance for our derivation, we devote the rest of this subsection to Axiom A.2.

Consider the following stronger version of Axiom A.2:

AXIOM A.2'-Certainty Independence: If f, g e T, x e X, and a E (0, 1), then

f g ' af + (1 - a)x ag + (1 - a)x.

Axiom A.2' is the original axiom of Gilboa and Schmeidler (1989). The next lemma shows how it strengthens Axiom A.2.

LEMMA 1: A binary relation - on F satisfies Axiom A.2' if and only if, for all f,g , x, y E X, anda, E (0, 1],

af + (1 - a)x > ag + (1 - a)x

= pf + (1 - P)y 3pg + (1 - 3)y.

Axiom A.2 is therefore the special case of Axiom A.2' in which the mixing coefficients a and /3 are required to be equal. At a conceptual level, Lemma 1 shows that Gilboa and Schmeidler's (1989) certainty independence axiom actu- ally involves two types of independence: independence relative to mixing with constants and independence relative to the weights used in such mixing. Our Axiom A.2 retains the first form of independence, but not the second. In other words, we allow for preference reversals in mixing with constants unless the weights themselves are kept constant.

This is a significant weakening of the certainty independence axiom, and its motivation is best seen when the weights a and /3 are very different, say a is close to 1 and p is close to 0. Intuitively, acts af + (1 - a)x and ag + (1 - a)x can then involve far more uncertainty than acts Pf + (1 - P)y and

p/g + (1 - P/3)y, which are almost constant acts. As a result, we expect that, at least in some situations, the ranking between the genuinely uncertain acts af + (1 - a)x and ag + (1 - a)x can well differ from that between the almost constant acts Pf + (1 - P)y and pg + (1 - P)y.

Page 10: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1455

Needless to say, even though we believe that such reversals can well occur (from both a positive and normative standpoint), the only way to test them, and so test the plausibility of Axioms A.2 and A.2', is by running experiments. This is possible because both Axioms A.2 and A.2' have clear behavioral implica- tions. For instance, the following (thought) experiment gives a simple testable way to compare Axioms A.2 and A.2', and running this type of experiments will be the subject of future research.

EXAMPLE 2: Consider an urn that contains 90 black and white balls in un- known proportion, and the bets (payoffs are in dollars)

t > 0 Black White

ft t t gt 3t 0.01t

that is, f, pays t dollars whatever happens, while g, pays 3t dollars if a black ball is drawn and t cents otherwise. For example,

10 Black White 104 Black White

fio 10 10 and f10o4 10,000 10,000 g10 30 0.1 g104 30,000 100

Assume the decision maker's preferences satisfy Axioms A.1 and A.3-A.6, and she displays constant relative risk aversion y E (0, 1).4 If her preferences satisfy Axiom A.2', then

either ft~, g, for all t or gt - f, for all t.

If, in contrast, her preferences only satisfy Axiom A.2 and not Axiom A.2', there might exist a threshold t such that

ft g, forallt>t and gt f, for all t < t.

This reversal is compatible with Axiom A.2, but it would reveal a violation of Axiom A.2'.

We close by observing that, in terms of preference functionals, by Theorem 3 and Proposition 19 all preference functionals (4) satisfy Axiom A.2 and violate Axiom A.2' unless they reduce to the multiple priors form (1).

4Notice that, for all t > - > 0 and y, there exist f, g, x, y, a, p as discussed in the text such that af + (1 - a)x - ft, ag + (1 - a)x - gt, Pf + (1 - P)y - f,, and pg + (1 - P)y - g,.

Page 11: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1456 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

3.2. Main Result

We can now state our main result, which characterizes preferences that sat- isfy Axioms A.1-A.6. Here A = A(2) denotes the set of all finitely additive probabilities on I endowed with the weak* topology5 and c: A -- [0, 00] is said to be grounded if its infimum value is zero.

THEOREM 3: Let >. be a binary relation on F. The following conditions are equivalent:

(i) The relation >- satisfies Axioms A.1-A.6. (ii) There exists a nonconstant affine function u : X - IR and a grounded, con-

vex, and lower semicontinuous function c: A -+ [0, oo] such that, for all f, g e F,

(6) f jg g

min(f u(f)dp + c(p) > min(f u(g) dp + c(p)).

For each u there is a (unique) minimal c* : A - [0, oo] that satisfies (6) and is given by

(7) c*(p) = sup(u(xf) - u(f) dp). fGT(

The representation (6) involves the minimization of a convex lower semicon- tinuous function, which is the most classic variational problem. This motivates the following definition:

DEFINITION 4: A preference - on F is called variational if it satisfies Ax-

ioms A.1-A.6.

By Theorem 3, variational preferences can be represented by a pair (u, c*). From now on, when we consider a variational preference, we will write u and c* to denote the elements of such a pair. Next we give the uniqueness properties of this representation.

COROLLARY 5: Two pairs (uo, c*) and (u, c*) represent the same variational preference

- as in Theorem 3 if and only if there exist a > 0 and P/ IR such that u = auo + p and c* =

aco. In Theorem 3 we saw that c* is the minimal nonnegative function on A for

which the representation (6) holds. More is true when u(X) = {u(x): x e X} is unbounded (either below or above):

5That is, the o(A(X,),

Bo(2)) topology where a net {PdideD converges to p if and only if pd(A) - p(A) for all A

• .

Page 12: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1457

PROPOSITION 6: Let > be a variational preference with u(X) unbounded. Then the function c* defined in (7) is the unique nonnegative, grounded, convex, and lower semicontinuous function on A for which (6) holds.

As shown in Lemma 29 in Appendix B, the assumption that u(X) is un- bounded (above or below) is equivalent to the following axiom (see Kopylov (2001)).

AXIOM A.7 -Unboundedness: There exist x >- y in X such that, for all a E (0, 1), there exists z E X that satisfies either y >- az + (1 - a)x or az + (1 - a)y >-x.

We call the variational preferences that satisfy Axiom A.7 unbounded.

3.3. Ambiguity Attitudes

We now study the ambiguity attitudes featured by variational preferences. We follow the approach proposed in Ghirardato and Marinacci (2002), to which we refer for a detailed discussion of the notions we use.

Begin with a comparative notion: given two preferences >1 and >2, say that (1 is more ambiguity averse than (2 if, for all f E F and x E X,

(8) f >- x = f f2 x. To introduce an absolute notion of ambiguity aversion, as in Ghirardato and Marinacci (2002) we consider subjective expected utility (SEU) preferences as benchmarks for ambiguity neutrality. We then say that a preference relation b is ambiguity averse if it is more ambiguity averse than some SEU preference.

We now apply these notions to our setting. The first thing to observe is that variational preferences are always ambiguity averse.

PROPOSITION 7: Each variational preference is ambiguity averse.

Since variational preferences satisfy Axiom A.5 and the choice rule that re- sults from (6) is a maxmin rule, intuitively it is not surprising that variational preferences always display a negative attitude toward ambiguity. Proposition 7 makes this intuition precise.

Next we show that comparative ambiguity attitudes for variational prefer- ences are determined by the function c*. Here u1 m u2 means that there exist a > 0 and /3 E R such that ul = au2 + / .

PROPOSITION 8: Given two variational preferences >- and -2, the following conditions are equivalent:

(i) The relation -1 is more ambiguity averse than j2.

Page 13: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1458 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

(ii) ul u2 and c <_

c* (provided ul = u2).

Given that u1a U2, the assumption ul = u2 is just a common normalization of the two utility indices. Therefore, Proposition 8 says that more ambiguity averse preference relations are characterized, up to a normalization, by smaller functions c*. Therefore, the function c* can be interpreted as an index of ambi- guity aversion.

We now give few simple examples that illustrate this interpretation of the function c*.

EXAMPLE 9: By Proposition 8, maximal ambiguity aversion is characterized by c*(p) = 0 for each p e A. In this case, (6) becomes

f g min u(f) dp > min f u(g)dp,

that is,

f -g min u(fS(s)) > min u(g(s)), sES seS

a form that clearly reflects extreme ambiguity aversion.

EXAMPLE 10: Minimal ambiguity aversion corresponds here to ambiguity neutrality because, by Proposition 7, all variational preferences are ambiguity averse. Therefore, the least ambiguity averse functions c* are those associated with SEU preferences. As will be shown in Corollary 20, when a SEU prefer- ence is unbounded, c* takes the stark form

c(p = , if p = q, =c0, otherwise,

where q is the subjective probability associated with the preference.

EXAMPLE 11: Denote by c* the ambiguity neutral index of Example 10 and denote by

cm the maximal ambiguity index of Example 9. Given a e (0, 1),

suppose the right-hand side of (6) is

(1 - a)u(f) dq+ amins u(f(s))

> (1 -a) u(g)dq + aminu(g(s)), f sSE which is the well known e-contaminated model. In this case,

c*(p) = min (1 - a)cq(P2)+ Cm(P):(1 - a)p2 + ap = p}

P1,P2P2EA

= 5(1-a)q+aA(P),

Page 14: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1459

and this is a simple example of an index c* that does not display extreme ambi- guity attitudes.

According to Proposition 8, variational preferences become more and more (less and less, resp.) ambiguity averse as their ambiguity indices become smaller and smaller (larger and larger, resp.). It is then natural to wonder what happens at the limit, when they go to 0 or to oc. The next result answers this question. In reading it, recall that

c1+• < c, implies dom c, dom c,+1 and arg min c, C arg min cn+1.6

PROPOSITION 12: (i) If c, 4 and limn cn~(p) = 0 whenever this limit is finite, then

(9) lim minf u(f)dp + cn(p) = min u(f ) dp n-+0o pEAj J pEUn dom-c CJ

for all f E 'F. (ii) If c , and limn cn (p) = 0c whenever this limit is not 0, then

(10) lim minf u(f)dp+cn(p) = min u(f ) dp n-+ c pEA I PEn argmincnJ

for all f E F.

Proposition 12 shows that the limit behavior of variational preferences is described by multiple priors preferences, but the size of the sets of priors they feature is very different. In fact, in (9) the relevant set of priors is given by U, dom c,, whereas in (10) it is given by the much smaller set n,, argmin c,. For example, Proposition 22 will show that for an important class of variational preferences, the set ", arg min c, is just a singleton, so that the limit preference in (10) is actually a SEU preference, while the set U, dom c, is very large.

We close with few remarks. First, observe that Lemma 32 in Appendix B shows that the set that Ghirardato and Marinacci (2002) call benchmark measures-those probabilities that correspond to SEU preferences less am- biguity averse than --is given here by arg min c* = {p E A: c*(p) = 0}.

Second, notice that by standard convex analysis results (see Rockafellar (1970)), Example 11 can be immediately generalized as follows: the ambigu- ity index of a convex combination of preference functionals that represent un- bounded variational preferences is given by the infimal convolution of their ambiguity indices.

6The term domc denotes the effective domain {c < oc} of c, whereas argminc, = {p E A: c,(p) = 0} because c, is grounded. Observe that dom cn represents the set of all probabilities that the decision maker considers to be relevant when ranking acts using the ambiguity index cn, whereas the smaller set argmin cn contains only the probabilities that are getting the highest weight by this decision maker.

Page 15: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1460 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

Finally, Proposition 12 is an example of a result about the limit behavior of sequences of variational preferences

V. (f) min u,(f)

dp + c,(p)),

a type of result that can be important in some applications to enable dealing with stability issues. These limit results involve the convergence of minima, a classic problem in variational analysis. A noteworthy feature of variational preferences is that the lower semicontinuity of the functions f u,(f) dp+c,(p) on A makes it possible to use the powerful de Giorgi-Wijsman theory of F con- vergence (often called epiconvergence; see, e.g., Dal Maso (1993)) to study the behavior of the sequence {V,(f)}, and, for example, to determine the conditions under which it converges to some preference functional V(f) =

minpEa(f u(f) dp + c(p)) for suitable limit functions u and c.

3.4. An Extension: Countable Additivity

In Theorem 3 we considered the set A of all finitely additive probabilities. In applications, however, it is often important to consider countably additive probabilities, which have very convenient analytical properties. For example, we will see momentarily that this is the case for divergence preferences, and therefore for the multiplier preferences of Hansen and Sargent (2001) and for the mean-variance preferences of Markowitz (1952) and Tobin (1958). For this reason, here we consider this technical extension.

Fortunately, in our setting we can still use the monotone continuity axiom introduced by Arrow (1970) to derive a SEU representation with a countably additive subjective probability (see Chateauneuf, Maccheroni, Marinacci, and Tallon (2005)).

AXIOM A.8-Monotone Continuity: If f, g E FT, x E X, {En,1> E with E D2 E2 D ... and n,,> E, = 0, then f >- g implies that there exists no > 1 such that xE,,f >- g.

Next we state the countably additive version of Theorem 3. Here Ad = AU (X) denotes the set of all countably additive probabilities defined on a ur-algebra Z, while A((q) denotes the subset of Ad that consists of all probabilities that are

absolutely continuous with respect to q; i.e., A'((q) = {p E d: p << q)}.

THEOREM 13: Let > be an unbounded variational preference. The following conditions are equivalent:

(i) The relation - satisfies Axiom A.8.

(ii) For each t > 0, {p E A : c*(p) < t} is a (weakly) compact subset of AU.

Page 16: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1461

In this case, there exists q E Az such that, for all f, g E F,

(11) f g

qmin

u(f)dp+c*(p)> Amin u(g)dp+c*(p). pEA'Or'(q) u(f PEAO"(q) (f~~~(S(R Lemma 30 in Appendix B shows that even when the preference is not un-

bounded, Axiom A.8 still implies the countable additivity of the probabilities involved in the representation.

In view of these results, we call the variational preferences that satisfy Ax- iom A.8 continuous.

3.5. Probabilistic Sophistication In this section we characterize variational preferences that are probabilis-

tically sophisticated, an important property of preferences introduced by Machina and Schmeidler (1992) that some authors (notably Epstein (1999)) identify with ambiguity neutrality (or absence of ambiguity altogether).

Our main finding is that what makes a variational preference probabilisti- cally sophisticated is a symmetry property of the ambiguity index. As a re- sult, all classes of variational preferences (for example, multiple priors and divergence preferences) contain a subclass of probabilistically sophisticated preferences characterized by a suitable symmetry property of the associated ambiguity indices.

Specifically, given a countably additive probability q on the o--algebra Z, a preference relation > is probabilistically sophisticated (with respect to q) if, given f and g in F,

q(s E S: f(s) = x) = q(s E S: g(s) = x) for all x E X = f ~ g.

For example, if S is finite and q is uniform, then probabilistic sophistication amounts to permutation invariance (i.e., each act is indifferent to all its per- mutations).

A preference - satisfies (first order) stochastic dominance (with respect to q)

if, given f and g in F,

q(s E S: f(s) -

x) < q(s c S:g(s) - x) for all x X E = f g.

Notice that a preference that satisfies stochastic dominance is probabilistically sophisticated.

To characterize probabilistic sophistication, we need to introduce a few well known notions from the theory of stochastic orders (see, e.g., Chong and Rice (1971) and Schmeidler (1979)). Define a partial order >-,, on AZ(q) by

p >cx p' iff Jf( dq> (dpf)dq f 'dq f' J' dq j~

Page 17: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1462 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

for every convex function 0 on R. This is the so-called convex order on proba- bility distributions, and by classic results of Rothschild and Stiglitz (1970) and Marshall and Olkin (1979) we have p >-~x p' when the "masses" dp(s) are more scattered with respect to dq(s) than the masses dp'(s).

An important property of the convex order >-c is that its symmetric part ~cx coincides with the identical distribution of the densities with respect to q. That is, given p and p' in A "(q),

(12) p ex, ' iff

q seS: dp (s)<t =q( sS: dp' (s)<t VteR.

dq dq

For p in AU(q), the set O(p) =- {p' e A (q): p' -,, p} is called the orbit of p. A function c: A -+ [0, oo] is rearrangement invariant (with respect to q) if

dom cC AO(q) and, given p and p' in Ad(q),

P cx pi c(p) = c(p'),

whereas it is Shur convex (with respect to q) if dom c C A'(q) and, given p and p' in A (q),

(13) p pc' p' = c(p) > c(p').

A Shur convex function is clearly rearrangement invariant, whereas the con- verse is, in general, false. A subset C of Ao(q) is Shur convex (with respect to q) if its indicator function 8c is Shur convex.7 Finally, we say that q in Ad is adequate if either q is nonatomic or S is finite and q is uniform.

We can now state our characterization result, which shows that rearrange- ment invariance is the symmetry property of ambiguity indices that character- izes probabilistically sophisticated preferences.

THEOREM 14: Let > be a continuous unbounded variational preference. If q E Ao is adequate, then the following conditions are equivalent (with respect to q):

(i) The relation - satisfies stochastic dominance. (ii) The relation > is probabilistically sophisticated. (iii) The function c* is rearrangement invariant. (iv) The function c* is Shur convex.

Moreover, for any variational preference >, (iv) implies (i) even if q is not ade- quate.

7That is, {p' e A•(q): p' -<cx p} C for every p E C.

Page 18: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1463

The proof of Theorem 14 builds on the results of Luxemburg (1967) and Chong and Rice (1971), as well as on some recent elaborations of these results provided by Dana (2005).

As a first straightforward application of Theorem 14, observe that if a set of priors is Shur convex, then the corresponding (continuous) multiple priors preferences are probabilistically sophisticated and the converse is true in the adequate case. More is true in this case: multiple priors preferences are prob- abilistically sophisticated if and only if the indicator functions Sc of their sets of priors C are rearrangement invariant, that is, if and only if the sets C are orbit-closed (O(p) c C for every p e C).8

To further illustrate Theorem 14, we now introduce a new class of variational preferences that plays an important role in the rest of the paper. As before, as- sume there is an underlying probability measure q E Ad. Given a X-measurable function w:S --+ R with infs,s w(s) > 0 and f w dq = 1, and a convex contin- uous function &P : R+ *R+ such that 0 (1) = 0 and limt,,, (t)/t = 00, the

w-weighted 0 divergence of p E A with respect to q is given by

(14) Dw(p II q)= w(s) (s)) dqr(s), ifspe

A(q),

00, otherwise.

Here w is the (normalized) weighting function, and in the special case of uni- form weighting w(s) = 1 for all s E S, we just write DO(p II q) and we get back to standard divergences, which are a widely used concept of "distance between distributions" in statistics and information theory (see, e.g., Liese and Vajda (1987)). The two most important divergences are the relative entropy (or Kullback-Leibler divergence) given by 4 (t) = t In t - t + 1 and the relative Gini concentration index (or X2 divergence) given by 0 (t) = 2-' (t - 1)2

The next lemma collects the most important properties of weighted diver- gences.

LEMMA 15: A weighted divergence Dw(. 1 q) : A -- [0, oo] is a grounded, con- vex, and lower semicontinuous function, and the sets

(15) {p E A: D (p II q) < t}

are (weakly) compact subsets of Ad(q) for all t E R. Moreover, Do(. II q) :A --+ R is Shur convex whenever w is uniform.

8This fact can be used to provide an alternative derivation of some of the results of Marinacci (2002), which showed that multiple priors preferences that are probabilistically sophisticated re- duce to subjective expected utility when there exists even a single nontrivial unambiguous event. The direct proofs in Marinacci (2002) are, however, shorter and more insightful for the problem with which that article was dealing.

Page 19: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1464 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

Thanks to the foregoing properties, preferences represented by the func- tional

(16) V(f) = minf u(f)dp +

OD(p I

q)}

where 0 > 0 and u: X -+ IR is an affine function, belong to the class of varia- tional preferences. In view of their importance, we call them divergence prefer- ences; that is, > on Y is a divergence preference if

ft >g min

u(f) dp + OD(p II q)

> minEm u(g)dp + OD(p | q) }

THEOREM 16: Suppose u(X) is unbounded. Then divergence preferences are continuous variational preferences with index of ambiguity aversion given by

c*(p) = OD=(p || q) Vp EA.

In particular, these preferences are probabilistically sophisticated whenever w is uniform.

Divergence preferences are a very important class of variational preferences and we further study them in the next section. Unlike multiple priors prefer- ences, they are in general smooth (see Proposition 23), a noteworthy feature for applications.

For a finite state space S, X = R, u(t) = t, and w uniform, some classes of divergence preferences have been considered by Ben-Tal (1985) and Ben-Tal, Ben-Israel, and Teboulle (1991). In the next section, we will study two signifi- cant examples of divergence preferences, corresponding to the relative entropy and to the relative Gini concentration index. Here it is important to observe that, by Theorem 16, all divergence preferences represented by

V(f)

=Iqmm

in u(f)dp+OD(p( lq)p ,

are examples of probabilistically sophisticated variational preferences.' However, the next example shows that even under minimal nonuniformi-

ties of the weighting function, divergence preferences are in general not prob- abilistically sophisticated and they exhibit Ellsberg-type behavior. Therefore,

'Analogously, Theorem 14 and Lemma 15 show that all multiple priors preferences with sets of priors {p e A'I(q):D4(p II q) < r}j are probabilistically sophisticated. These preferences include the constraint preferences of Hansen and Sargent (2001), mentioned in footnote 2.

Page 20: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1465

the probabilistic sophistication of divergence preferences crucially depends on the uniformity of the weight w.

EXAMPLE 17: Consider a standard Ellsberg three color urn, with 30 red balls and 60 balls either green or blue. As usual, consider the bets:

Red Green Blue

fR 1 0 0 fG 0 1 0

fRuB 1 0 1 fGUB 0 1 1

where fR pays one dollar if a red ball is drawn and nothing otherwise, fG pays one dollar if a green ball is drawn and nothing otherwise, and so on. As is well known, Ellsberg (1961) argued that most subjects rank these acts as

(17) fR >f and fRUB-<fGUB.

Consider a decision maker who has divergence preferences represented by the preference functional V given by (16). Here it is natural to consider a uniform q on the three states. Without loss of generality, set u(0) = 0 and u(1) = 1. By Theorem 16, when w is uniform, it cannot be the case that V(fR) > V(fG) and V(fRUB) < V(fGUB). However, consider the weighting func- tion w:{R, G, B} -+ IR given by

w(R) = 1.01, w(G) = 0.99, and w(B) = 1,

which is only slightly nonuniform. If we set 0 = 1 and take either the weighted relative entropy 0 (t) = t ln t - t + 1 or the weighted Gini relative index P(t) = 2-' (t - 1)2, then some simple computations (available upon request) show that

V(fR) > V(fG) and V(fRUB) < V(fGUB),

thus delivering the Ellsberg pattern (17).

3.6. Smoothness

Most economic models are based on the optimization of an objective func- tion. When this function is differentiable, solving the optimization problem is easier and the solution has appealing properties. There is a well established set of techniques (first-order necessary conditions, envelope theorems, implicit function theorems, and so on) that are extremely useful in both finding a solu- tion of an optimization problem and characterizing its properties. For example, they make it possible to carry out comparative statics exercises, a key feature in most economic models.

Page 21: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1466 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

In view of all this, it is important to study whether our variational preference functionals are differentiable. In this section (see Theorem 18), we fully char- acterize the differentiability properties of variational preference functionals and show that they are adequate for economic applications.

Throughout this section, we assume that X is the set of all monetary lot- teries, that is, the set of all finitely supported probability measures on R. An act f of F is monetary if f(s) is a degenerate lottery for every s e S, that is, if f(s) e IR (with the usual identification of z IR with the degenerate lottery dz E X). The set of all monetary acts can thus be identified with Bo(2).

We consider a variational preference functional

V(f) = min u(f) dp + c*(p)}

restricted to Bo(I), that is, restricted to monetary acts. We also make the stan- dard assumption that the associated utility function u is concave (thus reflect- ing risk aversion), strictly increasing, and differentiable on R.

To state our results, we need some standard notions of calculus in vector spaces (see Rockafellar (1970) and Phelps (1992)). Given f E Bo(I), the di- rectional derivative of V : Bo(E) -+ R at f is the functional V'(f; .): Bo()

-- R

defined by

V(f + th)- V(f) V'(f; h) = lim hV(f

+ th) - V(f Bo(). t4o t

The functional V is (Gateaux) differentiable at f if V'(f; -) is linear and sup- norm continuous on Bo(2). In this case, V'(f; .) is the (Gateaux) differential of V at f.

The superdifferential of V at f is the set 8V(f) of all linear and supnorm continuous functionals L: Bo(1) -*+ R such that

V'(f; h) < L(h) V heBo(X).

In particular, dV(f) is a singleton if and only if V is differentiable at f. In this case, dV(f) consists only of the differential V'(f; -) ; i.e., dV(f) = {V'(f; -)}.

We are now ready to state our result. It provides an explicit formula for the superdifferential dV(f) at every f E Bo(E) and a full characterization of differentiability, along with an explicit formula for the differential.

THEOREM 18: For all f e Bo(e),

(18) dV(f)-= u'(f)dr:reargmin u(f) dp+c*(p)).

Page 22: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1467

In particular, V is everywhere differentiable on Bo (X) if and only if c* is essentially strictly convex.'0 In this case,

(19) V'(f; h) = f hu'(f) dr,

where {r} = argminpEa(f u(f) dp + c*(p)).

The strict convexity of the ambiguity index thus characterizes everywhere differentiable variational preference functionals. For example, Proposition 23 will show that divergence preferences have a strictly convex ambiguity index, provided 0 is strictly convex. By Theorem 18, they are everywhere differen- tiable.

Variational preferences that feature an index c that is not strictly convex are, by Theorem 18, not everywhere differentiable in general. This is a large class of preferences, which includes, but is much larger than, that of multiple priors preferences.

However, although there are plenty of examples of variational preferences that are not everywhere differentiable, since the convex combination of a con- vex cost function and a strictly convex function is strictly convex, one can ap- proximate arbitrarily well any variational preference with another one that is everywhere differentiable.

In view of all this, the interest of Theorem 18 is both theoretical and prac- tical. In applications, the explicit formulas (18) and (19) are very impor- tant because they make possible the explicit resolution of optimal problems based on the variational preference functional V. It is worth observing that in Maccheroni, Marinacci, and Rustichini (2006) we show that a version of these formulas holds also for dynamic variational preferences, the intertemporal ver- sion of the variational preferences we are introducing in this paper.

On the theoretical side, Theorem 18 is interesting because in some economic applications of ambiguity aversion, the lack of smoothness (and in particu- lar the existence of "kinks" in the indifference curves) has played a key role. For example, it has been used to justify nonparticipation in asset markets (see Epstein and Wang (1994)). By fully characterizing the differentiability of vari- ational preferences, Theorem 18 clarifies the scope of these results. In partic- ular, it shows that although kinks are featured by some important classes of ambiguity averse preferences, they are far from being a property of ambiguity aversion per se. Indeed, we just observed in the preceding text that it is always possible to approximate any variational preference arbitrarily closely with an- other one that is everywhere differentiable.

10Here u'(f) dr denotes the functional on Bo(E) that associates f hu'(f) dr to every h E Bo(2), and c is essentially strictly convex if it is strictly convex on line segments in

UfeB0(V) argminpeA(f u(f) dp + c*(p)). Clearly, if c is strictly convex, then a fortiori it is essen- tially strictly convex.

Page 23: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1468 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

4. SPECIAL CASES

In this section we study in some more detail two important classes of varia- tional preferences: the multiple priors preferences of Gilboa and Schmeidler (1989) and the divergence preferences we just introduced. In particular, we show that two important classes of preferences-the multiplier preferences of Hansen and Sargent (2001) and the mean-variance preferences of Markowitz (1952) and Tobin (1958), are special cases of divergence, and therefore varia- tional, preferences.

4.1. Multiple Priors Preferences

Begin with the multiple priors choice model axiomatized by Gilboa and Schmeidler (1989). As we mentioned in Section 3.1, the multiple priors model is characterized by Axiom A.2', a stronger version of our independence Ax- iom A.2. Next we show in greater detail the relationship between Axiom A.2' and the variational formula (6).

In particular, when Axiom A.2' replaces Axiom A.2, the only probabilities in A that "matter" in the representation (6) are those to which the decision maker attributes "maximum weight," that is, those in argmin c*. The set of priors C used in the multiple priors model is then given by {p e A: c*(p) = 0}.

PROPOSITION 19: Let > be a variational preference. The following conditions are equivalent:

(i) The relation - satisfies Axiom A.2'. (ii) For all f E T,

(20) min u(f)dp+c*(p)p) = mm

inf=)

u(f)dp.

If, in addition, > is unbounded, then (ii) is also equivalent to the following condition:

(iii) The function c* takes on only values 0 and oc.

The characterization of the multiple priors model via Axioms A.1, A.2', and A.3-A.6 is due to Gilboa and Schmeidler (1989). Proposition 19 shows how the multiple priors model fits in the representation we established in Theorem 3.

As is well known, the standard SEU model is the special case of the multiple priors model characterized by the following stronger version of Axiom A.5.

AXIOM A.5'-Uncertainty Neutrality: If f, g e F and a e (0, 1),

f~g g af + (1 - a)g f.

In terms of our representation, by Theorem 3 and Proposition 19 we can make the following statement:

Page 24: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1469

COROLLARY 20: Let > be a variational preference. The following conditions are equivalent:

(i) The relation > satisfies Axiom A.5'. (ii) The relation is SEU. (iii) The relation -

satisfies Axiom A.2' and {p e A: c*(p) = 0} is a singleton. If, in addition, - is unbounded, then (iii) is also equivalent to the statement:

(iv) There exists q e A such that c*(q) = 0 and c*(p) = oc for every p 0 q.

4.2. Divergence Preferences

In the previous section we introduced divergence preferences so as to illus- trate our results on probabilistic sophistication. Here we discuss their ambigu- ity attitudes.

Recall that a preference > on T is a divergence preference if

ftg d min u(f)dp+ O6D'(p 1|q)}

> minf u(g) dp + OD (p q) , where 0 > 0, q is a countably additive probability on the o--algebra Z, u:X -* R is an affine function, and D q(. II q): A -> [0, oo] is the w-weighted 4 divergence given by (14).11

By Theorem 16, all divergence preferences are examples of continuous vari- ational preferences. As a result, to determine their ambiguity attitudes, we can invoke Propositions 7 and 8. By the former result, divergence preferences are ambiguity averse. As to comparative attitudes, the next simple consequence of Proposition 8 shows that they depend only on the parameter 0, which can therefore be interpreted as a coefficient of ambiguity aversion.

COROLLARY 21: Given two (w, 4) divergence preferences >- and >2, the fol- lowing conditions are equivalent:

(i) The relation >1 is more ambiguity averse than 2*. (ii) u, u2 and 0

<_ 02 (provided ul = u2).

According to Corollary 21, divergence preferences become more and more (less and less, resp.) ambiguity averse as the parameter 0 becomes closer and closer to 0 (closer and closer to oc, resp.).

The limit cases where 0 goes either to 0 or to oo are described by Proposi- tion 12, which takes an especially stark form for divergence preferences under

"1When we want to be specific about w and p, we speak of (w, 4) divergence preferences.

Page 25: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1470 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

some very mild assumptions. To see this, we need a piece of notation: given a simple measurable function ?p :S -- R, set

essmin •p(s)

= max{t e R:q({s e S: (s) > t}) = 1}. seS

For example, when q has a finite support supp(q), we have

(21) essmin~p(s)= min ?p(s). seS sEsupp(q)

PROPOSITION 22: (i) If w e Lo(q), then for all f F,

lim min fu(f) dp + ODW(p q) = ess min u(f(s)). OtO pe"W(q) Ij seS

(ii) If 0 is strictly convex, then for all f e F,

lim min{ u(f) dp+ODp(p jq) =u(f)dq. Otoo pEMO(q) fU0J

When w E L"(q), divergence preferences therefore tend more and more, as 0 - 0, to rank acts according to the very cautious criterion given by essmin,,s u(f(s)). In contrast, when ( is strictly convex, divergence prefer- ences tend more and more, as 0 -+ o, to rank acts according to the SEU cri- terion given by f u(f) dq.

The next result, a simple consequence of Theorem 18, shows that divergence preferences are smooth under the assumption that 0 is strictly convex (most examples of divergence Dw satisfy this condition; see, e.g., Liese and Vajda (1987)). This is an important feature of divergence preferences that makes them especially suited for optimization problems and that also differentiates them from multiple priors preferences, which are not everywhere differen- tiable (see, e.g., Epstein and Wang (1994, p. 295)).

As we did for Theorem 18, we assume that X is the set of all monetary lotteries and we regard Bo(.) as the collection of all monetary acts. We also assume that the utility function u is concave, strictly increasing, and differen- tiable on R.

PROPOSITION 23: If 4 is strictly convex, then Dw(.1 I q):A(X) - [0, oc] is strictly convex on its effective domain and the variational preference functional V : Bo(E) --+ R, given by

V(f=min u(f)dp+ OD(p II q)f Vfe

B0(B), PEA(X) f 0 1~piJYfBZ)

Page 26: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1471

is everywhere differentiable for all 0 > 0. In this case,

(22) V' (f; h) = f hu' (f) dr,

where {r} = arg minpEA(f u(f) dp + ODw(p II q)).

Consider, for example, multiplier preferences, a special class of divergence preferences in which c*(p) = OR(p II q). By some well known properties of the relative entropy (see Dupuis and Ellis (1997, p. 34)), formula (22) takes, in this case, the neat form

( f; h) hu'(f) exp(- (f) dq (f h) exp(-) dq f exp(- 0) dq

for all f, h E Bo(V). Unlike the multiple priors case, for divergence preferences, we do not have

an additional axiom that, on top of Axioms A.1-A.6, would deliver them (for multiple priors preferences, the needed extra axiom was Axiom A.2'). We hope that this will be achieved in later work and in this regard it is worth observing that the ambiguity index D'(p II q) is additively separable, a strong structural property.

After having established the main properties of divergence preferences, we now move on to discuss two fundamental examples of this class of variational preferences.

4.2.1. Entropic and multiplier preferences

We say that a preference >- on F. is an entropic preference if

f g (min u(f) dp +ORW(p | q))

min (f u(g) dp+ORW(p | q)),

where 0 > 0, q e A , u: X -- R is an affine function, and RW"( I q) : A -+ [0, oo] is the weighted relative entropy given by

w(s) -q(s)log q(s)

- -(s) + 1 dq(s), RW(p 11 q) = s 5 dq dq dq

if pE dAc (q), 00, otherwise.

Page 27: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1472 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

Since the entropy RW(p II q) is a special case of divergence Dv(p II q) defined in (14), where 4)(t) = t log t - t + 1, entropic preferences are an example of di- vergence preferences. Hence, by Theorem 16, they are continuous variational preferences, with index of ambiguity aversion given by

c*(p) = ORW(p II q) Vp E A.

When w is uniform, is probabilistically sophisticated and it features (a pos- itive multiple of) the standard relative entropy R(. 1I q) as the index of ambi- guity aversion. This is the case considered by Hansen and Sargent (see, e.g., Hansen and Sargent (2000, 2001)), who called this class of entropic prefer- ences in which acts are ranked according to

V(f) =min (f u(f)dp+ OR(p | q))

multiplier preferences. Although multiplier preferences are probabilistically so- phisticated, Example 17 shows that this not the case for general entropic preferences that have nonuniform weighting functions w. These "nonuniform" entropic preferences thus provide a specification of preferences that can, in general, produce Ellsberg-type behavior-and so are ambiguity averse accord- ing to all notions of ambiguity available in literature-but that also retain the good analytical tractability of multiplier preferences.

As to the ambiguity attitudes featured by entropic preferences, by Corol- lary 21 they are characterized by the parameter 0 as follows: the lower 0 is, the more ambiguity averse is the entropic preference. The parameter 0 can therefore be interpreted as a coefficient of ambiguity aversion.

We have shown how entropic, and hence multiplier, preferences are a special case of divergence preferences. As we already observed, for divergence pref- erences we do not have an additional axiom that, on top of Axioms A.1-A.6, would deliver them (even though we have been able to point out some strong structural properties that their ambiguity indices satisfy). On the other hand, we view entropic preferences as essentially an analytically convenient specifi- cation of variational preferences, much in the same way as, for example, Cobb- Douglas preferences are an analytically convenient specification of homothetic preferences. As a result, in our setting there might not exist behaviorally sig- nificant axioms that would characterize entropic preferences (as we are not aware of any behaviorally significant axiom that characterizes Cobb-Douglas preferences). Similar considerations apply to the Gini preferences that we will introduce momentarily.

For their macroeconomic applications, Hansen and Sargent are mostly in- terested in dynamic choice problems. Although our model is static, in the follow-up paper (Maccheroni, Marinacci, and Rustichini (2006)), we provide a dynamic version of it and, inter alia, we are able to provide some dynamic

Page 28: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1473

specifications of multiplier preferences that are time consistent. As is often the case in choice theory, also here the analysis of the static model is key to paving the way for its dynamic extension.

We close by discussing the related work of Wang (2003), who (in a quite different setting) recently proposed an axiomatization of a class of preferences that include multiplier preferences as special cases. He considered preferences over triplets (f, C, q), where f is a payoff profile, q is a reference probability, and C C A is a confidence region. For such preferences, he axiomatized the representation

(23) V(f, C, q)=min{u(f) dp +OR(p li q)J. peCIf His modeling is very different from ours: in our setup, preferences are defined only on acts, and we derive simultaneously both the utility index u and the am- biguity index c; that is, uncertainty is subjective. In Wang (2003), both C and q are exogenous, and so uncertainty is objective; moreover, agents' preferences are defined on the significantly larger set of all possible triplets that consist of payoff profile, confidence region, and reference model.

In any case, observe that when (23) is viewed as a preference functional on F, then it actually represents variational preferences that have as ambi- guity index the sum of Sc and OR(- II q). As a result, Wang's preferences are a special case of variational preferences once they are interpreted in our setting.

4.2.2. Gini and mean-variance preferences

We say that a preference ? on F is a Gini preference if

fg min(f u(f) dp +?OGW(p 11 q))

min u(g) dp +POGw(p||q)),

where 0 > 0, q e An, u: X -- R is an affine function, and G"(. II q): A -- [0, co] is the weighted relative Gini index given by

G"(p II q) = w(s)dp (s) - 1 dq(s), if p EA•(q), 00c,

otherwise.

Like the weighted relative entropy R"(p II q), the Gini index GW(p 11 q) also is a special case of divergence Dw(p II q) defined in (14), with 4(t) =

Page 29: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1474 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

2-1(t - 1)2.12 As a result, by Theorem 16, Gini preferences are continuous vari- ational preferences, with ambiguity index OG"(. II q).

In particular, - is probabilistically sophisticated when w is uniform. In this

case, when X is the set of all monetary lotteries and u(t) = t for all t E R, we call the restriction of these preferences to the collection Bo(X) of all mon- etary acts monotone mean-variance preferences, written >mmv; that is, for all f, ge Bo(1),

f mmV g Q i (ffdp+ OG(p I q))

>amin( gdp+OG(p

iq)) Since the Gini index is, along with Shannon's entropy, a classic concentration index, monotone mean-variance preferences are a natural example of diver- gence preferences. However, we are not considering them just for this: their main interest lies in the close connection they have with mean-variance pref- erences.

In fact, consider the classic mean-variance preferences of Markowitz (1952) and Tobin (1958) that are defined on Bo(X) by

(24) f -mv g f dq -

Var(f) g dq --

Var(g),

where Var is the variance with respect to q. These preferences are not monotone unless their domain is suitably restricted to the set M on which the (Gateaux) differential of the mean-variance functional f ? f fdq - (1/20) Var(f) is positive (as a linear functional). The convex set M, called the domain of monotonicity of >mv, is where these preferences do not violate the monotonicity Axiom A.4; that is, where they are economically meaningful.

THEOREM 24: The domain of monotonicity M of >-mv is the set

j f e Bo() : f - f f dq < 0,

q-a.s.}. Moreover,

ftmg (min f d (p+ G(q)

'min( gdp + OG(p I q))

12The classic Gini concentration index can be obtained by normalization from the relative index (with uniform w) in the same way Shannon's entropy can be obtained from relative entropy.

Page 30: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1475

for all f, g E M.

By Theorem 24, mean-variance preferences coincide with monotone mean- variance preferences once they are restricted on their domain of monotonic- ity M, which is where they are meaningful.

Inter alia, this result suggests that monotone mean-variance preferences are the natural adjusted version of mean-variance preferences that satisfy monotonicity. This insight is developed at length in Maccheroni, Marinacci, Rustichini, and Taboga (2004), and we refer the interested reader to that pa- per for a detailed analysis of this class of variational preferences.

Istituto di Metodi Quantitativi and IGIER, Universita Bocconi, Milano, 20100 Italy,

Dipartimento di Statistica e Matematica Applicata and Collegio Carlo Alberto, Universitui di Torino, Torino, 10122 Italy; [email protected],

and Dept. of Economics, University of Minnesota, Minneapolis, MN 55455, U S.A.

Manuscript received August, 2004; final revision received April, 2006.

APPENDIX A: NIVELOIDS

The set of all functions in Bo(2) (resp. B(2)) that take values in the interval K C R is denoted by B0(2, K) (resp. B(X, K)).

When endowed with the supnorm, Bo(Y) is a normed vector space and B(2?) is a Banach space. The norm dual of Bo0() (resp. B(.)) is the space ba(.) of all bounded and finitely additive set functions

:'X -- R endowed with the

total variation norm, the duality being (?P, /t) = f cp dlt for all ?c E B0(E) (resp. B(,Y)) and all / E ba(2) (see, e.g., Dundorf and Schwartz (1958, p. 258)).

As is well known, the weak* topologies o-(ba(.X), Bo(N)) and o-(ba(2), B(.)) coincide on A(Y). Moreover, a subset of A7(X) is weakly* compact if and only if it is weakly compact (i.e., compact in the weak topology of the Banach space ba (X)). For pI, I E B(s), we write ?p > tp if ?p(s) > f&(s) for all s E S.

A functional I: 0 -- R, defined on a nonempty subset P of B(2), is a niveloid if

I((p) - I($i) < sup((p - 0)

for all ?p, f E 0; see Dolecki and Greco (1995). Clearly a niveloid is Lipschitz- continuous in the supnorm. A niveloid I is normalized if I(kls) = k for all k e R such that k1s e c0. With a little abuse, we sometimes write k instead of k1s.

LEMMA 25: Let 0 E int(K).A functional I: BO0(., K) -+ R is a niveloid if and only if, for all cp, f E Bo(Z, K), k E K, and a E (0, 1),

(i) c > q implies I(pc) > I(ti), and (ii) I(atp + (1 - a)k) = I(acp) + (1 - a)k.

Page 31: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1476 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

In this case, I is concave if and only if (iii) I(tf) = I(Q?p) implies I(aqf + (1 - a) p) > I((p).

Properties (i) and (ii) are called monotonicity and vertical invariance, respec- tively. The proof of Lemma 25 can be found in Maccheroni, Marinacci, and Rustichini (2005). If I:(P -* R is a niveloid, then

I(I) = sup [I() + inf(qi(s) - p(s))] iV e B(I)

is the least niveloid on B(s) that extends I (see Dolecki and Greco (1995)). Moreover: * If ( is convex and I is concave, then I is concave. * If (P + R

_ Bo(),13 then I is the unique niveloid on B(X) that extends I (see

Maccheroni, Marinacci, and Rustichini (2005)). If 0 is convex and I: (P -- R is a concave niveloid, direct application of

the Fenchel-Moreau theorem (see, e.g., Phelps (1992, p. 42)) to I guarantees that I(p) = minLEba(5)((p, /-) - I*(Aj)), where I*(1i) = infqB(Y)((q, p ) - I(0)) is the

Fenchel conjugate of I. If At is not positive, there exists qp > 0 such that

(?, /t) < 0. Then (ap, ) - I(a?p) _

a((p, /t) - I(0) for all a >0, whence

I*(QL) = -00. If ,t(S) = 1, choose q B(2). Then (41 + b, F) - I(0q

+ b) =

(q, /-) - I(jq) + b(CL(S) - 1) for all b •R

and so I*(1i) = -oc. That is,

(25) (cp) = min ((?, p) - I*(p)); pEA(?)

see also Fllmer and Schied (2002, Theorem 4.12). Set I*(p) = I*(p) for each p E A(s) and set d,I(4o) = {p pe A():I(qj) - I(?p) < (q - n, p) for each

The next two results are proved in Maccheroni, Marinacci, and Rustichini (2005).

LEMMA 26: Let 0 be a convex subset of B(s) that contains at least one con- stant function, and let I : -- R be a concave and normalized niveloid. Then:

(i) For each p e A(E), I*(p) = infEP((I, p) - I(-q)). (ii) I* : A(X) -+ [-oo, 0] is concave and weakly* upper semicontinuous.

(iii) For each e E 0, ,I(p?) = argminpEA(z•(((p,

p) - I*(p)) and it is not empty.

(iv) dI(kls) = {I* = 0} = argmaxpeAJ()I*(p) for all k E R such that kls E (P.

13p? + R is the set {~c + b: •p 0E, b E R}. An important special case in which P + R = Bo(E) is

when 0 = Bo(2, K) and K is unbounded.

Page 32: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1477

(v) I* is the maximal functional R" A(X) - [-oc, 0] such that

(26) I(po)= inf ((p, p)- R(p)) V o E P. pEA(l)

Moreover, if P + R 2 Bo (2), 1* is the unique concave and weakly* upper semicontinuous function R :A()) - [-oo, 0] such that (26) holds.

(vi) If (26) holds and T _

0 is such that sups,,s (s) - infses ii(s) < b for all

iqe j , then

(27) I(f) = inf ((p , p) - R(p)) V i E T. {peAd():R(p)>-b}

PROPOSITION 27: Let I: Bo(2, K) --+ R be a normalized concave niveloid, with K unbounded and L a ur-algebra. Then the following conditions are equiva- lent:

(i) If c, q E Bo(X, K), k E K, {En•}>i

E . with E1 D E2 D . . and n,• En =

0, then I (p) > I ( /) implies that there exists no > 1 such that

I(klEn0 + (lE)> 1).

(ii) The set {p E A(X) : I*(p) > b} is a weakly compact subset of A((Y) for each b< 0.

(iii) There exists q E d"(X) such that {p E A() : I*(p) > b} is a weakly com- pact subset of Ao(X, q) for each b < 0, and, for each 4 E Bo(I, K),

(28) I(G) = mi ((, p) - *(p)). pEA0(.,q)

APPENDIX B: PROOFS OF THE RESULTS IN THE MAIN TEXT

The main result proved in this appendix is Theorem 3. Its proof proceeds as follows: using Lemma 28, we first show that if - satisfies Axioms A.1-A.6, then there exists a nonconstant affine function u: X -- R and a normalized and concave niveloid I: Bo(0, u(X))

-- R such that f >- g if and only if I(u(f)) >

I(u(g)). Then (25) obtained in Appendix A delivers the desired variational representation I(u(f)) = minp,,(n)(f u(f) dp + c(p)).

We now move to the proofs. The standard proof of Lemma 1 is omitted.

LEMMA 28: A binary relation > on F satisfies Axioms A.1-A.4 and A.6 if and only if there exist a nonconstant affine function u :X -- IR and a normalized niveloid I : Bo(., u(X)) --+ R such that

ft g I(u(f)) > I(u(g)).

PROOF: Assume that > on . satisfies Axioms A.1-A.4 and A.6. Let x, y E X be such that x - y. If there exists z E X such that x + z y + z, with-

Page 33: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1478 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

out loss of generality ix ?+ z >- y + 1z; by Axiom A.2 (we can replace z with x to obtain),lx + x > y + x and (we can replace z with y to ob-

tain) ix ?+ y >- y ?+ y, and we conclude that x >- y, which is absurd. Then the hypotheses of the mixture space theorem (Hernstein and Milnor (1953)) are satisfied and there exists an affine function u:X -+ R such that x 'e y if and only if u(x) > u(y). By Axiom A.6 there exist f, g e F such that f >- g. Let x, y E X be such that x >- f(s) and g(s) > y for all s e S. Then x > f >- g > y implies x >- y and u cannot be constant. Moreover, u is unique up to positive affine transformations and we can assume 0 E int(u(X)).

For all f e F, let x,y E X be such that x - f(s) >- y for all s E S. Then

x > f > y. By Axiom A.3, the sets {aE [0, 1]:ax + (1 - a)y >- f} and {a E [0, 1]: f f ax + (1 - a)y) are closed; they are nonempty because 1 belongs to the first and 0 belongs to the second; their union is the whole [0, 1]. Because [0, 1] is connected, their intersection is not empty; hence, there exists P3 [0, 1] such that fx + (1 - 3)y - f. In particular, any act f admits a certainty equiv- alent xf E X.

If f ~ xf, set U(f) = u(xf). The function U is well defined because f - xf

and f ~ y, with xf, yf EX implies xf - yf and u(xf) = u(yf). Clearly, f > g if and only if xf xg if and only if u(xf) > u(xg) if and only if U(f) > U(g). Therefore, U represents >-.

If f F, then u(f) e Bo(1, u(X)). Conversely, if p E B0o(, u(X)), then ?p(s) = u(xi) if s e Ai for suitable xl, ..., XN EX and a partition {(A, A2, S.., AN) of S in X. Therefore, setting f(s) = x, if s e Ai, we have ?q = u(f). We can conclude that Bo(1, u(X)) =

{u(f):'f E F). Moreover, u(f) = u(g) if

and only if u(f(s)) = u(g(s)) for all s E S if and only if f(s) - g(s) for all s e S, and, by Axiom A.4, f -- g or, equivalently, U(f) = U(g).

Define I('p) = U(f) if cp = u(f). By what we have just observed, I:Bo(, u(X)) -+ R is well defined. If ?c = u(f), 0f = u(g) e Bo(Z, u(X)), and p > q , then u(f(s)) > u(g(s)) for all s E S and f(s) - g(s) for all s e S, SO f g, U(f) > U(g), and I(?p) = I(u(f)) = U(f) > U(g) = I(u(g)) = I(iq). There- fore, I is monotonic. Take k e u(X), say k = u(x). Then I(k1s) = I(u(x)) = U(x) = u(x) = k. Therefore, I is normalized.

Take a E (0, 1), ?P = u(f) e Bo(., u(X)), and k = u(xk) e u(X); denote

by x0 an element in X such that u(xo) = 0. Choose x, y E X such that x> f (s) >- y for all s e S. Then ax +(1- a)xo > af (s) + (1- a)xo >- ay + (1- a)xo for all s e S. The technique used in the second paragraph of this proof yields the existence of p E [0, 1] such that P(ax + (1 - a)xo) + (1 - P)(ay + (1 - a)xo) - af + (1 - a)xo, i.e., az + (1 - a)xo - af + (1 - a)xo, where z =

pix + (1 - P)y E X. Then, by Axiom A.2, az + (1 - a)xk - af + (1 - a)xk and

I(ap? +(1 - a)k) = I(u(af + (1 - a)xk)) = u(az + (1 - a)Xk)

Page 34: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1479

= au(z) + (1 - a)k = au(z) + (1 - a)O + (1 - a)k

= u(az + (1 - a)xo) + (1 - a)k

= I(u(af + (1 - a)xo)) + (1 - a)k = I(ap?) + (1 - a)k.

By Lemma 25, I is a niveloid, and we already proved that it is normalized. Conversely, assume there exist a nonconstant affine function u: X -- R

and a normalized niveloid I: Bo(Z, u(X)) -W R such that f -

g if and only if I(u(f)) > I(u(g)).

Choose c E R such that 0 e int(u(X) + c) and set v = u + c. Define J:

Bo(,, v(X)) --+ R by J((p) = I('p - c) + c. Notice that J is a normalized

niveloidl4 and

f >g I(u(f)) > I(u(g))

SI(u(f)+ c-c)+ c > I(u(g) + c-c) + c

SI(v(f) - c) + c > I(v(g) - c) + c

SJ(v(f)) > J(v(g)).

Clearly, > satisfies Axiom A.1. If f, g E

•, x, y c X, and a E (0, 1), then av(h), (1 - a)v(z), av(h) + (1 -

a)v(z) e Bo(Y, v(X)) for h = f, g and z = x, y. Moreover,

af + (1 - a)x > ag + (1 - a)x

= J(av(f) + (1 - a)v(x)) > J(av(g) + (1 - a)v(x))

= J(av(f)) + (1 - a)v(x) > J(av(g)) + (1 - a)v(x)

= J(av(f)) + (1 - a)v(y) > J(av(g)) + (1 - a)v(y)

= J(av(f) + (1 - a)v(y)) > J(av(g) + (1 - a)v(y))

= af + (1 - a)y > ag + (1 - a)y

and Axiom A.2 holds. If f, g, h E , a E [0, 1], and there exists a, E [0, 1] such that ac, - a and

af + (1 - an)g 9 h for all n > 1, then v(anf + (1 - an)g) = av(f) + (1 - an)v(g) converges uniformly to av(f) + (1 - a)v(g) = v(af + (1 - a)g). The

inequality J(v(anf + (1 - an)g)) > J(v(h)) for all n > 1 and the continuity of J guarantee J(v(af + (1 - a)g)) > J(v(h)). Therefore, {a E [0, 1]: af +

14In fact, for all ?p, 4 E Bo(2, v(X)),

J(p) - J(f) = I(p( - c) + c - I(( - c) - c < sup((p - c) - (qf - c)) = sup(?p - qf).

Moreover, for all t E v(X), J(t) = I(t - c) + c = t - c + c = t.

Page 35: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1480 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

(1 - a)g - h} is closed. A similar argument shows that {a E [0, 1]: h > af + (1 - a)g} too is closed, and Axiom A.3 holds.

Given f, g E T, f(s) > g(s) for all s E S if and only if J(v(f(s))) > J(v(g(s))) for all s if and only if v(f(s)) > v(g(s)) for all s, then monotonicity of J yields J(v(f)) > J(v(g)). This shows Axiom A.4.

Finally, because v is not constant and it represents on X, there exist x >- y and Axiom A.6 holds too. Q.E.D.

PROOFS OF THEOREM 3 AND OF PROPOSITION 6: Assume > satisfies Ax- ioms A.1-A.6. By Lemma 28, there is a nonconstant affine function u:X - -R and a normalized niveloid I: Bo(X, u(X)) --+ R such that f

- g if and only if

I(u(f)) >_ I(u(g)). Next we show that Axiom A.5 implies that I: B0o(, u(X)) -* R is concave.

Let ?q, 1'fe B0o(, u(X)) be such that I(qp) = I((qi) and a E (0, 1). If f, g e .F are such that op = u(f) and ii = u(g), then f - g and, by Axiom A.5, af + (1 -

a)g > f, that is,

I(a•• + (1 - a)q/) = I(au(f) + (1 - a)u(g))

= I(u(af + (1 - a)g)) > I(u(f)) = I(p).

Lemma 25 guarantees the concavity of I. The functional I: Bo(X, u(X)) --+ R is, therefore, a concave and normalized

niveloid. For all p e A(I), set c*(p) = -I*(p). Lemma 26 guarantees that

c*(p) = inf ((i, p) -I(q))=sup(u(xf) - u(f)dp qEBoB( ', u(X )) fET

/

for all p E A(.) (where xf is a certainty equivalent for f), that c* is nonnega- tive, grounded, convex, and weakly* lower semicontinuous, and that

I((p) = min (p dp +pc*(p) Vp Y Bo (,,u(X));

see also (25). Let c: A(s) -+ [0, 00] be a grounded, convex, and weakly* lower semicon-

tinuous function such that

fig

p(min u(f)dp + c(p) > min u(g) dp + c(p))

Page 36: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1481

For all o = u(f) E Bo(X, u(X)),

I(w)= I(u(f)) = u(x) = mi (f u(x) dp +c(p)

= mm p(fu(f)dp+c(p))

= minm (fdp +c(p). PEA(-Y) (/ peA(sY) (J

Lemma 26 guarantees c* < c (this concludes the proof that (i) implies (ii)). Moreover, if u(X) is unbounded, again Lemma 26 guarantees c = c* (this proves Proposition 6).

For the converse, notice that u and I(cp) = minpeA(X)(f ~p dp + c(p)) are, respectively, a nonconstant affine function and a normalized niveloid that rep- resents >. By Lemma 28, > satisfies Axioms A.1-A.4 and A.6. Concavity of I guarantees Axiom A.5.15 Q.E.D.

PROOF OF COROLLARY 5: Let (uo, c;) represent > as in Theorem 3. If (u, c*) is another representation of > (as in Theorem 3), by (6), u and uo are affine representations of the restriction of - to X. Hence, by standard unique- ness results, there exist a > 0 and / E IR such that u = auo + P. By (7),

c*(p) = sup (u(xf) - u(f ) dp)

=sup •auo(xf)+1-f(auo(f)+1)dp) =ac•*(p),

fE-F(

as desired. The converse is trivial. Q.E.D.

LEMMA 29: Let > be a binary relation on X represented by an affine function u :X -- R. The set u(X) is unbounded (either below or above) if and only if j

satisfies Axiom A.7.

The standard proof is omitted. Proposition 7 is a consequence of Lemma 32.

PROOF OF PROPOSITION 8: (ii) trivially implies (i). As to the converse, let

(ui, c7) represent i as in Theorem 3, i = 1, 2, and set Ii(cp) = minpac() (f p dp -

c7(p)) for all co E Bo(0, ui(X)).

15If f - g and a E (0, 1), then

I(u(af + (1 - a)g)) = I(au(f) + (1 - a)u(g))

> aI(u(f)) + (1 - a)I(u(g)) = I(u(f)).

Page 37: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1482 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

By (8) and the fact that u1 and u2 are not constant, we can choose

ul = u2 = u. For all f e ?, if f ~1 x, then f >2 X; therefore, Ii(u(f)) = u(x) < I2(u(f)). This implies I < I12 and

cl(P)

= sup (Ip)

- ( dp) ? E Bo(1,u(X)) fJ/

< sup

I2(P)-fqDdp)=c2(p) ?PEBo(X, u(X)) \f)

for all p e A(?). Q.E.D.

PROOF OF PROPOSITION 12: Observe that the functions c, are weakly* lower semicontinuous on A and so is f u(f) dp + Cn(p) for each n. Using this observation, we now prove (i) and (ii).

(i) The decreasing sequence {f u(f) dp + c,(p)}n pointwise converges to

(29) f U(f) dp •Undomcn

(P)

Hence, by Dal Maso (1993, Proposition 5.7), this sequence F-converges to

/ u(f) dp + Un,dom

c, (P)

the weakly* lower semicontinuous envelope of (29). By Dal Maso (1993, The- orem 7.4), this implies

lim min u(f)dp+cn(p) =minf u(f)dp + 5U,,dom

cn

(ii) Since arg min c, = (p E A: cn(p) = 0), the increasing sequence If u(f) dp+ c,(p))}, pointwise converges (and so, by Dal Maso (1993, p. 47), F-converges) to

J u(f) dp + Sn argmincn (p)

Because A is weak-* compact, Dal Maso (1993, Theorem 7.8) implies

limmin u(f ) dp + cn(P) n pA I )d + S

=min u(f) dp + 8FnnargminCn(P) Q.E.D.

Page 38: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1483

PROOF OF THEOREM 13: Let (u, c*) represent >- as in Theorem 3 and set, for all cp e B0o(, u(X)), I(Qp) = minpeA( )(f •p

dp + c*(p)). It is easy to check that - satisfies Axiom A.8 on .F if and only if I satisfies condition (i) of Proposi- tion 27 on Bo(E, u(X)). Unboundedness of u(X) and the relation c* = -I* al- low us to apply Proposition 27 and to obtain the desired equivalence. Q.E.D.

LEMMA 30: Let >- be a variational preference that satisfies Axiom A.8. Then

Sg inf (f u(f) dp + c*(p)

>

Xinf(

u(f) dp+c*(p)) Vf, gEc.

PROOF: Let (u, c*) represent >- as in Theorem 3 and set I(p) =

minpEA(,)(f f dp + c*(p)) for all cp E Bo(E, u(X)). Let p e int(Bo( , u(X))), E, { 0, and e > 0 such that cp - se Bo(E, u(X)). Then

I(lEcp + (minp En- E)E,) - I(p) < I( - lE) - I( P)

< -ep(E,) < O

for all pe e dI(q(). Consider a sequence {kj}jl in u(X) such kj < I(?O) and kj f I(qo). By Axiom A.8, I satisfies (i) of Proposition 27 on Bo(Y, u(X)). Then, for all j > 1, there is no > 1 such that kj < I(cplE o + (min q E -

8)lEn0). Hence,

limn-o I(~plEc + (min m - e)lEn) > kj because the sequence I(?plEc + (min c -

e) 1En) is increasing. Passing to the limit for j -- oc, we obtain

(30) lim I(GplE + (minp - 8)1E,) = IE )

n---> oc

and p(E,) -- 0 (uniformly with respect to pe dI(?p)); that is, OdI(p) is a (weakly compact) subset of A0(Y). By Lemma 26, minPEA(X)(f ?p dp + c*(p)) is attained for all 'p e int(Bo(Z, u(X))) in (dI(c) and hence in) Az(E). Hence,

I(cp)= (minf ~pdp+c*(p)) and

J('p)= )in(f 'pdpc+ c*(p)

coincide on int(Bo(Z, u(X))) and, being continuous, they coincide on Bo( , u(X)). Q.E.D.

The results of Section 3.5 require some notation and preliminaries. For all cp e L'(, q), set G,(t) = q({Qp > t}) for each t e R and set T~(b) = inf{t e R: G,(t) < b} for each b e [0, 1].

Page 39: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1484 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

The function G, is the survival function of ?c, and 7, is the decreasing re- arrangement of cp. Two functions c', / e L'(Z, q) are equimeasurable if G,= GQ (if and only if 7, = F,). We refer to Chong and Rice (1971) for a compre- hensive study of equimeasurability. The preorder -<cx defined in Section 3.5 can be naturally regarded as a relation on L'(Z, q) by putting p -<,, ji if and

only if f P(cp) dq < f P(qj) dq for every convex 0 on R. Analogously, the definition of Shur convexity (resp. rearrangement invari-

ance) can be spelled for functions T: L(Z, q) -- (-oo, 0o] by requiring that p -<cx implies T(?9) < T(Iq) (resp. p cx implies T(o) = T(ql)).

Put ~p -<-<,cx q if and only if fo' 1,(b) db < fo' F(b) db for all t E [0, 1]. Then -, ,<c if and only if qp -<-<c , and f p dq = f qj dq (see, e.g., Chong (1976)).

Moreover, the equimeasurability relation is the common symmetric part -,cx of <-<cx and -<cx. Simple manipulation of the results of Luxemburg (1967, pp. 125-126) yields the following lemma:

LEMMA 31-Luxemburg: Let q be adequate. If T: Bo(Z, K) -+ (-00, 00] is

rearrangement invariant, then the function defined for all p e L' (Z, q) by H (S) =

supEBO(Z,K) (f (p/ dq - T(

q,)) is Shur convex. If H is monotonic, then p -<-<cx

5'

implies H (?) < H (p').

PROOF: Let c' E L1'(, q). By Luxemburg (1967, Theorem 9.1) for all qJ E Bo(Z1, K),

(31) maxJ pfi'dq: ~'~ f' e Bo(,, K) = fF,(t)Fq(t)

dt.

Therefore, H(p) < sup,B,,,(,K)( f0 ,(t)q,(t)dt - T(q,)). Again by (31),

for all q, e Bo(,, K) there exists fi' e Bo(V, K) with q,' -- q' such that

f; F,(t)FI(t) dt = f~,c'

dq. Since T(q') = T(q,),

then

J IF,(t)F,(t)dt- T(q) = p'dq- T(qj) = Jp'dq- T(')

and

H(•o) = sup F(t)Fq(t)dt- T(q) .

qEBo(1,K) 0

If ? -<,x 9?', an inequality of Hardy (see, e.g., Chong and Rice (1971, pp. 57-58)) delivers H(Sp) < H(?o'). Let cp -<-<,cx '. By Chong (1976, Theorem 1.1), there is a nonnegative Sc" e L' (2, q) such that Sc + Sc" -<,, S'. If H is monotonic, then H((p) < H(Sc + So") < H(cp'). Q.E.D.

Page 40: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1485

REMARK 1: Analogously, if q is adequate and T: L'(2, q) -+ (-oo, oo] is

rearrangement invariant (resp., q e A'z(2) and T is Shur convex), then the function defined for all cp e Bo(X) by H(cp) -

sup,eL1(2,q)(f cp dq - T(f)) is Shur convex. If H is monotonic, then ?o

-<-<• 9' implies H(?p)

_ H(p'). (If T is

Shur convex and q is not adequate, use Chong and Rice (1971, Theorem 13.8) rather than Luxemburg (1967, Theorem 9.1).) See also Dana (2005).

PROOF OF THEOREM 14: For f e F, qf denotes the finite support probabil- ity on X defined by qf(x) = q(f-'(x)) for all x E X. We prove (i) =4 (ii) = (iii)

=:= (iv) = (i).

(i) = (ii) This step can be proved by a routine argument. (ii) = (iii) Let >- be a continuous unbounded variational preference, which

is also probabilistically sophisticated with respect to an adequate q E A?(E). Let (u, c*) represent > as in Theorem 3 and set, for all p? Bo(2, u(X)), I(p?) = minp A)s,(f (p dp + c*(p)). For all p E A,

*(p) =sup u(xf) - u(f) dp) = sup (I(j)- f jdp). fET \ Jf / EBo(2,u(X)) \ Jf

Theorem 13 guarantees that {c* < oo} C Ao(2). If p e A'( )\A'(, q), there exists A E Y such that p(A) > 0 and q(A) = 0. If u(X) is unbounded be- low, without loss of generality assume u(X) 2 (-oo, 0]. Let x, e u-l(-n) and y E u-1(0). Consider the act f, = x,Ay and the constant act y. Since qf, = 6y =

qy, by probabilistic sophistication, f - y and I(-nlA) = I(u(fn)) = 0 for all n> 1; therefore,

c*(p) sup(I(-nlA) - lA dp) = 00. nal \(Jf

If u(X) is unbounded above, without loss of generality assume u(X) D [0, oo). Let x,, E u-l'(n) and y e u-'(0). Consider the act f, =

yAx,. Since qf, - 8x, - qx,, then fn ~ x, and I(nlAc) = I(u(fn)) = n for all n > 1; therefore,

c*(p) > sup (I(nlAc) -

nlA dp) = sup(n - n(1 - p(A))) = oo.

n>l / n>l

We conclude that {c* < oo} C Al( , q). In particular, 40 = ?o' q-a.s. implies I(p) = I(4').

Assume cp, E Bo(X, u(X)) are equimeasurable. Therefore (see, e.g., Chong and Rice (1971, p. 12)), there exist x1, ... , xe X, with u(x1) > ... > u(x,), and two partitions {A1, A2, ..., A,} and {B1, B2, ..., Bj} of S in 2, with q(Ai) = q(Bi) for all i = 1, 2, ..., n, such that q-a.s.

n n

= LU(Xi)1A, and

=•u(xYi)lBi. i=1 i=1

Page 41: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1486 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

If f(s) = xi for all s E Ai and g(s) = xi for all s E Bi, then f,g e IF and qf = qg. By probabilistic sophistication of -, we obtain f - g, whence I(?p) =

I(u(f)) = I(u(g)) = I(if) and I is rearrangement invariant. Setting T(f) = -I(-q#), T:Bo(2, -u(X)) -- R is rearrangement invariant too. For all p E Ao(E, q),

c* (p) sup - d p) 4EBo(X,u(X))

(J

sup ( dq - T()

,EBo(,,-u(X)) - dq

T(

By Lemma 31, H(p?) = supo(EBO,(x -ux(f (p dq - T(tI)) for all ?p c L'(V, q)

is Shur convex; therefore, c*(p) is Shur convex and a fortiori rearrangement invariant.

(iii) =? (iv) Consider T: L'(C, q) -+ [0, co] defined as

(32) T(f) = c*(p,), if J > 0 q-a.s., f/ dq= 1, 00, otherwise,

where p, is the element of AO(I, q) such that dp,/dq = . Denote by P the set

S4 L'(X, q): i> 0 q-a.s. and /idq= 1.

Let q/' ~cx i. Then q E P if and only if q' e P (see, e.g., Chong and Rice (1971, pp. 15-16)). If 4,, 4' P, then T(4) = oo = T(q'). If 4, 4' e P, then q and 4' are the Radon-Nikodym derivatives of p, and p,,. By the rearrangement in- variance of c*, T(4') = c*(po,) = c*(pq,) = T(qI) and so T is rearrangement invariant. By Theorem 13, (p A: c*(p) < t} is a weakly compact subset of A7 for each t > 0; therefore, {1 E L'(Z, q): T(q ) < t} is a weakly compact subset of L'(E, q) and a fortiori T is weakly lower semicontinuous. Since obviously T is convex, Luxemburg (1967, Theorem 13.3) guarantees that T (hence c*) is Shur convex.'6

(iv) =* (i) Let - be a variational preference. Let (u, c) represent > as in Theorem 3 and assume c is Shur convex (with respect to q).17 Set, for all p e

16Different definitions caveat: Luxemburg (1967, pp. 123-124) defined Shur convexity of T by rearrangement invariance, convexity, and weak lower semicontinuity. His Theorem 13.3 shows that if T has these features, then o -<,, so' implies T((o) < T(?p').

'7Notice that we are not assuming that > is unbounded or continuous, or that c = c*, or that q is adequate.

Page 42: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1487

B0(o), I(p) = minpeA(d)(f ?o dp + c(p)). Then

-I() =- inf -p pdq c(p)

sup ( pdql-T(q), q/ ELI .N, q) (fJ;,q}

where T is defined as in (32), replacing c* with c. Let I ' -<c, . If /, V P, then T(q) = oc > T(f'). Else if

/, e P, f q' dq = f if dq = 1 and ip' > 0 q-a.s. (see,

e.g., Chong and Rice (1971, p. 62)); that is, q' E P. The Shur convexity of c ensures that T(i') = c(py,) < c(pq,) = T(if). Thus, T is Shur convex. By Re- mark 1, H(cp) = -I(-?p) is Shur convex. Next we show that >- satisfies stochas- tic dominance. Assume that q({s E S: f(s) x})

_< q({s E S:g(s) x}) for all

x E X. For all t E u(X), q({s E S: u(f(s)) > t}) > q({s E S: u(g(s)) > tj), and this is a fortiori true if t V u(X). We conclude that Gu(f), Gu(g), F >F , F Uf) <

Fu(g (see, e.g., Chong and Rice (1971, pp. 30-31)), and -u(f) -<-<x -u(g). Since H is monotonic, Remark 1 yields -H(-u(f)) > -H(-u(g)), so that I(u(f)) > I(u(g)) and f ?- g. Q.E.D.

PROOF OF LEMMA 15: Groundedness (in particular, D (q II q) = 0) and convexity are trivial, so is Shur convexity if w is uniform. Weak* lower semicon- tinuity descends from the fact that the sets {p E A: Dw(p II q) < t} are weakly compact in AU(q) for all t E R. The remaining part of the proof is devoted to show this compactness feature.

We first consider a simple Y-measurable function w with min,,s w(s) > 0, without requiring f w dq = 1. Set w = minSEs w(s) and without loss of general- ity suppose t > 0. By definition, {p E A: D(p II q) < t}

_ A'(q), hence

{p A :Dw(p II q) < t} = {p E AU:D•(p 11 q) t}. 0 0-D4( [ )<t

Denote the set on the right-hand side by D. We show that (a) limq(B>+o p(B) = 0 uniformly with respect to p E D and (b) if {ppnl}l

_ D and p,(B) -- p(B) for

all B E Z, then p E D. Then a classical result of Bartle, Dunford, and Schwartz guarantees that D is weakly compact (see, e.g., Dunford and Schwartz (1958, Chapter IV)).s

(a) Notice that D'(p II q) < D'(p II q) for all p E A", so that

S= {p e d":D'(p II q) < t} {p E A: D(p II q) < t}= C.

1"Part (a) guarantees that D is relatively sequentially weakly compact; that is, every sequence {p,n in D admits a weakly convergent subsequence {p,j }. Part (b) guarantees that the limit of p,j belongs to D; that is, D is sequentially weakly compact. The Eberlein-Smulian theorem guaran- tees that D is weakly compact.

Page 43: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1488 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

Next we show that limq(B0>o

p(B) = 0 uniformly with respect to p E C and a fortiori with respect to p E D. Let p E C. Clearly,

w • dq< _ wo? - )dq<t+1 forallAeX. JA U'w \9/Ud q) f dq Y

For all e > 0, because lim,,,, _

= 0+, there exists 6 > 0 such that WO(T) b

0< <b if b> ,

w4 (b) 2(t + 1) 28'

in particular, 04(b) > 0. For all B such that q(B) < 8 and all p E C,

p(B) =

qdq-= dq + dq

Sdq W - 2nld dq nU S1 dq dp 18 dq

WC•Pdo < --q(B) + +w w dq

25 2 8nt > . fL) ( LP) dq

21 8Ed p <-8+

wI (ddq<e. 2 2(t+1)J n•>

- dq

This concludes the proof of (a). Let u (B) = fBw dq for all B E I. Notice that

[ fdp .d, if p E zAU(q), (33) D'(p II q) = I dq (q), 00, otherwise.

Denote by H,, the set of all finite partitions 7r of S in X finer than rw, = {w- (b) :b e w(S)}. For all r E Hw,, define on Ad the function

Dw(pl, II q,)=- q(A) (A), 0

A E,7 q(A)

with the conventions p -A) = 0 if p(A) = 0 and p( ) )t(A) =- o if p(A) > 0, when q(A) = 0. Notice that because rT is finer than r,,,, then for all A e :r, w(A) is a singleton (improperly denoted by w(A)) and

(A)= wdq = w(A)q(A).

CLAIM 1: We have Dl(p II q) = sup,,, D'(p, II q,) for all p E A'.

Page 44: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1489

PROOF: Let p << q. Take an increasingly finer sequence 7rn in IH such that dp/dq is (U, l irr,)-measurable. Then

1A ) IA --

dp q-a.e. q(A) dq

(see, e.g., Billingsley (1995, p. 470)), hence ctL-a.e. Continuity of 4 and the Fa- tou lemma imply

J ()dp\ < liminf O(p(A)N)

,Aq) q(A) (A AEi n

< sup : (p - (A).

AtE!!w

The converse inequality is guaranteed if f 4)(L)d-= coc. Else, if dq(L) is

/t-integrable, the conditional Jensen inequality guarantees that, for all iTr e ,,

EM? Em7dqI m-a.e., ?dq) \dq )

where m = Ix/It(S)

and E is the expectation. Hence,

Em( =-Em(Em (dp7 >Em kEm d V \) ((o dq) )

-((E (dp for all r E H,, but

dqdp 1 dp (dq E:m(A)O m(A)JA dq

AE~r: m(A)ZO

A(I (S) 1

dpd) Ar: L(A) (A) A(S)Adq

AE7T: q(A)0 (w(A)q(A) dq

- 1A w(A)q(A)w(A) (A))

AE-i: q(A) wO

= 1A

*(A) AET q

Page 45: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1490 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

Therefore,

f dp ()=d. - t(S)Em ( (dp) dq)(SE ( dq

>/-t(S)Em(C( A1A(P(A))) (q(A)

t (A), (p(A)A)

AE7T

so that

fS ) dp L sup j ( p() A). riH AE. q(A) )

If p E Ao\Ar(q), there exists B E . such that p(B) > 0 and q(B) = 0, and we can assume it belongs to some r 1Hw,. Therefore,

S(p(A)p(B) w supp P (A) i(A)> 4 p(B B)=

oc-= D(p 11 q). reHU' Ar, Kq(A) q(B))

This concludes the proof of the claim. Q.E.D.

(b) For all Ir E H,, set D, = {p E A':DI(p,1 II q,) < t}. We show that

{Pn}n •l D), and p,(B) - p(B) for all B E I imply p E D,. First notice that p E AO (for the Vitali-Hahn-Saks theorem). For all A e 7T such that q(A) = 0, then p,(A) = 0 for all n > 1 (else Dw((p,), 1I q,) = cx) and hence

Pnq(A) (0) = (kp (A)) q(A) q(A))

For all A E Tr such that q(A) > 0, clearly pn(A) p(A). We conclude that

h (p(A) j(A) = lim (Pn(A) YA)< t,

AE7T q(A) A q(A)

as wanted. Now (b) descends from Claim 1 and the observation that

p= {pe dA:Dw(p II q) < t}

= pe 6": sup D(p, II q,)5

t = N ,. "rrEHw rEn-w

Page 46: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1491

This completes the proof when w is simple. Suppose now that w is any 2-measurable function with w = infSEs w(s) > 0. Then there exists a sequence of simple 2-measurable functions w, such that w, T w and minses w,n(s) > w for all n > 1. By Levi's monotone convergence theorem, D "(p | q) T DW (p | q) for all p E A-(q), therefore

D= {pe A'(q):Dw(p II q) < t}

= p A'(q):supDn(pII q)< tJ n =

nl n>l

We conclude that D as well is weakly compact. Q.E.D.

PROOF OF THEOREM 16: Lemma 15 and Theorem 3 guarantee that diver-

gence preferences are variational. Unboundedness of u together with Proposi- tion 6 guarantees that c*(-) = OD'(- 11 q). Finally, Theorem 13 and Lemma 15 imply that > satisfies Axiom A.8. Q.E.D.

PROOF OF THEOREM 18: Set I(cp)= minpe,,A)(fpdp + c*(p)) for all p B0(Z). Because -c* coincides with the Fenchel conjugate of I (see

Lemma 26) on A(2) and I*(/t) = - 00 for all/t e ba(Y)\A(2), then

d(?p) = arg min ( dp + c*(p)) V Bo(),

U II((p)=

U argminfu(f)dp+c*(p) qpeBo(2) fEBo(2)

and I is Gateaux differentiable on Bo(2) if and only if P* is strictly concave on line segments in UEBO(l) I(cp), that is, if and only if c* is essentially strictly convex.19

Let f E Bo(X). Set A = {u'(f) drIr e arg minpEa(f u(f) dp + c*(p))} and no- tice that A is the image of argminpEa(f u(f) dp + c*(p)) = 8I(u(f)) through the map from A(2) to ba(2) that associates to p(.) in A(X) the bounded and finitely additive set function

4.) u'(f) dp in ba(2). Because this map is linear

and a(ba(2), Bo(2))-o-(ba(2), Bo(2))-continuous, and because dI(u(f)) is weakly* compact and convex, then A is or(ba(X), Bo(,))-compact and convex.

19This is proved in Bauschke, Borwein, and Combettes (2001). Notice that our definition of essential strict convexity is weaker than theirs and it guarantees that strict convexity implies es- sential strict convexity.

Page 47: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1492 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

By standard results,

(34) I'(u(f); u'(f)h)= mm(f

u'(f)hdp

= minf h dp, 'Vhe Bo(). I-A EA J

Let h E Bo(2) and s E S. If h(s) 0 O, then

u(f + th)(s) = u(f (s) + th(s)) = u(f(s)) + u'(f(s))th(s) + o(th(s))

for th(s) -+ 0, that is,

lim u(f (s) + th(s)) - u(f (s)) - u'(f (s))th(s) -

th(s)-0O th(s) and so

u(f (s) + th(s)) - u(f (s)) - u'(f (s))h(s)t (35) lim = 0.

t3,O t

Clearly (35) holds also if h(s) = 0. Because f and h are simple, the preceding limit is uniform with respect to s e S. Therefore, for t 4 0,

< I(u(f + th)) - I(u(f)) minhd -t /la n

< (I(u(f + th)) - I(u(f) + tu'(f)h)

+ I(u(f) + tu'(f)h) - I(u(f))))/t- minf

h

di I(u(f + th)) - I(u(f) + tu'(f)h)

t

? I(u(f) + tu'(f)h) - I(u(f)) minf hd t XEA

Ilu(f + th) - u(f) - tu'(f)hll + o(1), t

where the last inequality descends from the Lipschitz continuity of I and (34). Uniformity of limit (35) then delivers

im I(u(f + th)) - I(u(f)) -minm hd

tq0 t /eA d

Page 48: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1493

for all h E B0o() or V'(f; -) = min, Ef(.-) d/u, that is, OV(f) = A.

Now, if c* is essentially strictly convex, then dI(u(f)) is a singleton for every f in B0o() and dV(f) = {u'(f) dr I r E (u(f))} is a singleton too. Conversely, assume that V is Gateaux differentiable on Bo(0) and per contra that c* is not essentially strictly convex. Then there exists ?0 E B0o() such that I(40p) contains two distinct elements r, and r2. Because u: R - R is concave, then it is unbounded below and there is b E R such that p? + b E Bo(2, u(R)). In particular, there exists f E Bo(V) such that u(f) = ?? + b and

rl, r2E

I(4)-=-arg min (fdp+c*(p)

=-argmin (cf(+b)dp+c*(p))= d(u(f)).

It follows that u'(f) dr1, u'(f) dr2 e •V(f).

Since V is Gateaux differentiable on Bo(X), then u'(f) dr1 = u'(f) dr2, that is,

(36) f hu'(f) dr = f hu'(f) dr2 VhE Bo().

Because u is strictly monotonic, concave, and differentiable, then u'(z) # 0 for all z E R and (36) implies f f

dr- = f Ji dr2 for all & E Bo(Y), contradicting

rl r2. Q.E.D.

PROOF OF PROPOSITION 19: Let (u, c*) represent - as in Theorem 3 and

without loss of generality assume [-1, 1] _

u(X). (i) =, (ii) By Gilboa and Schmeidler (1989, Theorem 1), there is a weakly*

compact and convex set C _

A(2) such that u(xf) - minPEC f u(f) dp for all f E F and each xf - f. By Theorem 3,

(37) c*(p) = sup (minf u(f)dq - u(f) d) dp Vp A().

rf.T ?qECff

Suppose p E C. Then c*(p) < 0. Because c* is nonnegative, we have c*(p) = 0. Next, suppose po V C. By the separating hyperplane theorem, there is a simple measurable function p: S -+ u(X) such that f cp dp > f 1? dpo for each p E C. Hence, taking f E T such that 40 = u(f), minpc f u(f) dp - f u(f) dpo > 0, which in turn implies c*(po) > 0. We conclude that c*(p) = 0 if and only if p E C. Therefore, for all f E F,

min(fu(f) dp +c*(p)) = u() = minf u(f) dp pE(f) peCf

= min u(f) dp. pE{c*=0}

Page 49: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1494 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

(ii) =4 (i) and (iii) =- (ii) are trivial. Now assume u(X) is unbounded above (resp., below).

(ii) ?= (iii) For all pE A(E),

c*(p)= sup mm in u(f)dq- u(f)dp fey qEf{c*=O)f}

S sup (minf pdq - f pdp. oPeBo(X,u(X)) \q

({c*=0}J J

Suppose, c*(po) > 0. There exist a nonnegative (resp. nonpositive), simple, measurable function p: S -- u(X) and e > 0 such that minpeIc =01o f dp -

f po dpo > e, but nqp E Bo(V, u(X)) for all n E N and

c*(po)> minf n dp - n

•pdpo > n

for all n E N. We conclude c*(po) = 0 . Q.E.D.

The proof of Corollary 20 is omitted. Just notice that Axiom A.5' can be used to obtain affinity of the functional I that appears in Lemma 28 in the same way in which Axiom A.5 is used to obtain its concavity at the beginning of the proof of Theorem 3.

LEMMA 32: Let - be a variational preference represented by (u, c*) as in The- orem 3 and let q e A(2). The following conditions are equivalent:

(i) q corresponds to a SEU preference that is less ambiguity averse than >-. (ii) c*(q) = 0.

(iii) q e 81(k) for some (all) k e u(X), where I(?p) = minpa,,()(f(p dp + c*(p)) for all Op E B0o(, u(X)).

In particular, any variational preference is ambiguity averse.

PROOF: (i) =? (ii) Suppose to is a SEU preference, with associated subjec- tive probability q and nonconstant affine utility index uo such that > is more ambiguity averse than >o. By (8), we can assume uo = u. By Proposition 8, and c* < co, and by Corollary 20, co(q) = 0; hence, 0 < c*(q) < c(*(q) = 0.

(ii) =? (iii) We have c*(q) = 0 if and only if supEB,,0(z,,,,x(IX(p) - f ?p dq) = 0

if and only if I((p) Sf p dq for all 'p e Bo(2, u(X)) if and only if I(Gp) - I(k) <

f 'p dq - f k dq for all 'p e B0o(?, u(X)) and all k e u(X) if and only if q e lI(k) for all k e u(X).

(iii) =4 (i) If q e dl(k) for some k e u(X), then I('p) - I(k) < f p dq -

f k dq for all 'p e B0o(, u(X)) and then I('p) <f p dq for all 'p e B0o(, u(X)). Denote by o0 the SEU preference, with associated subjective probability q and utility index u. Notice that for all fe •F

and x e X, f x implies I(u(f)) > u(x), a fortiori f u(f) dq > u(x) and f -o x.

Page 50: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1495

Ambiguity aversion of >- now follows from the observation that arg minpEA() c*(p) is nonempty and minpEA(.) c*(p) =0. Q.E.D.

PROOF OF PROPOSITION 22: (i) Let 0, 4 0. By Proposition 12,

limmin f u(f) dp +p OnD(p 1 q

= minff u(f) dp + {D(-lIq)<oo}(P)J

If w E LO(q) and w = infs s w(s), then wDp(p I| q) < Dl(p II q) < IlwllK x

D4(p II q), which implies domDp(. II q) = domDw(. II q). In turn, this implies

limminf fu(f)dp + OnD(pl q) mDinf q)

u(f) dp. n peA IfJpedomD(.IlIq)

J

Next we show that, for all cp E Bo( ),

(38) inf f pdp = inf f dp. pedomD4(.I q)J peA" (C,q)J

Let cp = •,l

ajlAj, with a, > ... > a, and {Ai}il

a partition of S in . For

every p E A(X, q), define p, E Bo(2) by

p- q• (A> ) i:Ai

+ 1Ai i: q(Ai)>0 i:

q(Ai)=0 It is easy to see that q, is nonnegative and f rp dq = 1. Call p' the element of

A(~(, q) such that dp'/dq = fp,. Since p,

is a simple function, then DO(p' II q) = f 4(frjp)dq E R, so that p' E domDo(. I| q). Since f cp dp' = l p dp, we then have I{ fp dp:p e A(X, q)I} C {f p dp:p E domD(. II q)}, which yields (38) since the converse inclusion is trivial.

It remains to prove that

(39) inf f dp = ess

min•p(s).

pedAa(q) J seS

Here essmin,,s p(s) = min{ja :q(Ai) > 0}. Let i* E {1,..., n} be such that

aoi = min{ai: q(Ai) > 0} and let qA,. be the conditional distribution of q

on Ai,.

Then qAi, E A?(q), and f p dqA, = ess min,,s 4?(s). This proves (39). (ii) Let 6,, f oc. By Proposition 23, strict convexity of 4 implies that of

D (p 11 q) on its effective domain. Hence, argmin O,,D(p

1I q) - {q} for each n and so the result follows from Proposition 12. Q.E.D.

Page 51: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1496 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

PROOF OF PROPOSITION 23: By Theorem 18, it is enough to show that Dw(. I q): A(,) -- [0, 00] is strictly convex on its effective domain. Let p, p' E domD'(. II q), p

- p', and a e (0, 1). Let pt(B) =

fB wdq for all B E . The assumption infses w(s) > 0 guarantees that the measure pt is equiv- alent to the probability q. By (33), p, p' e domD(. II q) if and only if

f () \ d,, f (dp'' ) d < 00 if and only if d( L), •(

) e L1(X, ). Convex-

ity and nonnegativity of 0 guarantee that

< dp dp'\ (1 - a) (s) d ( a -q

dq /

( dp N d dp' <a(OP -(s) + (1- a)o4 -(s) )VseS;

dq /(dq

strict convexity, p : p', and a E (0, 1) imply that the second inequality is strict on a set E E X with q(E) > 0, and hence g(E) > 0. Therefore, 0(aL + (1 - a )) E L1 (L , g) and

-a dp (s) I (1 - a)4 d 0 (s)

S dp +(1-a)dp'\d>0,

S

dq dF q )-J.

that is, Dw(ap + (1 - a)p' II q) < aD (p II q) + (1 - a)Dw(p' II q), as de- sired. Q.E.D.

PROOF OF THEOREM 24: The functional J: Bo() -B + R defined by J(p) = f ~ dq - (20)-' Varq(Gp) is concave and Gateaux differentiable. Concavity is trivial. Moreover, for all p, I e Bo(V) and t E IR,

J( + to) = (s + tip) dq - 1Varq(?q + tf)

20

+ [J(1-i p Jpdq ))dq]t+J(p), +[f ( 0

so that

(40) J'(Gp; q) =

limt

J( + tq) -

- J( dq dq. t--+O ~t f 0

That is, for all p e B0o(2), the Gateaux differential of J at (p is represented by a measure with Radon-Nikodym derivative with respect to q given by 1 - 1 ('p -

Page 52: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

VARIATIONAL REPRESENTATION OF PREFERENCES 1497

f ?p dq). Therefore, J'(cp; -) is positive as a linear functional on B0o() if and only if

q({s S:p(s) -fJ dq })1. This relation characterizes the elements of the domain of monotonicity M. In Maccheroni, Marinacci, Rustichini, and Taboga (2004) we show that

mJ()-•=qin (f dp+?OG(p1lq))

V( M. Q.E.D.

REFERENCES

ANSCOMBE, E J., AND R. J. AUMANN (1963): "'A Definition of Subjective Probability," TheAnnals

of Mathematical Statistics, 34, 199-205. [1453] ARROW, K. (1970): Essays in the Theory of Risk-Bearing. Amsterdam: North-Holland. [1460] BAUSCHKE, H. H., J. M. BORWEIN, AND P. L. COMBETTES (2001): "Essential Smoothness, Essen-

tial Strict Convexity, and Legendre Functions in Banach Spaces," Communications in Contem- porary Mathematics, 3, 615-647. [1491]

BEN-TAL, A. (1985): "The Entropic Penalty Approach to Stochastic Programming," Mathematics

of Operations Research, 10, 263-279. [1464] BEN-TAL, A., A. BEN-ISRAEL, AND M. TEBOULLE (1991): "Certainty Equivalents and Informa-

tion Measures: Duality and Extremal Principles," Journal of Mathematical Analysis and Appli- cations, 157, 211-236. [1464]

BILLINGSLEY, P. (1995): Probability and Measure (Third Ed.). New York: Wiley. [1489] CHATEAUNEUF, A., E MACCHERONI, M. MARINACCI, AND J.-M. TALLON (2005): "Monotone

Continuous Multiple Priors," Economic Theory, 26, 973-982. [1460] CHEN, Z., AND L. G. EPSTEIN (2002): 'Ambiguity, Risk, and Asset Returns in Continuous Time,"

Econometrica, 70, 1403-1443. [1448] CHONG, K. M. (1976): "Doubly Stochastic Operators and Rearrangement Theorems," Journal of

Mathematical Analysis and Applications, 56, 309-316. [1484] CHONG, K. M., AND N. M. RICE (1971): "Equimeasurable Rearrangements of Functions,"

Queens Papers in Pure and Applied Mathematics, 28. [1461,1463,1484-1487] DAL MASO, G. (1993): An Introduction to F-Convergence. Boston: Birkhauser. [1460,1482] DANA, R.-A. (2005): "A Representation Result for Concave Schur Concave Functions," Mathe-

matical Finance, 15, 613-634. [1463,1485] DOLECKI, S., AND G. H. GRECO (1995): "Niveloids," Topological Methods in Nonlinear Analysis,

5, 1-22. [1475,1476] DUNFORD, N., AND J. T. SCHWARTZ (1958): Linear Operators, Part I. New York: Wiley. [1475,

1487] DUPUIS, P., AND R. S. ELLIS (1997): A Weak Convergence Approach to the Theory of Large Devia-

tions. New York: Wiley. [1471] ELLSBERG, D. (1961): "Risk, Ambiguity, and the Savage Axioms," Quarterly Journal of Eco-

nomics, 75, 643-669. [1447,1465] EPSTEIN, L. G. (1999): 'A Definition of Uncertainty Aversion," Review of Economic Studies, 66,

579-608. [1451,1454,1461] EPSTEIN, L. G., AND T. WANG (1994): "Intertemporal Asset Pricing under Knightian Uncer-

tainty," Econometrica, 62, 283-322. [1448,1467,1470] FOLLMER, H., AND A. SCHIED (2002): Stochastic Finance. Berlin: de Gruyter. [1476]

Page 53: Ambiguity Aversion, Robustness, and the Variational Representation of Preferencesstatic.luiss.it/hey/ambiguity/papers/Maccheroni_Marinac... ·  · 2009-10-06Econometrica, Vol. 74,

1498 E MACCHERONI, M. MARINACCI, AND A. RUSTICHINI

GHIRARDATO, P., AND M. MARINACCI (2002): '"Ambiguity Made Precise: A Comparative Foun- dation," Journal of Economic Theory, 102, 251-289. [1451,1454,1457,1459]

GILBOA, I., AND D. SCHMEIDLER (1989): "Maxmin Expected Utility with a Non-Unique Prior," Journal of Mathematical Economics, 18, 141-153. [1447,1449,1452,1454,1468,1493]

HANSEN, L., AND T. SARGENT (2000): "Wanting Robustness in Macroeconomics," Mimeo, Uni-

versity of Chicago and Stanford University. [1448,1472] (2001): "Robust Control and Model Uncertainty," American Economic Review, 91,

60-66. [1448,1460,1464,1468,1472] HERSTEIN, I. N., AND J. MILNOR (1953): "An Axiomatic Approach to Measurable Utility,"

Econometrica, 21, 291-297. [1478] Hsu, M., M. BHATT, R. ADOLPHS, D. TRANEL, AND C. E CAMERER (2005): "Neural Systems

Responding to Degrees of Uncertainty in Human Decision-Making," Science, 310, 1680-1683. [1452]

KEREN, G., AND L. E. GERRITSEN (1999): "On the Robustness and Possible Accounts of Ambi-

guity Aversion," Acta Psychologica, 103, 149-172. [1452] KOPYLOV, I. (2001): "Procedural Rationality in the Multiple Prior Model," Mimeo, University of

Rochester. [1457] KOHBERGER, A., AND J. PERNER (2003): "The Role of Competition and Knowledge in the Ells-

berg Task," Journal of Behavioral Decision Making, 16, 181-191. [1452] LIESE, E, AND I. VAJDA (1987): Convex Statistical Distances. Leipzig: Teubner. [1463,1470] LUXEMBURG, W. A. J. (1967): "Rearrangement Invariant Banach Function Spaces," Queens Pa-

pers in Pure and Applied Mathematics, 10, 83-144. [1463,1484-1486] MACCHERONI, E, M. MARINACCI, AND A. RUSTICHINI (2005): "Niveloids and Their Extensions,"

Mimeo. [1476] (2006): "Dynamic Variational Preferences," Journal of Economic Theory, 128, 4-44.

[1467,1472] MACCHERONI, E, M. MARINACCI, A. RUSTICHINI, AND M. TABOGA (2004): "Portfolio Selection

with Monotone Mean-Variance Preferences," ICER Working Papers, Applied Mathematics Series 27-2004. [1475,1497]

MACHINA, M. J., AND D. SCHMEIDLER (1992): "A More Robust Definition of Subjective Proba-

bility," Econometrica, 60, 745-780. [1461] MARINACCI, M. (2002): "Probabilistic Sophistication and Multiple Priors," Econometrica, 70,

755-764. [1463] MARKOWITZ, H. M. (1952): "Portfolio Selection," Journal of Finance, 7, 77-91. [1450,1460,1468,

1474] MARSHALL, A. W., AND I. OLKIN (1979): Inequalities: Theory of Majorization and Its Applications.

New York: Academic Press. [1462] PHELPS, R. R. (1992): Convex Functions, Monotone Operators and Differentiability. New York:

Springer-Verlag. [1466,1476] ROCKAFELLAR, R. T. (1970): Convex Analysis. Princeton, NJ: Princeton University Press. [1449,

1459,1466] ROTHSCHILD, M., AND J. E. STIGLITZ (1970): "Increasing Risk: I. A Definition," Journal of Eco-

nomic Theory, 2, 225-243. [1462] RUSTICHINI, A. (2005): "Emotion and Reason in Making Decisions," Science, 310, 1624-1625.

[1452] SCHMEIDLER, D. (1979): '"A Bibliographical Note on a Theorem of Hardy, Littlewood, and

Polya," Journal of Economic Theory, 20, 125-128. [1461] (1989): "Subjective Probability and Expected Utility Without Additivity," Econometrica,

57, 571-587. [1451,1454] TOBIN, J. (1958): "Liquidity Preference as Behavior Toward Risk," Review of Economic Studies,

25, 65-86. [1450,1460,1468,1474] WANG, T. (2003): "A Class of Multi-Prior Preferences," Mimeo, University of British Columbia.

[1473]