HOME

TheInfoList



OR:

In
cooperative game theory In game theory, a cooperative game (or coalitional game) is a game with competition between groups of Player (game), players ("coalitions") due to the possibility of external enforcement of cooperative behavior (e.g. through contract law). Those ...
and social choice theory, the Nakamura number measures the degree of rationality of preference aggregation rules (collective decision rules), such as voting rules. It is an indicator of the extent to which an aggregation rule can yield well-defined choices. *If the number of alternatives (candidates; options) to choose from is less than this number, then the rule in question will identify "best" alternatives without any problem. In contrast, *if the number of alternatives is greater than or equal to this number, the rule will fail to identify "best" alternatives for some pattern of voting (i.e., for some profile (
tuple In mathematics, a tuple is a finite ordered list (sequence) of elements. An -tuple is a sequence (or ordered list) of elements, where is a non-negative integer. There is only one 0-tuple, referred to as ''the empty tuple''. An -tuple is defi ...
) of individual preferences), because a voting paradox will arise (a ''cycle'' generated such as alternative a socially preferred to alternative b, b to c, and c to a). The larger the Nakamura number a rule has, the greater the number of alternatives the rule can rationally deal with. For example, since (except in the case of four individuals (voters)) the Nakamura number of majority rule is three, the rule can deal with up to two alternatives rationally (without causing a paradox). The number is named after (1947–1979), a Japanese game theorist who proved the above fact that the rationality of collective choice critically depends on the number of alternatives.


Overview

To introduce a precise definition of the Nakamura number, we give an example of a "game" (underlying the rule in question) to which a Nakamura number will be assigned. Suppose the set of individuals consists of individuals 1, 2, 3, 4, and 5. Behind majority rule is the following collection of ("decisive") ''coalitions'' (subsets of individuals) having at least three members: : A Nakamura number can be assigned to such collections, which we call ''simple games''. More precisely, a simple game is just an arbitrary collection of coalitions; the coalitions belonging to the collection are said to be ''winning''; the others ''losing''. If all the (at least three, in the example above) members of a winning coalition prefer alternative x to alternative y, then the society (of five individuals, in the example above) will adopt the same ranking (''social preference''). The Nakamura number of a simple game is defined as the minimum number of winning coalitions with empty intersection. (By intersecting this number of winning coalitions, one can sometimes obtain an empty set. But by intersecting less than this number, one can never obtain an empty set.) The Nakamura number of the simple game above is three, for example, since the intersection of any two winning coalitions contains at least one individual but the intersection of the following three winning coalitions is empty: \, \, \. Nakamura's theorem (1979) gives the following necessary (also sufficient if the set of alternatives is finite) condition for a simple game to have a nonempty "core" (the set of socially "best" alternatives) for all profiles of individual preferences: the number of alternatives is less than the Nakamura number of the simple game. Here, the core of a simple game with respect to the profile of preferences is the set of all alternatives x such that there is no alternative y that every individual in a winning coalition prefers to x; that is, the set of ''maximal'' elements of the social preference. For the majority game example above, the theorem implies that the core will be empty (no alternative will be deemed "best") for some profile, if there are three or more alternatives. Variants of Nakamura's theorem exist that provide a condition for the core to be nonempty (i) for all profiles of ''acyclic'' preferences; (ii) for all profiles of ''transitive'' preferences; and (iii) for all profiles of ''linear orders''. There is a different kind of variant (Kumabe and Mihara, 2011), which dispenses with ''acyclicity'', the weak requirement of rationality. The variant gives a condition for the core to be nonempty for all profiles of preferences that have ''maximal elements''. For ''ranking'' alternatives, there is a very well known result called "
Arrow's impossibility theorem Arrow's impossibility theorem, the general possibility theorem or Arrow's paradox is an impossibility theorem in social choice theory that states that when voters have three or more distinct alternatives (options), no ranked voting electoral syst ...
" in social choice theory, which points out the difficulty for a group of individuals in ranking three or more alternatives. For ''choosing'' from a set of alternatives (instead of ''ranking'' them), Nakamura's theorem is more relevant. An interesting question is how large the Nakamura number can be. It has been shown that for a (finite or) algorithmically computable simple game that has no veto player (an individual that belongs to every winning coalition) to have a Nakamura number greater than three, the game has to be ''non-strong''. This means that there is a ''losing'' (i.e., not winning) coalition whose complement is also losing. This in turn implies that nonemptyness of the core is assured for a set of three or more alternatives only if the core may contain several alternatives that cannot be strictly ranked.


Framework

Let N be a (finite or infinite) nonempty set of ''individuals''. The subsets of N are called coalitions. A simple game (voting game) is a collection W of coalitions. (Equivalently, it is a coalitional game that assigns either 1 or 0 to each coalition.) We assume that W is nonempty and does not contain an empty set. The coalitions belonging to W are ''winning''; the others are ''losing''. A simple game W is monotonic if S \in W and S\subseteq T imply T \in W. It is proper if S \in W implies N\setminus S \notin W. It is strong if S \notin W implies N\setminus S \in W. A veto player (vetoer) is an individual that belongs to all winning coalitions. A simple game is nonweak if it has no veto player. It is finite if there is a finite set (called a ''carrier'') T \subseteq N such that for all coalitions S, we have S \in W iff S\cap T \in W. Let X be a (finite or infinite) set of ''alternatives'', whose
cardinal number In mathematics, cardinal numbers, or cardinals for short, are a generalization of the natural numbers used to measure the cardinality (size) of sets. The cardinality of a finite set is a natural number: the number of elements in the set. T ...
(the number of elements) \# X is at least two. A (strict) preference is an ''asymmetric'' relation \succ on X: if x \succ y (read "x is preferred to y"), then y\not \succ x. We say that a preference \succ is ''acyclic'' (does not contain ''cycles'') if for any finite number of alternatives x_1, \ldots, x_m, whenever x_1 \succ x_2, x_2 \succ x_3,…, x_ \succ x_m, we have x_m \not\succ x_1. Note that acyclic relations are asymmetric, hence preferences. A profile is a list p=(\succ_i^p)_ of individual preferences \succ_i^p. Here x \succ_i^p y means that individual i prefers alternative x to y at profile p. A ''simple game with ordinal preferences'' is a pair (W, p) consisting of a simple game W and a profile p. Given (W, p), a ''dominance'' (social preference) relation \succ^p_W is defined on X by x \succ^p_W y if and only if there is a winning coalition S \in W satisfying x \succ_i^p y for all i \in S. The core C(W,p) of (W, p) is the set of alternatives undominated by \succ^p_W (the set of maximal elements of X with respect to \succ^p_W): :x \in C(W,p) if and only if there is no y\in X such that y \succ^p_W x.


Definition and examples

The Nakamura number \nu(W) of a simple game W is the size (cardinal number) of the smallest collection of winning coalitions with empty intersection: :\nu(W)=\min\ if \cap W = \cap_ S = \emptyset (no veto player); otherwise, \nu(W)= +\infty (greater than any cardinal number). it is easy to prove that if W is a simple game without a veto player, then 2\le \nu(W)\le \# N. Examples for finitely many individuals (N=\) (see Austen-Smith and Banks (1999), Lemma 3.2). Let W be a simple game that is monotonic and proper. *If W is strong and without a veto player, then \nu(W)=3. *If W is the majority game (i.e., a coalition is winning if and only if it consists of more than half of individuals), then \nu(W)=3 if n\ne 4; \nu(W)=4 if n=4. *If W is a q-rule (i.e., a coalition is winning if and only if it consists of at least q individuals) with n/2, then \nu(W)= /(n-q)/math>, where /math> is the smallest integer greater than or equal to x. Examples for at most countably many individuals (N=\). Kumabe and Mihara (2008) comprehensively study the restrictions that various properties (monotonicity, properness, strongness, nonweakness, and finiteness) for simple games impose on their Nakamura number (the Table "Possible Nakamura Numbers" below summarizes the results). In particular, they show that an algorithmically computable simple game See a section for Rice's theorem for the definition of a computable simple game. In particular, all finite games are computable. without a veto player has a Nakamura number greater than 3 only if it is proper and nonstrong.


Nakamura's theorem for acyclic preferences

Nakamura's theorem (Nakamura, 1979, Theorems 2.3 and 2.5). Let W be a simple game. Then the core C(W,p) is nonempty for all profiles p of acyclic preferences if and only if X is finite and \# X < \nu(W). Remarks *Nakamura's theorem is often cited in the following form, without reference to the core (e.g., Austen-Smith and Banks, 1999, Theorem 3.2): The dominance relation \succ_W^p is acyclic for all profiles p of acyclic preferences if and only if \# B< \nu(W) for all finite B \subseteq X (Nakamura 1979, Theorem 3.1). *The statement of the theorem remains valid if we replace "for all profiles p of ''acyclic'' preferences" by "for all profiles p of ''negatively transitive'' preferences" or by "for all profiles p of ''linearly ordered'' (i.e., transitive and total) preferences". *The theorem can be extended to \mathcal-simple games. Here, the collection \mathcal of coalitions is an arbitrary
Boolean algebra In mathematics and mathematical logic, Boolean algebra is a branch of algebra. It differs from elementary algebra in two ways. First, the values of the variables are the truth values ''true'' and ''false'', usually denoted 1 and 0, whereas i ...
of subsets of N, such as the \sigma-algebra of Lebesgue measurable sets. A \mathcal-''simple game'' is a subcollection of \mathcal. Profiles are suitably restricted to measurable ones: a profile p is ''measurable'' if for all x, y \in X, we have \ \in \mathcal.


A variant of Nakamura's theorem for preferences that may contain cycles

In this section, we discard the usual assumption of acyclic preferences. Instead, we restrict preferences to those having a maximal element on a given ''agenda'' (''opportunity set'' that a group of individuals are confronted with), a subset of some underlying set of alternatives. (This weak restriction on preferences might be of some interest from the viewpoint of behavioral economics.) Accordingly, it is appropriate to think of X as an ''agenda'' here. An alternative x \in X is a ''maximal'' element with respect to \succ_i^p (i.e., \succ_i^p has a maximal element x) if there is no y \in X such that y\succ_i^p x. If a preference is acyclic over the underlying set of alternatives, then it has a maximal element on every ''finite'' subset X. We introduce a strengthening of the core before stating the variant of Nakamura's theorem. An alternative x can be in the core C(W,p) even if there is a winning coalition of individuals i that are "dissatisfied" with x (i.e., each i prefers some y_i to x). The following solution excludes such an x: :An alternative x\in X is in the core C^+(W,p) without majority dissatisfaction if there is no winning coalition S\in W such that for all i \in S, x is ''non-maximal'' (there exists some y_i \in X satisfying y_i \succ_i^p x). It is easy to prove that C^+(W,p) depends only on the set of maximal elements of each individual and is included in the union of such sets. Moreover, for each profile p, we have C^+(W,p) \subseteq C(W,p). A variant of Nakamura's theorem (Kumabe and Mihara, 2011, Theorem 2). Let W be a simple game. Then the following three statements are equivalent: #\# X < \nu(W); #the core C^+(W,p) without majority dissatisfaction is nonempty for all profiles p of preferences that have a maximal element; #the core C(W,p) is nonempty for all profiles p of preferences that have a maximal element. Remarks *Unlike Nakamura's original theorem, X being finite is ''not a necessary condition'' for C^+(W,p) or C(W,p) to be nonempty for all profiles p. Even if an agenda X has infinitely many alternatives, there is an element in the cores for appropriate profiles, as long as the inequality \# X < \nu(W) is satisfied. *The statement of the theorem remains valid if we replace "for all profiles p of preferences that have a maximal element" in statements 2 and 3 by "for all profiles p of preferences that have ''exactly one'' maximal element" or "for all profiles p of ''linearly ordered'' preferences that have a maximal element" (Kumabe and Mihara, 2011, Proposition 1). *Like Nakamura's theorem for acyclic preferences, this theorem can be extended to \mathcal-simple games. The theorem can be extended even further (1 and 2 are equivalent; they imply 3) to ''collections'' W' \subseteq \mathcal' ''of winning sets'' by extending the notion of the Nakamura number.


See also

* Gibbard–Satterthwaite theorem * May's theorem * Voting paradox


Notes

{{Reflist, 30em Social choice theory Voting theory Cooperative games Economics theorems