Skip to main content
Log in

The algebraic geometry of perfect and sequential equilibrium: an extension

  • Research Article
  • Published:
Economic Theory Aims and scope Submit manuscript

Abstract

We extend the generic equivalence result of Blume and Zame (Econometrica 62:783–794, 1994) to a broader context of perfectly and sequentially rational strategic behavior (including equilibrium and nonequilibrium behavior) through a unifying solution concept of “mutually acceptable course of action” (MACA) proposed by Greenberg et al. (Econ Theory 40:91–112, 2009. https://doi.org/10.1007/s00199-008-0349-5). As a by-product, we show, in the affirmative, Dekel et al.’s (J Econ Theory 89:165–185, 1999) conjecture on the generic equivalence between the sequential and perfect versions of rationalizable self-confirming equilibrium.

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. By the Tarski–Seidenberg theorem, many equilibrium solution concepts can be defined by a first-order formula and thus, the equilibrium correspondences are semi-algebraic (cf. BZ94, Theorems 1 and 2). A semi-algebraic set has the following structure: Each semi-algebraic set has only a finite number of open connected components and has a well-defined dimension (see, e.g., Hardt 1980; Bochnak et al. 1987).

  2. The notion of extensive-form convex hull is designed to overcome the notorious problem of imperfection under subjective uncertainty over (behavioral) strategies in extensive-form games. Two important features of making use of the notion of the extensive-form convex hull are (i) it eliminates weakly dominated strategies in simultaneous-move games, and (ii) it yields the backward-induction strategy profile in a “generic” perfect-information game.

  3. Dekel et al. (1999, Footnote 4) made a claim without proof: “There are two closely related notions of optimality at off-path information sets that we consider: best replies to the limit of a sequence of trembles, namely sequential rationality, as in Kreps and Wilson (1982), and best replies to the sequence itself, as in Selten’s notion (1975) of trembling-hand perfection. We expect that, as in the relationship between sequential and perfect equilibrium, the difference is only in nongeneric games see Kreps and Wilson (1982) and Blume and Zame (1994) but verifying this takes us too far afield.”

  4. That is, different players are not required to hold the same beliefs on how players “tremble”. Fudenberg and Tirole (1991, p. 341) pointed out, “Why should all players have the same theory to explain deviations that, after all, are either probability-0 events or very unlikely, depending on one’s methodological point of view? The standard defense is that this requirement is in the spirit of equilibrium analysis, since equilibrium supposes that all players have common beliefs about the others’ strategies. Although this restriction is usually imposed, we are not sure that we find it convincing.”

  5. See also Aumann (1985) and van Damme (1998) for Robert Aumann’s insightful and thoughtful discussions on the importance of “relationships” in scientific research.

  6. More precisely, a statement is “generically” true if it is false only for a lower-dimensional subset of the payoff vector space. Besides full dimension we used here, there are other notions of “genericity” such as open and dense, residual, meager complement, almost surely, and almost everywhere.

  7. This example shows that there is no “generic” equivalence between Myerson’s (1978) proper equilibrium and perfect equilibrium: For “generic” payoffs \(u>1\), \(\left( S,r\right)\) is a perfect equilibrium but not a proper equilibrium. van Damme (1992, Theorem 2.6.1) presented an “almost all” theorem: In “almost all” normal form games, Nash equilibria are “regular” equilibria (hence proper equilibria). Nevertheless, as van Damme (1992, p. 45) pointed out, the analysis “is of limited value for the study of extensive-form games as any nontrivial such game gives rise to a nongeneric normal form.”

  8. Since the formal description of an extensive form is by now standard [see, for instance, Kreps and Wilson (1982) and Osborne and Rubinstein (1994)], we here include the necessary notation only. We note our approach can be extended to games with nature moves.

  9. A set \(X\subseteq {\mathbb {R}} ^{n}\) is semi-algebraic if it is the finite union of sets of the form \(\{x\in {\mathbb {R}}^{n}:f_{1}(x)=0,\ldots ,f_{k}(x)=0\) and \(g_{1}(x)>0,\ldots ,g_{m}(x)>0\}\), where the \(f_{i}\) and \(g_{j}\) are polynomials with real coefficients. A correspondence is semi-algebraic if and only if its graph is a semi-algebraic set.

  10. Player i’s expected payoff conditional on h is defined as \(v_{i}(y_{i}^{\prime },\left( y,\mu \right) ,u_{i}|h)=\Sigma _{z\in Z}u_{i}(z)\Pr \{z|\left( y_{i}^{\prime },\quad y_{-i}\right) ,\mu ,h\}\), where \(\Pr \{z|\left( y_{i}^{\prime },\quad y_{-i}\right) ,\mu ,h\}\) is the probability that z is reached conditionally on h under \(\left( y_{i}^{\prime },\quad y_{-i}\right)\) and \(\mu\).

  11. To relate to Selten’s (1975) perfectness, Kreps and Wilson (1982, Proposition 6) provided a useful characterization of sequential equilibrium in terms of “payoff perturbations” ; they relaxed Selten’s criterion by allowing some (vanishingly) small uncertainty on the part of players’ payoffs [cf. also Halpern (2009)].

  12. The notion is also related to Rubinstein and Wolinsky’s (1994) notion of a “rationalizable conjectural equilibrium” (RCE).

  13. It is worth noting that in the generic set \(U\backslash U^{0}\) (Corollary 3), there are games with strictly/weakly dominated strategies, so that the statement of Corollary 3 is not vacuously true. The major reason is that if a strategy is strictly dominated at a payoff vector \(u\in U\), then this is also true in a neighborhood of u; hence, strictly dominated strategies exist in a full-dimensional payoff subset (excluding the lower-dimensional set \(U^{0}\)). We thank an anonymous referee for drawing our attention to this point.

  14. This example shows that the notions of perfect Bayesian equilibrium and sequential equilibrium are generically distinct, because \((E_{1},C_{2},R_{3})\) is a perfect Bayesian equilibrium but not a sequential equilibrium.

  15. By allowing for distinct trembling sequences for different players, Aryal and Stauber (2014) introduced the notions of trembling-hand perfect equilibrium and robust sequential equilibrium in extensive games with ambiguity averse players.

  16. For any \(y_{i}^{\prime }\in {\mathbb {Y}}_{i}\), we define \(v_{i}\left( \left( y_{i}^{\prime },\quad x_{-i}^{t}\right) ,u_{i}^{t}|h\right) \equiv v_{i}\left( y_{i}^{\prime },\left( x_{-i}^{t},\mu _{i}^{t}\right) ,u_{i}^{t}|h\right)\).

  17. Chakrabarti and Topolyan (2016) adopted a similar backward approach to show the existence of sequential equilibrium. They backwardly constructed perturbations on strategies, while here we backwardly construct perturbations on payoffs.

References

  • Aryal, G., Stauber, R.: Trembles in extensive games with ambiguity averse players. Econ. Theory 57, 1–40 (2014). https://doi.org/10.1007/s00199-014-0828-9

    Article  Google Scholar 

  • Asheim, G.B., Perea, A.: Sequential and quasi-perfect rationalizability in extensive games. Games Econ. Behav. 53, 15–42 (2005)

    Article  Google Scholar 

  • Aumann, R.J.: What is game theory trying to accomplish? In: Arrow, K., Honkapohja, S. (eds.) Frontiers of Economics, pp. 28–76. Basil Blackwell, Oxford (1985)

    Google Scholar 

  • Battigalli, P.: On rationalizability in extensive form games. J. Econ. Theory 74, 40–61 (1997)

    Article  Google Scholar 

  • Battigalli, P., Cerreia-Vioglio, S., Maccheroni, F., Marinacci, M.: Self-confirming equilibrium and model uncertainty. Am. Econ. Rev. 105, 646–77 (2015)

    Article  Google Scholar 

  • Bernheim, B.D.: Rationalizable strategic behavior. Econometrica 52, 1007–1028 (1984)

    Article  Google Scholar 

  • Blume, L.E., Zame, W.R.: The algebraic geometry of competitive equilibrium. In: Neuefeind, W., Riezman, R.G. (eds.) Essays in General Equilibrium and International Trade: In Memoriam Trout Rader, pp. 53–66. Springer, Berlin (1993)

    Google Scholar 

  • Blume, L.E., Zame, W.R.: The algebraic geometry of perfect and sequential equilibrium. Econometrica 62, 783–794 (1994)

    Article  Google Scholar 

  • Chakrabarti, S.K., Topolyan, I.: An extensive form-based proof of the existence of sequential equilibrium. Econ. Theory Bull. 4, 355–365 (2016)

    Article  Google Scholar 

  • Bochnak, J., Coste, M., Roy, M.-F.: Geometie Algebrique Reelle, Ergebnisse der Mathematik und ihrer Grenzgebiete. Springer, Berlin (1987)

    Google Scholar 

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

    Article  Google Scholar 

  • Brandenburger, A., Friedenberg, A.: The relationship between rationality on the matrix and the tree. NYU Working paper (2003)

  • Demichelis, S., Ritzberger, K.: From evolutionary to strategic stability. J. Econ. Theory 113, 51–75 (2003)

    Article  Google Scholar 

  • Dekel, E., Fudenberg, D., Levine, D.K.: Payoff information and self-confirming equilibrium. J. Econ. Theory 89, 165–185 (1999)

    Article  Google Scholar 

  • Dekel, E., Fudenberg, D., Levine, D.K.: Subjective uncertainty over behavior strategies: a correction. J. Econ. Theory 104, 473–478 (2002)

    Article  Google Scholar 

  • Esponda, I.: Rationalizable conjectural equilibrium: a framework for robust predictions. Theor. Econ. 8, 467–501 (2013)

    Article  Google Scholar 

  • Fudenberg, D., Kamada, Y.: Rationalizable partition-confirmed equilibrium. Theor. Econ. 3, 775–806 (2015)

    Article  Google Scholar 

  • Fudenberg, D., Levine, D.K.: Self-confirming equilibrium. Econometrica 61, 523–545 (1993)

    Article  Google Scholar 

  • Fudenberg, D., Levine, D.K.: Superstition and rational learning. Am. Econ. Rev. 96, 630–651 (2006)

    Article  Google Scholar 

  • Fudenberg, D., Levine, D.K.: Self-confirming equilibrium and the Lucas critique. J. Econ. Theory 144, 2354–2371 (2009)

    Article  Google Scholar 

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

    Google Scholar 

  • Govindan, S., McLennan, A.: On the generic finiteness of equilibrium outcome distributions in game forms. Econometrica 69, 455–471 (2001)

    Article  Google Scholar 

  • Govindan, S., Wilson, R.: Direct proofs of generic finiteness of Nash equilibrium outcomes. Econometrica 69, 765–769 (2001)

    Article  Google Scholar 

  • Govindan, S., Wilson, R.: Sufficient conditions for stable equilibria. Theor. Econ. 1, 167–206 (2006)

    Google Scholar 

  • Govindan, S., Wilson, R.: Axiomatic equilibrium selection for generic two-player games. Econometrica 80, 1639–1699 (2012)

    Article  Google Scholar 

  • Greenberg, J., Gupta, S., Luo, X.: Mutually acceptable courses of action. Econ. Theory 40, 91–112 (2009). https://doi.org/10.1007/s00199-008-0349-5

    Article  Google Scholar 

  • Haller, H., Lagunoff, R.: Genericity and Markovian behavior in stochastic games. Econometrica 80, 1639–1699 (2012)

    Article  Google Scholar 

  • Halpern, J.Y.: A nonstandard characterization of sequential equilibrium, perfect equilibrium, and proper equilibrium. Int. J. Game Theory 38, 37–49 (2009)

    Article  Google Scholar 

  • Hardt, R.M.: Semi-algebraic local-triviality in semi-algebraic mappings. Am. J. Math. 102, 291–302 (1980)

    Article  Google Scholar 

  • Herings, P.J.J., Vannetelbosch, V.J.: Refinements of rationalizability for normal-form games. Int. J. Game Theory 28, 53–68 (1999)

    Article  Google Scholar 

  • Hillas, J., Kohlberg, E.: Foundations of strategic equilibrium. In: Aumann, R.J., Hart, S. (eds.) Handbook of Game Theory with Economic Applications, vol. 3, pp. 1597–1663. North-Holland, Amsterdam (2002)

    Google Scholar 

  • Kohlberg, E., Mertens, J.-F.: On the strategic stability of equilibria. Econometrica 54, 1003–1039 (1986)

    Article  Google Scholar 

  • Kohlberg, E., Reny, P.J.: Independence on relative probability spaces and consistent assessments in game trees. J. Econ. Theory 75, 280–313 (1997)

    Article  Google Scholar 

  • Kreps, D., Wilson, R.: Sequential equilibria. Econometrica 50, 1003–1038 (1982)

    Article  Google Scholar 

  • Kubler, K., Schmedders, K.: Competitive equilibria in semi-algebraic economies. J. Econ. Theory 145, 301–330 (2010)

    Article  Google Scholar 

  • Luo, X., Wang, B.: An epistemic characterization of MACA. NUS Working paper (2019)

  • McKelvey, R.D., McLennan, A.: Computation of equilibria in finite games. In: Amman, H.M., Kendrick, D.A., Rust, J. (eds.) Handbook of Computational Economics, vol. 1, pp. 87–142. North-Holland, Amsterdam (1996)

    Google Scholar 

  • Myerson, R.B.: Refinements of the Nash equilibrium concept. Int. J. Game Theory 15, 133–154 (1978)

    Article  Google Scholar 

  • Osborne, M.J., Rubinstein, A.: A Course in Game Theory. MIT Press, Cambridge (1994)

    Google Scholar 

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

    Article  Google Scholar 

  • Perea, A.: A note on the one-deviation property in extensive form games. Games Econ. Behav. 40, 322–338 (2002)

    Article  Google Scholar 

  • Perea, A.: Why forward induction leads to the backward induction outcome: a new proof for Battigalli’s theorem. Games Econ. Behav. 110, 120–138 (2018)

    Article  Google Scholar 

  • Perea, A., Jansen, M., Peters, H.: Characterization of consistent assessments in extensive form games. Games Econ. Behav. 21, 238–252 (1997)

    Article  Google Scholar 

  • Pimienta, C., Litan, C.: Conditions for equivalence between sequentiality and subgame perfection. Econ. Theory 35, 539–553 (2008)

    Article  Google Scholar 

  • Pimienta, C., Shen, J.: On the equivalence between (quasi-) perfect and sequential equilibria. Int. J. Game Theory 43, 395–402 (2014)

    Article  Google Scholar 

  • Ritzberger, K.: The theory of normal form games from the differentiable viewpoint. Int. J. Game Theory 23, 207–236 (1994)

    Article  Google Scholar 

  • Rubinstein, A., Wolinsky, A.: Rationalizable conjectural equilibrium: between Nash and rationalizability. Games Econ. Behav. 6, 299–311 (1994)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Shimoji, M.: On the equivalence of weak dominance and sequential best response. Games Econ. Behav. 48, 385–402 (2004)

    Article  Google Scholar 

  • van Damme, E.: Stability and Perfection of Nash Equilibria. Springer, Berlin (1992)

    Google Scholar 

  • van Damme, E.: On the state of the art in game theory: an interview with Robert Aumann. Games Econ. Behav. 24, 181–210 (1998)

    Article  Google Scholar 

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

    Article  Google Scholar 

Download references

Acknowledgements

We are grateful to the Editor, an Associate Editor, and two anonymous referees for valuable comments. We thank Yi-Chun Chen, Shravan Luckraz, Yongchuan Qiao, Yeneng Sun, and Satoru Takahashi for helpful comments and discussions. An earlier version of the paper was presented at the Asian Meeting of the Econometric Society, Singapore; and the SAET Conference in Faro, Portugal. Financial support from the National University of Singapore is gratefully acknowledged. The usual disclaimer applies.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yang Sun.

Additional information

Publisher's Note

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

Appendices

Appendix 1: Definitions

Consider an extensive-form game \(\Gamma \left( u\right)\) with perfect recall.

Definition 3

  1. (i)

    [Selten (1975); Osborne and Rubinstein (1994, Definition 251.1)] A strategy profile \(y\in {\mathbb {Y}}\) is a perfectly equilibrium if there exists a trembling sequence \(y^{t}\rightsquigarrow\) y such that, for all \(t\ge 0\), \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}}_{h}}v_{i}\left( \left( y_{h}^{\prime }, \quad y_{-h}^{t}\right) ,u_{i}\right)\) for all \(i\in N\) and \(h\in H_{i}\).

  2. (ii)

    [Kreps and Wilson (1982)] A strategy profile \(y\in {\mathbb {Y}}\) is a sequential equilibrium if there exists a consistent assessment \(\left( y,\mu \right)\) such that \(y_{i}\in \arg \max _{y_{i}^{\prime }\in {\mathbb {Y}}_{i}}v_{i}\left( y_{i}^{\prime },\left( y,\mu \right) ,u_{i}|h\right)\) for all \(i\in N\) and \(h\in H_{i}\).

  3. (iii)

    A strategy profile \(y\in {\mathbb {Y}}\) is a “weakly” perfect equilibrium if for each player \(i\in N\), there exists a trembling sequence \(y^{t}\rightsquigarrow\) y such that, for all \(t\ge 0\), \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}} _{h}}v_{i}\left( \left( y_{h}^{\prime },\quad y_{-h}^{t}\right) ,u_{i}\right)\) for all \(h\in H_{i}\).

  4. (iv)

    A strategy profile \(y\in {\mathbb {Y}}\) is a “weakly” sequential equilibrium if for each player \(i\in N\), there exists a consistent assessment \(\left( y,\mu \right)\) such that \(y_{i}\in \arg \max _{y_{i}^{\prime }\in {\mathbb {Y}}_{i}}v_{i}\left( y_{i}^{\prime },\left( y,\mu \right) ,u_{i}|h\right)\) for all information set \(h\in H_{i}\).

In the notions of perfect equilibrium and sequential equilibrium, all the players are required to have a (common) trembling sequence. In contrast, in the notions of “weakly” perfect equilibrium and “weakly” sequential equilibrium, different players are allowed to adopt distinct trembling sequences.Footnote 15

Definition 4

[Greenberg et al. (2009) and Dekel et al. (1999, 2002)] A product set \(Y=\Pi _{i\in N}Y_{i}\subseteq {\mathbb {Y}}\) is a perfectly (or sequentially) rationalizable set if for each player \(i\in N\), every \(y_{i}\in Y_{i}\) is perfect (or sequential) best response with respect to \(\left( \Pi _{j\ne i}co^{e}(Y_{j}),u_{i}\right)\). An element in a perfectly (or sequentially) rationalizable set is said to be a perfectly (or sequentially) rationalizable strategy profile.

That is, Y is a perfectly rationalizable set if, for each player i and \(y_{i}\in Y_{i}\), we can find a strategy profile \(y_{-i}\in \Pi _{j\ne i}co^{e}(Y_{j})\) and a trembling sequence which converges to \(\left( y_{i},y_{-i}\right)\) such that \(y_{i}\) is a best response along the trembling sequence. In simultaneous-move games, the perfectly rationalizable set is related to Herings and Vannetelbosch’s (1999) definition of “weakly perfect rationalizability.” Similarly, Y is a sequentially rationalizable set if for each player i and \(y_{i}\in Y_{i}\), we can find a consistent assessment \(\left( \left( y_{i},y_{-i}\right) ,\mu \right)\) where \(y_{-i}\in \Pi _{j\ne i}co^{e}(Y_{j})\) such that \(y_{i}\) is a sequential best response. The sequentially rationalizable set is associated with Dekel et al.’s (1999, 2002) notion of sequential rationalizability.

Definition 5

(Dekel et al. 1999, 2002). A strategy profile \(\widehat{y}\in {\mathbb {Y}}\) is a sequentially rationalizable self-confirming equilibrium (SRSCE) if there exists \(Y=\Pi _{i\in N}Y_{i}\subseteq {\mathbb {Y}}\), such that for each \(i\in N\) and \(y_{i}\in Y_{i}\), we can find a consistent assessment \(\left( y,\mu \right)\), with the restrictions that (i) \(y_{-i}\in \Pi _{j\ne i}co^{e}(Y_{j})\), (ii) y and \(\widehat{y}\) yield a same distribution over terminal nodes, and (iii) \(y_{i}\) is a sequential best response under \(\left( y,\mu \right)\).

Appendix 2: Proofs

For each player \(i\in N\) and \(Y_{-i}\subseteq {\mathbb {Y}}_{-i}\), let

$$\begin{aligned} PB_{i}(Y_{-i},u_{i})\equiv & {} \left\{ y_{i}\in {\mathbb {Y}}_{i}:\quad y_{i} \text { is a perfect best response with respect to }\left( Y_{-i},u_{i}\right) \right\} ; \\ SB_{i}(Y_{-i},u_{i})\equiv & {} \left\{ y_{i}\in {\mathbb {Y}}_{i}:\quad y_{i} \text { is a sequential best response with respect to }\left( Y_{-i},u_{i}\right) \right\} . \end{aligned}$$

Let \(cl({\mathcal {R}}_{i})\) and \(vcl_{U_{i}}({\mathcal {R}}_{i})\) denote the closure of the perfectly rational state set \({\mathcal {R}} _{i}\) and vertical closure of the perfectly rational state set \({\mathcal {R}}_{i}\) (on \(U_{i}\)), respectively; that is,

$$\begin{aligned} cl({\mathcal {R}}_{i})\equiv & {} \left\{ \left( u_{i},x,y_{i}\right) \in U_{i}\times {\mathbb {Y}}\times {\mathbb {Y}}_{i}:\exists \left\{ \left( u_{i}^{t},x^{t},y_{i}^{t}\right) \right\} _{t=1}^{\infty }\in {\mathcal {R}}_{i} \text { s.t. }\left( u_{i}^{t},x^{t},y_{i}^{t}\right) \rightarrow \left( u_{i},x,y_{i}\right) \right\} ; \\ vcl_{U_{i}}({\mathcal {R}}_{i})\equiv & {} \left\{ \left( u_{i},x,y_{i}\right) \in U_{i}\times {\mathbb {Y}}\times {\mathbb {Y}}_{i}:\exists \left\{ \left( u_{i},x^{t},y_{i}^{t}\right) \right\} _{t=1}^{\infty }\in {\mathcal {R}}_{i} \text { s.t. }\left( u_{i},x^{t},y_{i}^{t}\right) \rightarrow \left( u_{i},x,y_{i}\right) \right\} . \end{aligned}$$

Let

$$\begin{aligned} Q\left( Y_{-i}\right) \equiv \left\{ \left( u_{i},x,y_{i}\right) \in U_{i}\times {\mathbb {Y}}\times {\mathbb {Y}}_{i}:x_{i}=y_{i}\text { and } x_{-i}\in Y_{-i}\right\} \end{aligned}$$

denote the set of “consistent-belief” states at which player i’s belief about the opponent players’ behavior lies in the given set \(Y_{-i}\) and the belief about his own behavior is consistent with his strategy.

To prove Theorem 1, we need the following four lemmas:

Lemma 1

In an extensive form \(\Gamma\) with perfect recall, for each player \(i\in N\) and \(Y_{-i}\subseteq {\mathbb {Y}}_{-i}\) (a) \(y_{i}\in PB_{i}(Y_{-i},u_{i})\Leftrightarrow \exists \left( u_{i},x,y_{i}\right) \in vcl_{U_{i}}({\mathcal {R}}_{i})\cap Q\left( Y_{-i}\right)\); (b) \(y_{i}\in SB_{i}(Y_{-i},u_{i})\Leftrightarrow \exists \left( u_{i},x,y_{i}\right) \in cl({\mathcal {R}}_{i})\cap Q\left( Y_{-i}\right)\).

Proof

(a) Suppose \(y_{i}\in PB_{i}(Y_{-i},u_{i})\). Then, there exist \(y_{i}^{t}\rightsquigarrow\) \(y_{i}\) and \(x_{-i}^{t}\rightsquigarrow x_{-i}\) \(\in Y_{-i}\) such that \(\left( u_{i},\left( y_{i}^{t},x_{-i}^{t}\right) ,y_{i}\right) \in {\mathcal {R}} _{i}\) for all t. Since \(\left( u_{i},\left( y_{i}^{t},x_{-i}^{t}\right) ,y_{i}\right) \rightarrow \left( u_{i},\left( y_{i},x_{-i}\right) ,y_{i}\right)\), \(\left( u_{i},x,y_{i}\right) \in vcl_{U_{i}}({\mathcal {R}} _{i})\cap Q\left( Y_{-i}\right)\). Conversely, suppose \(\left( u_{i},x,y_{i}\right) \in vcl_{U_{i}}\left( {\mathcal {R}}_{i}\right)\), \(x_{-i}\in Y_{-i}\) and \(x_{i}=y_{i}\). Then, there exists a sequence \(\left( u_{i},x^{t},y_{i}^{t}\right) \in {\mathcal {R}}_{i}\) converging to \(\left( u_{i},x,y_{i}\right)\). Since \(\Gamma\) is finite and \(y_{i}^{t}\rightarrow y_{i}\), there is a sufficiently large T such that, for all \(t\ge T\) and \(h\in H_{i}\), \(a_{h}\in\)support \(\left( y_{h}\right)\) implies \(a_{h}\in\) support \(\left( y_{h}^{t}\right)\) and \(a_{h}\in \arg \max _{a_{h}^{\prime }\in A_{h}}v_{i}\left( \left( a_{h}^{\prime },\quad x_{-h}^{t}\right) ,u_{i}\right)\). Therefore, \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y }}_{h}}v_{i}\left( \left( y_{h}^{\prime },\quad x_{-h}^{t}\right) ,u_{i}\right)\) for all \(t\ge T\) and all \(h\in H_{i}\). That is, \(y_{i}\in PB_{i}(Y_{-i},u_{i})\).

(b) It suffices to show \(y_{i}\in SB_{i}(Y_{-i},u_{i})\) iff there exist \(u_{i}^{t}\rightarrow u_{i}\)and \(\left( y_{i}^{t},x_{-i}^{t}\right) \rightsquigarrow\) \(\left( y_{i},x_{-i}\right)\) such that \(x_{-i}\in Y_{-i}\), \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}}_{h}}v_{i}\left( \left( y_{h}^{\prime },\quad x_{-h}^{t}\right) ,u_{i}^{t}\right)\) for all t and all \(h\in H_{i}\).

\(\Leftarrow\)”: Let \(x^{t}=\left( y_{i}^{t},x_{-i}^{t}\right) \rightsquigarrow\) \(\left( y_{i},x_{-i}\right) =x\) such that \(x_{-i}\in Y_{-i}\). Without loss of generality, assume \(\left( x^{t},\mu _{i}^{t}\right) \rightarrow \left( x,\mu _{i}\right)\), where \(\mu _{i}^{t}\) is derived from \(x^{t}\) using Bayes’ rule. Suppose that there exist \(u_{i}^{t}\rightarrow u_{i}\) such that \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}}_{h}}v_{i}\left( \left( y_{h}^{\prime },\quad x_{-h}^{t}\right) ,u_{i}^{t}\right)\) for all t and all \(h\in H_{i}\). Then, for all \(h\in H_{i}\) and all t, \(v_{i}\left( (y_{h},\quad x_{-h}^{t}),u_{i}^{t}|h\right) \ge v_{i}\left( (y_{h}^{\prime },\quad x_{-h}^{t}),u_{i}^{t}|h\right)\) \(\forall y_{h}^{\prime }\in {\mathbb {Y}}_{h}\).Footnote 16 Since \(v_{i}(\left( y_{h},\cdot \right) ,\cdot |h)\) is continuous, \(v_{i}\left( \left( y_{h},\quad x_{-h}\right) ,u_{i}|h\right) \ge v_{i}\left( \left( y_{h}^{\prime },\quad x_{-h}\right) ,u_{i}|h\right)\). Since \(\Gamma\) is perfect recall, every sequential optimal strategy satisfies the one deviation property (see, Perea 2002), for all \(h\in H_{i}\), \(v_{i}\left( \left( y_{i},\quad x_{-i}\right) ,u_{i}|h\right) \ge v_{i}\left( \left( y_{i}^{\prime },\quad x_{-i}\right) ,u_{i}|h\right)\) \(\forall y_{i}^{\prime }\in {\mathbb {Y}}_{i}\). That is, \(y_{i}\in SB_{i}(Y_{-i},u_{i}).\)

\(\Rightarrow\)”: Let \(y_{i}\in SB_{i}(Y_{-i},u_{i})\). Then, there is \(\left( \left( y_{i}^{t},x_{-i}^{t}\right) ,\mu _{i}^{t}\right) \rightarrow \left( \left( y_{i},x_{-i}\right) ,\mu _{i}\right)\) such that \(x_{-i}\in Y_{-i}\) and \(y_{i}\) is sequentially optimal against the assessment \(\left( \left( y_{i},x_{-i}\right) ,\mu _{i}\right)\). Denote \(\left( y_{i},x_{-i}\right) =x\) and \(\left( y_{i}^{t},x_{-i}^{t}\right) =x^{t}\). Clearly, \(x^{t}\rightsquigarrow\) x. We proceed to construct a payoff sequence \(u_{i}^{t}\rightarrow u_{i}\) such that \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}}_{h}}v_{i}\left( \left( y_{h}^{\prime }, \quad x_{-h}^{t}\right) ,u_{i}^{t}\right)\) for all t and all \(h\in H_{i}\).

Since \(\Gamma\) is finite and perfect recall holds, we can define a (finite) partition \(\left\{ H_{i}^{l}\right\} _{l=1}^{L}\) of the set \(H_{i}\) as follows: \(H_{i}^{0}=\varnothing ,\) \(H_{i}^{l}\equiv \left\{ h\in H_{i}\backslash \cup _{\ell<l}H_{i}^{\ell }:\text {no }h^{\prime }\in \left[ H_{i}\backslash \cup _{\ell <l}H_{i}^{\ell }\right] \backslash h\text { is reached by }h\right\}\) for all \(l\ge 1\). Therefore, for all t and \(l=1,\ldots ,L\), we can define \(u_{i}^{t,l}\) recursively as follows: Let \(u_{i}^{t,0}\equiv u_{i}\),

$$\begin{aligned} u_{i}^{t,l}(z)\equiv \left\{ \begin{array}{ll} u_{i}^{t,l-1}(z)+\delta _{a_{h}^{*}}^{t}, &{} \hbox {if}\; z\; \hbox{is}\;\hbox{not}\; \hbox{precluded}\;\hbox{by}\; a_{h}^{*}\in \hbox { support}\;\left( y_{h}\right)\; \hbox { from }\; h \\ u_{i}^{t,l-1}(z), &{} \hbox { otherwise } \end{array} \right. , \end{aligned}$$

where \(h\in H_{i}^{l}\), support\(\left( y_{h}\right) =\left\{ a_{h}\in A_{h}:y_{h}\left( a_{h}\right) >0\right\}\) and

$$\begin{aligned} \delta _{a_{h}^{*}}^{t}=\max _{a_{h}\in A_{h}}v_{i}\left( \left( a_{h}, \quad x_{-h}^{t}\right) ,u_{i}^{t,l-1}|h\right) -v_{i}\left( \left( a_{h}^{*},\quad x_{-h}^{t}\right) ,u_{i}^{t,l-1}|h\right) . \end{aligned}$$

Therefore, for \(l=1,\ldots ,L\), \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}}_{h}}v_{i}\left( \left( y_{h}^{\prime }, \quad x_{-h}^{t}\right) ,u_{i}^{t,l}|h\right)\) \(\forall h\in H_{i}^{l}\). For all \(a_{h}\), \(a_{h}^{\prime }\in A_{h}\), \(v_{i}\left( \left( a_{h},\quad x_{-h}^{t}\right) ,u_{i}^{t,l+1}|h\right) -v_{i}\left( \left( a_{h}^{\prime },\quad x_{-h}^{t}\right) ,u_{i}^{t,l+1}|h\right) =v_{i}\left( \left( a_{h},\quad x_{-h}^{t}\right) ,u_{i}^{t,l}|h\right) -v_{i}\left( \left( a_{h}^{\prime },\quad x_{-h}^{t}\right) ,u_{i}^{t,l}|h\right)\). By induction on l, we have \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}} _{h}}v_{i}\left( \left( y_{h}^{\prime },\quad x_{-h}^{t}\right) ,u_{i}^{t,l}\right)\) \(\forall h\in \cup _{\ell =1}^{l}H_{i}^{\ell }\). Hence, \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}}_{h}}v_{i}\left( \left( y_{h}^{\prime },\quad x_{-h}^{t}\right) ,u_{i}^{t,L}\right)\) \(\forall h\in H_{i}\), which implies that \(\left( u_{i}^{t},\left( y_{i}^{t},x_{-i}^{t}\right) ,y_{i}\right) \in {\mathcal {R}}_{i}\).

It remains to show \(u_{i}^{t,L}\rightarrow u_{i}^{L}=u_{i}\). We prove this by induction on l. Clearly, \(u_{i}^{t,0}\rightarrow u_{i}\) trivially holds. Suppose \(u_{i}^{t,\ell }\rightarrow u_{i}^{\ell }=u_{i}\) for \(\ell \le l-1\). By construction of \(u_{i}^{t,l}\), it suffices to show \(\delta _{a_{h}^{*}}^{t}\rightarrow 0\) \(\forall h\in H_{i}^{l}\). Let \({\hat{a}} _{h}\in \arg \max _{a_{h}\in A_{h}}v_{i}\left( \left( a_{h},\quad x_{-h}^{t}\right) ,u_{i}^{t,l-1}|h\right)\). Because of the continuity of \(v_{i}\), for any \(\varepsilon >0\) there is a sufficiently large T such that, for all \(t>T\),

$$\begin{aligned} v_{i}\left( \left( {\hat{a}}_{h},\quad x_{-h}^{t}\right) ,u_{i}^{t,l-1}|h\right) -v_{i}\left( \left( {\hat{a}}_{h},\quad x_{-h}\right) ,u_{i}^{l-1}|h\right)< & {} \varepsilon ; \\ v_{i}\left( \left( a_{h}^{*},\quad x_{-h}\right) ,u_{i}^{l-1}|h\right) -v_{i}\left( \left( a_{h}^{*},\quad x_{-h}^{t}\right) ,u_{i}^{t,l-1}|h\right)< & {} \varepsilon . \end{aligned}$$

Since \(y_{h}\in \arg \max _{y_{h}^{\prime }\in {\mathbb {Y}}_{h}}v_{i}\left( \left( y_{h}^{\prime },\quad x_{-h}\right) ,u_{i}^{t}\right)\) and, by the induction hypothesis, \(u_{i}=u_{i}^{l-1}\),

$$\begin{aligned} v_{i}\left( \left( {\hat{a}}_{h},\quad x_{-h}\right) ,u_{i}^{l-1}|h\right) -v_{i}\left( \left( a_{h}^{*},\quad x_{-h}\right) ,u_{i}^{l-1}|h\right) \le 0. \end{aligned}$$

Therefore, \(v_{i}\left( \left( {\hat{a}}_{h},\quad x_{-h}^{t}\right) ,u_{i}^{t,l-1}|h\right) -v_{i}\left( \left( a_{h}^{*},\quad x_{-h}^{t}\right) ,u_{i}^{t,l-1}|h\right) <2\varepsilon\), i.e., \(\delta _{a_{h}^{*}}^{t}\rightarrow 0\).Footnote 17

Lemma 2

Let \(W\subseteq {\mathbb {R}}^{n+m}\) be a semi-algebraic set. (a) \(cl\left( W\right)\) and \(vcl_{\mathbb { R}^{n}}\left( W\right)\) are semi-algebraic. (b) There exists a closed, lower-dimensional semi-algebraic subset \(E\subset {\mathbb {R}}^{n}\) such that for all \(x\in {\mathbb {R}}^{n}\backslash E\), \(\{y\in {\mathbb {R}}^{m}:(x,y)\in cl\left( W\right) \}=\{y\in {\mathbb {R}}^{m}:(x,y)\in vcl_{{\mathbb {R}}^{n}}\left( W\right) \}\).

Proof

(a) \(vcl_{{\mathbb {R}}^{n}}\left( W\right)\) can be rewritten as

$$\begin{aligned} \left\{ \left( x,y\right) \in {\mathbb {R}}^{n}\times {\mathbb {R}}^{m}:\forall \varepsilon >0, \exists \left( x,y^{\prime }\right) \in W\text { s.t. } \left\| y-y^{\prime }\right\| <\varepsilon \right\} , \end{aligned}$$

where \(\left\| \cdot \right\|\) is the Euclidean norm. Since W is semi-algebraic, it follows from Tarski–Seidenberg theorem that \(vcl_{\mathbb { R}^{n}}\left( W\right)\) is also semi-algebraic. Similarly, \(cl\left( W\right)\) is semi-algebraic.

(b) This proof is similar to the proof of Theorem 4 in BZ94. Define f, g : \({\mathbb {R}}^{n}\times {\mathbb {R}}^{m}\rightarrow {\mathbb {R}}\) by

$$\begin{aligned} f\left( x,y\right)\equiv & {} \underset{\left( x^{\prime },y^{\prime }\right) \in W}{\inf }\left\| \left( x,y\right) -\left( x^{\prime },y^{\prime }\right) \right\| \\ g\left( x,y\right)\equiv & {} \underset{\left( x,y^{\prime }\right) \in W}{\inf } \left\| \left( x,y\right) -\left( x,y^{\prime }\right) \right\| , \end{aligned}$$

where x and \(x^{\prime }\) belong to \({\mathbb {R}}^{n}\), y and \(y^{\prime }\) belong to \({\mathbb {R}}^{m}\).

Define

$$\begin{aligned} E=\left\{ x\in {\mathbb {R}}^{n}:\exists \left( x,y\right) \in W, f\left( x,y\right) =0\text { and }g\left( x,y\right) >0\right\} . \end{aligned}$$

Since f and g are semi-algebraic functions, E is a semi-algebraic set. Note that \((x,y)\in cl\left( W\right)\) if and only if \(f(x,y)=0\); and \((x,y)\in vcl_{{\mathbb {R}}^{n}}\left( W\right)\) if and only if \(g(x,y)=0\). Thus, \((x,y)\in cl\left( W\right) \backslash vcl_{{\mathbb {R}}^{n}}\left( W\right)\) implies \(x\in E\). Therefore, for all \(x\in {\mathbb {R}} ^{n}\backslash E\), \(\{y\in {\mathbb {R}}^{m}:(x,y)\in cl\left( W\right) \}=\{y\in {\mathbb {R}}^{m}:(x,y)\in vcl_{{\mathbb {R}}^{n}}\left( W\right) \}\). Suppose E is not lower-dimensional in \({\mathbb {R}}^{n}\). Then, there is a semi-algebraic open set \(\mathcal {O}\subseteq {\mathbb {R}}^{n}\) and an \(\varepsilon >0\) with the property that for any \(x\in \mathcal {O}\) there exists \(\left( x,y\right) \in W\) such that \(\ f\left( x,y\right) =0\) and \(g\left( x,y\right) \ge \varepsilon\). The set

$$\begin{aligned} G=\left\{ \left( x,y\right) \in \mathcal {O}\times {\mathbb {R}}^{m}:f\left( x,y\right) =0\;{\text { and}}\;g\left( x,y\right) \ge \varepsilon \right\} \end{aligned}$$

is semi-algebraic, and its projection onto \({\mathbb {R}}^{n}\) is all of \(\mathcal {O}\). So we can choose a semi-algebraic selection \(\beta :\mathcal {O} \rightarrow {\mathbb {R}}^{m}\) with the property that \(\left( x,\beta \left( x\right) \right) \in G\). By BZ94’s (p. 786) Lemma, there exists a semi-algebraic open set \(\mathcal {O}^{\prime }\subset \mathcal {O}\) on which \(\beta\) is continuous. Since \(\left( x,\beta \left( x\right) \right) \in cl(W)\) for any \(w\in \mathcal {O}^{\prime }\), there is a sequence \(\left\{ \left( x^{t},y^{t}\right) \right\} _{t=1}^{\infty }\) in W with limit \(\left( x,\beta \left( x\right) \right)\). From the continuity of \(\beta\), \(\left\| \beta \left( x^{t}\right) -\beta \left( x\right) \right\| \rightarrow 0\). Thus, for t large enough, \(g\left( x^{t},\beta \left( x^{t}\right) \right) <\varepsilon\), which contradicts the construction of \(\beta\). Therefore, E is a lower-dimensional in \({\mathbb {R}}^{n}\) and the result follows. □

Lemma 3

Consider an extensive form \(\Gamma\) with perfect recall. For each player i, there is a closed, lower-dimensional semi-algebraic subset \(U_{i}^{0}\subset U_{i}\) such that, for any \(Y_{-i}\subseteq {\mathbb {Y}}_{-i}\), \(PB_{i}(Y_{-i},u_{i})=SB_{i}(Y_{-i},u_{i})\)\(\forall u_{i}\in U_{i}\backslash U_{i}^{0}.\)

Proof

Let \(Y_{-i}\subseteq {\mathbb {Y}}_{-i}\). By Lemma 1, \(SB_{i}(Y_{-i},u_{i})=\{y_{i}\in {\mathbb {Y}}_{i}:\left( u_{i},x,y_{i}\right) \in cl({\mathcal {R}}_{i})\cap Q\left( Y_{-i}\right) \}\) and \(PB_{i}(Y_{-i},u_{i})=\{y_{i}\in {\mathbb {Y}}_{i}:\left( u_{i},x,y_{i}\right) \in vcl_{U_{i}}({\mathcal {R}}_{i})\cap Q\left( Y_{-i}\right) \}\). By the Tarski–Seidenberg theorem, \({\mathcal {R}}_{i}\) is semi-algebraic. By Lemma 2, \(cl({\mathcal {R}}_{i})\) and \(vcl_{U_{i}}({\mathcal {R}}_{i})\) are generically equivalent on \(U_{i}\) and hence, \(cl({\mathcal {R}}_{i})\cap Q\left( Y_{-i}\right)\) and \(vcl_{U_{i}}({\mathcal {R}}_{i})\cap Q\left( Y_{-i}\right)\) are generically equivalent on \(U_{i}\). Therefore, there is a closed, lower-dimensional semi-algebraic subset \(U_{i}^{0}\subset U_{i}\) such that \(PB_{i}(Y_{-i},u_{i})=SB_{i}(Y_{-i},u_{i})\)\(\forall u_{i}\in U_{i}\backslash U_{i}^{0}.\)

Lemma 4

Let \(F:U\rightrightarrows {\mathbb {R}}^{n}\) and \(F^{\prime }:U\rightrightarrows {\mathbb {R}}^{n}\) . Suppose \(V^{0}\equiv \left\{ u^{0}\in U:F\left( u^{0}\right) \ne F^{\prime }\left( u^{0}\right) \right\}\) is a lower-dimensional subset of U . Then, \(F\left( u\right) \subseteq F^{\prime }\left( u\right)\) for all \(u\in U\) at which \(F\left( \cdot \right)\) is lower hemi-continuous and \(F^{\prime }\left( \cdot \right)\) is upper hemi-continuous.

Proof of Lemma 4

Since \(V^{0}\) is lower-dimensional, \(V^{0}\) contains no open set in U. Let \(u\in U\). Therefore, we can find a sequence \(\left\{ u^{t}\right\} _{t=1}^{\infty }\) in \(U\backslash V^{0}\) such that \(u^{t}\rightarrow u\) and \(F\left( u^{t}\right) =F^{\prime }\left( u^{t}\right)\) for all t. If \(y\in F\left( u\right)\), by lower hemi-continuity of \(F\left( \cdot \right)\), there exists a subsequence \(u^{t_{k}}\rightarrow u\) such that \(y^{k}\rightarrow y\) and \(y^{k}\in F\left( u^{t_{k}}\right) =F^{\prime }\left( u^{t_{k}}\right)\). Since the correspondence \(F^{\prime }\left( \cdot \right)\) is upper hemi-continuous, \(y\in F^{\prime }\left( u\right)\). That is, \(F\left( u\right) \subseteq F^{\prime }\left( u\right).\)

Proof of Theorem 1

By Lemma 3, for each player \(i\in N\), there is a closed, lower-dimensional semi-algebraic subset \(U_{i}^{0}\subset U_{i}={\mathbb {R}}^{\left| Z\right| }\) such that for all \(u_{i}\in U_{i}\backslash U_{i}^{0}\), \(PB_{i}(Y_{-i},u_{i})=SB_{i}(Y_{-i},u_{i})\) for all \(Y_{-i}\subseteq {\mathbb {Y}}_{-i}\); hence, \(PB_{i}(\Pi _{j\ne i}co^{e}(Y_{j}),u_{i})=SB_{i}(\Pi _{j\ne i}co^{e}(Y_{j}),u_{i})\) for all \(\Pi _{j\ne i}Y_{j}\subseteq {\mathbb {Y}}_{-i}\). Therefore, for all the payoffs \(u\in U\backslash U^{0}\) where \(U^{0}=\cup _{i\in N}\left( U_{i}^{0}\times U_{-i}\right)\), a course of action \(\sigma\) is perfectly supported by Y in \(\Gamma (u)\) if and only if \(\sigma\) is sequentially supported by Y in \(\Gamma (u)\). Thus, we find a closed, lower-dimensional semi-algebraic subset \(U^{0}\subset U\) such that \(\Sigma ^{\text {sequential} }(u)=\Sigma ^{\text {perfect}}(u)\) and \(\top _{\sigma }^{\text {sequential} }(u)=\top _{\sigma }^{\text {perfect}}(u)\) for all \(u\in U\backslash U^{0}\) and for an arbitrary given course of action \(\sigma\) in \(\Gamma\).

Because \(PB_{i}(\Pi _{j\ne i}co^{e}(Y_{j}),u_{i})=SB_{i}(\Pi _{j\ne i}co^{e}(Y_{j}),u_{i})\) for all \(i\in N\) and \(u\in U\backslash U^{0}\), \(\{u^{0}\in U:\Upsilon ^{\text {sequential}}\left( u^{0}\right) \ne \Upsilon ^{\text {perfect}}\left( u^{0}\right) \} \subseteq U^{0}\) is a lower-dimensional subset of U. By Lemma 4, \(\Upsilon ^{\text {sequential}}\left( u\right) \subseteq \Upsilon ^{\text {perfect}}\left( u\right)\) for all the payoffs \(u\in U\) at which \(\Upsilon ^{\text {sequential}}\left( \cdot \right)\) is lower hemi-continuous and \(\Upsilon ^{\text {perfect}}\left( \cdot \right)\) is upper hemi-continuous. But, since any perfect-MACA in \(\Gamma (u)\) is also a sequential-MACA in \(\Gamma (u)\), \(\Upsilon ^{\text { perfect}}\left( u\right) \subseteq \Upsilon ^{\text {sequential}}\left( u\right)\) for all \(u\in U\). Hence, \(\Upsilon ^{\text {sequential}}\left( u\right) =\Upsilon ^{\text {perfect}}\left( u\right)\) for all the payoffs \(u\in U\) at which \(\Upsilon ^{\text {sequential}}\left( \cdot \right)\) is lower hemi-continuous and \(\Upsilon ^{\text {perfect}}\left( \cdot \right)\) is upper hemi-continuous. Similarly, for an arbitrary given course of action \(\sigma\) in \(\Gamma\), \(\top _{\sigma }^{\text {sequential}}(u)=\top _{\sigma }^{\text {perfect}}(u)\) for all the payoffs \(u\in U\) at which \(\top _{\sigma }^{\text {sequential}}(\cdot )\) is lower hemi-continuous and \(\top _{\sigma }^{\text {perfect}}(\cdot )\) is upper hemi-continuous. □

Proof of Corollary 1

Define

$$\begin{aligned} \overset{o}{{\mathcal {R}}}\mathcal {=}\left\{ \left( u,x,y\right) \in U\times {\mathbb {Y}}\times {\mathbb {Y}}:\quad \left( u_{i},x,y_{i}\right) \in \mathcal { R}_{i}\quad \forall i\right\} , \end{aligned}$$

that is, \(\overset{o}{{\mathcal {R}}}\) is the set of “joint” perfectly rational states in which the players have consistently aligned beliefs about the opponent players’ behavior. Clearly, \(\left( u,x,y\right) \in vcl_{U}\left( \overset{o }{{\mathcal {R}}}\right)\) iff there exists a common sequence \(x^{t}\rightsquigarrow x\) such that for each player i, \(y_{i}\) is the best response along the trembling sequence \(x^{t}\). By Definition 3(i), \(\left( u,y,y\right) \in vcl_{U}\left( \overset{o}{\mathcal {R}}\right)\) iff y is a perfect equilibrium; by Kreps and Wilson’s (1982) Proposition 6, \(\left( u,y,y\right) \in cl\left( \overset{o}{\mathcal {R}}\right)\) iff y is a sequential equilibrium strategy profile. By the Tarski–Seidenberg theorem, \(\overset{o}{\mathcal {R}}\) is semi-algebraic. By Lemma 2, \(cl\left( \overset{o }{\mathcal {R}}\right)\) and \(vcl_{U}\left( \overset{o}{\mathcal {R}}\right)\) are generically equivalent. Consequently, there is a closed, lower-dimensional semi-algebraic subset \(U^{0}\subset U\) such that for all \(u\in U\backslash U^{0}\), any sequential equilibrium strategy profile is a perfect equilibrium in game \(\Gamma (u)\).

Now, suppose that the sequential equilibrium correspondence \(SE\left( \cdot \right)\) is lower hemi-continuous and the perfect equilibrium correspondence \(PE\left( \cdot \right)\) is upper hemi-continuous at \(u\in U\). Since \(\left\{ u^{0}\in U:SE\left( u^{0}\right) \ne PE\left( u^{0}\right) \right\} \subseteq U^{0}\) is a lower-dimensional subset, by Lemma 4, \(SE\left( u\right) \subseteq PE\left( u\right)\). Thus, \(SE\left( u\right) =PE\left( u\right).\)

Proof of Corollary 2

(i) By Definition 3(iii)-(iv), a strategy profile \(y\in {\mathbb {Y}}\) is a weakly perfect (or weakly sequential) equilibrium in \(\Gamma (u)\) iff for each player i, \(y_{i}\in PB_{i}\left( y_{-i},u_{i}\right) =PB_{i}\left( \Pi _{j\ne i}co^{e}(y_{j}),u_{i}\right)\) (or \(y_{i}\in SB_{i}\left( y_{-i},u_{i}\right) =SB_{i}\left( \Pi _{j\ne i}co^{e}(y_{j}),u_{i}\right)\) ). By Definition 2, the strategy profile y is a weakly perfect (or weakly sequential) equilibrium in \(\Gamma (u)\) iff y is a perfect (or sequential) MACA in \(\Gamma (u)\). By Theorem 1, for generic \(u\in U\backslash U^{0}\), the set of weakly sequential equilibria in \(\Gamma (u)\) coincides with the set of weakly perfect equilibria in \(\Gamma (u)\).

(ii) By Definition 4, a product set \(Y=\Pi _{i\in N}Y_{j}\subseteq {\mathbb {Y }}\) is a perfectly (or sequentially) rationalizable set in \(\Gamma (u)\) iff for each player i, \(Y_{i}\subseteq PB_{i}\left( \Pi _{j\ne i}co^{e}(Y_{j}),u_{i}\right)\) (or \(Y_{i}\subseteq SB_{i}\left( \Pi _{j\ne i}co^{e}(Y_{j}),u_{i}\right)\)). By Definition 2, the set Y is a perfectly (or sequentially) rationalizable set in \(\Gamma (u)\) iff Y perfectly (or sequentially) supports the null MACA in \(\Gamma (u)\). By Theorem 1, for generic payoffs \(u\in U\backslash U^{0}\), the union of sequentially rationalizable sets in \(\Gamma (u)\) coincides with the union of perfectly rationalizable sets in \(\Gamma (u)\); that is, the set of sequentially rationalizable strategy profiles in \(\Gamma (u)\) coincides with the set of perfectly rationalizable strategy profiles in \(\Gamma (u)\).

(iii) This result follows directly from the following lemma: □

Lemma 5

A path resulting from an SRSCE in Definition A3is a path sequential-MACA in Definition 2and vice versa.

Proof of Lemma 5

\(\Rightarrow\) ” Suppose \(\widehat{y}\) is an SRSCE in \(\Gamma (u)\). By Definition 5, there exists \(Y=\Pi _{i\in N}Y_{i}\subseteq {\mathbb {Y}}\), for each \(i\in N\) and each \(y_{i}\in Y_{i}\), we can find \(y_{-i}\in \Pi _{j\ne i}co^{e}(Y_{j})\) such that \(y_{h}=\widehat{y}_{h}\) if \(h\in H\) is reachable under \(\widehat{y}\) and \(y_{i}\in SB_{i}\left( y_{-i},u_{i}\right)\). Construct a path course of action \(\sigma\) (associated with \(\widehat{y}\)) as follows: For any \(h\in H\),

$$\begin{aligned}\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \sigma _{h}=\left\{ \begin{array}{ll} \widehat{y}_{h}, &{} \text {if }h\text { is reachable under }\widehat{y} \\ \emptyset , &{} \text {otherwise} \end{array} \right. . \end{aligned}$$

By the construction of \(\sigma\), if \(\sigma _{h}\ne \emptyset\), then \(y_{h}=\sigma _{h}\) for all \(y\in Y\). Moreover, since \(y_{i}\in SB_{i}\left( y_{-i},u_{i}\right)\) and \(y_{-i}\in \Pi _{j\ne i}co^{e}(Y_{j})\), \(y_{i}\in SB_{i}\left( \Pi _{j\ne i}co^{e}(Y_{j}),u_{i}\right)\). By Definition 2, the path course of action \(\sigma\) (associated with \(\widehat{y}\)) is a path sequential-MACA in \(\Gamma (u)\) supported by Y.

\(\Leftarrow\)” Suppose \(\sigma\) is a path sequential-MACA in \(\Gamma (u)\) supported by \(Y=\Pi _{i\in N}Y_{i}\subseteq {\mathbb {Y}}\). By Definition 2, for each player i and each \(\widehat{y}_{i}\in\) \(Y_{i}\), there exists \(\widehat{y}_{-i}\in \Pi _{j\ne i}co^{e}(Y_{j})\) such that \(\widehat{y}_{i}\in SB_{i}\left( \widehat{y} _{-i},u_{i}\right)\); moreover, if \(\sigma _{h}\ne \emptyset\), \(y_{h}=\sigma _{h}\) for all \(y\in Y\) and hence \(\widehat{y}_{h}=\sigma _{h}\). That is, \(\widehat{y}_{h}=\sigma _{h}\) if h is reachable under \(\widehat{y}\). By Definition 5, \(\widehat{y}\) is an SRSCE that results in the same path \(\sigma\).

To show the last part of Corollary 2, let \(\sigma\) be a course of action in \(\Gamma\). Notice that the player-by-player union of the sequentially/perfectly \(\sigma\)-supporting sets in \(\Gamma (u)\) is again a sequentially/perfectly \(\sigma\)-supporting set in \(\Gamma (u)\). Thus, \(\Upsilon _{\sigma }^{\text {sequential}}(u)\) is the largest sequentially \(\sigma\)-supporting set, and \(\Upsilon _{\sigma }^{\text {perfect}}(u)\) is the largest perfectly \(\sigma\)-supporting set. By Theorem 1, the largest sequentially \(\sigma\)-supporting set coincides with the largest perfectly \(\sigma\)-supporting set for generic payoffs \(u\in U\backslash U^{0}.\)

Proof of Corollary 3

Consider a finite normal form \(\Gamma =(N,\left\{ A_{i}\right\} _{i\in N})\). Let \(\left( W^{k}(u)\right) _{k=0}^{K}\) be an IEWDS procedure in \(\Gamma (u)\), where \(W^{k}(u)=\Pi _{i\in N}W_{i}^{k}(u)\) such that \(W_{i}^{0}(u)=A_{i}\) and \(W_{i}^{k}(u)\subseteq A_{i}\) includes the set of i’s surviving weakly undominated actions in the \(\left( k-1\right)\)-th round of elimination for all \(k\ge 1\). Let \(i\in N\) and \(k=1,2,\ldots ,K\). Because \(\Gamma\) is a normal form, by Pearce’s (1984) Lemma 3, \(a_{i}\in W_{i}^{k}(u)\) is not strictly dominated in \(W^{k}(u)\) iff \(a_{i}\in SB_{i}\left( \Delta \left( W_{-i}^{k}(u)\right) ,u_{i}\right)\); by Pearce’s (1984) Lemma 4, \(a_{i}\in W_{i}^{k}(u)\) is not weakly dominated in \(W^{k}(u)\) iff \(a_{i}\in PB_{i}\left( \Delta \left( W_{-i}^{k}(u)\right) ,u_{i}\right)\). By Lemma 3, there is a closed, lower-dimensional subset \(U_{i}^{0}\subset U_{i}\) such that for all \(u_{i}\in U_{i}\backslash U_{i}^{0}\), \(SB_{i}\left( Y_{-i},u_{i}\right) =PB_{i}\left( Y_{-i},u_{i}\right)\) \(\forall Y_{-i}\subseteq \Delta (A_{-i})\). Letting \(Y_{-i}=\{y_{-i}\in \Delta (A_{-i}):y_{-i}\) has full support on \(W_{-i}^{k}(u)\}\), we have \(SB_{i}\left( \Delta \left( W_{-i}^{k}(u)\right) ,u_{i}\right) =PB_{i}\left( \Delta \left( W_{-i}^{k}(u)\right) ,u_{i}\right)\) for all \(u_{i}\in U_{i}\backslash U_{i}^{0}\). Define \(U^{0}=\cup _{i\in N}\left( U_{i}^{0}\times U_{-i}\right)\). Therefore, \(U^{0}\subset U\) is a closed, lower-dimensional semi-algebraic subset such that for all \(u\in U\backslash U^{0}\), \(a\in W^{k}(u)\backslash W^{k+1}(u)\) iff \(a\in W^{k}(u)\) is strictly dominated in \(W^{k}(u)\), and hence, \(\left( W^{k}(u)\right) _{k=0}^{K}\) is an IESDS procedure in \(\Gamma (u)\). Since IESDS is order-independent in finite games, IEWDS is generically an order-independent procedure. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Luo, X., Qian, X. & Sun, Y. The algebraic geometry of perfect and sequential equilibrium: an extension. Econ Theory 71, 579–601 (2021). https://doi.org/10.1007/s00199-020-01259-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-020-01259-z

Keywords

JEL Classification

Navigation