Skip to main content
Log in

The refined best-response correspondence in normal form games

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

Abstract

This paper provides an in-depth study of the (most) refined best-response correspondence introduced by Balkenborg et al. (Theor Econ 8:165–192, 2013). An example demonstrates that this correspondence can be very different from the standard best-response correspondence. In two-player games, however, the refined best-response correspondence of a given game is the same as the best-response correspondence of a slightly modified game. The modified game is derived from the original game by reducing the payoff by a small amount for all pure strategies that are weakly inferior. Weakly inferior strategies, for two-player games, are pure strategies that are either weakly dominated or are equivalent to a proper mixture of pure strategies. Fixed points of the refined best-response correspondence are not equivalent to any known Nash equilibrium refinement. A class of simple communication games demonstrates the usefulness and intuitive appeal of the refined best-response correspondence.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

Notes

  1. What we here call a best-response-like dynamics or correspondence is what in Balkenborg et al. (2013) is formally defined and termed a generalized best-response dynamics or correspondence.

  2. A proper mixture of pure strategies places positive weight on at least two pure strategies.

  3. The function \(u\) will also denote the expected utility function in the mixed extension of the game \(\Gamma \).

  4. Balkenborg (1992) calls strategies \(s_i \in \mathcal {S}_i(x)\) semi-robust best responses. This is inspired by Okada (1983) who calls a strategy a robust best response to strategy profile \(x\) if it is a best response for an open neighborhood of \(x\). One could call a strategy robust if it is a robust best response against some strategy profile. Any pure strategy that is either a robust best response or a semi-robust best response against some strategy profile \(x\) is, thus, a robust strategy. Note that, while every strategy profile \(x \in \Theta \) has a semi-robust best response for all players, it may not have a robust best response.

  5. Our notions of strict and weak inferiority are motivated by, but not identical to, the notion of inferior choices in Harsanyi and Selten (1988).

  6. Theorem 2 does also not extend to games with more than two players. One direction is, of course, true. That is that any pure strategy that is weakly dominated or equivalent to a proper mixture of pure strategies is weakly inferior. But there may well be additional weakly inferior strategies. The reason is well-known. In three player games an undominated strategy may still be never a best response (as players here always know, or believe, if you will, that opponents cannot correlate their strategy choices). For a textbook example see Ritzberger (2002, Example 5.7).

  7. Note that the best-response correspondence in the former case is typically not identical to the best-response correspondence in the latter case.

  8. Note that this condition is, for instance, satisfied, for the mixed equilibrium in the two-player game of matching pennies.

  9. In fact, this can be seen directly from the observation that player \(2\)’s strategy \((1/2,0,1/2)\) is weakly dominated by strategy \(E\).

  10. Strategy \(C\) yielding the lowest possible payoff must be played with much smaller probability than strategy \(A\) in any \(\epsilon \)-proper equilibrium.

  11. That \((B,B)\) is not strictly perfect also follows from the fact, shown in Vermeulen (1996), that in \(3 \times 3\)-games strictly perfect equilibria are proper.

  12. The games here are much simpler than those of Crawford and Sobel (1982). For a discussion of communication in simple games see Farrell and Rabin (1996). See Gordon (2006) for a discussion of persistent retracts in the cheap-talk games of Crawford and Sobel (1982).

  13. Note that even though players in this game have own-payoff equivalent pure strategies, this game is in the class \(\mathcal {G}^*\).

  14. For this procedure see e.g. Dekel and Fudenberg (1990), Börgers (1994), and Ben Porath (1997).

  15. This transversality theorem is a straightforward consequence of Sard’s theorem. If one assumes an algebraic map and uses in its proof in Guillemin and Pollack (1974) the algebraic version of Sard’s theorem in Bochnak et al. (1998) one obtains a stronger version of the transversality theorem where the conclusion of the theorem holds outside a lower dimensional semi-algebraic set.

  16. The result is well known, e.g., in algebraic geometry: \(q_{i}\) defines the so-called Segre-embedding. The result is needed in algebraic geometry to show that the product of projective spaces can itself be embedded into a projective space, i.e., is projective-algebraic.

References

  • Balkenborg D (1992) The properties of persistent retracts and related concepts. Ph.D. thesis, University of Bonn, Bonn

  • Balkenborg D, Hofbauer J, Kuzmics C (2013) Refined best-response correspondence and dynamics. Theor Econ 8:165–192

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Ben Porath E (1997) Rationality, Nash equilibrium and backwards induction in perfect information games. Rev Econ Stud 64:23–46

    Article  Google Scholar 

  • Bochnak J, Coste M, Roy M (1998) Real algebraic geometry. Springer, New York

    Book  Google Scholar 

  • Börgers T (1994) Weak dominance and approximate common knowledge. J Econ Theory 64:265–276

    Article  Google Scholar 

  • Brandenburger A, Friedenberg A, Keisler HJ (2008) Admissibility in games. Econometica 76:307–353

    Article  Google Scholar 

  • Crawford V, Sobel J (1982) Strategic information transmission. Econometrica 50:1431–1451

    Article  Google Scholar 

  • Dekel E, Fudenberg D (1990) Rational behavior with payoff uncertainty. J Econ Theory 52:243–267

    Article  Google Scholar 

  • Farrell J, Rabin M (1996) Cheap talk. J Econ Perspect 10(3):103–118

    Article  Google Scholar 

  • Fudenberg D, Tirole J (1991) Game theory. MIT Press, Cambridge

    Google Scholar 

  • Gilboa I, Matsui A (1991) Social stability and equilibrium. Econometrica 59:859–867

    Article  Google Scholar 

  • Gordon S (2006) Iteratively stable cheap talk equilibria (unpublished manuscript)

  • Guillemin V, Pollack A (1974) Differential topology. Prentice Hall, Englewood Cliffs

    Google Scholar 

  • Harsanyi J, Selten R (1988) A general theory of equilibrium selection in games. MIT Press, Cambridge

    Google Scholar 

  • Hendon E, Jacobson HJ, Sloth B (1996) Fictitious play in extensive form games. Games and Economic Behavior 15:177–202

    Article  Google Scholar 

  • Hillas J (1990) On the Definition of the Strategic Stability of Equilibria. Econometrica 58(6):1365–1390

    Article  Google Scholar 

  • Hofbauer J (1995) “Stability for the best response dynamics”, Unpublished manuscript

  • Jansen M, Jurg A, Borm P (1994) On strictly perfect sets. Games and Economic Behavior 6:400–415

    Article  Google Scholar 

  • Jansen M, Jurg P, Vermeulen D (2002) On the set of equilibria of a bimatrix game: A survey. In: Borm P, Peters H (eds) Chapters in Game Theory. Kluwer, Amsterdam, pp 121–142

    Google Scholar 

  • Kalai E, Samet D (1984) Persistent equilibria in strategic games. Int J Game Theory 13:129–144

    Article  Google Scholar 

  • Kohlberg E, Mertens J-F (1986) On the strategic stability of equilibria. Econometrica 54:1003–1037

    Article  Google Scholar 

  • Matsui A (1992) Best response dynamics and and socially stable strategies. J Econ Theory 57:343–362

    Article  Google Scholar 

  • Mertens J-F (1991) Stable equilibria: a reformulation. Part II. Discussion of the definition, and further results. Math Oper Res 16:694–753

    Article  Google Scholar 

  • Myerson RB (1978) Refinements of the Nash equilibrium concept. Int J Game Theory 7:73–80

    Article  Google Scholar 

  • Myerson RB (1991) Game theory: analysis of conflict. Harvard University Press, Cambridge

    Google Scholar 

  • Okada A (1983) Robustness of equilibrium points in strategic games. Working Paper B-137, Tokyo Center for Game Theory

  • Pearce DG (1984) Rationalizable strategic behavior and the problem of perfection. Econometrica 52:1029–1051

    Article  Google Scholar 

  • Ritzberger K (2002) Foundations of non-cooperative game theory. Oxford University Press, Oxford

    Google Scholar 

  • Rockafellar RT (1970) Convex analysis. Princeton University Press, Princeton

    Google Scholar 

  • Selten R (1975) Re-examination of the perfectness concept for equilibrium points in extensive games. Int J Game Theory 4:25–55

    Article  Google Scholar 

  • Swinkels JM (1993) Adjustment dynamics and rational play in games. Games Econ Behav 5:455–484

    Article  Google Scholar 

  • van Damme EEC (1991) Stability and perfection of Nash equilibria. Springer, Berlin, Heidelberg

    Book  Google Scholar 

  • van Damme EEC (2002) Strategic equilibrium. In: Aumann R, Hart S (eds) Handbook in game theory, vol 14, III edn. North Holland, Amsterdam

    Google Scholar 

  • Vermeulen AJ (1996) Stability in noncooperative game theory. Department of Mathematics, University of Nijmegen, PhD Thesis

  • Voorneveld M (2005) Persistent retracts and preparation. Games Econ Behav 51:228–232

    Article  Google Scholar 

Download references

Acknowledgments

We would like to thank Carlos Alos-Ferrer, Pierpaolo Battigalli, Eddie Dekel, Christian Ewerhart, Amanda Friedenberg, Drew Fudenberg, Klaus Ritzberger, Karl Schlag, Mark Voorneveld, and Jörgen Weibull, as well as editor Vijay Krishna and two anonymous referees for helpful comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christoph Kuzmics.

Appendices

Appendix

On the generic equivalence of best responses and refined best responses

This appendix provides a proof of Theorem 4, which is organized in a number of steps: We will first fix some notations for the mappings and various submanifolds to be considered. Step 1 argues that the embedding of the uncorrelated strategy combinations into the set of beliefs has nice differentiability properties. Step 2 invokes the transversality theorem (see Guillemin and Pollack 1974) to show that for generic payoff functions we obtain the needed transversality conditions.Footnote 15 Step 3 argues that we can restrict attention to completely mixed strategy combinations of the opponents. If the player is indifferent between several of his strategies against a given completely mixed strategy combination, step 4 shows how we can construct an arbitrarily nearby strategy combination, against which the player strictly prefers a given one among these strategies. Step 5 completes the argument.

For any finite set \(M\) let \(\mathbb {R}^{M}\) be the vector space of all mappings \(y:M \rightarrow \mathbb {R} \). The dimension of \(\mathbb {R}^{M}\) is the number of elements in \(M\).

Let \(q_{i}:\prod _{j\ne i}\mathbb {R}^{S_{j}} \rightarrow \mathbb {R}^{S_{-i}}\) be the mapping defined by

$$\begin{aligned} \left( q_{i}\left( x_{-i}\right) \right) \left( s_{-i}\right) :=\prod _{j\ne i}x_{j}\left( s_{j}\right) . \end{aligned}$$

The multilinear function \(q_{i}\) describes the first step in the computation mentioned above. While \(x_{-i}\in \prod _{j\ne i}\mathbb {R}^{S_{j}}\) denotes the usual strategy combinations of the opponents, we use \(\chi _{-i}\in \mathbb {R}^{S_{-i}}\) to describe a “correlated strategy of the opponents”, i.e., a belief over the set \(S_{-i}\) of pure strategy combinations of the opponents. \(q_{i}\) maps mixed strategy combinations to such beliefs.

Let \(L_{i}\) be the vector space of all linear mappings

$$\begin{aligned} v_{i}:\mathbb {R}^{S_{-i}}\rightarrow \mathbb {R}^{S_{i}}. \end{aligned}$$

If \(\chi _{-i}\in \mathbb {R}^{S_{-i}}\) is a belief of player \(i\) and \(s_{i}\in S_{i}\) a pure strategy player \(i\) chooses, then \(\left( v_{i}\left( \chi _{-i}\right) \right) \left( s_{i}\right) \) is the payoff player \(i\) expects with his strategy choice. In this context a vector \(z\in \mathbb {R}^{S_{i}}\) represents the various gains a player could make, not probabilities. The linear function \(v_{i}\) describes for every \(s_{i}\) the second step in the computation of the expected payoff. Any \(v_{i}\in L_{i}\) corresponds via \(u_{i}\left( s_{-i} ,s_{i}\right) =\left( v_{i}\left( s_{-i}\right) \right) \left( s_{i}\right) \) uniquely to a payoff function

$$\begin{aligned} u_{i}:S\rightarrow \mathbb {R} \end{aligned}$$

in the standard notation (and this relation is a homeomorphism).

For \(T_{i}\subseteq S_{i}\) set \(Z_{i}\left( T_{i}\right) =\{z\in \mathbb {R}^{S_{i}}\mid \forall s_{i},t_{i}\in T_{i}:z\left( s_{i}\right) =z\left( t_{i}\right) \}\). \(v_{i}\left( \chi _{-i}\right) \in Z_{i}\left( T_{i}\right) \) means that player \(i\) is indifferent between all his strategies in \(T_{i}\) when his belief is \(\chi _{-i}\). Let \(X_{j} :=\{x_{j}\in \mathbb {R}^{S_{j}}\mid \sum _{s_{j}\in S_{j}}x_{j}\left( s_{j}\right) =1\}\) be the affine space generated by player \(j^{\prime }s\) strategy simplex for \(j\ne i \) and let \(X_{-i}:=\prod _{j\ne i}X_{j}\).

For \(T_{j}\subseteq S_{j}\) \(\left( j\ne i\right) \) and \(T_{-i} :=\prod _{j\ne i}T_{j}\) set

$$\begin{aligned} X_{j}\left( T_{j}\right) :=\{x_{j}\in X_{j}\mid \forall s_{j}\notin T_{j}:x_{j}\left( s_{j}\right) =0\} \end{aligned}$$

and

$$\begin{aligned} X_{-i}\left( T_{-i}\right) :=\prod _{j\ne i}X_{j}\left( T_{j}\right) . \end{aligned}$$

The sets \(\Theta _{-i}\cap X_{-i}\left( T_{-i}\right) \) describe the various faces of the polyhedron \(\Theta _{-i}\). The strategies of player \(j\) with support in \(T_{j}\) have \(X_{j}\left( T_{j}\right) \) as their affine hull.

Step 1: For all \(T_{-i}\) the mapping \(q_{i} :X_{-i}\left( T_{-i}\right) \rightarrow \mathbb {R}^{S_{-i}} \setminus \{0\}\) is a diffeomorphism onto its image (in particular \(q_{i}\left( X_{-i}\left( T_{-i}\right) \right) \) is a closed submanifold of \(\mathbb {R}^{S_{-i}} \setminus \{0\}\)).

Proof

\(X_{-i}\left( T_{-i}\right) \) is a closed affine submanifold in \(\prod _{j\ne i}\left( \mathbb {R}^{S_{j}}\setminus \{0\}\right) \). It is straightforward to check that

$$\begin{aligned} q_{i|X_{-i}\left( T_{-i}\right) }:X_{-i}\left( T_{-i}\right) \rightarrow \mathbb {R}^{S_{-i}}\setminus \{0\} \end{aligned}$$

is well defined, is one-to-one, maps \(X_{-i}\left( T_{-i}\right) \) onto a closed set, and has a derivative \(dq_{i}|_{x_{-i}}\) with maximal rank everywhere.Footnote 16 \(\square \)

Step 2: Let \(Z\subseteq \mathbb {R}^{S_{i}}\) and \(X\subseteq \mathbb {R}^{S_{-i} }\setminus \{0\}\) be submanifolds. Then for almost every \(v_{i}\in L_{i}\) the mapping \(v_{i}|_{X}:X\rightarrow \mathbb {R}^{S_{-i} }\setminus \{0\}\) is transversal to \(Z\).

Proof

The family of linear maps \(L_{i}\) defines a mapping

$$\begin{aligned} V_{i}:L_{i}\times \mathbb {R}^{S_{-i}}&\rightarrow \mathbb {R}^{S_{i}}\end{aligned}$$
(1)
$$\begin{aligned} \left( v_{i},\chi _{-i}\right)&\mapsto v_{i}\left( \chi _{-i}\right) . \end{aligned}$$
(2)

The derivative of \(V_{i}\) at \(\left( v_{i},\chi _{-i}\right) \) can be computed as

$$\begin{aligned} dV_{i}|_{\left( v_{i},\chi _{-i}\right) }:T_{v_{i}}L_{i}\times T_{\chi _{-i}} \mathbb {R}^{S_{-i}} \cong L_{i} \times \mathbb {R}^{S_{-i}}&\rightarrow \mathbb {R}^{S_{i}}\end{aligned}$$
(3)
$$\begin{aligned} \left( \nu _{i},\xi _{-i}\right)&\mapsto \nu _{i} \left( \chi _{-i}\right) + v_{i}\left( \xi _{-i}\right) . \end{aligned}$$
(4)

If \(\chi _{-i} \ne 0\) we can find for every \(\zeta _{i} \in \mathbb {R}^{S_{i}}\) some \(\nu _{i}\in L_{i}\) with \(\nu _{i}\left( \chi _{-i}\right) =\zeta _{i}\). Then \(dV_{i}|_{\left( v_{i},\chi _{-i}\right) }\left( \nu _{i},0\right) =\zeta _{i}\).

Because for \(\chi _{-i} \in X\) the tangent space \(T_{\chi _{-i}} X \subseteq \mathbb {R}^{S_{-i}}\) contains the \(0\)-vector, \(dV_{i}|_{\left( v_{i},\chi _{-i}\right) }: T_{v_{i}}L_{i}\times T_{\chi _{-i} }X\rightarrow \mathbb {R}^{S_{i}}\) is surjective. Thus \(V_{i}:L_{i}\times X\rightarrow \mathbb {R}^{S_{i}}\) is transversal to \(Z\) and our claim follows from the transversality theorem.

By step 1 and step 2 almost every \(v_{i} \in L_{i}\) satisfies: \(\square \)

\(\otimes \) :

For all subsets \(T_j \subseteq S_j\) \(\left( 1 \le j \le n\right) \) the mapping \(\left( v_i \circ q_i\right) |_{X_{-i}\left( T_{-i}\right) }\) is transversal to \(Z_i\left( T_i\right) \).

For given \(v_{i}\) define \(Y\left( T_{i}\right) =\{x_{-i}\in X_{-i}\mid \left( v_{i}\circ q_{i}\right) \left( x_{-i}\right) \in Z\left( T_{i}\right) \}\). \(Y\left( T_{i}\right) \cap \Theta _{-i}\) is the set of strategy combinations of the opponents such that player \(i\) is indifferent between the strategies in \(T_{i}\) (i.e., they give the same payoff). If \(T_{i}\) is a set of best replies against \(x_{-i}\), then \(x_{-i}\in Y\left( T_{i}\right) \cap \Theta _{-i}\).

Step 3: Suppose \(v_{i}\) satisfies \(\otimes \). For \(T_{i}\subseteq S_{i}\) let \(x_{-i}\in Y\left( T_{i}\right) \cap \Theta _{-i}\) and let \(O_{-i}\) be a neighborhood of \(x_{-i}\). Then \(O_{-i}\cap Y\left( T_{i}\right) \) contains a point in the interior of \(\Theta _{-i}\).

Proof

Suppose \(x_{-i}\) is in the boundary of \(\Theta _{-i}\). For each \(j\ne i\) define \(T_{j}:=\{s_{j}\in S_{j}\mid x_{j}\left( s_{j}\right) \ne 0\}\). Thus \(x_{-i}\in X\left( T_{-i}\right) \cap \Theta _{-i}\). If \(T_{j}=S_{j}\), \(x_{j}\) is in the relative interior of \(\Theta _{j}\). By assumption \(T_{j}\ne S_{j}\) for at least one opponent \(j\). Fix \(j^{*}\ne i\) with \(T_{j*}\ne S_{j^{*}}\) and \(t_{j*}\notin T_{j*}\). Set \(\tilde{T}_{j}:=T_{j}\) for \(i\ne j\ne j^{*}\) and \(\tilde{T}_{j*}:=T_{j*}\cup \{t_{j*}\}\). We show that \(O_{-i}\cap Y\left( T_{i}\right) \) contains some \(y_{-i}\in \Theta _{-i}\cap X\left( \tilde{T}_{-i}\right) \) such that \(\tilde{T}_{j}=\{s_{j}\in S_{j}\mid y_{j}\left( s_{j}\right) \ne 0\}\) for all \(j\ne i\). In other words: \(y_{-i}\) is in the relative interior of the face \(\Theta _{-i}\cap X\left( \tilde{T}_{-i}\right) \). The claim then follows by induction.

The transversality conditions imply that the submanifolds \(X_{-i}\left( T_{-i}\right) \) and \(Y\left( T_{i}\right) \left( \tilde{T} _{-i}\right) \) meet transversally in \(X_{-i}\left( \tilde{T}_{-i}\right) \) (see Guillemin and Pollack 1974 , Exercise 1.6.7). Since \(X_{-i}\left( T_{-i}\right) \) has codimension 1 in \(X_{-i}\left( \tilde{T}_{-i}\right) \) it follows with arguments as in the next step that \(X_{-i}\left( \tilde{T}_{-i}\right) \cap Y\left( T_{i}\right) \cap \{y_{-i}\mid y_{j*}\left( t_{j*}\right) >0\}\cap O_{-i}\) intersects the relative interior of \(X_{-i}\left( \tilde{T}_{-i}\right) \cap \Theta _{-i}\). \(\square \)

Step 4: Suppose \(v_{i}\) satisfies \(\otimes \). For \(T_{i}\subseteq S_{i}\) with \(\#T_{i}\ge 2\) let \(x_{-i}\in Y\left( T_{i}\right) \) be in the interior of \(\Theta _{-i}\) and let \(O_{-i}\) be a neighborhood of \(x_{-i}\). Then we can find for every \(s_{i}\in T_{i}\) some \(y_{-i}\in O_{-i}\cap \Theta _{-i}\) such that

$$\begin{aligned} \left( v_{i}\circ q_{i}\right) \left( y_{-i}\right) \left( s_{i}\right) >\left( v_{i}\circ q_{i}\right) \left( y_{-i}\right) \left( t_{i}\right) \quad \text{ for } \text{ all } \ t_{i}\in T_{i}\setminus \{s_{i}\}. \end{aligned}$$

Proof

Because \(v_{i}\circ q_{i}:X_{-i} \rightarrow \mathbb {R}^{S_{i}}\) is transversal to both \(Z\left( T_{i}\right) \) and \(Z\left( T_{i}\setminus \{s_{i}\}\right) \) it follows that \(v_{i}\circ q_{i}:Y\left( T_{i}\setminus \{s_{i}\}\right) \rightarrow Z\left( T_{i}\setminus \{s_{i}\}\right) \) is transversal to \(Z\left( T_{i}\right) \). From this we can deduce the existence of a tangent vector \(\xi \in T_{x_{-i}}\left( Y\left( T_{i}\setminus \{s_{i}\}\right) \right) \) with \(d\lambda _{|x_{-i}}\left( \xi \right) =1\), where \(\lambda \) is the function

$$\begin{aligned} \lambda :Y_{i}\left( T_{i}\setminus \{s_{i}\}\right) \cap X_{-i}&\rightarrow \mathbb {R} \end{aligned}$$
(5)
$$\begin{aligned} y_{-i}&\rightarrow \left( v_{i}\circ q_{i}\right) \left( y_{-i}\right) \left( s_{i}\right) -\left( v_{i}\circ q_{i}\right) \left( y_{-i}\right) \left( t_{i}\right) \end{aligned}$$
(6)

defined for arbitrary but fixed \(t_{i}\in T_{i}\setminus \{s_{i}\}\). We can therefore select a differentiable curve

$$\begin{aligned} c:\left( -\epsilon ,\epsilon \right) \rightarrow Y_{i}\left( T_{i} \setminus \{s_{i}\}\right) \end{aligned}$$

with \(c\left( 0\right) =x_{-i}\) and \(\left( \lambda \circ c\right) ^{\prime }\left( 0\right) =1\). For sufficiently small \(0<\gamma <\epsilon \quad \)the vector \(y_{-i}:=c\left( \gamma \right) \) has the required properties.

Step 5: Suppose \(s_{i}\) is a pure best response against \(x_{-i}\). For every neighborhood \(O_{-i}\) of \(x_{-i}\) the continuity of the payoff function and the two steps above can be used to find \(y_{-i}\in O_{-i}\) such that \(s_{i}\in T_{i}\) is the unique best response against \(y_{-i}\). Shrinking the open sets we can find a sequence of such \(y_{-i}\)’s converging to \(x_{-i}\). Continuity yields an open set around each element in the sequence, where \(s_{i}\) is the unique best response. \(s_{i}\) is the unique best response on the union of these sets, which is again open. Thus \(s_{i}\) is a refined best response against \(x_{-i}\). \(\square \)

Refined best responses in two-player games

This appendix provides proofs of Theorems 1 and 2. In the case of two player games the payoff function is linear in the mixed strategy choice of the opponent. This allows the use of convex analysis (see Rockafellar 1970) to study the best-response correspondence of a player. The most direct consequence is the convexity of the region where a strategy is a best response. From this Theorem 1 follows immediately, the arguments are given after Lemma 1 below. More work is needed to obtain Theorem 2. We use conjugate functions and provide the proof after Lemma 2.

We will restrict attention to the best responses of player 1. Suppose player 2 has \(K\ge 2\) strategies \(s_{2}^{1},\cdots ,s_{2}^{K}\). It will be convenient to identify the mixed strategies \(x_{2}\in \Theta _{2}\) with the vectors

$$\begin{aligned} x_{2}=\left( x_{2}^{1},x_{2}^{2},\cdots ,x_{2}^{K-1}\right) \in \mathbb {R}^{K-1} \end{aligned}$$
(7)

for which \(x_{2}^{k}\ge 0\) for all \(1\le k\le K-1\) and \(x_{2}^{K} :=1-\sum _{k=1}^{K-1}x_{2}^{k}\le 0\). Notice that the zero vector corresponds to pure strategy \(s_{2}^{K}.\)

We define the function \(f:\mathbb {R}^{K-1}\rightarrow \mathbb {R}\) by

$$\begin{aligned} f\left( x_{2}\right) =\left\{ \begin{array} [c]{lcc} {\max }_{s_{1}\in S_{1}}u_{1}\left( s_{1},x_{2}\right) &{} \text{ for } &{} x_{2} \in \Theta _{2}\\ +\infty &{} \text{ else } &{} \end{array} \right. \end{aligned}$$
(8)

Because \(u_{1}\) is linear in \(x_{2}\), \(f\) is, in the terminology of Rockafellar (1970) a proper convex polyhedral function. A key idea explored in the following is that the strategies of player 1 that are refined best responses, correspond to the maximal compact faces of the epigraph

$$\begin{aligned} F=\left\{ \left( x_{2},\alpha \right) \in \mathbb {R}^{K-1}\times \mathbb {R}\, |\, f\left( x_{2}\right) \le \alpha \right\} \end{aligned}$$

of \(f\), which is a convex (but not compact) polyhedron. Duality theory allows us to identify these faces with the extreme points of the epigraph \(F^{*}\) of the conjugate function \(f^{*}\) of \(f\). This will be used in the proof of theorem Theorem 2.

Each strategy \(x_{1}\in \Theta _{1}\) defines an affine function \(a:\mathbb {R}^{K-1}\rightarrow \mathbb {R}\) by \(a\left( x_{2}\right) =u_{1}\left( x_{1},x_{2}\right) \), which, for all \(x_{2}\in \Theta _{2}\), satisfies the inequality \(a\left( x_{2}\right) \le f\left( x_{2}\right) \) and \(a\left( x_{2}\right) =f\left( x_{2}\right) \) holds if and only if \(x_{1}\in \beta _{1}\left( x_{2}\right) \).

For a strategy \(x_{1}\in \Theta _{1}\) we define the set

$$\begin{aligned} G\left( x_{1}\right) =\left\{ \left( x_{2},\alpha \right) \in \Theta _{2}\times \mathbb {R}\,|\,x_{1} \in \beta \left( x_{2}\right) \text{ and } \alpha =u_{1}\left( x_{1}, x_{2}\right) \right\} \end{aligned}$$
(9)

and the set \(H\left( x_{1}\right) =\left\{ x_{2}\in \Theta _{2} \,|\, x_{1}\in \beta \left( x_{2}\right) \right\} \), the projection of \(G\left( x_{1}\right) \) onto \(\Theta _{2}\). \(H\left( x_{1}\right) \) is the region where \(x_{1}\) is a best response.

Lemma 1

The region \(H\left( x_{1}\right) \) is a convex polyhedron.

Proof

\(G\left( x_{1}\right) \) is a face of the epigraph \(\left\{ \left( x_{2},\alpha \right) \in \Theta _{2}\times \mathbb {R}\,|\,f\left( x_{2}\right) \le \alpha \right\} \) of the function \(f\), which is a convex polyhedron. \(G\left( x_{1}\right) \) is hence a convex polyhedron. \(H\left( x_{1}\right) \) is the image of the convex polyhedron \(G\left( x_{1}\right) \) under the linear projection mapping and hence also a convex polyhedron. \(\square \)

Clearly \(x_{1}\) is not weakly inferior if and only if the convex polyhedron \(H\left( x_{1}\right) \) has non-empty interior \(H^{\circ }\left( x_{1}\right) \). Moreover, \(H\left( x_{1}\right) \) is the closure of \(H^{\circ }\left( x_{1}\right) \) if \(H^{\circ }\left( x_{1}\right) \) is not empty. Therefore, if \(x_{1}\) is not weakly inferior and a best response against \(x_{2},\) then \(x_{2}\) is in the closure of the open set \(H^{\circ }\left( x_{1}\right) \) and so \(x_{1}\) is a refined best response against \(x_{2}\). Given Definition 1 this implies immediately Theorem 1.

The remainder of this section aims at proving Theorem 2. We consider again the epigraph \(F\) of the map \(f\) defined above. We notice that \(F\) is a polyhedral convex set whose compact faces are precisely the sets \(G\left( x_{1}\right) \) with \(x_{1}\in \Theta _{1}\). The non-compact faces are of the form \(F\cap \left( \Theta _{1}^{\prime }\times \mathbb {R}\right) \), where \(\Theta _{1}^{\prime }\) is a face of \(\Theta _{1}\).

The conjugate function \(f^{*}:\mathbb {R}^{K-1} \rightarrow \mathbb {R}\) of \(f\) is defined by

$$\begin{aligned} f^{*}\left( x_{2}^{*}\right) =\sup _{x_{2}\in \mathbb {R}^{K-1}}\left\{ x_{2}^{*}\bullet x_{2}-f\left( x_{2}\right) \right\} =\max _{x_{2}\in \Theta _{2}}\left\{ x_{2}^{*}\bullet x_{2}-f\left( x_{2}\right) \right\} <\infty , \end{aligned}$$
(10)

where \(x_{2}^{*}\bullet x_{2}\) denotes the usual scalar product \(\sum _{_{k=1}}^{K-1}x_{2}^{*k}x_{2}^{k}\). As shown for any convex polyhedral function in Rockafellar (1970), the conjugate is again a convex polyhedral function and one has \(f^{**}\left( x_{2}\right) =f\left( x_{2}\right) \).

Any two strategies \(x_{1},x_{1}^{\prime }\in \Theta _{1}\) define the same affine function if and only if the two strategies are own-payoff equivalent. Without loss of generality we can thus identify \(\Theta _{1}\) up to own-payoff equivalence with a subset of the affine functions on \(\mathbb {R}^{K-1}\).

Any vector \(\left( x_{2}^{*},\alpha \right) \) with \(x_{2}^{*} \in \mathbb {R}^{K-1}\) and \(\alpha \in \mathbb {R}\) defines one and only one affine function on \(\mathbb {R}^{K-1}\) by

$$\begin{aligned} a\left( x_{2}\right) =-\alpha +\sum _{k=1}^{K-1}x_{2}^{*k}x_{2}^{k} \end{aligned}$$
(11)

We will identify affine functions with such vectors. For instance, \(e=\left( 1,\ldots ,1\right) \) corresponds to the function \(-x_{2}^{K+}=-1+\sum _{k=1}^{K-1}x_{2}^{k}\) that assigns \(0\) to the first \(K-1\) pure strategies and \(-1\) to the last pure strategy of player 2.

Let \(F^{*}\) be the epigraph of \(f^{*}\).

Lemma 2

\(F^{*}\) is a polyhedral convex set generated by extreme points \(x_{1}\) that are refined best responses in \(\Theta _{1}\) and the directions

$$\begin{aligned} -e_{k}=\left( -e_{k}^{1},\ldots ,-e_{k}^{K}\right) \in \mathbb {R}^{K} \text{ with } e_{k}^{l}=\left\{ \begin{array} [c]{ccc} -1 &{} \text{ for } &{} k=l\\ 0 &{} \text{ else } &{} \end{array} \right. \end{aligned}$$
(12)

for \(k=1,\ldots ,K-1\) and

$$\begin{aligned} e=\left( 1,\ldots ,1\right) \in \mathbb {R}^{K} \end{aligned}$$
(13)

Proof

By definition \(\left( x_{2}^{*},\alpha ^{*}\right) \in F^{*}\) if and only if \(\alpha ^{*}\ge x_{2}^{*}\bullet x_{2}-f^{*}\left( x_{2}\right) \) for all \(x_{2}\in \Theta _{2}\). \(v\in \mathbb {R}^{K}\) is a direction in \(F^{*}\) if and only if there exists \(\left( x_{2}^{*},\alpha ^{*}\right) \in F^{*}\) such that all vectors \(\left( x_{2}^{*},\alpha ^{*}\right) +\lambda v\) with \(\lambda \ge 0\) are in \(F^{*}\). We can write \(v=-\sum _{k=1}^{K-1}\rho _{k}e_{k}+\rho _{K}e\) with \(\rho _{1},\ldots ,\rho _{K}\in \mathbb {R}\) since \(-e_{1},\ldots ,-e_{K-1},e\) form a vector basis of \(\mathbb {R}^{K}\). We must show that \(v\) is a direction in \(F^{*}\) if and only if all \(\rho _{i}\) are non-negative. Suppose that \(v\) is a direction in \(F^{*}\). Let \(x_{2}=\left( 0,\ldots ,0\right) \in \Theta _{2}\). The condition that \(\left( x_{2}^{*},\alpha ^{*}\right) +\lambda v\in F^{*}\) for all \(\lambda \ge 0\) yields for this \(x_{2}\) that \(\alpha ^{*}+\lambda \rho _{K}\ge -f\left( x_{2}\right) \) holds for all \(\lambda \ge 0\). This can be true only if \(\rho _{K}\ge 0\). For \(e_{k}\in \Theta _{2}\) (\(1\le k\le K-1\)) we obtain similarly \(\alpha ^{*}+\lambda \rho _{K}\ge x_{2}^{*k}-\lambda \rho _{k}+\lambda \rho _{K}-f\left( e_{k}\right) \) for all \(\lambda \ge 0\), which can hold only if \(\rho _{k}\ge 0\). Thus only positive combinations of \(-e_{1},\ldots ,-e_{K-1},e\) can be directions in \(F^{*}\). For every combination \(v=-\sum _{k=1}^{K-1}\rho _{k}e_{k}+\rho _{K}e\) with \(\rho _{1} ,\ldots ,\rho _{K}\ge 0\), every \(\lambda \ge 0,\) every \(\left( x_{2}^{*},\alpha ^{*}\right) \in F^{*}\) and every \(x_{2}\in \Theta _{2}\) we have conversely

$$\begin{aligned} \alpha ^{*}+\lambda \rho _{K}\ge x_{2}^{*\prime }x_{2}-\sum _{k=1} ^{K-1}\lambda \rho _{k}x_{2}^{k}+\lambda \rho _{K}-f\left( x_{2}\right) \end{aligned}$$
(14)

which proves that \(v\) is a direction in \(F^{*}\).

We have characterized the directions of \(F^{*}\) and must now determine the extremal points of \(F^{*}\). Suppose \(\left( \hat{x}_{2}^{*},\hat{\alpha }^{*}\right) \) is an extremal point. Because \(F^{*}\) has only finitely many extremal points, these are exposed points by Straszewick’s theorem (Theorem 18.6 in Rockafellar 1970). Therefore we can find \(x_{2}\in \Theta _{2}\) such that the hyperplane \(\left\{ x_{2}^{*}\bullet x_{2}=f\left( x_{2}\right) \right\} \) is a supporting hyperplane that meets \(F^{*}\) only in \(\left( \hat{x}_{2}^{*},f^{*}\left( \hat{x} _{2}^{*}\right) \right) \). Because \(F^{*}\) has only finitely many extreme points and directions there exists an open neighborhood \(U\) of \(x_{2} \) in \(\Theta _{2}\) for which the hyperplanes \(\left\{ x_{2}^{*}\bullet y_{2}=f\left( y_{2}\right) \right\} \) are for all \(y_{2}\in U\) supporting hyperplanes that intersect \(F^{*}\) only in \(\left( \hat{x}_{2}^{*},f^{*}\left( \hat{x}_{2}^{*}\right) \right) \). This implies that the graph of the affine function \(\left( \hat{x}_{2}^{*},f^{*}\left( \hat{x}_{2}^{*}\right) \right) \) intersects \(F\) in a \(K-1\) dimensional face. It is therefore identical to an affine function defined by a strategy \(x_{1}\) in \(\Theta _{1}\) for which \(H\left( x_{1}\right) \) is full dimensional. Given our identification, \(\left( \hat{x}_{2}^{*},f^{*}\left( \hat{x}_{2}^{*}\right) \right) \) is consequently a not weakly inferior strategy in \(\Theta _{1}\), which was to be shown. \(\square \)

Proof of Theorem 2

The lemma implies that all extreme points and hence all the points in the compact faces of \(F^{*}\) are in \(\Theta _{1}\).

However, no points on the compact faces of \(F^{*}\) apart from the extreme points are not weakly inferior strategies. To see this, notice that a proper mixture \(x_{1}=\sum _{l=1}^{L}\rho _{l}x_{1k}\) (\(L>2,\rho _{l}>0\), \(\sum _{l=1}^{L}\rho _{l}=1\)) of non-equivalent not weakly inferior strategies in \(\Theta _{1}\) is weakly inferior. Otherwise there would be an open set in \(\Theta _{2}\) on which \(x_{1}\) and hence all strategies \(x_{1k}\) were best responses. They would yield identical payoffs on an open set and were hence (by Kalai and Samet Kalai and Samet 1984, Lemma 4) all own-payoff equivalent, contradicting the assumption. Per construction such a mixture is own-payoff equivalent to a proper mixture of strategies that are pairwise not own-payoff equivalent.

It remains to consider strategies in \(\Theta _{1}\) that are not on a compact face of \(F^{*}\). Such a strategy can be written as \(x_{1}^{\prime } =x_{1}-\sum _{k=1}^{K}\rho _{k}e_{k}+\rho _{K}e\) where \(x_{1}\) is on one of the compact faces of \(F^{*}\) and, hence, in \(\Theta _{1}\), and the \(\rho _{k}\) are all non-negative and at least some of them are strictly positive. We obtain

$$\begin{aligned} u_{1}\left( x_{1}^{\prime },x_{2}\right) =u_{1}\left( x_{1},x_{2}\right) -\sum _{k=1}^{K-1}\rho _{k}x_{2}^{k}-\rho _{K}\left( 1-\sum _{k=1}^{K-1}x_{2} ^{k}\right) \le u_{1}\left( x_{1},x_{2}\right) , \end{aligned}$$
(15)

where this inequality holds as a strict one for the \(k\)-th pure strategy of player 2 whenever \(\rho _{k}>0\). Thus \(x_{1}^{\prime }\) is weakly dominated. It is a weakly inferior strategy because it is a best response only on a proper face of \(\Theta _{1}\) (see Pearce 1984).

In summary, the only robust strategies in \(\Theta _{1}\) are the extreme points of \(F^{*}\). All other strategies are proper mixtures of not own-payoff equivalent not weakly inferior strategies or are weakly dominated and therefore weakly inferior. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Balkenborg, D., Hofbauer, J. & Kuzmics, C. The refined best-response correspondence in normal form games. Int J Game Theory 44, 165–193 (2015). https://doi.org/10.1007/s00182-014-0424-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00182-014-0424-z

Keywords

JEL codes

Navigation