Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Mathematical definition  





2 Cooperative game theory definition  





3 Subgames  





4 Properties for characterization  



4.1  Superadditivity  





4.2  Monotonicity  





4.3  Properties for simple games  







5 Relation with non-cooperative theory  





6 Solution concepts  



6.1  The stable set  



6.1.1  Properties  







6.2  The core  



6.2.1  Properties  







6.3  The core of a simple game with respect to preferences  





6.4  The strong epsilon-core  





6.5  The Shapley value  





6.6  The kernel  





6.7  Harsanyi dividend  





6.8  The nucleolus  



6.8.1  Properties  









7 Convex cooperative games  



7.1  Properties  





7.2  Similarities and differences with combinatorial optimization  







8 The relationship between cooperative game theory and firm  





9 See also  





10 References  



10.1  Further reading  







11 External links  














Cooperative game theory






العربية
Català
Čeština
Deutsch
Español
فارسی
Français

Italiano
עברית
Nederlands

Português
Русский
Українська


 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 




In other projects  



Wikimedia Commons
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 


Ingame theory, a cooperative game (orcoalitional game) is a game with groups of players who form binding “coalitions” with external enforcement of cooperative behavior (e.g. through contract law). This is different from non-cooperative games in which there is either no possibility to forge alliances or all agreements need to be self-enforcing (e.g. through credible threats).[1]

Cooperative games are analysed by focusing on coalitions that can be formed, and the joint actions that groups can take and the resulting collective payoffs.[2][3]

Mathematical definition[edit]

A cooperative game is given by specifying a value for every coalition. Formally, the coalitional game consists of a finite set of players , called the grand coalition, and a characteristic function [4] from the set of all possible coalitions of players to a set of payments that satisfies . The function describes how much collective payoff a set of players can gain by forming a coalition.

Cooperative game theory definition[edit]

Cooperative game theory is a branch of game theory that deals with the study of games where players can form coalitions, cooperate with one another, and make binding agreements. The theory offers mathematical methods for analysing scenarios in which two or more players are required to make choices that will affect other players wellbeing.[5]

Common interests: In cooperative games, players share a common interest in achieving a specific goal or outcome. The players must identify and agree on a common interest to establish the foundation and reasoning for cooperation. Once the players have a clear understanding of their shared interest, they can work together to achieve it.[citation needed]

Necessary information exchange: Cooperation requires communication and information exchange among the players. Players must share information about their preferences, resources, and constraints to identify opportunities for mutual gain. By sharing information, players can better understand each other's goals and work towards achieving them together.[citation needed]

Voluntariness, equality, and mutual benefit: In cooperative games, players voluntarily come together to form coalitions and make agreements. The players must be equal partners in the coalition, and any agreements must be mutually beneficial. Cooperation is only sustainable if all parties feel they are receiving a fair share of the benefits.[citation needed]

Compulsory contract: In cooperative games, agreements between players are binding and mandatory. Once the players have agreed to a particular course of action, they have an obligation to follow through. The players must trust each other to keep their commitments, and there must be mechanisms in place to enforce the agreements. By making agreements binding and mandatory, players can ensure that they will achieve their shared goal.[citation needed]

Subgames[edit]

Let be a non-empty coalition of players. The subgame on is naturally defined as

In other words, we simply restrict our attention to coalitions contained in . Subgames are useful because they allow us to apply solution concepts defined for the grand coalition on smaller coalitions.

Properties for characterization[edit]

Superadditivity[edit]

Characteristic functions are often assumed to be superadditive (Owen 1995, p. 213). This means that the value of a union of disjoint coalitions is no less than the sum of the coalitions' separate values:

whenever satisfy .

Monotonicity[edit]

Larger coalitions gain more:

.

This follows from superadditivity. i.e. if payoffs are normalized so singleton coalitions have zero value.

Properties for simple games[edit]

A coalitional game v is considered simple if payoffs are either 1 or 0, i.e. coalitions are either "winning" or "losing".[6]

Equivalently, a simple game can be defined as a collection W of coalitions, where the members of W are called winning coalitions, and the others losing coalitions. It is sometimes assumed that a simple game is nonempty or that it does not contain an empty set. However, in other areas of mathematics, simple games are also called hypergraphsorBoolean functions (logic functions).

A few relations among the above axioms have widely been recognized, such as the following (e.g., Peleg, 2002, Section 2.1[7]):

More generally, a complete investigation of the relation among the four conventional axioms (monotonicity, properness, strongness, and non-weakness), finiteness, and algorithmic computability[8] has been made (Kumabe and Mihara, 2011[9]), whose results are summarized in the Table "Existence of Simple Games" below.

Existence of Simple Games[10]
Type Finite Non-comp Finite Computable Infinite Non-comp Infinite Computable
1111 No Yes Yes Yes
1110 No Yes No No
1101 No Yes Yes Yes
1100 No Yes Yes Yes
1011 No Yes Yes Yes
1010 No No No No
1001 No Yes Yes Yes
1000 No No No No
0111 No Yes Yes Yes
0110 No No No No
0101 No Yes Yes Yes
0100 No Yes Yes Yes
0011 No Yes Yes Yes
0010 No No No No
0001 No Yes Yes Yes
0000 No No No No

The restrictions that various axioms for simple games impose on their Nakamura number were also studied extensively.[11] In particular, a computable simple game without a veto player has a Nakamura number greater than 3 only if it is a proper and non-strong game.

Relation with non-cooperative theory[edit]

Let G be a strategic (non-cooperative) game. Then, assuming that coalitions have the ability to enforce coordinated behaviour, there are several cooperative games associated with G. These games are often referred to as representations of G. The two standard representations are:[12]

Solution concepts[edit]

The main assumption in cooperative game theory is that the grand coalition will form.[13] The challenge is then to allocate the payoff among the players in some way. (This assumption is not restrictive, because even if players split off and form smaller coalitions, we can apply solution concepts to the subgames defined by whatever coalitions actually form.) A solution concept is a vector (or a set of vectors) that represents the allocation to each player. Researchers have proposed different solution concepts based on different notions of fairness. Some properties to look for in a solution concept include:

An efficient payoff vector is called a pre-imputation, and an individually rational pre-imputation is called an imputation. Most solution concepts are imputations.

The stable set[edit]

The stable set of a game (also known as the von Neumann-Morgenstern solution (von Neumann & Morgenstern 1944)) was the first solution proposed for games with more than 2 players. Let be a game and let , be two imputationsof. Then dominates if some coalition satisfies and . In other words, players in prefer the payoffs from to those from , and they can threaten to leave the grand coalition if is used because the payoff they obtain on their own is at least as large as the allocation they receive under .

Astable set is a set of imputations that satisfies two properties:

Von Neumann and Morgenstern saw the stable set as the collection of acceptable behaviours in a society: None is clearly preferred to any other, but for each unacceptable behaviour there is a preferred alternative. The definition is very general allowing the concept to be used in a wide variety of game formats.[citation needed]

Properties[edit]

The core[edit]

Let be a game. The coreof is the set of payoff vectors

In words, the core is the set of imputations under which no coalition has a value greater than the sum of its members' payoffs. Therefore, no coalition has incentive to leave the grand coalition and receive a larger payoff.

Properties[edit]

The core of a simple game with respect to preferences[edit]

For simple games, there is another notion of the core, when each player is assumed to have preferences on a set of alternatives. A profile is a list of individual preferences on. Here means that individual prefers alternative to at profile . Given a simple game and a profile , a dominance relation is defined on by if and only if there is a winning coalition (i.e., ) satisfying for all . The core of the simple game with respect to the profile of preferences is the set of alternatives undominated by (the set of maximal elements of with respect to ):

if and only if there is no such that .

The Nakamura number of a simple game is the minimal number of winning coalitions with empty intersection. Nakamura's theorem states that the core is nonempty for all profiles ofacyclic (alternatively, transitive) preferences if and only if is finite and the cardinal number (the number of elements) of is less than the Nakamura number of . A variant by Kumabe and Mihara states that the core is nonempty for all profiles of preferences that have a maximal element if and only if the cardinal number of is less than the Nakamura number of . (See Nakamura number for details.)

The strong epsilon-core[edit]

Because the core may be empty, a generalization was introduced in (Shapley & Shubik 1966). The strong -core for some number is the set of payoff vectors

In economic terms, the strong -core is the set of pre-imputations where no coalition can improve its payoff by leaving the grand coalition, if it must pay a penalty of for leaving. may be negative, in which case it represents a bonus for leaving the grand coalition. Clearly, regardless of whether the core is empty, the strong -core will be non-empty for a large enough value of and empty for a small enough (possibly negative) value of . Following this line of reasoning, the least-core, introduced in (Maschler, Peleg & Shapley 1979), is the intersection of all non-empty strong -cores. It can also be viewed as the strong -core for the smallest value of that makes the set non-empty (Bilbao 2000).

The Shapley value[edit]

The Shapley value is the unique payoff vector that is efficient, symmetric, and satisfies monotonicity.[14] It was introduced by Lloyd Shapley (Shapley 1953) who showed that it is the unique payoff vector that is efficient, symmetric, additive, and assigns zero payoffs to dummy players. The Shapley value of a superadditive game is individually rational, but this is not true in general. (Driessen 1988)

The kernel[edit]

Let be a game, and let be an efficient payoff vector. The maximum surplus of player i over player j with respect to xis

the maximal amount player i can gain without the cooperation of player j by withdrawing from the grand coalition N under payoff vector x, assuming that the other players in i's withdrawing coalition are satisfied with their payoffs under x. The maximum surplus is a way to measure one player's bargaining power over another. The kernelof is the set of imputations x that satisfy

for every pair of players i and j. Intuitively, player i has more bargaining power than player j with respect to imputation xif, but player j is immune to player i's threats if , because he can obtain this payoff on his own. The kernel contains all imputations where no player has this bargaining power over another. This solution concept was first introduced in (Davis & Maschler 1965).

Harsanyi dividend[edit]

The Harsanyi dividend (named after John Harsanyi, who used it to generalize the Shapley value in 1963[15]) identifies the surplus that is created by a coalition of players in a cooperative game. To specify this surplus, the worth of this coalition is corrected by the surplus that is already created by subcoalitions. To this end, the dividend of coalition in game is recursively determined by

An explicit formula for the dividend is given by . The function is also known as the Möbius inverseof.[16] Indeed, we can recover from by help of the formula .

Harsanyi dividends are useful for analyzing both games and solution concepts, e.g. the Shapley value is obtained by distributing the dividend of each coalition among its members, i.e., the Shapley value of player in game is given by summing up a player's share of the dividends of all coalitions that she belongs to, .

The nucleolus[edit]

Let be a game, and let be a payoff vector. The excessof for a coalition is the quantity ; that is, the gain that players in coalition can obtain if they withdraw from the grand coalition under payoff and instead take the payoff . The nucleolusof is the imputation for which the vector of excesses of all coalitions (a vector in ) is smallest in the leximin order. The nucleolus was introduced in (Schmeidler 1969).

(Maschler, Peleg & Shapley 1979) gave a more intuitive description: Starting with the least-core, record the coalitions for which the right-hand side of the inequality in the definition of cannot be further reduced without making the set empty. Continue decreasing the right-hand side for the remaining coalitions, until it cannot be reduced without making the set empty. Record the new set of coalitions for which the inequalities hold at equality; continue decreasing the right-hand side of remaining coalitions and repeat this process as many times as necessary until all coalitions have been recorded. The resulting payoff vector is the nucleolus.

Properties[edit]

Convex cooperative games[edit]

Introduced by Shapley in (Shapley 1971), convex cooperative games capture the intuitive property some games have of "snowballing". Specifically, a game is convex if its characteristic function issupermodular:

It can be shown (see, e.g., Section V.1 of (Driessen 1988)) that the supermodularityof is equivalent to

that is, "the incentives for joining a coalition increase as the coalition grows" (Shapley 1971), leading to the aforementioned snowball effect. For cost games, the inequalities are reversed, so that we say the cost game is convex if the characteristic function is submodular.

Properties[edit]

Convex cooperative games have many nice properties:

Similarities and differences with combinatorial optimization[edit]

Submodular and supermodular set functions are also studied in combinatorial optimization. Many of the results in (Shapley 1971) have analogues in (Edmonds 1970), where submodular functions were first presented as generalizations of matroids. In this context, the core of a convex cost game is called the base polyhedron, because its elements generalize base properties of matroids.

However, the optimization community generally considers submodular functions to be the discrete analogues of convex functions (Lovász 1983), because the minimization of both types of functions is computationally tractable. Unfortunately, this conflicts directly with Shapley's original definition of supermodular functions as "convex".

The relationship between cooperative game theory and firm[edit]

Corporate strategic decisions can develop and create value through cooperative game theory.[17] This means that cooperative game theory can become the strategic theory of the firm, and different CGT solutions can simulate different institutions.

See also[edit]

References[edit]

  1. ^ Shor, Mike. "Non-Cooperative Game - Game Theory .net". www.gametheory.net. Retrieved 2016-09-15.
  • ^ Chandrasekaran, R. "Cooperative Game Theory" (PDF).
  • ^ Brandenburger, Adam. "Cooperative Game Theory: Characteristic Functions, Allocations, Marginal Contribution" (PDF). Archived from the original (PDF) on 2016-05-27.
  • ^ denotes the power setof.
  • ^ Javier Muros, Francisco (2019). Cooperative Game Theory Tools in Coalitional Control Networks (1 ed.). Springer Cham. pp. 9–11. ISBN 978-3-030-10488-7.
  • ^ Georgios Chalkiadakis; Edith Elkind; Michael J. Wooldridge (25 October 2011). Computational Aspects of Cooperative Game Theory. Morgan & Claypool Publishers. ISBN 978-1-60845-652-9.
  • ^ Peleg, B. (2002). "Chapter 8 Game-theoretic analysis of voting in committees". Handbook of Social Choice and Welfare Volume 1. Vol. 1. pp. 395–423. doi:10.1016/S1574-0110(02)80012-1. ISBN 9780444829146.
  • ^ See a section for Rice's theorem for the definition of a computable simple game. In particular, all finite games are computable.
  • ^ Kumabe, M.; Mihara, H. R. (2011). "Computability of simple games: A complete investigation of the sixty-four possibilities" (PDF). Journal of Mathematical Economics. 47 (2): 150–158. arXiv:1102.4037. Bibcode:2011arXiv1102.4037K. doi:10.1016/j.jmateco.2010.12.003. S2CID 775278.
  • ^ Modified from Table 1 in Kumabe and Mihara (2011). The sixteen types are defined by the four conventional axioms (monotonicity, properness, strongness, and non-weakness). For example, type 1110 indicates monotonic (1), proper (1), strong (1), weak (0, because not nonweak) games. Among type 1110 games, there exist no finite non-computable ones, there exist finite computable ones, there exist no infinite non-computable ones, and there exist no infinite computable ones. Observe that except for type 1110, the last three columns are identical.
  • ^ Kumabe, M.; Mihara, H. R. (2008). "The Nakamura numbers for computable simple games". Social Choice and Welfare. 31 (4): 621. arXiv:1107.0439. doi:10.1007/s00355-008-0300-5. S2CID 8106333.
  • ^ Aumann, Robert J. "The core of a cooperative game without side payments." Transactions of the American Mathematical Society (1961): 539-552.
  • ^ Peters, Hans (2008). Game theory: a multi-leveled approach. Springer. pp. 123. doi:10.1007/978-3-540-69291-1_17. ISBN 978-3-540-69290-4.
  • ^ Young, H. P. (1985-06-01). "Monotonic solutions of cooperative games". International Journal of Game Theory. 14 (2): 65–72. doi:10.1007/BF01769885. ISSN 0020-7276. S2CID 122758426.
  • ^ Harsanyi, John C. (1982). "A Simplified Bargaining Model for the n-Person Cooperative Game". Papers in Game Theory. Theory and Decision Library. Springer, Dordrecht. pp. 44–70. doi:10.1007/978-94-017-2527-9_3. ISBN 9789048183692.
  • ^ Set Functions, Games and Capacities in Decision Making | Michel Grabisch | Springer. Theory and Decision Library C. Springer. 2016. ISBN 9783319306889.
  • ^ Ross, David Gaddis (2018-08-01). "Using cooperative game theory to contribute to strategy research". Strategic Management Journal. 39 (11): 2859–2876. doi:10.1002/smj.2936. S2CID 169982369.
  • Further reading[edit]

    External links[edit]


    Retrieved from "https://en.wikipedia.org/w/index.php?title=Cooperative_game_theory&oldid=1223531109"

    Categories: 
    Cooperative games
    Game theory
    Mathematical and quantitative methods (economics)
    Hidden categories: 
    Articles with short description
    Short description matches Wikidata
    Articles lacking in-text citations from April 2024
    All articles lacking in-text citations
    All articles with unsourced statements
    Articles with unsourced statements from April 2024
    Pages containing links to subscription-only content
    Webarchive template wayback links
     



    This page was last edited on 12 May 2024, at 19:28 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki