Play by rules and beyond the rules.

Game theory is the study of mathematical models of conflict and cooperation among rational decision-makers. {Myerson1991} It is an advanced form of decision theory: while decision theory formulate optimization problems as if only one agent makes the decision, game theory analyzes decentralized/independent decision makers with interacting interests, who only care about their own objectives. That's how game theory get its other names, such as strategic decision making, interactive decision theory, and the logical side of decision science.

In essence, an optimization problem (or decision making problem) presents a real-valued function and asks to find the optimum set in a typically compact domain; while a game is a vector-valued function \( G: \prod_{i=i}^n S_i \to \mathbb{R}^n \), with domain the Cartesian product of \(n\) sets and range the \(n\)-dimensional Euclidean space, and the problem asks to find the equilibrium set in the domain, where an equilibrium by a certain definition has components optimizing their corresponding functions. Although multi-attribute decision making problems also take vector values, it imposes its own definitions of optimality. As the most commonly used equilibrium concept in game theory, pure strategy Nash equilibrium can be stated as: \( G_i( s^* ) = \max_{s_i} G_i(s_i, s_{-i}^* ), \forall i \in N \). The distinction in solution concepts also shows up in solution algorithms: gradient methods in optimization, and best-response dynamics in games (which sequentially move along player strategy sets).

The mathematical theory of games can be understood in two distinct ways: as rules and as models. As rules of artificial setups, games exactly specify the strategy sets and payoffs of each player, as is the case for chess and auction. As models of real-world cooperation and conflicts, games need meticulous care to truthfully represent the possible actions and retribution of participating parties, as exemplified in Tenancy Game.

Charles Waldegrave in 1713 provided a minimax mixed strategy solution to a two-person version of the card game le Her. Antoine Augustin Cournot in 1838 considered a duopoly competition problem and provided a stable equilibrium solution (sink of a discrete dynamical system). John von Neumann in 1928 proved the existence of mixed strategy equilibria (mutually consistent solutions) in two-person zero-sum games with perfect information, known as minimax theorem, using Brouwer fixed-point theorem. von Neumann and Morgenstern in 1944 published a collection of work establishing the field of game theory, providing methods for finding equilibria, extending minimax theorem for games with imperfect information and cooperative games of more than two players; its second edition in 1947 appended an axiomatic theory of utility, as an independent discipline. {von-Neumann1947} declared that economic theory needed to use functional analytic methods, especially convex sets and topological fixed-point theorem, rather than the traditional differential calculus, because the maximum-operator did not preserve differentiable functions. Merrill Flood and Melvin Dresher in 1950 at RAND Corporation developed a game theoretical model of cooperation and conflict known as the Prisoner's Dilemma, for possible applications to global nuclear strategy. Also in 1950, John Nash developed a criterion for mutual consistency of players' strategies known as Nash equilibrium, and proved its existence in n-player non-cooperative games using Kakutani fixed-point theorem.

Basic Concepts

Game \(G = \{N, S, \succeq\} \) is a collection of players, strategy space, and preferences:

  1. Player \(i \in N = \{1, \dots, n\}\): a participant in a game who has a nontrivial action space and choose strategies based on payoffs;
  2. Strategy \(s_i \in S_i\): a complete set of actions a player takes at all possible positions of a game;
    • Action (or allocation, move) \(a_i \in A_i\): one of the possible alternatives a player can choose;
    • Strategy profile (or outcome) \(s \in S = \prod_i S_i\) is a collection of all players' strategies;
  3. Preference \(\succeq_i\) or payoff \(u_i\): the reward for a player given an outcome at the end of a game;

In static games of complete information, strategy and action are equivalent. A strategy is pure if only one action is chosen at each position of the game. A strategy is mixed if the player chooses probabilistically from the action space of at least one position. When a player's strategy space \( S_i \) is finite, the mixed strategy space of the player is a simplex, which is compact and convex, with vertices being all the pure strategies: \( \Delta_i = \Delta^{|S_i| - 1} \). The notation for mixed strategy space \( \Delta = \prod_i \Delta_i \) of a game is different from the pure strategy space \(S\).

As vector-valued functions, a no-conflict game is \( G_i(s) = P_i(s_i) \), which is equivalent to independent optimization; a dummy game is \( G_i(s) = Q_i(s_{-i}) \); a strategically separable game is \( G_i(s) = P_i(s_i) + Q_i(s_{-i}) \), the linear combination of a no-conflict and a dummy game; an identical interest game (or perfect coordination) is \( G_i(s) = P(s) \); a coordination-dummy separable game is \( G_i(s) = P(s) + Q_i(s_{-i}) \) {Slade1994}. Condition \( \forall i \in N \) is implicit in all previous propositions.

Number of players

The simplest games are defined as two-player games. An n-player game is a game consistently defined for two or more players. A game is atomic if it has a finite number of players; it is non-atomic (large, oceanic) if it has a continuum of players.

Two-player games are mathematically simpler, and have been extensively studied. Many results for two-player games are readily extensible to n-player situations, just like extending pairs of random variables to random vectors.

Many intuitive ideas can be examined in two-player case and then generalized, like fixed-point theorem and its generalization under super-modularity. But a few important results only hold for games with at least three players, just like chaos in 3D dynamic systems, although harder to find.

Preference and utility

Every player in a game has a preference ordering over the strategy space, which is modeled mathematically as a totally ordered set or a real-valued function. In practice, two types of utility functions are used: ordinal utility and cardinal utility. Utility is a mapping of preferences to real numbers. Comparing probability space \( (U,\Sigma,P) \) and utility \( (N,S,u) \), both are extrinsically assigned rather than inherent attributes.

An ordinal utility function represents a preference ordering over the pure strategy space, without imposing a metric: \( \forall a, b \in S, a \preceq b \iff u(a) \leq u(b) \). Ordinal utility suffice for computing pure strategy Nash equilibrium. Debreu theorems claim the existence of continuous ordinal utility function for any preference relation that is complete, transitive and continuous {Debreu1954}; and the existence of additive ordinal utility function for any preference relation with preferentially independent subsets {Debreu1960}. For a given preference ordering, the ordinal utility function is unique up to strictly increasing transformation.

A cardinal utility function represents a preference ordering over the generated cone of the strategy space, associating a metric: \( \forall a, b \in \mathbb{R}_{\ge 0}^{|S|}, a \cdot S \preceq b \cdot S \iff u(a) \leq u(b) \). Cardinal utility must be employed for mixed strategy calculations. von Neumann–Morgenstern (VNM) utility theorem {von-Neumann1947} constructively claims the existence of cardinal utility function for a player satisfying the four axioms of VNM-rationality (completeness, transitivity, continuity, and independence). For a given preference ordering, the cardinal utility function is unique up to a strictly increasing linear transformation. In other words, the set of cardinal utility functions representing a decision maker's preference form an open half-plane, with positive scale and arbitrary shift.

A set of games with the same strategy space are fully equivalent in mixed strategy (strategically equivalent?) if their (cardinal) payoff functions for each player represent the same preference. {Myerson1991} In other words, given a strategy space \( S \) and player preferences \( \succeq \), the set of fully equivalent games (in mixed strategy) \( \Gamma_{S, \succeq} \) form a product space of \(n\) open half-planes (or the product space of a positive cone and an Euclidean space, both \(n\)-dimensional): \( \forall G \in \Gamma_{S, \succeq} \subset \{ G: S \to \mathbb{R}^n \} \), \( \forall a \in \mathbb{R}^n_+, b \in \mathbb{R}^n, \text{diag}(a)~G + b \in \Gamma \).

While homogeneous preferences are naturally comparable, heterogeneous (multi-dimensional) preferences may only be partially ordered, just like positive semi-definite matrices. And sometimes we may aggregate heterogeneous axes into one axis and hence completely ordering the strategy space. But any such aggregate function over heterogeneous measures is subjective/arbitrary and shall not expect universal acceptance. In economics, the convention has been using monetary value as the universal measure of use values across goods and individuals.

Perturbation by small numbers don't change real-world phenomena significantly; it only does in the realm of theoretical models. Real data have noises and human mind do round-ups; sharp categorical transitions are caused by model simplification. Moreover, some solution concepts such as mixed strategy Nash equilibrium are susceptible to extreme values of utility. In such cases, the choice of solution concepts should be reconsidered.

As a side note, the meanings of "ordinal" and "cardinal" in utility theory are very different from those in set theory. An ordinal is a set strictly well-ordered with respect to set membership \( \in \), and every element of it is also a subset of it. {von-Neumann1923} Finite ordinals can be denoted by a natural number; the smallest infinite ordinal \( \omega \) is the order type of natural numbers. The cardinal number of a well-ordered set is the smallest ordinal number equinumerous (等势) to the set. Finite cardinals can be denoted by a natural number; the smallest infinite cardinal \( \aleph_0 \) is the cardinality of natural numbers.

Information set

(First order) mutual knowledge among a set of players is a piece of information that all players know. A piece of information is \(m\)-th order mutual knowledge among a set of players if the \((m-1)\)-th order mutual knowledge about this information is also mutual knowledge. Common knowledge (共有知识) among a set of players is a mutual knowledge of infinite order of recursion. {Binmore and Brandenburger, 1988}

Table: Information space of two players with common knowledge (\(I\) denotes the common knowledge; Each lowercase letter represents the player of the acquired information; each sequence represents a piece of compound information.)

Order of recursion 0 1 2 ...
Player A aI abI abaI ...
Player B bI baI babI ...

That all players are rational is typically assumed to be a common knowledge, known as the common knowledge of rationality.

A game is of complete information if its definition is common knowledge to all players. Otherwise, the game is of incomplete information, which is first studied in {Harsanyi1967} where the game's payoffs depend upon states not all players know.

Cooperative vs. Non-cooperative

A game is cooperative if players may negotiation contracts enforceable through an outside party before the game. The default model of a game is non-cooperative, where entities not explicitly included cannot enforce contracts for the players. The distinction between cooperative and non-cooperative games is unrelated to the mathematical description of a game, but rather the possibility of pre-play communication, coalition, and side-payment (post-game exchange). {Nash1951}

A cooperative game is given by stating a value for every subset of players, called coalition: \(v: 2^N \to \mathbb{R} \). The value of a coalition is then shared among its members. A payoff vector is efficient if it exactly splits the total value \(v(N)\). A payoff vector is individually rational if no player receives less than what he could get on his own, i.e. \(v(\{i\})\). An efficient and individually rational payoff vector is called an imputation.

The main assumption in cooperative game theory is that the grand coalition \(N\) will form. If players split off and form smaller coalitions, we can apply solution concepts to the subgames defined by whatever coalitions actually form. The challenge of solving cooperative games is then to find stable imputations.

A coalition is α-effective for an outcome if its members have strategies to force that outcome of the game, regardless of its complement's strategy. A coalition is β-effective for an outcome if for any strategies of its complement, its members can respond with strategies that ensure that outcome.

Solution Concepts

In game theory, a solution of a game is a strategy profile that may be adopted by players. A solution concept specifies for any game a subset of its strategy space that are considered solutions: \( \forall G \in \Gamma, F(G) \subseteq S_G \).

The value of a game is the payoff vector given some solution concept, if determinable.

Minimax

Brouwer fixed-point theorem {Brouwer1911}: Any continuous function from a compact convex set into itself has a fixed point.

Minimax solution of two-player zero-sum games (backward induction) {von-Neumann1928}.

A two-person zero-sum game can be considered solved on several levels: {Allis1994}

  1. Ultra-weak: for the initial position, the game-theoretic value (win/lose, draw) has been determined;
  2. Weak: for the initial position, a strategy has been determined to obtain at least the game-theoretic value for both players, under reasonable resources;
  3. Strong: for all legal positions, such a strategy has been determined;

Nash equilibrium

A Nash equilibrium (NE) is a strategy profile that no player has incentive to deviate unilaterally {Nash1950}. A pure strategy Nash equilibrium (PSNE) is a Nash equilibrium in the pure strategy space, while a mixed strategy Nash equilibrium (MSNE) is a Nash equilibrium in the mixed strategy simplex that is not a vertex. Symbolically, a pure strategy Nash equilibrium of a game is a strategy profile \( s^* \in S \) such that \( u_i( s_i^* , s_{-i}^* ) \ge u_i( s_i, s_{-i}^* ), \forall i \in N, s_i \in S_i \); Mixed strategy Nash equilibrium can be similarly formalized, replacing \( S \) with \( \Delta \). A (pure strategy) Nash equilibrium is strict if the inequalities are strict.

Nash equilibrium has nothing to do with saddle points: saddle point is a local property, but Nash equilibrium is a component-wise argument-wise global property.

Kakutani fixed-point theorem {Kakutani1941}: Any correspondence on a compact convex set with convex images and a closed graph has a fixed point.

A correspondence between two sets is any subset of their Cartesian product: \( R \subseteq X \times Y \). The image of an element of the former set with respect to a correspondence is the subset of the latter set whose elements are paired with the element in the correspondent: \( \text{Im}_R(x) = R \cap (\{x\} \times Y) = \{ y \in Y | (x, y) \in R \} \) The best-response correspondence of a player is a correspondence from the opponents' strategy space into the player's strategy space, such that the image of any opponent strategy with respect to the correspondence is the set of the player's optimal strategies in response: \(B_i \subseteq S_{-i} \times S_i = S \), \( \text{Im}_{B_i} (s_{-i}) = \arg \max_{s_i} u_i (s_i, s_{-i}) \). The best-response correspondence of a game is a correspondence from its strategy space into itself, such that the image of any strategy profile with respect to the correspondence is the Cartesian product of each player's best-response set given the opponents' strategy: \(B \subseteq S \times S \), \( \text{Im}_B (s) = \prod_i \text{Im}_{B_i}(s_{-i})\).

Games with the same strategy space are best-response equivalent if they have the same best-response correspondence.

Finite games have mixed strategy Nash equilibrium {Nash1950}: the mixed strategy space of a finite game is the Cartesian product of simplices (probability distribution over finite set), thus compact and convex; mixed strategy payoffs are polylinear forms, linear in each argument, so every mixed strategy profile has convex best-responses and thus convex images with respect to best-response correspondence; since mixed strategy payoffs are polylinear thus continuous, best-response correspondence has a closed graph; by Kakutani fixed-point theorem, the previous conditions implies existence of fixed point, thus mixed strategy Nash equilibrium.

A continuous game is a game with compact player strategy spaces and continuous payoff functions. Continuous games have mixed strategy Nash equilibrium, by Kakutani-Glicksberg-Fan theorem {Glicksberg1952, Fan1952}. A convex game is a game where each player has a convex strategy space and a concave payoff function \( u_i(s_i; s_{-i}) \) for all opponent strategies. Convex games have pure strategy Nash equilibrium, if it has a compact strategy space and continuous payoff functions {Nikaido1955}; the reasoning directly follows that of {Nash1950}. Strictly convex/concave games are convex games with diagonally strictly concave payoffs: \(\sum_i (x'_i - x_i) \cdot (\nabla_i u_i(\mathbf{x}) - \nabla_i u_i(\mathbf{x'})) > 0\). Strictly convex/concave games have unique pure strategy Nash equilibrium, which is also globally asymptotically stable in a system of differential equations {Rosen1965}.

Refinements

Equilibrium refinement (narrowing) shrinks the space of equilibria as much as possible, ideally to a unique equilibrium.

A focal point is a solution with features prominent to all players, which all players also believe to be a 3rd order mutual knowledge {Schelling1960}. Thus focal points are most likely to be chosen by the players in the absence of communication.

subgame perfect equilibrium {Selten1965} trembling hand perfect equilibrium {Selten1975} proper equilibrium {Myerson1978}

Bayesian equilibrium {Harsanyi1967} perfect Bayesian equilibrium

A Nash equilibrium is strong if no coalition can cooperatively deviate to benefit all its members, given opponents' strategies {Aumann1959}. Strong Nash equilibrium are Pareto-efficient. A Nash equilibrium is coalition-proof if no backward coalition can cooperatively deviate to benefit all its members, fixing opponent strategies one by one {Bernheim1987}.

Stable equilibrium is a refinement of Nash equilibrium so that every game has a unique set solution that is connected, not dominated, containing backwards induction equilibrium, invariant in strategically equivalent games, and containing a common solution of games eliminating any dominated strategy {Mertens1986}.

Generalizations

A correlated equilibrium is a joint probability distribution over the strategy space where no player has incentive to deviate unilaterally {Aumann1974}. Correlated equilibrium is more general than mixed strategy Nash equilibrium, because the latter can only be independent distributions over each player's strategy space.

The rationalizable set of actions in a normal-form game is the set of actions remained after progressive elimination of actions which are never a best-response to the opponents' any rational action {Bernheim1984, Pearce1984}. An outcome is Pareto efficient (or Pareto optimal) if it is not dominated under any player's preference. In comparison, an outcome is Hicks optimal if it has the largest sum of payoffs. For one player, a strategy is strictly dominated by another if the latter is strictly better than the former under any strategy of the opponents: \( u_i(s_i, s_{-i}) > u_i(s_i', s_{-i}), \forall s_{-i} \in S_{-i} \). If the inequality is only strict for some strategy of the opponents, the strategy is called weakly dominated by the other. For two-player games, the set of all rationalizable strategies can be found by progressive/iterated elimination of strictly dominated strategies (ISEDS).

Learning in Games

Nash equilibrium is one thing; game dynamics is another. Instead of studying the dynamics of a game, Nash equilibrium dictate the outcome of the game. The interpretation of Nash equilibrium is tricky in static games.

The set of Nash equilibria shrinks under various refinements but real-world behaviors may never converge to any of these equilibria. When game theory correctly predicts reality without including all factors in the model, it is only because the factors excluded do not have crucial influence on the outcome, which may become crucial beyond this range.

Solution concepts may be formally (and elusively) defined for different models of games, but the essential question is what kind of knowledge you would like to know about the conflict and cooperation for self-interests. The positive view asks mass action: What would be the average behavior if players learn the structure of a game while playing it repeatedly? The normative view asks rational prediction: What should we expect to observe in a game played by rational players who are fully knowledgeable about the structure of the game and each others' rationality? The positive view suggests massive simulation of a game and see the successful strategies selected by the rules. The normative view have made game theory dauntingly complicated, while still struggles to predict real-world problems. To be clear, it is totally acceptable that many resource allocation situations in real life do not have a unique strategy profile with a special property of any sense. In such cases, players can only resort to negotiation/bargaining and power play.

Two classes of evolutionary dynamics {Sandholm2008}:

  • imitative dynamics: EGT replicator dynamics;
  • belief-based dynamics: best reply dynamics; Logit dynamics {Fudenberg1998}

Rule of thumb behavior

{Brown1951} proposed fictitious play ("belief-based" learning rule) to justify mixed strategy Nash equilibrium, where players learn their optimal strategies by simulating the game with arbitrary initial mixed strategy profile and best-response dynamics. Player strategies are assumed to be stationary.

  1. Paths generated by fictitious play terminate at strict Nash equilibrium.
  2. Paths generated by fictitious play converge at pure strategy Nash equilibrium.
  3. Paths generated by fictitious play converge in time average at mixed strategy Nash equilibrium.

Several classes of games converge in time average under fictitious play: two-player zero-sum games, two-player two-strategy games, games solvable by iterated strict dominance, and potential games.

continuous-time fictitious play (CTFP)

perturbed CTFP

Evolutionary game theory

A population game is a symmetric game with finite player strategies and a large number of players. Since player payoff in symmetric games only depends on self strategy and the set of opponent strategies, which is equivalent to the distribution of strategies give a finite player strategy set, a population game can be fully specified by the payoff to each strategy given the strategy distribution. A strategy distribution is monomorphic if it is pure; otherwise it is polymorphic. The strategy distribution space approximates the simplex over the player strategy space as the number of players increases, so a population game can be formally expressed as a continuous vector-valued function on a simplex (of the lower dimension): \( F: \Delta^{|S|-1} \to \mathbb{R}^{|S|} \).

Evolutionary game theory was originated by {Maynard-Smith1973, 1982}, who analyzed the evolution of population strategy in a repeated matching symmetric matrix game, where every player is hard-wired with a pure strategy and reproduces proportionally to its payoff. These games are linear population games: \( F(x) = A x \), where \(A\) is the payoff matrix. Congestion games is another example of population games.

Since each player has a negligible influence on the strategy distribution, a Nash equilibrium of population games is a strategy distribution under which all chosen strategies have the same payoff no worse than the other strategies: \( F_i( x^* ) \ge F_j( x^* ) \), \( \forall i,j \in S, x_i^* > 0 \). A strategy distribution is a Nash equilibrium if and only if the payoff vector at that point is in the polar cone at that point: \( x \in \text{NE}(F) \Leftrightarrow F(x) \in \text{cone}_x (X)^* \), where \( X = \Delta^{|S|-1} \) and \( \text{cone}_x (X)^* = \{ z \mid (y - x) \cdot z \le 0, \forall y \in X \} \). Population games have Nash equilibrium.

A strategy distribution of a population game is an evolutionarily stable state (ESS) if the following three equivalent conditions hold:

  1. It is a Nash equilibrium that is a better-response to all nearby best-responses {Maynard-Smith1973}: \( \forall y \in X, x \cdot F(x) \ge y \cdot F(x) \), and \( \exists \varepsilon > 0: \forall y \in \check{U}_\varepsilon (x) \), \( y \cdot F(x) = x \cdot F(x) \Rightarrow y \cdot F(y) < x \cdot F(y) \);
  2. It has a uniform invasion barrier {Taylor1978}: \( \exists c > 0, \forall \varepsilon \in (0,c), \forall y \ne x \), \( x' = (x + \varepsilon y) / (1 + \varepsilon): y \cdot F(x') < x \cdot F(x') \);
  3. It is an (infinitesimal) local invader {Hofbauer1979}: \( \exists \varepsilon > 0: \forall y \in \check{U}_\varepsilon (x), y \cdot F(y) < x \cdot F(y) \);

The equivalence of these three definitions means that for population games: An ESS is a Nash equilibrium; A strict Nash equilibrium is an ESS. ESS may not exist for some population games.

A state is a regular ESS {Taylor1978} if it is a quasi-strict Nash equilibrium: \( F_i(x) > F_j(x), \forall i,j \in S, x_i > 0, x_j = 0 \); additionally, \( z^\text{T} \nabla F(x) z < 0 \), \( \forall z \ne 0, \sum_{x_i > 0} z_i = 0, \sum_{x_j = 0} z_j^2 = 0 \).

The replicator dynamics can be formalized as such {Taylor1978}: the strategy distribution evolves as \( x_s^{t+1} = {x_s^t \mathbf{E}u (s, \sigma^t)}/{\mathbf{E}u(\sigma^t, \sigma^t)} \).

In replicator dynamics, stationary state means all individuals of the population have the same reproductive fitness; asymptotically stable state means no perturbation could establish itself in the population. For symmetric games, a Nash equilibrium is a stationary state; an asymptotically stable state is a Nash equilibrium; ESS are asymptotically stable.

stochastically stable equilibrium {Foster1990}

Adjustment process

Without stability, game-theoretic equilibrium is meaningless. Equilibrium stability is defined by an adjustment process, which takes expectation of strategy profile as input and can take place in discrete or continuous time:

  1. Time scale
    • discrete time: \( \mathbf{s}_{t+1} = \mathbf{g}(\mathbf{s}_t) \)
    • continuous time: \( \dot{\mathbf{s}} = \mathbf{g}(\mathbf{s}) \)
  2. Concurrence
    • simultaneous: \( \mathbf{g}(\mathbf{s}) = (g_i(\mathbf{s}_{-i}))_i \)
    • sequential: \( \mathbf{g}(\mathbf{s}) = \mathbf{g}_n \circ \cdots \circ \mathbf{g}_1 (\mathbf{s}) \), with \( \mathbf{g}_i (\mathbf{s}) = (s_1, \dots, g_i (\mathbf{s}_{-i}), \dots, s_n) \)
  3. Expectation
    • Cournot: \( \mathbf{f}_{i, t+1} = \mathbf{s}_{-i, t} \)
    • adaptive: \( \mathbf{f}_{i, t+1} = \mathbf{f}_{i, t} + m_i (\mathbf{s}_{-i, t} - \mathbf{f}_{i, t}) \), where \( m_i \) is the speed of adjustments;
    • extrapolative: \( \mathbf{f}_{i, t+1} = \sum_{k = 0}^l a_k \mathbf{s}_{-i, t-k} \), where \( \sum_{k = 0}^l a_k = 1 \);

Table: Comparison of Common Adjustment Processes

Timing Best-response Better-response
Discrete-time sequential sequential best-response sequential improvement
Discrete-time simultaneous discrete-time best-response -
Continuous-time continuous-time best-response gradient

In discrete-time sequential adjustment processes, a path is a sequence of strategy profiles \( \{ s^k \}, k \in \mathbb{N} \) with only one player changes strategy at each step; a cycle is a path with identical start and end.

Under sequential improvement dynamics, an improvement path is a path where at each step the deviator has improved payoff: \( \forall k \in \mathbb{N}, \exists i \in N: s^{k+1}_{-i} = s^k_{-i}, u_i(s^{k+1}) > u_i(s^k) \). A non-deteriorating path is a path where at each step the deviator has non-deteriorating payoff. A weak improvement cycle is a non-deteriorating cycle with some improvement steps. Improvement path terminates at a Nash equilibrium.

Under sequential best-response dynamics, a path is best-response compatible if at each step the deviator moves to a best-response. A path is strict best-response compatible if at each step the deviator moves to a best-response if not already. A (strict) best-response cycle is a (strict) best-response compatible cycle with some improvement steps. Strict best-response compatible path terminates at a Nash equilibrium.

Simultaneous best-response adjustment processes includes discrete-time best-response dynamics: \( \mathbf{s}_{t+1} = B(\mathbf{s}_t) \); and continuous-time best-response dynamics: \( \dot{\mathbf{s}} = M (B(\mathbf{s}) - \mathbf{s}) \).

Gradient process {Arrow1960}: \( \dot{\mathbf{s}} = S \nabla \mathbf{u}(\mathbf{s}) \), where \(S\) is a positive diagonal matrix of adjustment speeds, and pseudo-gradient \( \nabla \mathbf{u}(\mathbf{s}) = ( \nabla_i \mathbf{u}_i(\mathbf{s}) )_{i=1}^n \).

Stability

An equilibrium is locally (asymptotically) stable under a specific adjustment process if any trajectory of the dynamical system starting within a neighborhood of the equilibrium converges to the equilibrium. Similarly, an equilibrium is globally (asymptotically) stable if all trajectories of the dynamical system on the strategy space converges to the equilibrium. Note that global equilibrium implies uniqueness of equilibrium. Currently, most results on the stability of game theoretic equilibrium are local; while global stability is studied for evolutionarily stable strategies, games with linear best-responses, or specific cases.

A game can have different stability under different adjustment processes. For example, Cournot game with linear marginal rent has a unique pure strategy Nash equilibrium. As an exact potential game {Monderer1996b}, it is globally asymptotically stable under sequential better-response dynamics (with Cournot expectation). But under discrete-time simultaneous best-response dynamics (with Cournot expectation), it is asymptotically stable (thus a sink) only when it has two players; with three players, it asymptotically alternates between two points centered at the equilibrium along the eigen-direction of total production; with over three players, it is unstable in the aforementioned eigen-direction (thus a saddle point) {Theocharis1960}. With strategic substitutes, contraction stability is comparably restrictive due to simultaneous adjustments {Hefti2016}. Note that with adaptive expectation, the game can be asymptotically stable under discrete-time simultaneous best-response dynamics as long as the speed of adjustments is small enough, which also means the convergence will be slow {Okuguchi1999}. The stability properties in this example is global, since it is a linear dynamical system.

There appears to be no universal ordering among stability relations in general regular games, except for some special cases:

  • Under discrete-time best-response dynamics, local contraction stability (Jacobian of best-response has spectral radius less than one) implies local asymptotic stability.
  • sequential best-response
  • Contraction stability implies local stability under continuous-time best-response dynamics with a homogeneous diagonal cofactor {Hefti2016}: \( M(\mathbf{s}) = (m(s_i))_i \).
  • For one-dimensional games with weak strategic complements at an equilibrium, continuous-time best-response and gradient stabilities are equivalent {Hefti2016}.

Classes of Games

A game is finite if its strategy space \(S\) is finite, which means it has a finite set of (non-trivial) players each with a finite strategy space. (Finite games are not very applicable.)

A game is zero-sum (aka strictly competitive) if the total payoff of all players under any outcome is a constant. Zero-sum games only have \(n-1\) independent payoff functions.

A game is symmetric if the payoff function is permutation invariant: \( \forall \mathbf{s} \in S, P \in \mathcal{S}_n: \mathbf{u}(P\mathbf{s}) = P\mathbf{u}(\mathbf{s}) \). This implies that player strategy spaces are identical, only one component payoff function is independent, and that function is also independent from the order of opponent strategies. For two-person matrix games, symmetry means the two payoff matrices are transpose of each other. While a strategy profile is symmetric if all players use the same strategy, symmetric equilibrium is of special interest to symmetric games. Finite symmetric games have symmetric mixed strategy Nash equilibrium {Nash1951}.

A game has identical interests (aka perfect coordination) if it is best-response equivalent in mixed strategies to a game with identical payoff functions: \( \{ N, S, u \mathbf{1} \} \). {Monderer1996a} Identical-interest games apparently have only one independent payoff function.

A game is regular if its pseudo-gradient on the boundary points into the strategy space, and zeros of the pseudo-gradient have non-singular Jacobian. Regular games have a finite number of Nash equilibria, which are the zeros of the pseudo-gradient (critical points); uniqueness of equilibrium is equivalent to that on the set of critical points \( \det(-\nabla (\nabla \mathbf{u}(\mathbf{s}))) > 0 \); uniqueness of equilibrium is implied if every critical point has some type of stability (contraction, continuous-time best response, gradient) {Hefti2016}.

Simultaneous game

A game is static (or simultaneous) if all players choose their strategies without knowing each other strategies. no information about the decisions of others

A game in normal form (or strategic form) \(u: S \to \mathbb{R}^n\) is a matrix/array representation of a finite static game.

Sequential game

A game is sequential if players take actions in a predefined order, where at least some players can observe other players' previous moves.

A sequential game is of perfect information if players can observe others' moves in turn and can recall the entire history of play. In other words, every information set contains exactly one node. Otherwise, the game is of imperfect information.

Extensive form (or game tree) is a tree representation of a sequential game, where each internal node including the root is a possible position of the game where a player chooses an edge, and the game ends when a leaf node is reached which corresponds to an outcome. {Kuhn1953}

A subgame of a sequential game is the game corresponding to a subtree. Subgame perfect Nash equilibrium (SPNE, perfect equilibrium) is a strategy profile that yields a NE in every subgame of the original game. SPNE can be found by backward induction, iterative reasoning from eventual outcomes backwards to the present choice problem.

Repeated game

A game is repeated (or dynamic) if a stage game is played multiple times. Depending on the times repeated, the stage game could be finitely repeated or infinitely repeated.

"Folk theorems" (general feasibility theorem) refers to a class of theorems about possible Nash equilibrium payoff profiles in repeated games.

Mechanism design

The power of game theory lies in mechanism design (aka reverse game theory), which changes rules of the game to make players play in an intended way. Leonid Hurwicz, Eric Maskin and Roger Myerson are recognized 2007 Nobel Memorial Prize "for having laid the foundations of mechanism design theory". {Hurwicz1960, 1972}

Institution, in the sense of institutional arrangements ("rules of the game") rather than organizations ("artificial players"), defines legal actions of each player and the consequences of possible outcomes, which is mathematically equivalent to a function from the (legal) strategy space to outcomes, referred to as game-form in game theory or mechanism in economics. A game-form is not a game, in that an outcome describes what happened which need not have simple mathematical formulation, while payoffs are how players are rewarded given the outcome. Mechanism design separates game-form from payoff functions and private information such as endowments and technologies: legislators have full control of the former but no control of the latter.

When you write the rules of the game you take into account that the players will try to cheat. {Hurwicz}

The true strategy space is the set of all feasible actions, which is a superset of the (legal) strategy space. Legal strategies are those prescribed by the mechanism governing the system; other feasible actions are illegal. The true game has the true strategy space and true outcome.

Implementation is an essential part of an institution, which means to have the money and the information to run the institutions, and there is legislation authorizing this. Enforcement of an institution is successful if the outcomes of the true game ensure that illegal strategies are (weakly) dominated by legal strategies. More broadly, implementation of an institution is successful if the equilibrium outcomes of the true game correspond to those envisaged by the legislation. Nash equilibrium is neither self-enforcing nor self-implementing.

The greed process has the desired optimality properties for all environments from which so-called external economies or diseconomies are absent; but it lacks the stability properties known to hold for perfect competition at least in certain special cases.

informationally decentralized

The perfectly competitive market is the only efficient mechanism if there are: large numbers of buyers and sellers so that no single agent has significant market power; and (no significant externalities) an agent's consumption, production, and information does not affect others' production or consumption.

Implementation theory: {Maskin1977}

  1. If a social choice rule is implementable, it must be monotonic.
  2. A social choice rule is implementable if it is monotonic, no individual has veto power, and there are more than two players.

Social choice rule is a correspondence between states and strategy space: \( f \subseteq \Theta \times S \). Mechanism \(g\) implements social choice rule \(f\) in Nash equilibrium if the Nash equilibrium outcomes coincide with social optimum outcomes under all states: \( f(\theta) = \text{NE}_g(\theta), \forall \theta \). A social choice rule is monotonic if a social optimum in one state remains optimum in another as long as it does not decrease in any player's preference: \( \forall \theta \in \Theta, a \in f(\theta) \), if \( u_i(a, \theta) \ge u_i(b, \theta) \Rightarrow u_i(a, \theta') \ge u_i(b, \theta') \), \( \forall i \in N, b \in A \), then \( a \in f(\theta') \).

Information asymmetry {Arrow1963, Akerlof1970} can take the form of either hidden information (incomplete information) or hidden action (imperfect information). As of hidden information, adverse selection refers to the situation when the type of product is hidden from one party in a transaction. As of hidden action, moral hazard refers to the situation when individuals take greater risks when the cost is shared with other parties than taken alone.

The revelation principle: {Myerson1979, 1982, 1983, 1984} incentive-compatible direct mechanisms

In mechanism design, a process is incentive-compatible {Hurwicz2006} if all of the participants fare best when they truthfully reveal any private information asked for by the mechanism.

“information efficient”

Samuelson's conjectures:

  1. Truthful revelation of preferences is not a Nash equilibrium in the Lindahl game.
  2. There may be no successful implementation for public goods under decentralization.

Principal-agent (委托-代理) problem is a class of games with information asymmetry where one player (the principal) attempts to offer incentives to the other (the agent) to encourage the agent to act in the principal's best interest.

optimal mechanism

Efficient mechanism (Vickrey-Clarke-Groves mechanism)

Classic Examples

The most minimal class of games are two-person two-strategy games, which is fully defined by eight numbers, or equivalently two 2-by-2 matrices. von Neumann started with a smaller set: zero-sum games, which can be defined by four numbers, or a 2-by-2 matrix. These games have finite strategy space, and can be called matrix games.

The strategy space becomes convex if each player's preference admits metric and mixed strategies are allowed. "games on the square"

Hawk-Dove (Prisoner's Dilemma)

Table: Normal form of Hawk-Dove (symmetric game, showing player 1's payoff)

Player 1\2 Hawk Dove
Hawk \((v-c)/2\) \(v\)
Dove 0 \({v}/{2}\)

When \(v > c\), Hawk-Dove is strategically equivalent to prisoner's dilemma (PD), which has one pure strategy NE: (Hawk, Hawk), aka (Defect, Defect).

When \(v < c\), Hawk-Dove two pure strategy Nash equilibria (Hawk, Dove) and (Dove, Hawk), and a mixed strategy symmetric equilibrium.

Strategy for the iterated prisoners' dilemma:

  • 报复:成功的策略必须不是一个盲目乐观者。要始终报复。一个非报复策略的例子是始终合作。这是一个非常糟糕的选择,因为“下流”策略将残酷地剥削这样的傻瓜。
  • 宽恕:成功策略的另一个品质是必须要宽恕。如果对手不继续背叛,它们会一再退却到合作。这停止了报复和反报复的长期进行,最大化了得分点数。
  • 不嫉妒:最后一个品质是不嫉妒,就是说不去争取得到高于对手的分数(对于“友善”的策略来说这也是不可能的,也就是说“友善”的策略永远无法得到高于对手的分数)。【但也不一定就比对手低。】

Zero-sum game

Matching pennies  Rock–paper–scissors
1,0 0,1           1,1 0,2 2,0
0,1 1,0           2,0 1,1 0,2
                  0,2 2,0 1,1

Neither have pure strategy NE, but have one mixed strategy NE: random.

Coordination game

A coordination game is a game where players' payoffs are maximized by them taking the same action. In a Pareto coordination game such as Hi-Lo, one coordination outcome dominates other coordination outcomes. In a pure coordination game, multiple coordination outcomes are rationalizable.

Hi-Lo    Pure     Stag hunt  Battle of the sexes
2,2 0,0  1,1 0,0  2,2 0,1    2,1 0,0
0,0 1,1  0,0 1,1  1,0 1,1    0,0 1,2

Aggregative game

Cournot game is the n-player game based on Cournot competition: \( \{ N, S, u \} \) where \( S = \mathbb{R}_+^n \) and \( u_i(s)=s_i P(\sum_j s_j) - C_i(s_i) \). Here, \(P(q)\) is the price function of total production, aka the inverse demand function; \(C_i(q_i)\) is each player's total production cost. Cournot equilibrium {Cournot1838} (or Cournot-Nash equilibrium) is the stable focus of best-response dynamics, thus stronger than Nash equilibrium. Cournot games have pure strategy Nash equilibrium in without assuming quasi-concavity of payoff functions {Novshek1985} and convexity of strategy spaces {Kukushkin1994}. As a special form, quasi-Cournot game have an affine price function. Constant marginal costs can be used to simplify analysis. Homogeneous firms have identical cost functions, leading to a symmetric game, with symmetric \(n\)-firm Cournot equilibrium. Symmetric quasi-Cournot game is a strictly convex game, thus with a unique stable pure strategy Nash equilibrium.

A game with one-dimensional player strategy spaces is aggregative if every player's payoff is a function of the player's own strategy and the sum of all players' strategies {Selten1970}: \( u_i = u_i(s_i, \sum_j s_j) \). The game is fully aggregative if every player's payoff and marginal payoff are functions of the player's own strategy and an aggregate of all players' strategies {Cornes2012}: \( u_i(s_i, t(\mathbf{s})) \), \( \frac{\partial u_i}{\partial s_i}(s_i, t(\mathbf{s})) \).

A game with finite dimensional player strategy spaces is quasi-aggregative, if each player has an interaction function \(\sigma_i: S_{-i} \to \mathbb{R}\) and a shift function \( F_i: S_i \times \mathbb{R} \to \mathbb{R} \), both continuous, such that payoffs can be written as \( u_i = u_i(s_i, \sigma_i(s_{-i})) \) and the game admits an aggregator \( g: S \to \mathbb{R} \), \( g(s) = F_i(s_i, \sigma_i(s_{-i})) \). The game is generalized quasi-aggregative if instead \( g(s) = F_i(s_i, \sigma_i(s_{-i})) + v_i(s_{-i}) \), where \( v_i \) are arbitrary functions. {Jensen2010}

Characteristics of fully aggregative games: {Cornes2012} If exists continuously differentiable real functions \( h_0 \) and \( h_i, i \in N \), \(h_0\) strictly increasing and \( t(\mathbf{s}) = h_0(\sum_i h_i(s_i)) \), then the game is fully aggregative. The converse is true if \( t(\mathbf{s}) \) is twice continuously differentiable and has positive first partial derivatives with form \( \frac{\partial t}{\partial s_i}(s_i, t) \), and the game has at least three players.

Properties of aggregative games:

  • Aggregative games have strong comparative statics properties of its Nash equilibria {Corchon1994, Acemoglu2013};
  • Quasi-aggregative games are pseudo-potential games, and are best-response potential games if best-response correspondences are monotone {Jensen2010};
  • In fully aggregative games, every player has a backward reply correspondence (or replacement correspondence {Cornes2012});

replacement correspondence... \( R_i(\tau) = \arg\max \)

Supermodular games

A game is supermodular (submodular) {Topkis1979} if the player strategy spaces are complete lattice (partially ordered set with a top and a bottom) subsets of Euclidean spaces, and the payoffs are upper semicontinuous in self strategy, continuous in opponent strategy, and supermodular (submodular): \( f(x) + f(y) \le f(x \land y) + f(x \lor y) \). If a function is twice continuously differentiable, supermodularity (submodularity) is equivalent to nonnegative (nonpositive) mixed partial derivatives: \( \frac{\partial^2 f(x, y)}{\partial x \partial y} \ge 0, \forall x, y\).

Conventional substitutes and complements are defined by whether more aggressive strategies of a firm (lower price, increased advertising, greater quantity, etc.) lowers or raises another firm's total profits: \( \frac{\partial u_i}{\partial s_j} < 0 \) for substitutes; \( \frac{\partial u_i}{\partial s_j} > 0 \) for complements. In comparison, {Bulow1985} proposed strategic substitutes and complements in their study of multi-market oligopoly: the decisions of several players are strategic substitutes (strategic complements) if they mutually offset (reinforce) one another; that is, marginal utility of a player's strategy drops (raises) with increases in the other players' strategies: \( \frac{\partial^2 u_i}{\partial s_j \partial s_i} < 0 \) for strategic substitutes (submodular); \( \frac{\partial^2 u_i}{\partial s_j \partial s_i} > 0 \) for strategic complements (supermodular).

Methods used to study supermodular games are non-topological but algebraic, specifically lattice theory which exploits order properties:

  • Tarski's fixed point theorem {Tarski1955}: Any non-decreasing function from a complete lattice into itself has fixed points which also form a complete lattice.
  • Topkis' theorem {Topkis1979}: For a real-valued function \(f(x,t)\) from the Cartesian product of a compact subset of real numbers \(X \in \mathbb{R}\) and a partially ordered set \(T\), if it is supermodular and is upper semicontinuous in \(x\) for all \(t\), best-response \( B_x(t) = \arg \max_{x \in X} f(x, t) \) has a top \(\overline{x}(t)\) and a bottom \(\underline{x}(t)\), both non-decreasing functions.

By Topkis' theorem, supermodular games have non-decreasing tops and bottoms of players' best-responses. By Tarski's fixed point theorem, supermodular games have pure strategy Nash Equilibria.

For supermodular games, pure strategy Nash equilibria, rationalizable strategies (iterated elimination of strictly dominated strategies), and correlated equilibria have identical tops and bottoms; strategy profile converges within these bounds under dynamic adaptive choice behavior (best-response dynamics and Bayesian learning); and these bounds vary monotonically with certain exogenous parameters {Milgrom1990}.

Potential game

A selfish routing game \( (G, K, c) \) consists of a directed graph \(G = (V, E)\), traffic demand \( K = \{ (s_i, t_i, r_i) | s_i, t_i \in V, r_i > 0, i = 1, \dots, k \} \), and edge cost functions \( c_e: \mathbb{R}_{\ge 0} \to \mathbb{R}_{\ge 0}, e \in E \). Cost of a path under a flow over the network is the sum of edge costs incurred along the path: \( C(P, f) = \sum_{e \in P} c_e (f_e) \).

The nonatomic instance of a selfish routing game has a large population of players, each contributing a negligible amount to the demand of a single O/D pair in \(K\). For nonatomic instances, the O/D pairs in \(K\) are distinct. A (feasible) flow for an O/D pair is a distribution of its demand over the possible paths: \( f_i \in r_i \Delta^{| \mathcal{P}_i | - 1} \), where \( \Delta^n \) denotes the \(n\)-dimensional standard simplex and \( \mathcal{P}_i \) is the set of all possible paths in graph \(G\) with O/D pair \( (s_i, t_i) \). Nonatomic selfish routing was inspired by the toll road example of social cost in {Pigou1920} which was first stated as traffic equilibrium in {Wardrop1952} and first formalized in {Beckmann1956}.

For an atomic instance of a selfish routing game, \(K\) is the set of players, where the O/D pairs may have duplicates; \( \mathcal{P} = \prod_{i=1}^k \mathcal{P}_i \) is the strategy space, where each strategy profile represents a flow of the players. An atomic instance is unweighted if every player controls the same amount of traffic: \( r_i = r, i = 1, \dots, k \). Atomic selfish routing was first formalized in {Roughgarden2002}.

Congestion game {Rosenthal1973} is a special case of atomic selfish routing games. A congestion game is a game where each player chooses a combination from a finite pool of factors, with an additive payoff of factor costs each only depends on the number of players choosing that factor. Symbolically, given factors \( T = \{ 1, \dots, t \} \) and player-specific strategy spaces \( S_i \subseteq 2^T, i \in N \), each player has payoff \( u_i = \sum_{k=1}^t s_{ik} c_k(s_k) \). Congestion games always have pure-strategy Nash equilibria. {Rosenthal1973} Rosenthal's proof constructed an optimization problem with objective function (sometimes called Rosenthal's potential) \( \Phi(a) = \sum_k \sum_{q=0}^{s_k} c_k(q) \) and a finite feasible set, whose optimal points are pure-strategy Nash equilibria of the congestion game.

For the nonatomic instance, an equilibrium flow is a flow such that for every O/D pair, the paths in use all have the minimum cost among the possible paths: \( \{P | f_{iP}^* > 0 \} \subseteq \arg\min_{P \in \mathcal{P}_i} C( P, f^* ) \), \( \forall i \in \{1, \dots, k \} \). For an atomic instance, an equilibrium flow is a pure strategy Nash equilibrium: \( C( P_i^* , f^* ) \le C( P_i, f^* ) \), \( \forall i \in \{1, \dots, k \}, P_i \in \mathcal{P}_i \).

For the nonatomic instance, equilibrium flows exist and induce the same edge costs. For unweighted atomic instances, equilibrium flows exist. If edge cost functions are affine, equilibrium flows also exist for any atomic instance. {AGT2007}

Potential games are generalizations of congestion games and nonatomic selfish routing games (as in {Beckmann1956}). A potential game is a game where the payoffs depend on a real-valued function, called the potential function. {Monderer1996b} first defined four types of potential games: exact potential game, weighted potential game, ordinal potential game, generalized ordinal potential game. With potential function \( \Phi: S \rightarrow \mathbb{R} \), \( \forall {s_{-i} \in s_{-i}}, \forall {s'_i, s''_i \in S_i} \):

\[\begin{aligned} \Phi(s'_i, s_{-i}) - \Phi(s''_i, s_{-i}) &= u_i(s'_i, s_{-i}) - u_i(s''_i, s_{-i}) &&\text{(Exact)} \\ \Phi(s'_i, s_{-i}) - \Phi(s''_i,s_{-i}) &= (u_i(s'_i, s_{-i}) - u_i(s''_i, s_{-i})) w_i &&\text{(Weighted)} \\ u_i(s'_i, s_{-i}) > u_i(s''_i, s_{-i}) &\Leftrightarrow \Phi(s'_i, s_{-i}) > \Phi(s''_i, s_{-i})&&\text{(Ordinal)} \\ u_i(s'_i, s_{-i}) > u_i(s''_i, s_{-i}) &\Rightarrow \Phi(s'_i, s_{-i}) > \Phi(s''_i, s_{-i}) &&\text{(Generalized Ordinal)} \end{aligned}\]

{Voorneveld2000} and {Schipper2004} generalized ordinal potential games to best-response potential game and pseudo-potential game, respectively. The potential function satisfies \( \forall i \in N, \forall s_{-i} \in S_{-i} \):

\[\begin{aligned} B_i(s_{-i}) &= \arg\max_{s_i \in S_i} \Phi(s_i,s_{-i}) &&\text{(Best-response)} \\ B_i(s_{-i}) &\supset \arg\max_{s_i \in S_i} \Phi(s_i,s_{-i}) &&\text{(Pseudo)} \end{aligned}\]

q-potential games {Monderer2007} and nested potential games {Uno2007}.

It can be shown that congestion game is a special type of exact potential game, and the other types of potential games are progressively more general, except that generalized ordinal potential games are not necessarily best-response potential games. {Lã2016}

Characterizations of potential games:

  1. Exact potential games:
    • with real intervals as player strategy spaces and twice continuously differentiable payoffs are equivalent to games with identical mixed partial derivatives of payoffs: \( \frac{\partial^2 u_i}{\partial s_i \partial s_j} = \frac{\partial^2 u_j}{\partial s_i \partial s_j} \) {Monderer1996b};
    • equivalent to coordination-dummy separable games {Facchini1997};
    • include strategically separable, bilateral symmetric interaction {Ui2000}, and general symmetric observations games {La2016};
    • preserved when shrinking the strategy space;
    • have unique potential functions determined up to a constant;
    • that differ only in dummy payoffs have the same potential function;
  2. Weighted potential games are preserved by linear combinations of payoff functions {Monderer1996b}.
  3. Ordinal potential games: {Voorneveld1997}
    • equivalent to games with no weak improvement cycle and strategy space properly ordered by non-deteriorating paths;
    • the order condition is automatically satisfied if the strategy space is countable;
  4. Best-response potential games: {Voorneveld2000}
    • equivalent to games with no best-response cycle and strategy space properly ordered by best-response compatible paths;
    • the order condition is automatically satisfied if the strategy space is countable;
  5. Pseudo-potential games: {Schipper2004}
    • include games with no strict best-response cycle and strategy space properly ordered by strict best-response compatible paths;
    • their identical interest version must satisfy these properties;

Nash equilibrium of potential games:

  • The identical-interest game \( \{ N, S, \Phi \mathbf{1} \} \) resulted from replacing payoffs of a potential game by its potential function has a pure strategy Nash equilibrium being the potential function's global maxima (if exist), and is unique if the potential function is strictly concave;
  • A best-response potential game has the same Nash equilibrium as that of its identical-interest version {Monderer1996b, Voorneveld2000};
  • A pseudo-potential game includes the Nash equilibrium of its identical-interest version in its own {Dubey2006};

Convergence (stability) of Nash equilibrium of potential games:

  • (Generalized ordinal) potential games converge to a stable Nash equilibrium in better-response dynamics, because every improvement path gradually increases the potential function;
    • For finite ordinal potential games, every improvement path is finite {Monderer1996b};
    • For continuous ordinal potential games with bounded payoffs, every \(\epsilon\)-improvement path is finite, so every improvement path compatible with \(\epsilon\)-improvement converges to an \(\epsilon\)-Nash equilibrium {Monderer1996b};
  • Best-response potential games converge to a stable Nash equilibria in best-response dynamics, because every strict best-response compatible path gradually increases the potential function;
  • For finite pseudo-potential games, every best-response compatible path converges to a Nash equilibrium in a finite number of steps {Dubey2006};

Miscellaneous

Topics:

  • Extensive form games: bargaining, contracts;
  • Games with incomplete information: reputation, bargaining, global games, signaling, cheap talk, auctions;
  • Miscellaneous: axiomatic social choice theory;

forward induction; machina triangle; Berge's theorem of maximum; Zermelo's theorem; Kuhn's theorem; belief system; signaling games (perfect Bayesian equilibrium); unknown player payoffs; unknown other player payoffs; stochastic games: Markov perfect equilibrium;

Computation of equilibrium

Regret-matching algorithms; Lemke-Howson algorithm; Scarf's algorithm;

Any two-person game with a finite number of positions can be solved by a minimax algorithm that would exhaustively traverse the game tree.

Algorithmic game theory

Algorithmic game theory is the intersection of algorithms and game theory. Major advances are in algorithmic mechanism design {Nisan1999} and quantifying the inefficiency of equilibria {Roughgarden2002}. Inefficiency of equilibria (aka "price of anarchy") refers to the reduced social value in non-cooperative equilibria compared with the social optimum in cooperative settings.

Resources:


🏷 Category=Economics