Skip to main content
Log in

Nash equilibria in random games with right fat-tailed distributions

  • Original Paper
  • Published:
International Journal of Game Theory Aims and scope Submit manuscript

Abstract

We study the distribution of the number of mixed strategy Nash equilibria in two-player games where each player’s payoffs are independently drawn from an identical distribution. When the payoff distributions are sufficiently right fat-tailed, we characterize the Nash equilibria by best reply cycles of pure strategies, and we show that the expected number of Nash equilibria is approximately \(\sqrt{\pi mn/\left( m+n\right) }\) in a random \(m\times n\) asymmetric game and approximately n/2 in a random \(n\times n\) symmetric game. We also provide new lower bounds for the expected number of Nash equilibria in a random game with any type of payoff distribution.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. Indeed, finding just one Nash equilibrium is considered to be a difficult problem. A standard algorithm to find one Nash equilibrium, the Lemke-Howson algorithm developed by Lemke and Howson (1964), comprises an exponential number of steps in the number of pure strategies in the worst case (Savani and von Stengel 2006).

  2. Weinstein’s (2016) Proposition 4 states that when players are extremely risk loving, two sets support a Nash equilibrium if and only if they have the point best reply property. Our result and proof follow those of Weinstein (2016). We refine the result to show that the supported Nash equilibrium is unique.

  3. The results in McLennan (2005) also apply to games with more than two players. Since we use the relation between supports of Nash equilibria and best reply cycles, our results apply only to two-player games.

  4. In a random two-player asymmetric game, Bárány et al. (2007) show that the expected number of Nash equilibria in a random \(n\times n\) asymmetric game with uniformly distributed payoffs is bounded from below by a function of the order \(\left( \log n\right) ^{n-1}\) while the expected number of Nash equilibria in a random \(n\times n\) asymmetric game with normally distributed payoffs is bounded from below by a function of the order \(\left( \log n\right) ^{\left( n-1\right) /2}\).

  5. It follows from Bezout’s theorem that the number of Nash equilibria in a generic game with given finite sets of strategies is bounded above; see, for example, McKelvey and McLennan (1997, pp. 411–412). For a generic \(m\times n\) game, since each Nash equilibrium has the same size of supports and a given pair of such supports admits at most one Nash equilibrium, the number of Nash equilibria is bounded above by \(\sum _{x=1}^{\min (m,n)}{m \atopwithdelims ()x}{n \atopwithdelims ()x}={m+n \atopwithdelims ()m}-1\). A tighter bound is given by Keiding (1997), which is asymptotically equal to \(2.6^{n}\) in the case of \(m=n\).

  6. We write \(f\left( m,n\right) \sim g\left( m,n\right)\) if \(\lim _{m,n\rightarrow \infty }\frac{f\left( m,n\right) }{g\left( m,n\right) }=1\).

  7. The results in this subsection depend only on the genericity assumption that each pure strategy has a unique best reply. Such an assumption is satisfied with probability one in random games with continuous payoff distributions.

  8. Weinstein (2016) states the result for generic two-player games but does not specify the uniqueness of the solution. We show in the Appendix that the Nash equilibrium supported by \(\left( T_{1},T_{2}\right)\) with the point best reply property is unique.

  9. The Nash equilibrium supported by \(\left( \left\{ 1\right\} ,\left\{ 1\right\} \right)\) is \(\sigma _{1}=\left( 1,0,0,0,0\right)\), \(\sigma _{2}=\left( 1,0,0,0,0,0\right) .\) The Nash equilibrium supported by \(\left( \left\{ 2,3,4\right\} ,\left\{ 3,4,5\right\} \right)\) is \(\sigma _{1}=\left( 0,\frac{535}{2123},\frac{899}{2123},\frac{689}{2123},0\right)\), \(\sigma _{2}=\left( 0,0,\frac{2067}{7822},\frac{1287}{3911},\frac{3181}{7822},0\right)\). And the Nash equilibrium supported by \(\left( \left\{ 1,2,3,4\right\} ,\left\{ 1,3,4,5\right\} \right)\) is \(\sigma _{1}=\left( \frac{41923}{214976},\frac{41477}{214976},\frac{159219}{429952},\frac{103933}{429952},0\right)\), \(\sigma _{2}=\left( \frac{125965}{650267},0,\frac{147342}{650267},\frac{171599}{650267},\frac{205361}{650267},0\right)\).

  10. Note that both players have the same number of point rationalizable strategies.

  11. Scrutinizing the construction of best reply cycles allows us to identify not only the number of Nash equilibria, but also the number of symmetric Nash equilibria. In fact, as in the proof above, each odd length permutation cycle can form one symmetric best reply cycle whereas each even length permutation cycle can form two asymmetric best reply cycles in a pair. A pair of action sets with the point best reply property, i.e., a union of best reply cycles, is symmetric if and only if it consists of symmetric best reply cycles and pairs of asymmetric best reply cycles. Therefore, the distribution of the number of symmetric pairs of action sets with the point best reply property conditional on \({{\,\textrm{PR}\,}}=k\) under \({\mathbb {P}}_{n}^{F}\) is the same as the distribution of \(2^{{{\,\textrm{PC}\,}}_{o}+{{\,\textrm{PC}\,}}_{e}}-1\) under \({\mathbb {Q}}^{k}\). As in Proposition 2, we can show that if the outstanding best reply condition is satisfied, each symmetric pair of action sets with the point best reply property supports a unique symmetric Nash equilibrium. Then by Lemma 4 and calculations in Appendix A.3 (especially Eq. (4)), the expected number of symmetric Nash equilibria is asymptotically \(\sqrt{\pi n/2}\).

  12. A random permutation game is a random game where one player’s payoff matrix is the identity matrix and the other player’s payoff matrix is a random permutation of the identity matrix.

  13. Another lower bound can be derived from Sun (2020). He provides a lower bound for the number of Nash equilibria in a class of generic \(n\times n\) games. In particular, if a game with a “(restricted) even order distribution” has \(z\le n\) pure strategy Nash equilibria, then it has at least \(2^{z}-1\) Nash equilibria. Since the “(restricted-)even-order-distribution” condition is imposed on ordinal rankings of payoffs, this lower bound is independent of the payoff distribution in a random game. However, conditioning on the game having z pure strategy Nash equilibria, the probability of a game satisfying the “(restricted-)even-order-distribution” condition is \(1/\left( \left( n-1\right) !\right) ^{2\left( z-1\right) }\), which is an extremely small value. Therefore, the lower bound calculated for \({\mathbb {E}}{}_{n,n}^{F}\left[ {{\,\textrm{E}\,}}\right]\) is \(1+o\left( 1\right)\).

  14. The Bell number \(B_{w}\) counts the number of possible partitions of a set with cardinality w. The first few Bell numbers are \(B_{1}=1\), \(B_{2}=2\), \(B_{3}=5\), \(B_{4}=15\), \(B_{5}=52,...\) Refer to chapter II.3 of Sedgewick and Flajolet (2009) for more details.

  15. The Chen–Stein method is also applied to show that the number of pure strategy Nash equilibria asymptotically follows a Poisson distribution in random asymmetric games (Rinott and Scarsini 2000). Rinott and Scarsini (2000) apply the Chen–Stein method to the index set, which is the set of all action profiles.

References

  • Alon N, Rudov K, Yariv L (2021) Dominance solvability in random games. Preprint at https://arxiv.org/abs/2105.10743v1

  • Arratia R, Goldstein R, Gordon L (1989) Two moments suffice for Poisson approximations: the Chen–Stein method. Ann Probab 17(1):9–25

    Article  Google Scholar 

  • Bárány I, Vempala S, Vetta A (2007) Nash equilibria in random games. Random Struct Algorithms 31(4):391–405

    Article  Google Scholar 

  • Basu K, Weibull JW (1991) Strategy subsets closed under rational behavior. Econ Lett 36(2):141–146

    Article  Google Scholar 

  • Bernheim BD (1984) Rationalizable strategic behavior. Econometrica 52(4):1007–1028

    Article  Google Scholar 

  • Comtet L (1974) Advanced combinatorics: the art of finite and infinite expansions. D. Reidel Publishing Company, Dordrecht, Holland

  • Goldberg K, Goldman AJ, Newman M (1968) The probaiblity of an equilibrium point. J Res Natl Bureau Stand 72(2):93–101

    Article  Google Scholar 

  • Gül F, Pearce D, Stacchetti E (1993) A bound on the proportion of pure strategy equilibria in generic games. Math Oper Res 18(3):548–552

    Article  Google Scholar 

  • Keiding H (1997) On the maximal number of nash equilibria in an \(n\times n\) bimatrix game. Games Econ Behav 21(1):148–160

    Article  Google Scholar 

  • Lemke CE, Howson JT Jr (1964) Equilibrium points of bimatrix games. J Soc Ind Appl Math 12(2):413–423

    Article  Google Scholar 

  • McKelvey RD, McLennan A (1997) The maximal number of regular totally mixed nash equilibria. J Econ Theory 72(2):411–425

    Article  Google Scholar 

  • McLennan A (2005) The expected number of nash equilibria of a normal form game. Econometrica 73(1):141–174

    Article  Google Scholar 

  • McLennan A, Berg J (2005) Asymptotic expected number of Nash equilibria of two-player normal form games. Games Econ Behav 51(2):264–295

    Article  Google Scholar 

  • Pei T, Takahashi S (2019) Rationalizable strategies in random games. Games Econ Behav 118:110–125

    Article  Google Scholar 

  • Rinott Y, Scarsini M (2000) On the number of pure strategy nash equilibria in random games. Games Econ Behav 33(2):274–293

    Article  Google Scholar 

  • Roberts DP (2005) Pure Nash equilibria of coordination matrix games. Econ Lett 89(1):7–11

    Article  Google Scholar 

  • Savani R, von Stengel B (2006) Hard-to-solve bimatrix games. Econometrica 74(2):397–429

    Article  Google Scholar 

  • Savani R, von Stengel B (2016) Unit vector games. Int J Econ Theory 12(1):7–27

    Article  Google Scholar 

  • Sedgewick R, Flajolet P (1996) An introduction to the analysis of algorithms. Addison-Wesley, Boston

    Google Scholar 

  • Sedgewick R, Flajolet P (2009) Analytic combinatorics. Cambridge University Press, Cambridge

    Google Scholar 

  • Stanford W (1995) A note on the probability of k pure Nash equilibria in matrix games. Games Econ Behav 9(2):238–246

    Article  Google Scholar 

  • Stanford W (1996) The limit distribution of pure strategy Nash equilibria in symmetric bimatrix games. Math Oper Res 21(3):726–733

    Article  Google Scholar 

  • Sun CJ (2020) A sandwich theorem for generic \(n\times n\) two person games. Games Econ Behav 120:86–95

    Article  Google Scholar 

  • Takahashi S (2008) The number of pure Nash equilibria in a random game with nondecreasing best responses. Games Econ Behav 63(1):328–340

    Article  Google Scholar 

  • Weinstein J (2016) The effect of changes in risk attitude on strategic behavior. Econometrica 84(5):1881–1902

    Article  Google Scholar 

  • Weinstein J (2020) Best-reply sets. Econ Theory Bull 8:105–112

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ting Pei.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 Proof of Proposition 1

Without loss of generality, we assume that \(m\ge n\). The proof of Proposition 1 consists of two parts: the derivation of the expected number of Nash equilibria, \({{\,\textrm{E}\,}}\), and an asymptotic analysis. By Theorem 1, the expectation of \({{\,\textrm{E}\,}}\) can be derived from the distribution of the number of \({{\,\textrm{PR}\,}}\), the number of point rationalizable strategies, and the conditional expectation of \({{\,\textrm{E}\,}}\) given \({{\,\textrm{PR}\,}}\). The distribution of \({{\,\textrm{PR}\,}}\) in the following lemma is given by Theorems 3 and 10 in Pei and Takahashi (2019).

Lemma 3

In a random \(m\times n\) asymmetric game, we have for any \(1\le k\le \min \left( {m,n}\right)\)

$$\begin{aligned} {\mathbb {P}}_{m,n}\left( {{\,\textrm{PR}\,}} \ge k\right)&=\frac{m!}{\left( m-k\right) !m^{k}}\frac{n!}{\left( n-k\right) !n^{k}}. \end{aligned}$$

In a random \(n\times n\) symmetric game, for any \(1\le k\le n\), we have

$$\begin{aligned} {\mathbb {P}}_{n}\left( {{\,\textrm{PR}\,}} =k\right)&=k!\left( {\begin{array}{c}n\\ k\end{array}}\right) kn^{-k-1},\\ {\mathbb {P}}_{n}\left( {{\,\textrm{PR}\,}} \ge k\right)&=\frac{n!}{\left( n-k\right) !n^{k}}. \end{aligned}$$

In the rest of this proof, we first show the conditional expectation of E given PR, which is the Permutation Cycle Index. We then complete the proof with an asymptotic analysis.

1.1.1 Permutation cycle index

Denote the expectation under \({\mathbb {Q}}^{k}\) as \({\mathbb {E}}^{k}\). We compute the expectation of \(2^{ BC }-1\) conditional on each player having exactly k point rationalizable strategies, i.e., \({\mathbb {E}}^{k}\left[ 2^{{{\,\textrm{PC}\,}}_{o}+{{\,\textrm{PC}\,}}_{e}}-1\right]\) for asymmetric case and \({\mathbb {E}}^{k}\left[ 2^{{{\,\textrm{PC}\,}}_{o}+2{{\,\textrm{PC}\,}}_{e}}-1\right]\) for symmetric case.

Lemma 4

\({\mathbb {E}}^{k}\left[ 2^{{{\,\textrm{PC}\,}}_{o}+{{\,\textrm{PC}\,}}_{e}}-1\right] =k\) and \({\mathbb {E}}^{k}\left[ 2^{{{\,\textrm{PC}\,}}_{o}+2{{\,\textrm{PC}\,}}_{e}}-1\right] =k^{2}/4+k+\left( -1\right) ^{k}/8-1/8\).

Proof

Let \({\mathfrak {p}}\left( k,l;c_{1},c_{2},\ldots \right)\) be the number of k-permutations that have l permutation cycles, where \(c_{1},c_{2},\ldots\) represent the number of permutation cycles with length \(1,2,\ldots\) Clearly, \(c_{1}+c_{2}+\cdots =l\) and \(c_{1}+2c_{2}+\cdots =k\). By Comtet (1974, Theorem C in Section 6.2), we have the generating function:

$$\begin{aligned} \sum _{k,l,c_{1},c_{2},\ldots \ge 0}{\mathfrak {p}}\left( k,l;c_{1},c_{2},\ldots \right) \frac{t^{k}}{k!}x_{1}^{c_{1}}x_{2}^{c_{2}}\ldots&=\exp \left( x_{1}t+x_{2}\frac{t^{2}}{2}+x_{3}\frac{t^{3}}{3}+\cdots \right) . \end{aligned}$$

Note that

$$\begin{aligned} {\mathbb {E}}^{k}\left[ 2^{PC_{o}+PC_{e}}\right]&=\frac{1}{k!}\sum _{l,c_{1},c_{2},\ldots \ge 0}{\mathfrak {p}}\left( k,l;c_{1},c_{2},\ldots \right) 2^{l}, \end{aligned}$$

which is the coefficient of \(t^{k}\) of the generating function with \(x_{1}=x_{2}=\cdots =2\), i.e., \(\exp \left( 2t+\frac{2t^{2}}{2}+\frac{2t^{3}}{3}+\cdots \right) =\left( 1-t\right) ^{-2}\). Hence,

$$\begin{aligned} {\mathbb {E}}^{k}\left[ 2^{PC_{o}+PC_{e}}-1\right]&=\frac{1}{k!}\frac{\partial ^{k}}{\partial t^{k}}\left( 1-t\right) ^{-2}|_{t=0}-1=k+1-1=k. \end{aligned}$$

Similarly,

$$\begin{aligned} {\mathbb {E}}^{k}\left[ 2^{PC_{o}+2PC_{e}}\right]&=\frac{1}{k!}\sum _{l,c_{1},c_{2},\ldots \ge 0}{\mathfrak {p}}\left( k,l;c_{1},c_{2},\ldots \right) 2{}^{c_{1}}2^{2c_{2}}\ldots , \end{aligned}$$

which is the coefficient of \(t^{k}\) of the generating function with \(x_{1}=x_{3}=\cdots =2\) and \(x_{2}=x_{4}=\cdots =4\), i.e., \(\exp \left( 2t+\frac{4t^{2}}{2}+\frac{2t^{3}}{3}+\frac{4t^{4}}{4}+\cdots \right) =\left( 1-t\right) ^{-3}\left( 1+t\right) ^{-1}\). Hence,

$$\begin{aligned} {\mathbb {E}}^{k}\left[ 2^{{{\,\textrm{PC}\,}}_{o}+2{{\,\textrm{PC}\,}}_{e}}-1\right]&=\frac{1}{k!}\left. \frac{\partial ^{k}}{\partial t^{k}}\left( 1-t\right) ^{-3}\left( 1+t\right) ^{-1}\right| _{t=0}-1\\&=\frac{1}{k!}\left. \sum _{i=0}^{k}\left( {\begin{array}{c}k\\ i\end{array}}\right) \frac{\partial ^{i}\left( 1-t\right) ^{-3}}{\partial t^{i}}\frac{\partial ^{k-i}\left( 1+t\right) ^{-1}}{\partial t^{k-i}}\right| _{t=0}-1\\&=\frac{1}{k!}\sum _{i=0}^{k}\frac{k!}{\left( k-i\right) !i!}\frac{\left( i+2\right) !}{2}\left( -1\right) ^{k-i}\left( k-i\right) !-1\\&=k^{2}/4+k+\left( -1\right) ^{k}/8-1/8. \end{aligned}$$

\(\square\)

1.1.2 The expectation of E

By Corollary 1 and Lemmas 3 and 4, for the asymmetric cases, we have

$$\begin{aligned} \lim _{r\rightarrow \infty }{\mathbb {E}}_{m,n}^{F_{r},G_{r}}\left[ {{\,\textrm{E}\,}}\right]&=\lim _{r\rightarrow \infty }{\mathbb {E}}_{m,n}^{F_{r},G_{r}}\left[ {\mathbb {E}}_{m,n}^{F_{r},G_{r}}\left[ {{\,\textrm{E}\,}}| {{\,\textrm{PR}\,}} \right] \right] \\&=\lim _{r\rightarrow \infty }{\mathbb {E}}_{m,n}^{F_{r},G_{r}}\left[ 2^{BC}-1|{{\,\textrm{PR}\,}}\right] \\&=\lim _{r\rightarrow \infty }{\mathbb {E}}_{m,n}^{F}\left[ {{\,\textrm{PR}\,}}\right] \\&=\sum _{k=1}^{n}\frac{m!}{\left( m-k\right) !m^{k}}\frac{n!}{\left( n-k\right) !n^{k}}. \end{aligned}$$

And for the symmetric cases, we have

$$\begin{aligned} \lim _{r\rightarrow 0}{\mathbb {E}}_{n}^{F_{r}}\left[ {{\,\textrm{E}\,}}\right]&=\lim _{r\rightarrow 0}{\mathbb {E}}_{n}^{F_{r}}\left[ {\mathbb {E}}_{n}^{F_{r}}\left[ {{\,\textrm{E}\,}}|{{\,\textrm{PR}\,}}\right] \right] \\&=\frac{1}{4}{\mathbb {E}}_{n}\left[ {{\,\textrm{PR}\,}}^{2}\right] +\mathbb {{\mathbb {E}}}_{n}\left[ {{\,\textrm{PR}\,}} \right] +\frac{1}{8}{\mathbb {E}}_{n}\left[ \left( -1\right) ^{ {{\,\textrm{PR}\,}} }-1\right] \\ {}&=\sum _{k=1}^{n}\frac{n!\left( k^{2}/4+k+\left( -1\right) ^{k}/8-1/8\right) }{\left( n-k\right) !n^{k+1}}. \end{aligned}$$

1.1.3 Asymptotic analysis

We use the Laplace method in the asymptotic analysis. Details of the theorems and techniques used in this subsection can be found in Sections 4.5, 4.6, and 4.7 of Sedgewick and Flajolet (1996). To make this part self-contained, we list these results as lemmas here:

The analysis is closely related to Ramanujan \(Q-\)distribution:

$$\begin{aligned} \frac{n!}{\left( n-k\right) !n^{k}} \end{aligned}$$

and Ramanujan Q-function:

$$\begin{aligned} \sum _{k=n}^{n}&\frac{n!}{\left( n-k\right) !n^{k}}. \end{aligned}$$

The following lemma, which is Theorem 4.4 of Section 4.6 in Sedgewick and Flajolet (1996), gives the approximation for the Ramanujan Q-distribution.

Lemma 5

As \(n\rightarrow \infty\), the following (relative) approximation holds for \(k=o\left( n^{2/3}\right)\):

$$\begin{aligned} \frac{n!}{\left( n-k\right) !n^{k}}=&\exp \left( -\frac{k^{2}}{2n}\right) \left( 1+O\left( \frac{k}{n}\right) +O\left( \frac{k^{3}}{n^{2}}\right) \right) . \end{aligned}$$

In addition, the following (absolute) approximation holds for all k:

$$\begin{aligned} \frac{n!}{\left( n-k\right) !n^{k}}=&\exp \left( -\frac{k^{2}}{2n}\right) +O\left( \frac{1}{\sqrt{n}}\right) . \end{aligned}$$

In Page 181 of Sedgewick and Flajolet (1996), a simpler form of the Euler–Maclaurin summation formula is provided. We summarize the result in the following lemma:

Lemma 6

Let \(f\left( x\right)\) be a function defined on an interval \(\left[ a,b\right]\) where a and b are integers. Suppose that the derivatives \(f^{\prime \prime }\left( x\right)\) exist and are continuous. Then

$$\begin{aligned} \sum _{k=a}^{k=b}f\left( k\right)&=\int _{a}^{b}f\left( x\right) dx+\frac{f\left( a\right) +f\left( b\right) }{2}+\frac{1}{12}\left( f^{\prime }\left( b\right) -f^{\prime }\left( a\right) \right) +R, \end{aligned}$$

where R is a remainder term satisfying

$$\begin{aligned} \left| R\right|&<\frac{1}{\pi ^{2}}\int _{a}^{b}\left| f^{\prime \prime }\left( x\right) \right| dx. \end{aligned}$$

Now, we are ready to derive the asymptotic analysis of Proposition 1. We focus on the asymmetric cases first. Recall that

$$\begin{aligned} \lim _{r\rightarrow \infty }{\mathbb {E}}_{m,n}^{F_{r},G_{r}}\left[ {{\,\textrm{E}\,}}\right]&=\sum _{k=1}^{n}\frac{m!}{\left( m-k\right) !m^{k}}\frac{n!}{\left( n-k\right) !n^{k}}. \end{aligned}$$

For the asymptotic analysis, following the proof of Lemma 5 in Sedgewick and Flajolet (1996), here we use a more accurate form of the relative approximation of the Ramanujan \(Q-\)distribution for \(k=o\left( n^{2/3}\right)\):

$$\begin{aligned} \frac{n!}{\left( n-k\right) !n^{k}}=&\exp \left( \sum _{j=1}^{k-1}\ln \left( 1-\frac{j}{n}\right) \right) \nonumber \\ =&\exp \left( \sum _{j=1}^{k-1}\left( -\frac{j}{n}-\frac{j^{2}}{2n^{2}}-\frac{j^{3}}{3n^{3}}-\cdots \right) \right) \nonumber \\ =&\exp \left( -\frac{k\left( k-1\right) }{2n}-\frac{k\left( k-1\right) \left( 2k-1\right) }{12n^{2}}+O\left( \frac{k^{4}}{n^{3}}\right) \right) \nonumber \\ =&\exp \left( -\frac{k^{2}}{2n}\right) \exp \left( \frac{k}{2n} -\frac{k^{3}}{6n^{2}}+o\left( \frac{k^{3}}{n^{2}}\right) \right) \nonumber \\ =&\exp \left( -\frac{k^{2}}{2n}\right) \left( 1+\frac{k}{2n}-\frac{k^{3}}{6n^{2}} +o\left( \frac{k^{3}}{n^{2}}\right) \right) . \end{aligned}$$
(1)

Note that \(\frac{m!}{\left( m-k\right) !m^{k}}\frac{n!}{\left( n-k\right) !n^{k}}\) is decreasing with k. Let \(k_{0}\) be the nearest integer to \(\sqrt{n}\log n\), we have \(k_{0}=o\left( n^{2/3}\right)\). Note that \(m\ge n\). Then by Eq. (1),

$$\begin{aligned}&\frac{m!}{\left( m-k_{0}\right) !m^{k_{0}}}\frac{n!}{\left( n-k_{0}\right) !n^{k_{0}}}\nonumber \\&\quad = \exp \left( -\frac{k_{0}^{2}}{2m}-\frac{k_{0}^{2}}{2n}\right) \left( 1+\frac{k_{0}}{2n}-\frac{k_{0}^{3}}{6n^{2}}+o\left( \frac{k_{0}^{3}}{n^{2}}\right) \right) \left( 1+\frac{k_{0}}{2m}-\frac{k_{0}^{3}}{6m^{2}}+o\left( \frac{k_{0}^{3}}{m^{2}}\right) \right) \nonumber \\&\quad = \exp \left( -\frac{\left( m+n\right) \left( \log n\right) ^{2}}{2m}\right) \left( 1+O\left( \frac{\left( \log n\right) ^{3}}{\sqrt{n}}\right) \right) \nonumber \\&\quad = n^{-\frac{\left( m+n\right) \log n}{2m}}\left( 1+O\left( \frac{\left( \log n\right) ^{3}}{\sqrt{n}}\right) \right) . \end{aligned}$$
(2)

Hence, for all \(k\ge k_{0}\), since each term is exponentially small, the contribution to the sum is \(o\left( 1\right)\), as well as the error terms. Therefore, we can use the relative approximation to all k from 1 to \(k_{0}\), and the error terms are determined by

$$\begin{aligned} \exp \left( -\frac{\left( m+n\right) k^{2}}{2mn}\right) \left( \frac{k}{2n}-\frac{k^{3}}{6n^{2}}\right) . \end{aligned}$$

By applying the Euler–Maclaurin summation to the functions \(x\exp \left( -\left( m+n\right) x^{2}/2m\right)\) and \(x^{3}\exp \left( -\left( m+n\right) x^{2}/2m\right)\) with step \(1/\sqrt{n}\), we find that both \(\sum _{k=1}^{k_{0}}\left( k/n\right) \exp \left( -\left( m+n\right) k^{2}/\left( 2mn\right) \right)\) and \(\sum _{k=1}^{k_{0}}\left( k^{3}/n^{2}\right) \exp \left( -\left( m+n\right) k^{2}/\left( 2mn\right) \right)\) are \(O\left( 1\right)\). Hence,

$$\begin{aligned} {\mathbb {E}}_{m,n}^{F_{r},G_{r}}\left[ {{\,\textrm{E}\,}}\right]&=\sum _{k=1}^{n}\frac{m!}{\left( m-k\right) !m^{k}}\frac{n!}{\left( n-k\right) !n^{k}}+o\left( 1\right) \\&=\sum _{k=1}^{n}\exp \left( -\frac{\left( m+n\right) k^{2}}{2mn}\right) +O\left( 1\right) . \end{aligned}$$

Then we apply the Euler–Maclaurin summation to the remaining sum, i.e., the sum of the values of the function \(\exp \left( -\left( m+n\right) x^{2}/\left( 2m\right) \right)\) at regularly spaced points with step \(1/\sqrt{n}\). The approximation is

$$\begin{aligned} \sum _{k=1}^{n}\exp \left( -\frac{\left( m+n\right) k^{2}}{2m}\right) =&\sqrt{n}\int _{0}^{\sqrt{n}}\exp \left( \frac{-\left( m+n\right) k^{2}}{2m}\right) dk+O\left( 1\right) \nonumber \\ =&\sqrt{n}\int _{0}^{\infty }\exp \left( \frac{-\left( m+n\right) k^{2}}{2m}\right) dk+O\left( 1\right) \nonumber \\ =&\sqrt{\frac{\pi mn}{2\left( m+n\right) }}+O\left( 1\right) . \end{aligned}$$
(3)

The second equality holds as the integrand for \(k>\sqrt{n}\) is exponentially small, which means that the integral for \(k>\sqrt{n}\) is O(1).

For random \(n\times n\) symmetric games, recall that

$$\begin{aligned} \lim _{r\rightarrow 0}{\mathbb {E}}_{n}^{F_{r}}\left[ {{\,\textrm{E}\,}}\right]&=\lim _{r\rightarrow 0}{\mathbb {E}}_{n}^{F_{r}}\left[ {\mathbb {E}}_{n}^{F_{r}}\left[ {{\,\textrm{E}\,}}|{{\,\textrm{PR}\,}}\right] \right] \\&=\frac{1}{4}{\mathbb {E}}_{n}\left[ {{\,\textrm{PR}\,}}^{2}\right] +\mathbb {{\mathbb {E}}}_{n}\left[ {{\,\textrm{PR}\,}} \right] +\frac{1}{8}{\mathbb {E}}_{n}\left[ \left( -1\right) ^{ {{\,\textrm{PR}\,}} }-1\right] . \end{aligned}$$

Note that \({\mathbb {E}}_{n}\left[ \left( -1\right) ^{ {{\,\textrm{PR}\,}} }-1\right]\) is \(O\left( 1\right)\). And \({\mathbb {E}}_{n}\left[ {{\,\textrm{PR}\,}} \right] =\sum _{k=1}^{n}n!/\left( \left( n-k\right) !n^{k}\right)\) is the Ramanujan Q-function. Hence, by Theorem 4.8 in Sedgewick and Flajolet (1996),

$$\begin{aligned} {\mathbb {E}}_{n}\left[ PR \right] =&\sqrt{\frac{\pi n}{2}}+O\left( 1\right) . \end{aligned}$$
(4)

Now we have

$$\begin{aligned} {\mathbb {E}}_{n}\left[ {{\,\textrm{PR}\,}} ^{2}\right]&=\sum _{k=1}^{n}k^{2}k!\left( {\begin{array}{c}n\\ k\end{array}}\right) kn^{-k-1}=\sum _{k=1}^{n}\frac{n!}{\left( n-k\right) !n^{k}}\frac{k^{3}}{n}. \end{aligned}$$

Similarly, for the asymptotic analysis in asymmetric cases, we have

$$\begin{aligned} \sum _{k=1}^{n}\frac{n!}{\left( n-k\right) !n^{k}}\frac{k^{3}}{n}&=\sum _{k=1}^{n}\frac{k^{3}}{n}\exp \left( -\frac{k^{2}}{2n}\right) +O\left( 1\right) . \end{aligned}$$

We apply the Euler–Maclaurin summation to the remaining sum, i.e., sum of values of the function \(x^{3}\exp \left( -x^{2}/2\right)\) at regularly spaced points with step \(1/\sqrt{n}\) to obtain:

$$\begin{aligned} \sum _{k=1}^{n}\frac{k^{3}}{\sqrt{n^{3}}}\exp \left( -\frac{k^{2}}{2n}\right)&=2\sqrt{n}+O\left( 1\right) . \end{aligned}$$

Hence, \(\sum _{k=1}^{n}\frac{k^{3}}{n}\exp \left( -\frac{k^{2}}{2n}\right) =2n+O\left( \sqrt{n}\right)\). And we have

$$\begin{aligned} {\mathbb {E}}_{n}^{F_{r}}\left[ {{\,\textrm{E}\,}}\right]&=\frac{1}{4}{\mathbb {E}}_{n}\left[ {{\,\textrm{PR}\,}}^{2}\right] +\mathbb {{\mathbb {E}}}_{n}\left[ {{\,\textrm{PR}\,}}\right] +\frac{1}{8}{\mathbb {E}}_{n}\left[ \left( -1\right) ^{{{\,\textrm{PR}\,}}}-1\right] \\&=\frac{n}{2}+O\left( \sqrt{n}\right) . \end{aligned}$$

1.2 A.2 Proof of Proposition 2

For the sake of completeness, we provide a self-contained proof of Proposition 2. In addition, we would like to thank an anonymous referee for suggesting a simpler proof of the uniqueness of the Nash equilibrium.

For player p’s mixed strategy \(\sigma _p\) and \(\sigma _p^\prime\), given a subset T of the opponent’s strategy space, we write \(\sigma _p >_T\sigma _p^\prime\) if \(\sigma _p\) weakly dominates \(\sigma _p^\prime\) when opposing profiles are restricted to T. Then the following lemma, which is used in the proof, is Corollary 2 in Weinstein (2020):

Lemma 7

A two-player game has a Nash equilibrium with supports \(\left( T_{1},T_{2}\right)\), if and only if there is no pair \(\sigma _1\in \Delta \left\{ 1,\ldots ,m\right\} ,\sigma _1^\prime \in \Delta (T_1)\) such that \(\sigma _1>_{T_{2}}\sigma _1 ^\prime\) and no pair \(\sigma _2\in \Delta \left\{ 1,\ldots ,n\right\} ,\sigma _2^\prime \in \Delta (T_2)\) such that \(\sigma _2>_{T_{1}}\sigma _2 ^\prime\). When such pairs do exist, they can be chosen with disjoint support, that is, \(\sigma _{1i}\sigma _{1i}^\prime =0\) for all \(i\in \left\{ 1,\ldots ,m\right\}\).

To show that the Nash equilibrium supported by \(\left( T_{1},T_{2}\right)\) with the point best reply property is unique if the game satisfies the outstanding best reply condition, we present the following Lemma.

Lemma 8

In a generic \(m\times n\) game \(\left( U,V\right)\) that satisfies the outstanding best reply condition, each of player p’s pure best replies to \(\sigma _{-p}\) must be a best reply to some strategy in the support of \(\sigma _{-p}\), i.e., the pure best reply to \(\sigma _{-p}\) belongs to \(B_{p}\left( {{\,\textrm{Supp}\,}}\left( \sigma _{-p}\right) \right)\).

Proof

For player 1 and the mixed strategy \(\sigma _2\), let \(S_2={{\,\textrm{Supp}\,}}(\sigma _2)\). Consider the mixed strategy \(\sigma _1\) that has equal weight on all strategies in \(B_1(S_2)\). Then for any pure strategy \(i\notin B_1(S_2)\), as the game satisfies the outstanding best reply condition,

$$\begin{aligned} \sum _{j\in S_2}\sigma _{2j}(u_{ij}-{\underline{u}})<\sum _{j\in S_2}\sigma _{2j}\frac{{\bar{u}}_j-{\underline{u}}}{m-1}\le \sum _{j\in S_2}\sigma _{2j}\frac{{\bar{u}}_j-{\underline{u}}}{\left| B_1(S_2)\right| }\le \sum _{j\in S_2} \sum _{i^\prime \in B_1(S_2)} \sigma _{2j} \sigma _{1i^\prime } (u_{i^\prime j}-{\underline{u}}) \end{aligned}$$

i.e., the payoff for i against \(\sigma _{2}\) is strictly less than the payoff for i against \(\sigma _1\). Hence, the pure best reply to \(\sigma _2\) belongs to \(B_1({{\,\textrm{Supp}\,}}(\sigma _{2}))\).

A similar argument applies to player 2. And we conclude the proof. \(\square\)

Proof of Proposition 2

Suppose \(\left( T_{1},T_{2}\right)\) has the point best-reply property. We first show that there is some Nash equilibrium supported by \(\left( T_{1},T_{2}\right)\). If \(T_1=\left\{ 1,\ldots ,m\right\}\) or \(T_2=\{1,\ldots ,n\}\), then the condition following ‘if and only if’ in Lemma 7 is satisfied. Therefore, the lemma yields a Nash equilibrium. Now suppose \(T_1\subsetneq \{1,\ldots ,m\}\) and \(T_2\subsetneq \{1,\ldots ,n\}\). Then \(\left| T_1\right| \le m-1\) and \(\left| T_2\right| \le n-1\). For any \(i\in T_1\), it is a best reply to some \(j\in T_2\). As the game satisfies the definition of the outstanding best reply condition, we have

$$\begin{aligned} {u}_{ij}-{\underline{u}}&>\left( m-1\right) \left( u_{i^\prime j}-{\underline{u}}\right) \,\text {for all}\,\, i^\prime \ne i. \end{aligned}$$
(5)

Suppose \(\sigma _1>_{T_2}\sigma _1^\prime\) with \(\sigma _1\in \Delta \left\{ 1,\ldots ,m\right\} ,\sigma _1^\prime \in \Delta (T_1)\) having disjoint support, as in Lemma 7. Suppose \(\sigma _{1i}^\prime \ge 1/(m-1)\). Disjoint support implies \(\sigma _{1i}=0\). Then the payoff of \(\sigma _{1i}^\prime\) against j is strictly higher than the payoff of \(\sigma _{1i}\) against j as

$$\begin{aligned} \sigma _{1i}^\prime ({u}_{ij}-{\underline{u}})+\sum _{i^\prime \ne i}\sigma _{1i^\prime }^\prime ({u}_{i^\prime j}-{\underline{u}})&\ge \frac{1}{m-1}({u}_{ij}-{\underline{u}})>\sum _{i^\prime \in T}\sigma _{1i}^\prime ({u}_{i^\prime j}-{\underline{u}}), \end{aligned}$$
(6)

which is a contradiction. Hence, \(\sigma _{1i}^\prime < 1/(m-1)\) for all \(i\in T_1\), which contradicts \(\sum _{i\in T_1}\sigma _{1i}^\prime =1\). The analogous conditions hold for the reversed role. Therefore, we have proven the existence of the Nash equilibrium with supports \(\left( T_{1},T_{2}\right)\).

Now we show that the equilibrium supported by \(\left( T_{1},T_{2}\right)\) is unique. Suppose \(\left( T_{1},T_{2}\right)\) supports two different Nash equilibria \(\left( \sigma _{1},\sigma _{2}\right)\) and \(\left( \sigma _{1}^{\prime },\sigma _{2}^{\prime }\right)\). Without loss of generality, assume \(\sigma _{1}\ne \sigma _{1}^{\prime }\). Then all elements of \(T_2\) are best replies to both \(\sigma _1\) and \(\sigma _1^\prime\), and to any point on the line containing these points, including a \(\sigma _1^{\prime \prime }\) where this line leaves the simplex. However, some element of \(T_2\) is not a pure best response to any element of the support of \(\sigma _1^{\prime \prime }\), contrary to Lemma 8.

Lastly, note that by Lemma 8, if a game satisfies the outstanding best reply condition, then the support of any Nash equilibrium must have the point best-reply property. Hence, the second part of the proposition holds. We conclude the proof.

1.3 Random symmetric games with more than two players

In this appendix, we discuss the lower bounds for the expected number of Nash equilibria in random symmetric games with more than two players. Consider a w-player random symmetric game with each player having n strategies. To calculate lower bounds based on Proposition 3, we need to know the expected number of pure strategy Nash equilibria, denoted as \({\mathbb {E}}_{n}^{w}\left[ {{\,\textrm{PE}\,}}\right]\), and the probability of the game having no Nash equilibrium, denoted as \({\mathbb {P}}_{n}^{w}\left( {{\,\textrm{PE}\,}}=0\right)\).

We characterize strategy profiles by the partition of the w players. For example, when \(w=3\), there are five partitions: \(\left\{ \left\{ 1,2,3\right\} \right\}\), \(\left\{ \left\{ 1\right\} ,\left\{ 2,3\right\} \right\}\), \(\left\{ \left\{ 1,2\right\} ,\left\{ 3\right\} \right\}\), \(\left\{ \left\{ 1,3\right\} ,\left\{ 2\right\} \right\}\), and \(\left\{ \left\{ 1\right\} ,\left\{ 2\right\} ,\left\{ 3\right\} \right\}\). Each partition characterizes a group of strategy profiles where players in the same subset play the same strategy, and players in different subsets play different strategies, e.g., \(\left\{ \left\{ 1,2\right\} ,\left\{ 3\right\} \right\}\) is the group of strategies where player 1 and 2 choose the same strategy and player 3 chooses a different strategy.

Proposition 5

\({\mathbb {E}}_{n}^{w}\left[ {{\,\textrm{PE}\,}}\right] \rightarrow B_{w}\) as \(n\rightarrow \infty\), where \(B_{w}\) is the Bell number.Footnote 14

Proof

Suppose A is a partition of the w players. For any strategy profile s characterized as A, the probability that s is a pure strategy Nash equilibrium is \(1/n^{\left| A\right| }\). There are \(n!/\left( n-\left| A\right| \right) !\) strategy profiles that induce partition A. Therefore, the expected number of pure strategy Nash equilibria that induce partition A is \(n!/\left( n^{\left| A\right| }\left( n-\left| A\right| \right) !\right) \rightarrow 1\) as \(n\rightarrow \infty\). Hence, the expected number of pure strategy Nash equilibria equals the number of partitions of the w players, \(B_{w}\). \(\square\)

We now provide a method to calculate \({\mathbb {P}}_{n}^{w}\left( {{\,\textrm{PE}\,}}=0\right)\) as \(n\rightarrow \infty\), which is an application of the Chen–Stein method as indicated by Arratia et al. (1989).Footnote 15 Note that in a symmetric game, if an action profile is a pure strategy Nash equilibrium, then all permutations of the action profile are also pure strategy Nash equilibria. Hence, we cannot directly apply the Chen–Stein method to the set of all action profiles. Instead, we replicate their application with a slight change to the index set. Below, we use \(w=3\) as an example. For \(w>3\), the logic is the same, but the computation is more tedious.

When \(w=3\), we divide action profiles into three sets. Set \(A_{1}\) consists of element \(\left\{ \left( x,x,x\right) \right\}\): all three players employ the same action. The expected number of elements in \(A_{1}\) that are pure strategy Nash equilibria is one. Set \(A_{2}\) consists of element \(\left\{ \left( x,x,y\right) ,\left( x,y,x\right) ,\left( y,x,x\right) \right\}\): only two players employ the same action. The expected number of elements in \(A_{2}\) that are pure strategy Nash equilibria is \(n\left( n-1\right) /n^{2}\rightarrow 1\). Set \(A_{3}\) consists of element \(\left\{ \left( x,y,z\right) ,\left( x,z,y\right) ,\left( y,x,z\right) ,\left( y,z,x\right) ,\left( z,x,y\right) ,\left( z,y,x\right) \right\}\): all three players employ different actions. The expected number of elements in \(A_3\) that are pure strategy Nash equilibria is \(n\left( n-1\right) \left( n-1\right) /\left( 6n^{3}\right) \rightarrow 1/6\). We then apply the Chen–Stein method to the union of all elements in \(A_1\), \(A_2\), and \(A_3\). Therefore, the number of elements that are pure strategy Nash equilibria asymptotically follows a Poisson distribution with mean \(1+1+1/6=13/6\). And we have \({\mathbb {P}}_{n}^{3}\left( {{\,\textrm{PE}\,}}=0\right) \rightarrow e^{-13/6}\).

Hence, by Proposition 3, we obtain a lower bound for random symmetric games with three players \(2B_{3}-1+2e^{-13/6}\approx 9.23\).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pei, T., Takahashi, S. Nash equilibria in random games with right fat-tailed distributions. Int J Game Theory 52, 1153–1177 (2023). https://doi.org/10.1007/s00182-023-00863-2

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00182-023-00863-2

Keywords

Navigation