Skip to main content
Log in

Essential equilibria of large generalized games

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

Abstract

We characterize the essential stability of games with a continuum of players, where strategy profiles may affect objective functions and admissible strategies. Taking into account the perturbations defined by a continuous mapping from a complete metric space of parameters to the space of continuous games, we prove that essential stability is a generic property and every game has a stable subset of equilibria. These results are extended to discontinuous large generalized games assuming that only payoff functions are subject to perturbations. We apply our results in an electoral game with a continuum of Cournot-Nash equilibria, where the unique essential equilibrium is that only politically engaged players participate in the electoral process. In addition, employing our results for discontinuous games, we determine the stability properties of competitive prices in large economies.

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.

Similar content being viewed by others

Notes

  1. That is, for every player, objective functions and correspondences of admissible strategies are continuous.

  2. A subset of a metric space is residual if it contains the intersection of a countable family of dense and open sets.

  3. That is, the correspondence that associates games with the set of its pure strategy equilibria.

  4. See Schmeidler (1973) and Rath (1992) for continuous large games, Balder (1999, 2002) for continuous large generalized games, and Carmona and Podczeck (2014) for discontinuous large generalized games.

  5. That is, for any Borelian set \(E\subseteq {\mathbb {R}}^m, \{t \in T_1: H(t,f(t))\in E\}\) belongs to \({\mathcal {A}}\).

  6. Given \(t \in T_1\), continuity of \(\varGamma _t:\widehat{M} \times \widehat{{\mathcal {F}}}^2 \twoheadrightarrow K_{t}\) requires that it be both upper hemicontinuous and lower hemicontinuous. Upper hemicontinuity is satisfied at \((m,a)\in \widehat{M} \times \widehat{{\mathcal {F}}}^2\) when for any open set \(A\subseteq K_{t}\) such that \(\varGamma _t(m,a) \subseteq A\), there is an open neighborhood \(U \subseteq \widehat{M} \times \widehat{{\mathcal {F}}}^2\) of \((m,a)\) such that \(\varGamma _t(m',a') \subseteq A\) for every \((m',a') \in U\). Lower hemicontinuity is satisfied at \((m,a)\in \widehat{M} \times \widehat{{\mathcal {F}}}^2\) when for any open set \(A\subseteq K_{t}\) such that \(\varGamma _t(m,a) \cap A \ne \emptyset \), there is an open neighborhood \(U \subseteq \widehat{M} \times \widehat{{\mathcal {F}}}^2\) of \((m,a)\) such that \(\varGamma _t(m',a') \cap A \ne \emptyset \) for every \((m',a') \in U\). Same definitions apply for the correspondences of admissible strategies associated with atomic players \((\varGamma _t)_{t \in T_2}\).

  7. Under our assumptions, for every \((m,a)\in \widehat{M} \times \widehat{\mathcal {F}}^2\) the correspondences \(t\in T_1 \twoheadrightarrow K_t\) and \(t\in T_1 \twoheadrightarrow \varGamma _t(m,a)\) have measurable graph [see Aliprantis and Border (2006, Lemma 18.2 and Theorem 18.6)]. On the other hand, Balder’s result requires that players have a common universal action space. However, his result can be extended allowing different universal action spaces for atomic and non-atomic players [see Balder (2002), page 448, remark (v)].

  8. For instance, consider an electoral game with a continuum of non-atomic players \(T_1=[0,1]\), which vote for a party in \(\{a,b\}\). Let \(x_t\) be the action of player \(t \in T_1\) and assume that his objective function \(u_t\) only takes into account the benefits that he receives from the election of parties, given by \(\{v_t(a), v_t(b)\}\), weighted by the support that each party has in the population, i.e., \(u_t \equiv v_t (a) \mu (\{s \in T_1: x_s=a\}) + v_t (b) (1-\mu (\{s \in T_1: x_s=a\})),\) where \(\mu \) denotes the Lebesgue measure in \([0,1]\), that is, the utility level of a player \(t \in T_1\) in unaffected by his own action and, therefore, any measurable profile \(x:[0,1]\rightarrow \{a,b\}\) constitutes a Nash equilibrium of the game. Hence, the set of Nash equilibria is not compact. However, if we consider that each player receives as a message the support that party \(a\) has in the population, \(m=\mu (\{s \in T_1: x_s=a\})\), then the set of equilibrium messages is equal to \([0,1]\), which is a compact set.

  9. Since action profiles are coded using the function \(H\), there may exist several Cournot-Nash equilibria inducing a same message. Even that, this indetermination does not have real effects on players utility levels.

  10. A subset of \({\mathbb {G}}'\) is residual if it contains the intersection of a countable family of dense and open subsets of \({\mathbb {G}}'\).

  11. For general results of strategic approximations of continuous games by finite games, see Reny (2011).

  12. Indeed, \(B\) is non-empty and compact. Also, for any open set \(O \subseteq {\widehat{M}}\times {\widehat{\mathcal {F}}}^2\) such that \(B \subseteq O\), we have that \(A \subset O\). Thus, the essentiality of \(A\) with respect to \({\mathbb {G}}'\) ensures that \(B\) is essential too.

  13. By definition, components are non-empty. Since a component is a union of connected sets with at least one common element, it is connected too. Since the closure of a connected set is connected, components of compact sets are closed and, therefore, compact [for more details, see Berge (1997, page 98)].

  14. For instance, when there are personalized perturbations on players’ characteristics, as an example, fix a game \({\mathcal {G}}={\mathcal {G}}((K_{t},\varGamma _{t},u_{t})_{t\in T_1 \cup T_2}) \in {\mathbb {G}}\). Given \(i \in \{1,2\}\), let \(T_i^a, T_i^b, T_i^c\subseteq T_i\) be, respectively, the subsets of players in \(T_i\) for which we allow perturbations on objective functions, on strategy sets, and on the correspondences of admissible strategies. Let \({\mathbb {G}}_{\mathcal {G}}\subseteq {\mathbb {G}}\) be the set of generalized games \(\widetilde{{\mathcal {G}}}((\widetilde{K}_{t},\widetilde{\varGamma }_{t}, \widetilde{u}_{t})_{t\in T_1 \cup T_2})\) such that (1) for any \(t\in (T_1 {\setminus } T^a_1)\cup (T_2 {\setminus } T^a_2) , \widetilde{u}_t=u_t\); (2) for any \(t\in (T_1 {\setminus } T^b_1)\cup (T_2 {\setminus } T^b_2), \widetilde{K}_t=K_t\); and (3) for any \(t\in (T_1 {\setminus } T^c_1)\cup (T_2 {\setminus } T^b_2), \widetilde{\varGamma }_t=\varGamma _t\). Since \({\mathbb {G}}_{\mathcal {G}}\) is \(\rho \)-closed, it follows that \(({\mathbb {G}}_{\mathcal {G}},\rho )\) is a complete metric space. Therefore, since the immersion \(\iota : {\mathbb {G}}_{\mathcal {G}} \hookrightarrow {\mathbb {G}}\) is continuous, \((({\mathbb {G}}_{\mathcal {G}},\rho ), \iota )\) is a parametrization of \({\mathbb {G}}\).

  15. Suppose that \({\mathcal {G}}_1={\mathcal {G}}_1((K^1_{t}, \varGamma ^1_t, u^1_t)_{t\in T_1 \cup T_2}) \) and \({\mathcal {G}}_2={\mathcal {G}}_2((K^2_{t}, \varGamma ^2_t, u^2_t)_{t\in T_1 \cup T_2}) \). Since \(\widehat{K}\) and \(\widehat{K}_t\), where \(t\in T_2\), are convex subsets of normed spaces with metrics induced by norms, for each \(\lambda \in [0,1]\), the convex combination \(\lambda {\mathcal {G}}_1 + (1-\lambda ) {\mathcal {G}}_2\) is well defined and given by the game \({\mathcal {G}}((\lambda K^1_t + (1-\lambda )K^2_t, \lambda \varGamma ^1_t + (1-\lambda )\varGamma ^2_t, \lambda u^1_t + (1-\lambda )u^2_t )_{t\in T_1 \cup T_2}).\) Recall that, given subsets \(A\) and \(B\) of a vectorial space, \(\lambda A + (1-\lambda )B:=\{\lambda a+ (1-\lambda ) b: (a,b)\in A \times B\}\).

  16. It is sufficient to prove that any \({\mathcal {T}}\)-essential component contains a minimal \({\mathcal {T}}\)-essential set. Fix an \({\mathcal {T}}\)-essential component \(C \subseteq \varLambda (\kappa (\mathcal {X}))\). Let \(\mathcal {S}_C\) be the family of \({\mathcal {T}}\)-essential subsets of \(\varLambda (\kappa (\mathcal {X}))\) contained in \(C\), endowed with the partial order determined by set inclusion. Since essential sets are non-empty and compact, any totally ordered subset of \(\mathcal {S}_C\) has a lower bound. By Zorn’s Lemma, \(\mathcal {S}_C\) has a minimal element, which concludes the proof.

  17. More precisely, given a large game \({\mathcal {G}}\) with only non-atomic players, for any \(\epsilon >0\) define an \(\epsilon \)-perturbed game \({\mathcal {G}}_\epsilon \) where every player perceives that he, but no other, has a positive small impact on the social choice. Then, following our notation, \((f,a)\in \widehat{\mathcal {F}}^1 \times \widehat{\mathcal {F}}^2\) is a strategic equilibrium for a game \({\mathcal {G}}\) if there exists \(\{\epsilon _k\}_{k \in {\mathbb {N}}}\subset (0,1)\) decreasing to zero, and a sequence \(\{(f_k,a_k)\}_{k \in {\mathbb {N}}} \subset \widehat{\mathcal {F}}^1 \times \widehat{\mathcal {F}}^2\) converging to \((f,a)\), such that \((f_k,a_k)\) is a Cournot-Nash equilibrium for \({\mathcal {G}}_{\epsilon _k}\) for any \(k \in {\mathbb {N}}\).

  18. Perturbations on actions sets for non-atomic players, or on any atomic player characteristic, may change the underlying institutional structure, destroying the electoral dimension of the game. However, a natural perturbation in action sets is to forbid the voluntary vote, by changing \(K_t\) to \(\left\{ (x_1,\ldots ,x_{\overline{p}}) \in {\mathbb {Z}}^{\overline{p}}_+: \sum _{p=1}^{\overline{p}} x_p=1\right\} .\) In this case, in any Cournot-Nash equilibrium for \({\mathcal {E}}_0\), all voters participate in the election. In addition, the \({\mathcal {T}}\)-essential Cournot-Nash equilibria of \({\mathcal {E}}_0\) are those in which politically engaged players support their favorite party.

  19. Given a topological space \(X, u:X\rightarrow {\mathbb {R}}\) is upper semicontinuous if \(\{x \in X: u(x)\ge a\}\) is closed for any \(a\in {\mathbb {R}}\).

  20. For every \(m \in {\widehat{M}}\) and \(\epsilon >0\), generalized payoff security holds by choosing \(\alpha \equiv 0\).

  21. Taking \(m=1\) and \(\epsilon \in (0,1)\), if generalized payoff security holds for \(\overline{\mathcal {G}}\), then Definition 11(i) implies that \(\alpha (t)\le 0, \forall t \in T_1\). On the other hand, Definition 11(ii) ensures that there exists a positive measure set \(T'\subseteq T_1\) such that \(\alpha (t)+\epsilon \ge \sup _{x \in \varGamma _t(1)} u_t(x,1)\), which in turn implies that \(\epsilon \ge 1\). A contradiction.

  22. Yu (1999) also allows for perturbations on action sets and on correspondences of admissible strategies, but only for continuous games. Thus, we recover his results in Theorem 1.

  23. Scalzo (2013) extends the stability results of Carbonell-Nicolau (2010) to a space of finite-player discontinuous games where an aggregator of payoff functions satisfies a property called generalized positively quasi-transfer continuity. This property relaxes both generalized payoff security and upper semicontinuity. Although we focus on large generalized games where atomic players satisfy the assumptions imposed by Carbonell-Nicolau (2010), we presume that the same arguments of Scalzo (2013) may be applied to relax these assumptions in our context.

  24. Utility functions \((u_t)_{t\in T_1}\) are continuous and take nonnegative values. Since for every \(t \in T_1\) the initial endowment \(w(t)\in \text{ int }(\widehat{K})\), the budget set correspondence \(p \in \varDelta \twoheadrightarrow B_t(p)\) is continuous and has non-empty and compact values. Therefore, Berge’s Maximum Theorem [see Aliprantis and Border (2006, Theorem 17.31, page 570)] guarantees that, given \((m,p) \in \widehat{K} \times \varDelta \) and \(\epsilon >0\), generalized payoff security holds by choosing a sufficient small neighborhood \(U\) of \((m,p)\) and mappings \(\alpha :T_1\cup \{a\} \rightarrow {\mathbb {R}}\) and \((\varphi _t)_{t \in T_1 \cup \{a\}}\) such that \(\alpha (a)=V_a(\varphi _0(m,p),m)-\epsilon , \varphi _a(m',p')=\mathop {\hbox {argmax}}\nolimits _{p''\in \varDelta } V_a(p'',m'),\,\forall (m',p')\in U\), and for each non-atomic player \(t\in T_1, \alpha (t)=u_t(\varphi _t(m,p))-\epsilon \) and \(\varphi _t(m',p')=\mathop {\hbox {argmax}}\nolimits _{x(t)\in B_t(p')} u_t(x(t)),\,\forall (m',p')\in U\). The existence of finitely many types of non-atomic agents guarantees that there is a common neighborhood \(U\) for every \(t\in T_1\cup \{a\}\) and also ensures that for every \((m',p')\in U\), the map \(t\in T_1 \rightarrow (\alpha (t), \varphi _t(m',p'))\) is measurable.

  25. Since \({\mathcal {G}}((u_t)_{t \in T_1})\in {\mathbb {G}}_d\), it has a non-empty set of Cournot-Nash equilibria. As functions \((u_t)_{t \in T_1}\) take nonnegative values and are strictly increasing, \((\overline{m},\overline{p})\in \varLambda ({\mathcal {G}}((u_t)_{t \in T_1})) \) if and only if for some \(\overline{x}\in \widehat{\mathcal {F}}^1\) we have that

    1. (i)

      There exists a full measure set \(T_1'\subseteq T_1\) such that \(v_t(\overline{x}(t), \overline{p})=u_t(\overline{x}(t))=\max _{x(t)\in B_t(\overline{p})}\,\,u_t(x(t)),\,\forall t \in T_1'\).

    2. (ii)

      \(\overline{p}\gg 0\) and \(\overline{p}(\overline{x}(t) - w(t))=0,\,\forall t \in T_1'\).

    3. (iii)

      \(\overline{m}=\int _{T_1} \overline{x}(t)d\mu =\int _{T_1} w(t)d\mu .\)

    Thus, \((\overline{m},\overline{p})\in \varLambda ({\mathcal {G}}((u_t)_{t \in T_1})) \) if and only if there exists \(\overline{x}\in \widehat{\mathcal {F}}^1\) such that \((\overline{p},\overline{x})\) is an equilibrium for \({\mathcal {E}}((u_t)_{t \in T_1})\).

  26. Since \((S,d)\) is a compact metric space, it follows from Aliprantis and Border (2006, Theorem 3.85-(3) and Theorem 3.88-(2), pages 116 and 119, respectively) that \(A(S)\) is a complete metric space under the Hausdorff distance induced by \(d\). When \(S\) is a compact subset of a normed vector space, \((A_c(S),d_H)\) remains a complete metric space, since the Hausdorff limit of a sequence of compact convex sets is still a compact convex set.

  27. Since \(\widehat{M} \times \widehat{\mathcal {F}}^2\) is compact and \((A(\widehat{K}), d_H)\) is complete, for every \(t \in T_1\), the continuity of the correspondence \(\overline{\varGamma }_t\) follows from the completeness of the space of continuous functions \(\nu :\widehat{M} \times \widehat{\mathcal {F}}^2 \rightarrow A(\widehat{K})\) under the uniform metric induced by the Haussdorf distance. Indeed, any correspondence \(\varGamma :\widehat{M} \times \widehat{\mathcal {F}}^2 \twoheadrightarrow \widehat{K}\) with non-empty and compact values can be identified with the function \(B_\varGamma : \widehat{M} \times \widehat{\mathcal {F}}^2 \rightarrow A(\widehat{K})\) given by \(B_\varGamma (m,a)=\varGamma (m,a)\), in such form that \(\varGamma \) is continuous if and only if \(B_\varGamma \) is continuous [see Aliprantis and Border (2006, Lemma 3.97 and Theorem 17.15, pages 124 and 563)].

  28. Although maps \(\{g_n\}_{n \in {\mathbb {N}}}\) can take negative values, they are uniformly bounded from below (since \(\widehat{K}\) and \( T_1\) are compact sets, and \(H\) is continuous). Thus, as \(T_1\) has finite Lebesgue measure, we can apply the Fatou’s Lemma.

  29. In other words, for any \(t \in \tilde{T}_1\), there is at least one subsequence of \(\{g_n(t)\}_{n \in {\mathbb {N}}}\) converging to \(g(t)\).

  30. It is a direct consequence of the fact that, for any \(n \in {\mathbb {N}},\) we have

    $$\begin{aligned} |u_t^n(m_n,a_{n})-\overline{u}_t(\overline{m},\overline{a})|&\le \rho (\mathcal {G}_n,\mathcal {G})+ |\overline{u}_t(m_n,a_{n})-\overline{u}_t(\overline{m},\overline{a})|. \end{aligned}$$
  31. The function \(G\) is well defined, because \(\widehat{K}\) and \(\widehat{K}_t\), where \(t\in T_2\), are convex subsets of normed spaces with metrics induced by norms (see footnote 15).

  32. These properties are the core of the proof of equilibrium existence of Riascos and Torres-Martínez (2013).

  33. Remember that \(\lambda (z)=\frac{d(z,\overline{U}_2)}{d(z,\overline{U}_1)+d(z,\overline{U}_2)}.\)

  34. The connected components of a locally connected and compact space determine a partition of it into disjoint open sets. By compactness, this partition has finitely many elements [see Berge (1997, pages 98–100)].

  35. Indeed, since \((\varLambda (\kappa ({\mathcal {X}})){\setminus } A)\subset \varLambda (\kappa ({\mathcal {X}}))\), it is sufficient to ensure that it is closed. Let \(\{(m_n,a_n)\}_{n\in {\mathbb {N}}} \subset (\varLambda (\kappa ({\mathcal {X}})){\setminus } A)\) be a sequence that converges to \((m_0,a_0)\in \widehat{M}\times \widehat{\mathcal {F}}_2\). For any \(n \in {\mathbb {N}}, (m_n,a_n)\in \varLambda (\kappa ({\mathcal {X}}))\) and \((m_n,a_n)\notin A\). Thus, \((m_0,a_0) \in \varLambda (\kappa ({\mathcal {X}}))\). Furthermore, if \((m_0,a_0)\in A\), then for \(n\) large enough \((m_n,a_n)\in B[\pi , A]\), which is a contradiction with \(B[\pi , \varLambda (\kappa ({\mathcal {X}})){\setminus } A] \cap B[\pi , A] =\emptyset \). Therefore, \((m_0,a_0)\in \varLambda (\kappa ({\mathcal {X}})) {\setminus } A\).

  36. Following the notation used in the proof of Lemma 2, we can take \(\xi =\frac{0.5 \,\delta }{\# T_2}.\)

References

  • Aliprantis, C., Border, K.: Infinite Dimensional Analysis. Springer, Berlin, (2006)

  • Al-Najjar, N.: Strategically stable equilibria in games with infinitely many pure strategies. Math. Soc. Sci. 29, 151–164 (1995)

    Article  Google Scholar 

  • Aubin, J.P.: Mathematical Methods of Games and Economic Theory. North-Holland, Amsterdam (1982)

  • Aumann, R.J.: Integrals of set-valued functions. J. Math. Anal. App. 12, 1–12 (1965)

    Article  Google Scholar 

  • Balder, E.J.: A unifying pair of Cournot-Nash equilibrium existence results. J. Econ. Theory 102, 437–470 (2002)

    Article  Google Scholar 

  • Balder, E.J.: On the existence of Cournot-Nash equilibria in continuum games. J. Math. Econ. 32, 207–223 (1999)

    Article  Google Scholar 

  • Barelli, P., Meneghel, I.: A note on the equilibrium existence problem in discontinuous games. Econometrica 81, 813–824 (2012)

    Google Scholar 

  • Barelli, P., Soza, I.: On the existence of Nash equilibria in discontinuous and qualitative games, mimeo. University of Rochester,   (2009)

  • Barlo, M., Carmona, G.: Strategic behavior in non-atomic games, working paper, Munich Personal REPEC Arquive, http://mpra.ub.uni-muenchen.de/35549 (2011)

  • Berge, C.: Topological Spaces. Dover Publications, New York (1997)

  • Carbonell-Nicolau, O.: Essential equilibria in normal-form games. J. Econ. Theory 145, 421–431 (2010)

    Article  Google Scholar 

  • Carmona, G., Podczeck, K.: Existence of Nash equilibrium in games with a measure space of players and discontinuous payoff functions. J. Econ. Theory 152, 130–178 (2014)

    Article  Google Scholar 

  • Fort, M.K.: Essential and non-essential fixed points. Am. J. Math. 72, 315–322 (1950)

    Article  Google Scholar 

  • Fort, M.K.: A unified theory of semi-continuity. Duke Math. J. 16, 237–246 (1949)

    Article  Google Scholar 

  • Hildenbrand, W.: Core and Equilibria in a Large Economy. Princeton University Press, Princeton, (1974)

  • Hillas, J.: On the definition of the strategic stability of equilibria. Econometrica 58, 1365–1390 (1990)

    Article  Google Scholar 

  • Jiang, J.H.: Essential component of the set of fixed points of the multivalued mappings and its application to the theory of games. Sci. Sin. XII, 951–964 (1963)

    Google Scholar 

  • Jiang, J.H.: Essential fixed points of the multivalued mappings. Sci. Sin. XI, 293–298 (1962)

    Google Scholar 

  • Kinoshita, S.: On essential components of the set of fixed points. Osaka Math. J. 4, 19–22 (1952)

    Google Scholar 

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

    Article  Google Scholar 

  • Rath, K.P.: A direct proof of the existence of pure strategy equilibria in games with a continuum of players. Econ. Theory 2, 427–433 (1992)

    Article  Google Scholar 

  • Reny, P.: Strategic approximations of discontinuous games. Econ. Theory 48, 17–29 (2011)

    Article  Google Scholar 

  • Reny, P.: On the existence of pure and mixed strategy Nash equilibria in discontinuous games. Econometrica 67, 1029–1056 (1999)

    Article  Google Scholar 

  • Riascos, A.J., Torres-Martínez, J.P.: On pure strategy equilibria in large generalized games, Munich Personal REPEC Archive, http://mpra.ub.uni-muenchen.de/46840 (2013)

  • Scalzo, V.: Essential equilibria of discontinuous games. Econ. Theory 54(1), 27–44 (2013)

    Article  Google Scholar 

  • Schmeidler, D.: Equilibrium point of non-atomic games. J. Stat. Phys. 17, 295–300 (1973)

    Article  Google Scholar 

  • Yu, J.: Essential equilibrium of n-person noncooperative games. J. Math. Econ. 31, 361–372 (1999)

    Article  Google Scholar 

  • Yu, J., Yang, H.: Essential components of the set of equilibrium points for set-valued maps. J. Math. Anal. Appl. 300, 334–342 (2004)

    Article  Google Scholar 

  • Yu, J., Yang, H., Xiang, S.: Unified approach to existence and stability of essential components. Nonlinear Anal. 63, 2415–2425 (2005)

    Article  Google Scholar 

  • Yu, X.: Essential components of the set of equilibrium points for generalized games in the uniform topological space of best reply correspondences. Int. J. Pure Appl. Math. 55, 349–357 (2009)

    Google Scholar 

  • Wu, W., Jiang, J.H.: Essential equilibrium points of n-person noncooperative games. Sci. Sin. XI, 1307–1322 (1962)

    Google Scholar 

  • Zhou, Y., Yu, J., Xiang, S.: Essential stability in games with infinitely many pure strategies. Int. J. Game Theory 35, 493–503 (2007)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Juan Pablo Torres-Martínez.

Additional information

This work has greatly benefited from the suggestions of two anonymous referees, who encouraged us to include the results of essential stability for discontinuous games. Correa acknowledges financial support from Conicyt (Chilean Research Council) and University of Chile through graduate fellowships.

Torres-Martínez acknowledges the financial support of Conicyt through Fondecyt project 1120294.

Appendix

Appendix

Lemma 1

The set of messages \(\widehat{M}\) is non-empty and compact.

Proof

Since \(t \in T_1 \twoheadrightarrow K_t\) is measurable, from Aliprantis and Border (2006, Lemma 18.2, and Theorem 18.6), we know that this correspondence has an \({\mathcal {A}}\times {\mathcal {B}}(\widehat{K})\)-measurable graph. It follows from Aumann’s Selection Theorem [see Aliprantis and Border (2006, Theorem 18.26, page 608)] that there exists an \({\mathcal {A}}\)-measurable function \(g:T_{1}\rightarrow \widehat{K}\) such that, \(g(t)\in K_t,\,\forall t \in T_1\). Hence, the compactness of \(T_1\) and \(\widehat{K}\) and the continuity of \(H\) guarantee that the map \(t \rightarrow H(t, g(t))\) is bounded and measurable. Therefore, \(\widehat{M}\) is non-empty.

Since \(H\) is continuous, \(\mu \) is a finite measure, and the sets \(T_1\) and \( \widehat{K}\) are compact, the correspondence \(t\in T_1 \twoheadrightarrow H(t,\widehat{K})\) is integrable bounded and has closed values. Thus, as \(\widehat{M}={\int _{T_1} } H(t, \widehat{K}) d\mu \), it follows from Aumann (1965, Theorem 4) that \(\widehat{M}\) is compact. \(\square \)

Let us define some notations. Given a metric space \((S,d)\), consider the sets

$$\begin{aligned} A(S)&= \{K\subseteq S: K\,\,\text{ is } \text{ non-empty } \text{ and } \text{ compact }\},\\ A_c(S)&= \{C\subseteq A(S): C\,\,\text{ is } \text{ convex }\}. \end{aligned}$$

Denote by \(d_H\) the Hausdorff metric induced by the metric of \(S\). If \(S\) is compact, then \((A(S), d_H)\) is a complete metric space. Also, when \(S\) is compact and convex, \((A_c(S),d_H)\) is complete.Footnote 26

Given a set \(X\), let \({\mathcal {U}}(X)\) be the collection of bounded functions \(u:X\rightarrow {\mathbb {R}}\) endowed with the sup norm topology, i.e., the topology determined by the metric \(d(u_1,u_2)=\sup _{x\in X} \vert u_1(x) - u_2(x)\vert .\)

Proof of Proposition 1

Let \(\{{\mathcal {G}}_n\}_{n\in {\mathbb {N}}}\), with \({\mathcal {G}}_n= {\mathcal {G}}_n((K_{n,t},\varGamma _{n,t},u_{n,t})_{t\in T_1 \cup T_2})\), be a Cauchy sequence on \(({\mathbb {G}},\rho )\). It follows from definition of \({\mathbb {G}}\) and \(\rho \) that for any non-atomic player \(t\in T_1, \{K_{n,t}\}_{n\in {\mathbb {N}}}\) is a Cauchy sequence on \((A(\widehat{K}),d_H)\). Also, for any atomic player \(s \in T_2, \{K_{n,s}\}_{n\in {\mathbb {N}}}\) is a Cauchy sequence on \((A_c(\widehat{K}_s),d_{H,s})\). Hence, there are sets \(\{\overline{K}_t\}_{t\in T_1\cup T_2}\) such that (i) \((\overline{K}_t,\overline{K}_s)\in A(\widehat{K}) \times A_c(\widehat{K}_s),\,\forall (t,s)\in T_1\times T_2;\) and (ii) for any \((t,s)\in T_1\times T_2,\) we have that \(\lim _{n\rightarrow +\infty } d_H(K_{n,t},\overline{K}_t)= \lim _{n\rightarrow +\infty } d_{H,s}(K_{n,s},\overline{K}_s)=0.\)

The definition of the metric \(\rho \) ensures that, for any \(t\in T_1\) and \((m,a)\in \widehat{M}\times \widehat{\mathcal {F}}^2\), the sequence \(\{\varGamma _{n,t}(m,a)\}_{n\in {\mathbb {N}}}\subseteq A(\widehat{K})\) is Cauchy and, therefore, there exists a set \(K_t(m,a)\in A(\widehat{K})\) such that \(d_H(\varGamma _{n,t}(m,a),K_t(m,a))\) converges to zero as \(n\) goes to infinity. Let \(\overline{\varGamma }_t: \widehat{M}\times \widehat{\mathcal {F}}^2 \twoheadrightarrow \widehat{K}\) be the set-valued mapping defined by \(\overline{\varGamma }_t(m,a)=K_t(m,a).\) It follows that correspondences \((\overline{\varGamma }_t)_{t\in T}\) are continuous.Footnote 27 By analogous arguments, we can obtain that for any \(s\in T_2\), there is a continuous correspondence \(\overline{\varGamma }_s: \widehat{M}\times \widehat{\mathcal {F}}^2_{-s} \twoheadrightarrow \widehat{K}_s\) such that, for each \((m,a_{-s})\in \widehat{M}\times \widehat{\mathcal {F}}^2_{-s} \) both \(\overline{\varGamma }_{s}(m,a_{-s})\in A_c(\widehat{K}_s)\) and \(d_{H,s}(\varGamma _{n,s}(m,a_{-s}),\overline{\varGamma }_s(m,a_{-s}))\) converges to zero as \(n\) increases.

Since \(\{{\mathcal {G}}_n\}_{n\in {\mathbb {N}}}\) is Cauchy on \(({\mathbb {G}},\rho )\), there is a bounded function \(U:T_1\times \widehat{K}\times \widehat{M}\times \widehat{\mathcal {F}}^2 \rightarrow {\mathbb {R}}\) such that, for each \(t\in T_1\), the sequence \(\{u_{n,t}\}_{n\in {\mathbb {N}}}\subseteq {\mathcal {U}}(\widehat{K}\times \widehat{M}\times \widehat{\mathcal {F}}^2)\) converges uniformly to \(\overline{u}_t:=U(t,\cdot )\) and, therefore, \(\overline{u}_t\) is continuous. Analogously, for any \(t\in T_2\), the sequence \(\{u_{n,t}\}_{n\in {\mathbb {N}}}\subseteq {\mathcal {U}}( \widehat{M}\times \widehat{\mathcal {F}}^2)\) converges to some continuous function \(\overline{u}_t\in {\mathcal {U}}(\widehat{M}\times \widehat{\mathcal {F}}^2)\) that is quasi-concave on \(a_t\).

Let \(\overline{\mathcal {G}}= \overline{\mathcal {G}}((\overline{K}_{t},\overline{\varGamma } _{t},\overline{u}_{t})_{t\in T_1 \cup T_2})\). It follows from arguments above that \(\lim _{n\rightarrow +\infty } \rho ({\mathcal {G}}_n, \overline{\mathcal {G}})=0\). Thus, to conclude that \(({\mathbb {G}},\rho )\) is complete, it is sufficient to guarantee that

  1. (i)

    for each \((m,a)\in \widehat{M} \times \widehat{\mathcal {F}}^2, (t,x) \in T_1 \times \widehat{K} \rightarrow \overline{u}_t(x,m,a) \) is measurable;

  2. (ii)

    for each \((m,a)\in \widehat{M}\times \widehat{\mathcal {F}}^2, t\in T_1 \twoheadrightarrow \overline{\varGamma }_t(m,a)\) is measurable;

  3. (iii)

    the correspondence \(t\in T_1 \twoheadrightarrow \overline{K}_t\) is measurable.

Fix \((m,a) \in \widehat{M} \times \widehat{\mathcal {F}}^2\), the definition of \(\rho \) ensures that measurable functions \((t,x) \in T_1\times \widehat{K} \rightarrow u_{n,t}(x,m,a)\) converge to the mapping \((t,x) \in T_1 \times \widehat{K} \rightarrow \overline{u}_t(x,m,a) \). Since \((T_1\times \widehat{K}, {\mathcal {A}}\times {\mathcal {B}}(\widehat{K})) \) is a measurable space, item (i) holds [see Aliprantis and Border (2006, Lemma 4.29, page 142)]. Furthermore, since for every \(n\in {\mathbb {N}}\) the correspondence \(t\in T_1 \twoheadrightarrow \varGamma _{n,t}(m,a)\) is measurable, it follows from Aliprantis and Border (2006, Theorem 18.10, page 598) that the function \(\varTheta _{n,(m,a)}:T_1\rightarrow A(\widehat{K})\) defined by \(\varTheta _{n,(m,a)}(t)=\varGamma _{n,t}(m,a)\) is Borel measurable. Also, the sequence \(\{\varTheta _{n,(m,a)}\}_{n\in {\mathbb {N}}}\) converges to \(\overline{\varTheta }_{(m,a)}:T_1\rightarrow A(\widehat{K})\), where \(\overline{\varTheta }_{(m,a)}(t)=\overline{\varGamma }_t(m,a)\). By Aliprantis and Border (2006, Lemma 4.29), \(\overline{\varTheta }_{(m,a)}\) is a Borel measurable function. Thus, \(t\in T_1 \twoheadrightarrow \overline{\varGamma }_t(m,a)\) is measurable [see Aliprantis and Border (2006, Theorem 18.10)]. By analogous arguments, we obtain item (iii). \(\square \)

Proof of Theorem 1

The proof is a direct consequence of the following steps.

Step 1. \(\varLambda :{\mathbb {G}} \twoheadrightarrow \widehat{M} \times \widehat{\mathcal {F}}^2\) is upper hemicontinuous with compact values.

Since \( \widehat{M}\times \widehat{\mathcal {F}}^2\) is compact and non-empty, we only need to prove that \(\text{ Graph }(\varLambda )\) is closed, where \(\text{ Graph }(\varLambda ):=\left\{ (\mathcal {G},(m,a))\in \mathbb {G}\times \widehat{M}\times \widehat{\mathcal {F}}^2: (m,a)\in \varLambda ({\mathcal {G}})\right\} \). Let \(\{(\mathcal {G}_n,(m_n,a_n))\}_{n \in {\mathbb {N}}}\subset \text{ Graph }(\varLambda )\) be a sequence converging to \((\overline{\mathcal {G}},(\overline{m},\overline{a}))\in \mathbb {G}\times \widehat{M}\times \widehat{\mathcal {F}}^2\), where \(\overline{\mathcal {G}}= \overline{\mathcal {G}}((\overline{K}_{t},\overline{\varGamma }_{t},\overline{u}_{t})_{t\in T_1 \cup T_2})\) and, for every \(n \in {\mathbb {N}}, {\mathcal {G}}_n= {\mathcal {G}}_n((K^n_{t},\varGamma ^n_{t},u^n_{t})_{t\in T_1 \cup T_2})\).

We aim to ensure that \((\overline{m},\overline{a}) \in \varLambda ({\overline{\mathcal {G}}})\). Since for any \(n \in {\mathbb {N}}, (m_n,a_n)\in \varLambda ({\mathcal {G}_n})\), there exists \(f_n \in \widehat{\mathcal {F}}^1\) such that (i) the function \(g_n:T_1\rightarrow {\mathbb {R}}^m\) given by \(g_n(t)=H(t,f_n(t))\) is measurable and \(m_n=m(f_n)\); and (ii) for almost all \(t \in T_1\) both \(f_n(t)\in \varGamma ^n_t(m_n,a_n)\) and

$$\begin{aligned} u_t^n(f_n(t), m_n,a_{n})=\max \limits _{x\in \varGamma _t^n(m_n, a_n)}{u_t^n(x, m_n,a_{n})}. \end{aligned}$$

Claim A

There exists \(\overline{f}\in \widehat{\mathcal {F}}^1\) such that \(\overline{m}=\int _{T_1}{H(t,\overline{f}(t))d\mu }\).

Proof

Since \(H\) is continuous, \(T_1\) is compact and \(\{f_n\}_{n \in {\mathbb {N}}} \subset \widehat{\mathcal {F}}^1\), it follows that the sequence \(\{g_n\}_{n\in \mathbb {N}}\) is uniformly integrable [see Hildenbrand (1974, page 52)]. In addition, \(\{\int _{T_1}{g_n(t)d\mu }\}_{n \in {\mathbb {N}}} \subset {\mathbb {R}}^m\) converges to \(\overline{m}\) as \(n\) goes to infinity , and therefore, the multidimensional version of Fatou’s Lemma [see Hildenbrand (1974, page 69)] guarantees that there is \(g:T_1 \rightarrow {\mathbb {R}}^m\) integrable such that, Footnote 28

  1. (1)

    \(\overline{m}=\lim \limits _{n\rightarrow \infty }{\int _{T_1}{g_n(t)d\mu }}=\int _{T_1}{g(t)}d\mu \);

  2. (2)

    there exists a full measure set \(\tilde{T}_1\subseteq T_1\) such that, for any \(t \in \tilde{T}_1, g(t)\in L_S(g_n(t)),\) where \(L_S(g_n(t))\) is the set of cluster points of \(\{g_n(t)\}_{n\in \mathbb {N}}\).Footnote 29

Fix \(t\in \tilde{T}_1\). Then, there is a subsequence \(\left( g_{n_k}(t)\right) _k\) converging to \(g(t)\). Since \(\{f_{n_k}(t)\}_{k\in \mathbb {N}}\subseteq \widehat{K}\), by taking a subsequence if it is necessary, we can ensure that there exists \(f(t)\in \widehat{K}\) such that both \(f_{n_k}(t) \rightarrow f(t)\) and \(g(t)=\lim \limits _{k\rightarrow \infty }{H(t,f_{n_k}(t))}=H(t,f(t))\) hold.

Let \(\overline{f}:T_1 \rightarrow {\widehat{K}}\) be a function such that

$$\begin{aligned} \overline{f}(t)\in \left\{ \begin{array}{ll} \{ f(t)\},&{}\quad \text{ if }\;\;t\in \;\tilde{T}_1, \\ \overline{\varGamma }_t(\overline{m},\overline{a}),&{}\quad \text{ if }\;\; t\in \; T {\setminus } \tilde{T}_1. \end{array} \right. \end{aligned}$$

Then, it follows that

$$\begin{aligned} \overline{m}=\lim _{n \rightarrow \infty } m_n= \lim \limits _{n\rightarrow \infty }{\int \limits _{T_1} g_n(t) d\mu }=\int \limits _{T_1} g(t) d\mu =\int \limits _{T_1}H(t,\overline{f}(t))d\mu , \end{aligned}$$

where the last equality follows from the fact that \(T_1 {\setminus } \tilde{T}_1\) has zero measure. \(\square \)

Claim B

For almost all \(t\in T_1, \overline{f}(t)\in \overline{\varGamma }_t(\overline{m},\overline{a})\). In addition, for any \(t \in T_2, \overline{a}_t \in \overline{\varGamma }_t(\overline{m},\overline{a}_{-t})\).

Proof

Following the notation of the proof of the previous claim, fix \(t \in \tilde{T}_1\) and let \(\{f_{n_k}(t)\}_{k \in {\mathbb {N}}}\) be the sequence converging to \(\overline{f}(t)\) and that was obtained in the previous claim. We know that, for any \(k \in \mathbb {N}, f_{n_k}(t)\in \varGamma _t^{n_k}(m_{n_k},a_{n_k})\). Therefore,

$$\begin{aligned} d(\overline{f}(t),\overline{\varGamma }_t(\overline{m},\overline{a}))&\le \widehat{d}(\overline{f}(t), f_{n_k}(t)) +d_H(\varGamma _t^{n_k}(m_{n_k},a_{n_k}),\overline{\varGamma }_t(m_{n_k},a_{n_k}))\\&+d_H(\overline{\varGamma }_t(m_{n_k},a_{n_k}),\overline{\varGamma }_t(\overline{m},\overline{a}))\\&\le \widehat{d}(\overline{f}(t) , f_{n_k}(t))+\rho ({\mathcal {G}}_{n_k}, \overline{\mathcal {G}}) + d_H(\overline{\varGamma }_t(m_{n_k},a_{n_k}),\overline{\varGamma }_t(\overline{m},\overline{a})), \end{aligned}$$

where \(\widehat{d}\) denotes the metric of the compact metric space \(\widehat{K}\). Since \(\overline{\varGamma }_t\) is continuous, by taking the limit as \(k\) goes to infinity, we obtain the first property.

On the other hand, for any \((t,n) \in T_2 \times {\mathbb {N}}, a_{n,t}\in \varGamma _t^n(m_n,a_{n,-t})\), which implies that

$$\begin{aligned} d(\overline{a}_t,\overline{\varGamma }_t(\overline{m},\overline{a}_{-t}))&\le \widehat{d}_t(\overline{a}_t ,a_{n,t})+d_{H,t}(\varGamma _t^n(m_n,a_{n,-t}),\overline{\varGamma }_t(m_n,a_{n,-t}))\\&+d_{H,t}(\overline{\varGamma }_t(m_n,a_{n,-t}),\overline{\varGamma }_t(\overline{m},\overline{a}_{-t}))\\&\le \widehat{d}_t(\overline{a}_t , a_{n,t})+\rho ({\mathcal {G}}_n, \overline{\mathcal {G}}) + d_{H,t}(\overline{\varGamma }_t(m_n,a_{n,-t}),\overline{\varGamma }_t(\overline{m},\overline{a}_{-t})), \end{aligned}$$

where \(\widehat{d}_t\) denotes the metric of \(\widehat{K}_t\). Taking the limit as \(n\) goes to infinity, we obtain the result. \(\square \)

Claim C

The following properties hold:

  1. (i)

    For almost all \(t \in T_1, \overline{f}(t) \in \mathop {\hbox {argmax}}\nolimits _{x \in \overline{\varGamma }_t(\overline{m},\overline{a})}\,\,\overline{u}_t(x,\overline{m},\overline{a}).\)

  2. (ii)

    For any \(t\in T_2, \overline{a}_t\in \mathop {\hbox {argmax}}\nolimits _{x\in \overline{\varGamma }_t(\overline{m},\overline{a}_{-t})}\overline{u}_t(\overline{m},x,\overline{a}_{-t}).\)

Proof

(i) Given \(t \in \tilde{T}_1\), we have that

$$\begin{aligned} d_H(\varGamma _t^{n_k}(m_{n_k},a_{n_k}),\overline{\varGamma }_t(\overline{m},\overline{a}))&\le \rho (\mathcal {G}_{n_k},\overline{\mathcal {G}})+d_H(\overline{\varGamma }_t(m_{n_k},a_{n_k}),\overline{\varGamma }_t(\overline{m},\overline{a})). \end{aligned}$$

Then, \(\varGamma _t^{n_k}(m_{n_k},a_{n_k})\longrightarrow _{k}\,\, \overline{\varGamma }_t(\overline{m},\overline{a})\). Since \(u^{n_k}_t\) converges uniformly to \(\overline{u}_t\), it follows from Yu (1999, Lemma 2.5) and Aubin (1982, Theorem 3, page 70) that

$$\begin{aligned} u_t^{n_k}(f_{n_k}(t),m_{n_k},a_{n_k})=\max _{x\in \varGamma _t^{n_k}(m_{n_k},a_{n_k})}{u_t^{n_k}(x,m_{n_k},a_{n_k})}\longrightarrow _k\,\, \max _{x\in \overline{\varGamma }_t(\overline{m},\overline{a})}{\overline{u}_t(x,\overline{m},\overline{a})} \end{aligned}$$

On the other hand,

$$\begin{aligned} |u_t^{n_k}(f_{n_k}(t),m_{n_k},a_{n_k})-\overline{u}_t(\overline{f}(t),\overline{m},\overline{a})| \\ \le \rho (\mathcal {G}_{n_k},\overline{\mathcal {G}})+ |\overline{u}_t(f_{n_k}(t),m_{n_k},a_{n_k})-\overline{u}_t(\overline{f}(t),\overline{m},\overline{a})|. \end{aligned}$$

Taking the limit as \(k\) goes to infinity, we obtain that \(u_t^{n_k}(f_{n_k}(t),m_{n_k},a_{n_k})\rightarrow \overline{u}_t(\overline{f}(t),\overline{m},\overline{a})\). Hence, it follows from Claim B that \(\overline{f}(t) \in \mathop {\hbox {argmax}}\limits _{x\in \overline{\varGamma }_t(\overline{m},\overline{a})}{\overline{u}_t(x,\overline{m},\overline{a})}\).

(ii) Given \(t \in T_2\), analogous arguments to those made in the previous item ensure that

$$\begin{aligned} d_{H,t}(\varGamma _t^n(m_n,a_{n,-t}),\overline{\varGamma }_t(\overline{m},\overline{a}_{-t}))&\le \rho (\mathcal {G}_n,\mathcal {G})+d_{H,t}(\overline{\varGamma }_t(m_n,a_{n,-t}),\overline{\varGamma }_t(\overline{m},\overline{a}_{-t})), \end{aligned}$$

which implies that \(\varGamma _t^n(m_n,a_{n,-t})\) converges to \(\overline{\varGamma }_t(\overline{m},\overline{a}_{-t})\) as \(n\) goes to infinity. Hence, Yu (1999, Lemma 2.5) ensures that

$$\begin{aligned} u_t^n(m_n,a_{n})=\max _{x\in \varGamma _t^n(m_n,a_{n,-t})}{u_t^n(m_n,x,a_{n,-t})}\longrightarrow \max _{x\in \overline{\varGamma }_t(\overline{m},\overline{a}_{-t})}{\overline{u}_t(\overline{m},x,\overline{a}_{-t})}. \end{aligned}$$

Since \(\lim \limits _{n \rightarrow +\infty }\,\,u_t^n(m_n,a_{n})=\overline{u}_t(\overline{m},\overline{a})\),Footnote 30 \(\overline{a}_t\in \mathop {\hbox {argmax}}\limits _{x\in \overline{\varGamma }_t(\overline{m},\overline{a}_{-t})}\overline{u}_t(\overline{m},x,\overline{a}_{-t}).\) \(\square \)

It follows from Claims A and C that \((\overline{m},\overline{a})\in \varLambda ({\overline{\mathcal {G}}})\). Thus, we ensure that \(\varLambda \) is an upper hemicontinuous correspondence with compact values.

Step 2. There is a dense residual set \(Q\subseteq {\mathbb {G}}'\) where \(\varLambda \) is lower hemicontinuous.

Since \({\mathbb {G}}'\) is a closed subset of \({\mathbb {G}}\), it follows that \(({\mathbb {G}}',\rho )\) is a complete metric space and, therefore, it is a Baire space. Since the correspondence \(\varLambda \) is non-empty, compact-valued, and upper hemicontinuous, it follows from Lemmas 5 and 6 in Carbonell-Nicolau (2010) [see also Fort (1949) and Jiang (1962)] that there exists a dense residual subset \(Q\) of \(\mathbb {G}'\) in which \(\varLambda \) is lower hemicontinuous.

Step 3. If \(\mathcal {G}\in {\mathbb {G}}'\) is a point of lower hemicontinuity of \(\varLambda \), then \(\mathcal {G}\) is essential with respect to \({\mathbb {G}}'\).

Fix \((f^*,a^*) \in \text{ CN }({\mathcal {G}})\). Then, for any open neighborhood \(O\subseteq \widehat{M}\times \widehat{\mathcal {F}}^2\) of \( (m(f^*),a^*)\), we have \(\varLambda (\mathcal {G})\cap O\ne \emptyset \), and therefore, by the lower hemicontinuity, we have that \(\left\{ \mathcal {G}'\in \mathbb {G}'\,:\;\;\varLambda (\mathcal {G}')\cap O\ne \emptyset \right\} \) contains a neighborhood of \({\mathcal {G}}\), that is, for some \(\epsilon >0\) and for any \({\mathcal {G}}'\in {\mathbb {G}}'\) such that \(\rho ({\mathcal {G}}', {\mathcal {G}})<\epsilon \), we have \(\varLambda (\mathcal {G}')\cap O\ne \emptyset \). Hence, all Cournot-Nash equilibria of \({\mathcal {G}}\) are essential with respect to \({\mathbb {G}}'\).

It follows from Steps 2 and 3 that any game in \(Q\) is essential.

Finally, suppose that for some \({\mathcal {G}}\in {\mathbb {G}}'\) the set \(\varLambda (\mathcal {G}) \) is a singleton. Then, the upper hemi-continuity of \(\varLambda \) guarantees that it is continuous at \(\mathcal {G}\). Finally, Step 3 implies that \({\mathcal {G}}\) is an essential generalized game with respect to \({\mathbb {G}}'\). \(\square \)

Proof of Theorem 2

(i) Existence of a minimal essential set.

Fix \(\mathcal {G}\in \mathbb {G}'\). Let \(\mathcal {S}\) be the family of essential subsets of \(\varLambda (\mathcal {G})\) with respect to \({\mathbb {G}}'\) ordered by set inclusion. Since \(\varLambda \) is upper hemicontinuous, \(\varLambda (\mathcal {G})\in \mathcal {S}\) and, hence, \(\mathcal {S}\ne \emptyset \). As any essential set is non-empty and compact, each totally ordered subset of \(\mathcal {S}\) has a lower bound. By Zorn’s Lemma, \(\mathcal {S}\) has a minimal element and, by definition, it is an essential set of \(\varLambda (\mathcal {G})\) with respect to \({\mathbb {G}}'\).

(ii) If there are connected essential sets, then there are essential components.

Suppose that there is a connected essential set of \(\varLambda (\mathcal {G})\) with respect to \({\mathbb {G}}'\), denoted by \(c({\mathcal {G}})\). Since \(c({\mathcal {G}})\) is non-empty, fix \((\widehat{m},\widehat{a})\in c(\mathcal {G})\) and consider the set \(\varLambda _{(\widehat{m},\widehat{a})}(\mathcal {G})\) defined as the union of all connected subsets of \(\varLambda (\mathcal {G})\) that contains \((\widehat{m},\widehat{a})\). By definition, \(\varLambda _{(\widehat{m},\widehat{a})}(\mathcal {G})\) is a component of \(\varLambda (\mathcal {G})\). As the closure of a connected set is connected and \(\varLambda ({\mathcal {G}})\) is compact, it follows that \(\varLambda _{(\widehat{m},\widehat{a})}(\mathcal {G})\) is compact. Hence, the essentiality of \(c(\mathcal {G})\subseteq \varLambda _{(\widehat{m},\widehat{a})}(\mathcal {G})\) with respect to \({\mathbb {G}}'\) ensures that the component \(\varLambda _{(\widehat{m},\widehat{a})}(\mathcal {G})\) is also an essential subset of \(\varLambda (\mathcal {G})\) with respect to \({\mathbb {G}}'\).

(iii) Connectedness of minimal essential sets in normed spaces.

Suppose that \(\widehat{K}\) is a convex subset of a normed space and it is equipped with a metric induced by a norm. Fix a minimal essential set of \(\varLambda ({\mathcal {G}})\) with respect to \({\mathbb {G}}'\), denoted by \(m(\mathcal {G})\). Suppose, by contradiction, that \(m(\mathcal {G})\) is disconnected.

Claim A

There are open sets \(U_1, U_2 \subseteq \widehat{M} \times \widehat{\mathcal {F}}^2\) such that \(m({\mathcal {G}})\subset U_1 \cup U_2\) and \(\overline{U}_1\cap \overline{U}_2=\emptyset \).

Proof

Since \(m(\mathcal {G})\) is disconnected, there are closed and non-empty subsets \(A_1, A_2 \subseteq \varLambda (\mathcal {G})\) such that \(A_1 \cap A_2 = \emptyset \) and \(m(\mathcal {G})=A_1\cup A_2\). Since \(m(\mathcal {G})\) is minimal, neither \(A_1\) nor \(A_2\) are essential with respect to \({\mathbb {G}}'\). Hence, for each \(i \in \{1,2\}\), there exists an open set \(U_i\) such that \(A_i\subset U_i\) and for all \(\epsilon >0\), there exists \(\mathcal {G}^i_\epsilon \in \mathbb {G}'\) such that both \(\rho (\mathcal {G},\mathcal {G}^i_\epsilon )<\epsilon \) and \(\varLambda (\mathcal {G}^i_\epsilon )\cap U_i=\emptyset \). Since \(A_i\) is compact, we can assume that \(\overline{U}_1\cap \overline{U}_2=\emptyset \). \(\square \)

Claim B

There are large generalized games \({\mathcal {G}}_1, {\mathcal {G}}_2 \in {\mathbb {G}}'\) such that

$$\begin{aligned} \varLambda ({\mathcal {G}}_i) \cap U_i =\emptyset \quad \wedge \quad \varLambda ({\mathcal {G}}_i) \cap U_j \ne \emptyset ,\,\,\,\,\quad \forall i,j \in \{1,2\}: i \ne j. \end{aligned}$$

In addition, there is a continuous map \(G:\widehat{M}\times \widehat{\mathcal {F}}^2\rightarrow {\mathbb {G}}\) such that, for every \((m,a)\in \widehat{M}\times \widehat{\mathcal {F}}^2\),

$$\begin{aligned} \varLambda (G(m,a))\cap (U_1 \cup U_2)\ne \emptyset ,\quad \quad (G(m,a)={\mathcal {G}}_i \Longleftrightarrow (m,a)\in \overline{U}_i),\,\forall i \in \{1,2\}. \end{aligned}$$

Proof

As \(m(\mathcal {G})\) is essential with respect to \({\mathbb {G}}'\), there exists \(\nu >0\) such that for every \(\mathcal {G}'\in \mathbb {G}'\) with \(\rho (\mathcal {G},\mathcal {G}')<\nu \), we have \(\varLambda (\mathcal {G}')\cap (U_1 \cup U_2)\ne \emptyset \). Following the notation of the previous claim, for each \(i \in \{1,2\}\), set \({\mathcal {G}}_i={\mathcal {G}}^i_{\nu /3}\). Hence, for \(i \ne j\), we obtain \(\varLambda (\mathcal {G}_i)\cap U_j\ne \emptyset \).

Let \(G:\widehat{M}\times \widehat{\mathcal {F}}^2\rightarrow {\mathbb {G}}\) be the functionFootnote 31

$$\begin{aligned} G(m,a)=\lambda (m,a) \mathcal {G}_1+(1-\lambda (m,a))\mathcal {G}_2,\,\,\,\,\, \forall (m,a)\in \widehat{M}\times \widehat{\mathcal {F}}^2, \end{aligned}$$

where \(\lambda :\widehat{M}\times \widehat{\mathcal {F}}^2\rightarrow [0,1]\) is the continuous function given by,

$$\begin{aligned} \lambda (m,a)=\frac{d((m,a),\overline{U}_2)}{d((m,a),\overline{U}_1)+d((m,a),\overline{U}_2)}. \end{aligned}$$

By construction, for each \(i \in \{1,2\}, G(m,a)=\mathcal {G}_i\) if and only if \((m,a)\in \overline{U}_i\).

Since metric spaces \(\widehat{K}\) and \(\{\widehat{K}_t\}_{t \in T_2}\) are contained in normed vectorial spaces and their metrics are induced by norms, for any \((m,a) \in \widehat{M}\times \widehat{\mathcal {F}}^2\), we can ensure that \(G(m,a)\) is well defined and

$$\begin{aligned} \rho (G(m,a), {\mathcal {G}}_1)&= \rho \left( \lambda (m,a) {\mathcal {G}}_1 + (1-\lambda (m,a)) {\mathcal {G}}_2, \lambda (m,a) {\mathcal {G}}_1 + (1-\lambda (m,a)) {\mathcal {G}}_1\right) \\&= \lambda (m,a) \rho ({\mathcal {G}}_1, {\mathcal {G}}_1) + (1-\lambda (m,a)) \rho ({\mathcal {G}}_2, {\mathcal {G}}_1) \\&\le \rho ( \,{\mathcal {G}}_2, {\mathcal {G}}_1) \le \rho ( \,{\mathcal {G}}_2, {\mathcal {G}}) + \rho ( \,{\mathcal {G}}, {\mathcal {G}}_1) < \frac{2\nu }{3}, \end{aligned}$$

which implies that \(\rho ({\mathcal {G}}, G(m,a))\le \rho ( \,{\mathcal {G}}, {\mathcal {G}}_1) + \rho ( \,{\mathcal {G}}_1, G(m,a)) < \nu .\) Hence, for each \((m,a) \in \widehat{M}\times \widehat{\mathcal {F}}^2, \varLambda (G(m,a))\cap (U_1 \cup U_2)\ne \emptyset \). \(\square \)

Given a large generalized game \({\mathcal {G}}\in {\mathbb {G}}\), let \(\varPhi _{\mathcal {G}}:\widehat{M}\times \widehat{\mathcal {F}}^2 \twoheadrightarrow \widehat{M}\times \widehat{\mathcal {F}}^2\) be the correspondence defined by \(\varPhi _{{\mathcal {G}}}(m,a)=\left( \varOmega ^{{\mathcal {G}}}(m,a),(B^{_{{\mathcal {G}}}}_t(m,a_{-t}))_{t\in T_2}\right) \), where

$$\begin{aligned} \varOmega ^{{\mathcal {G}}}(m,a)&:= \left\{ \int \limits _{T_1} H(t,f(t)) d\mu : H(\cdot , f(\cdot )) \,\text{ is } \text{ integrable } \wedge f(t)\in B^{\mathcal {G}}_t(m,a),\,\forall t \in T_1\right\} \!; \\ B^{{\mathcal {G}}}_t(m,a)&:= \mathop {\hbox {argmax}}\limits _{x_t\in \varGamma _t(m,a)}{u_t(x_t,m,a)},\,\,\,\,\,\,\forall t \in T_1;\\ B^{{\mathcal {G}}}_t(m,a_{-t})&:= \mathop {\hbox {argmax}}\limits _{x_t\in \varGamma _t(m,a_{-t})}{u_t(x_t, m,a_{-t})},\,\,\,\,\,\,\forall t \in T_2. \end{aligned}$$

We affirm that \(\varPhi _{{\mathcal {G}}}\) is upper hemicontinuous and has non-empty, compact, and convex values. Note that, Berge’s Maximum Theorem [see Aliprantis and Border (2006, Theorem 17.31, page 570)] guarantees that for any \(t \in T_1 \cup T_2\), the correspondence \(B^{\mathcal {G}}_t\) is upper hemicontinuous and has non-empty and compact values. Also, as atomic players have convex strategy sets and quasi-concave payoff functions, correspondences \((B^{\mathcal {G}}_t)_{t\in T_2}\) have convex values. Hence, we want to prove that \(\varOmega ^{\mathcal {G}}(\cdot )=\int _{T_1} H(t,B^{\mathcal {G}}_t(\cdot ))d\mu \) is upper hemicontinuous and has non-empty, compact, and convex values. Aumann (1965, Theorem 1) guarantees that \(\varOmega ^{\mathcal {G}}\) has convex values. Fix \((m,a) \in \widehat{M}\times \widehat{\mathcal {F}}^2 \). Since \(t\in T_1 \twoheadrightarrow \varGamma _t(m,a)\) is measurable, it follows from Aliprantis and Border (2006, Lemma 18.2, page 593) that \(t\in T_1 \twoheadrightarrow \varGamma _t(m,a)\) is weakly measurable. The Measurable Maximum Theorem (Aliprantis and Border (2006, Theorem 18.19, page 605)) implies that \(t\in T_1 \twoheadrightarrow H(t,B^{\mathcal {G}}_t(m,a))\) has a measurable selector. Since \(H\) is continuous, the compactness of \(T_1\) and \(\widehat{K}\) guarantees that \(t\in T_1 \rightarrow H(t,B^{\mathcal {G}}_t(m,a))\) is bounded and, therefore, its measurable selectors are integrable. We conclude that \(\varOmega ^{\mathcal {G}}(m,a)\) is non-empty. Since \(T_1\) has finite measure and \(t\in T_1 \twoheadrightarrow H(t,B^{\mathcal {G}}_t(m,a))\) is bounded and has closed values, it follows from Aumann (1965, Theorem 4) that \(\varOmega ^{\mathcal {G}}(m,a)\) is compact. Finally, as \(H\) is continuous, \(\widehat{K}\) is compact, and for every \(t \in T_1\) the correspondence \((m,a) \twoheadrightarrow B^{\mathcal {G}}_t(m,a)\) has closed graph; it follows that \((m,a) \twoheadrightarrow H(t,B^{\mathcal {G}}_t(m,a))\) is upper hemicontinuous for each \(t \in T_1\). Thus, it follows from Aumann (1965, Corollary 5.2) that \(\varOmega ^{\mathcal {G}}\) is upper hemicontinuous.

Therefore, Kakutani’s Fixed Point Theorem implies that the set of fixed points of \(\varPhi _{{\mathcal {G}}}\) is non-empty and compact. Note that \((f^*,a^*)\) is a Cournot-Nash equilibrium of \({\mathcal {G}}\) if and only if \((m^*,a^*)\in \widehat{M}\times \widehat{\mathcal {F}}^2\) is a fixed point of \(\varPhi _{{\mathcal {G}}}\), where \(m^*=\int _{T_{1}}H(t,f^{*}(t))d\mu \).Footnote 32

Claim C

There exists \((\overline{m},\overline{a}) \in U_1\) such that \((\overline{m},\overline{a}) \in \varLambda (G(\overline{m},\overline{a}))\).

Proof

Given a compact, convex, and non-empty set \(\tilde{A}_1\subset U_1\), let \(\varTheta : \tilde{A}_1 \times \tilde{A}_1 \twoheadrightarrow \tilde{A}_1 \times \tilde{A}_1\) be the correspondence defined by \(\varTheta ((m_1,a_1), (m_2,a_2))=\left( \varPhi _{G(m_1,a_1)}(m_2,a_2) \cap \tilde{A}_1\right) \times \{(m_1,a_1)\}.\) If the set-valued map \(\varTheta _1: \tilde{A}_1 \times \tilde{A}_1 \twoheadrightarrow \tilde{A}_1\) given by \(\varTheta _1((m_1,a_1), (m_2,a_2))=\varPhi _{G(m_1,a_1)}(m_2,a_2) \cap \tilde{A}_1\) has closed graph, then \(\varTheta \) is upper hemicontinuous and has non-empty, compact, and convex values. Thus, applying Kakutani’s Fixed Point Theorem, we could find \((\overline{m},\overline{a}) \in \tilde{A}_1 \subset U_1\) such that \((\overline{m},\overline{a}) \in \varLambda (G(\overline{m},\overline{a}))\).

Therefore, to prove the claim, it is sufficient to ensure that \(\varTheta _1\) has closed graph. Fix a sequence \(\{(z^n_1,z^n_2, (m^n,a^n))\}_{n \in {\mathbb {N}}}\subset \text{ Graph }(\varTheta _1)\) that converges to \((\tilde{z}_1, \tilde{z}_2, (\tilde{m},\tilde{a})) \). We aim to guarantee that \((\tilde{m},\tilde{a}) \in \varTheta _1(\tilde{z}_1, \tilde{z}_2)\).

For convenience of notations, assume that \({\mathcal {G}}_i={\mathcal {G}}_i((K^i_{t},\varGamma ^i _{t},u^i_{t})_{t\in T_1 \cup T_2}),\,\forall i \in \{1,2\}\). Given \(t \in T_2\), let \(\gamma _t: (\widehat{M} \times {\widehat{\mathcal {F}}}^2_{-t})\times \tilde{A}_1 \twoheadrightarrow \widehat{K}_t\) be the correspondence defined by

$$\begin{aligned} \gamma _t((m, a_{-t}), z )=\mathop {\hbox {argmax}}\limits _{x \in \varPsi ((m,a_{-t}),z) } v_t(x,(m,a_{-t}),z), \end{aligned}$$

whereFootnote 33

$$\begin{aligned}&\varPsi ((m,a_{-t}),z)=\lambda (z)\varGamma _t^1(m, a_{-t})+ (1-\lambda (z)) \varGamma _t^2(m,a_{-t}), \\&v_t(x,(m,a_{-t}),z)= \lambda (z) u_t^1(m,x,a_{-t})+ (1-\lambda (z)) u_t^2(m,x,a_{-t}). \end{aligned}$$

It follows that \(\gamma _t\) is upper hemicountinuous with non-empty and compact values.\(^{33}\) Therefore, the correspondence \(\gamma : (\widehat{M} \times {\widehat{\mathcal {F}}}^2)\times \tilde{A}_1 \twoheadrightarrow \Pi _{t\in T_2 \,}\widehat{K}_t\) given by \(\gamma ((m,a),z_2)=\prod _{t \in T_2} \gamma _t((m,a_{-t}), z)\) is upper hemicontinuous with compact and non-empty values. In particular, \(\gamma \) has closed graph.

Since \(\{(z^n_1,z^n_2,a^n)\}_{n \in {\mathbb {N}}}\subset \text{ Graph }(\gamma )\), it follows that \(\tilde{a} \in \gamma (\tilde{z}_1,\tilde{z}_2)\).

On the other hand, for each \(n \in {\mathbb {N}}\), there exists \(f_n:T_1 \rightarrow \widehat{K}\) such that, \(m_n=\int _{T_1} H(t,f_n(t))d\mu \) and, for almost all \(t \in T_1, f_n(t) \in \xi _t(z^n_1,z^n_2):=\mathop {\hbox {argmax}}\nolimits _{x \in \varPsi (z^n_1,z^n_2) } v_t(x,z^n_1,z^n_2),\) where we use notations analogous to those described above. Note that, for all \(t \in T_1, \xi _t\), it has a closed graph.

Since \(m^n \rightarrow \tilde{m}\), analogous arguments to those made in Theorem 1 (Claim A) ensure that, applying the multidimensional Fatou’s Lemma [see Hildenbrand (1974, page 69)], there exists a full measure set \(\dot{T}_1\subseteq T_1\) and a function \(\overline{f}:T_1 \rightarrow \widehat{K}\) such that,

  1. (i)

    For any \(t \in \dot{T}_1\), there is a subsequence of \(\{f_n(t)\}_{n \in {\mathbb {N}}}\) that converges to \(\overline{f}(t)\);

  2. (ii)

    For any \(t \in T_1 {\setminus } \dot{T}_1, \overline{f}(t)\in \xi _t(\tilde{z}_1,\tilde{z}_2)\);

  3. (iii)

    \(\tilde{m}=\int _{T_1} H(t, \overline{f}(t)) d\mu .\)

Since for any \(t\in \dot{T}_1\) the correspondence \(\xi _t\) is closed, it follows from item (i) that \(\overline{f}(t) \in \xi _t(\tilde{z}_1,\tilde{z}_2)\). Items (ii) and (iii) jointly with the fact that \(\tilde{a} \in \gamma (\tilde{z}_1,\tilde{z}_2)\) imply that \((\tilde{m }, \tilde{a}) \in \varTheta _1(\tilde{z}_1, \tilde{z}_2)\). \(\square \)

Since \((\overline{m},\overline{a}) \in U_1\), it follows that \(G(\overline{m},\overline{a})={\mathcal {G}}_1\). Hence, Claim B implies that \(\varLambda (G(\overline{m},\overline{a}))\cap U_1 =\emptyset \). This is a contradiction, since both \((\overline{m},\overline{a}) \in U_1\) and \((\overline{m},\overline{a}) \in \varLambda (G(\overline{m},\overline{a}))\). Therefore, the minimal essential set \(m({\mathcal {G})}\) is connected.

(iv) If \(\varLambda ({\mathcal {G}})\) is finite, then \({\mathcal {G}}\) has at least one essential equilibrium.

Suppose that \(\widehat{K}\) is a convex subset of a normed space with a metric induced by a norm. It follows from (iii) that for every \({\mathcal {G}}\in {\mathbb {G}}'\) there is a minimal essential set of \(\varLambda ({\mathcal {G}})\) that is connected. On the other hand, as \(\varLambda ({\mathcal {G}})\) is finite, minimal essential sets are singletons. \(\square \)

Proof of Theorem 4

Given \({\mathcal {X}}\in {\mathbb {X}}\), the \({\mathcal {T}}\)-essential subsets of \(\varLambda (\kappa ({\mathcal {X}}))\) are stable.

It follows from Definition 4 that it suffices to guarantee that minimal essential sets are stable in the sense of Definition 8. Let \(\varLambda _m({\mathcal {T}}, \mathcal {X})\) be the collection of minimal \({\mathcal {T}}\)-essential subsets of \(\varLambda (\kappa ({\mathcal {X}}))\).

By contradiction, assume that there is \(A\in \varLambda _m({\mathcal {T}}, \mathcal {X})\) and \(\epsilon _0>0\) such that, for any \(\delta >0\), there is \(\mathcal {X}_\delta \in \mathbb {X}\) with \(\tau (\mathcal {X},\mathcal {X}_\delta )<\delta \) and \(A'\cap B[\epsilon _0, A]^c \ne \emptyset ,\,\,\forall A'\in \varLambda _m({\mathcal {T}}, \mathcal {X}_\delta )\), where \( B[\epsilon _0, A]^c:= (\widehat{M} \times \widehat{\mathcal {F}}^2) {\setminus } B[\epsilon _0, A]\). Since \(A\) is \({\mathcal {T}}\)-essential, there is \(\delta _0>0\) such that, for any \(\mathcal {X}'\in \mathbb {X}\) with \(\tau (\mathcal {X},\mathcal {X}')<\delta _0\), we have that \(\varLambda (\kappa ({\mathcal {X}}')) \cap C(\epsilon _0, A)\ne \emptyset \), where \(C(\epsilon _0,A)=\{(m,a)\in \widehat{M}\times \widehat{\mathcal {F}}^2:\,\inf \limits _{(m',a')\in A} \widehat{\sigma }((m,a),(m',a'))<\epsilon \}.\) It follows that \(\varLambda (\kappa ({\mathcal {X}}_{\delta _0})) \cap B[\epsilon _0, A]\) is a non-empty and closed set contained in \(B[\epsilon _0, A]\). Therefore, \(\varLambda (\kappa ({\mathcal {X}}_{\delta _0})) \cap B[\epsilon _0, A]\) is not an essential subset of \(\varLambda (\kappa ({\mathcal {X}}_{\delta _0}))\).

Hence, there exists \(\epsilon _1>0\) such that, for any \(n\in {\mathbb {N}}\), there is \({\mathcal {X}}_n \in {\mathbb {X}}\) with \(\tau (\mathcal {X}_{\delta _0},\mathcal {X}_n)<\frac{\delta _1}{n}\) and \( C(\epsilon _1, \varLambda (\kappa ({\mathcal {X}}_{\delta _0})) \cap B[\epsilon _0, A]) \cap \varLambda (\kappa ({\mathcal {X}}_n))=\emptyset \), where \(\delta _1>0\) is chosen in such form that, for any \({\mathcal {X}}''\in {\mathbb {X}}\), if \(\tau ({\mathcal {X}}_{\delta _0}, {\mathcal {X}}'')<\delta _1\), then \( \tau ({\mathcal {X}}, {\mathcal {X}}'')<\delta _0\). The last property ensures that \(\tau ({\mathcal {X}}, {\mathcal {X}}_n)<\delta _0\) for any \(n \in {\mathbb {N}}\), which implies that \(\varLambda (\kappa ({\mathcal {X}}_n)) \cap C(\epsilon _0, A)\) is non-empty. Take a sequence \(\{(m_n,a_n)\}_{n\in {\mathbb {N}}}\) such that \( (m_n,a_n) \in \varLambda (\kappa ({\mathcal {X}}_n)) \cap C(\epsilon _0, A),\,\forall n \in {\mathbb {N}}\). Without the loss of generality, there is \((m_0,a_0)\in B[\epsilon _0,A]\) such that \((m_n,a_n)\rightarrow _{n} (m_0,a_0)\). The upper hemicontinuity of \((\varLambda \circ \kappa )\) ensures that \((m_0,a_0) \in \varLambda (\kappa ({\mathcal {X}}_{\delta _0}))\), that is, \((m_0,a_0) \in \varLambda (\kappa ({\mathcal {X}}_{\delta _0}))\cap B[\epsilon _0,A].\) However, as for any \(n \in {\mathbb {N}}\), we have that \( (m_n,a_n) \in \varLambda (\kappa ({\mathcal {X}}_n))\) and \( C(\epsilon _1, \varLambda (\kappa ({\mathcal {X}}_{\delta _0})) \cap B[\epsilon _0, A]) \cap \varLambda (\kappa ({\mathcal {X}}_n))=\emptyset \), it follows that \((m_n,a_n)\notin C(\epsilon _1, \varLambda (\kappa ({\mathcal {X}}_{\delta _0})) \cap B[\epsilon _0, A]),\,\forall n\in {\mathbb {N}}\). Thus, \((m_0,a_0) \notin \varLambda (\kappa ({\mathcal {X}}_{\delta _0}))\cap B[\epsilon _0,A],\) which is a contradiction.

If \(\mathbb {X}\) is a convex subset of a normed space and \(\tau \) is induced by a norm, then for each \({\mathcal {X}}\in {\mathbb {X}}\), the \({\mathcal {T}}\)-essential components of \(\varLambda (\kappa ({\mathcal {X}}))\) are strongly stable.

Since \(\widehat{M}\subset {\mathbb {R}}^m\) is compact and \(\widehat{K}_t\), with \(t\in T_2\), are compact subsets of normed spaces with metrics induced by norms, it follows that \(\varLambda (\kappa ({\mathcal {X}}))\subseteq \widehat{M} \times \widehat{\mathcal {F}}^2\) is a locally connected and compact space. Therefore, \(\varLambda (\kappa ({\mathcal {X}})\) as a finite number of connected components.Footnote 34 For this reason, given a \({\mathcal {T}}\)-essential component \(A\subseteq \varLambda (\kappa (\mathcal {X}))\), there exists \(\pi >0\) such that \(B[\pi ,A ]\cap B[\pi ,\varLambda (\kappa (\mathcal {X})){\setminus } A]=\emptyset \).

Furthermore, it follows from the proof of Theorem 2(i) that there is \(A_m\in \varLambda _m({\mathcal {T}}, \mathcal {X})\) such that \(A_m \subseteq A\). By the previous item, for each \(\epsilon >0\), there is \(\delta _1>0\) such that, given \(\mathcal {X}'\in \mathbb {X}\) with \(\tau (\mathcal {X},\mathcal {X}')<\delta _1\), there exists \(A'_m\in \varLambda _m({\mathcal {T}}, \mathcal {X}')\) for which \(A'_m\subseteq B[\epsilon , A_m]\subseteq B[\epsilon , A]\). Since \(\mathbb {X}\) is a convex subset of a normed space and \(\tau \) is induced by a norm, it follows from Theorem 3(iv) that minimal essential sets are connected, and therefore, following analogous arguments to those made in the proof of Theorem 2(ii), we can ensure that for any \(\mathcal {X}'\in \mathbb {X}\) with \(\tau (\mathcal {X},\mathcal {X}')<\delta _1\), there is an essential component \(A' \in \varLambda _c({\mathcal {T}},{\mathcal {X}}')\) which contains \(A'_m\), where \(\varLambda _c({\mathcal {T}},{\mathcal {X}}') \) is the set of \({\mathcal {T}}\)-essential components of \(\varLambda (\kappa ({\mathcal {X}}'))\). We want to prove that, for \({\mathcal {X}}'\) closely enough to \({\mathcal {X}}, A' \subseteq B[\epsilon , A].\)

Since the correspondence \(\varLambda \circ \kappa \) is upper hemicontinuous, there is \(\delta _2>0\) such that for any \(\mathcal {X}'\in \mathbb {X}\) with \(\tau (\mathcal {X},\mathcal {X}')<\delta _2\) we have that \(\varLambda (\kappa ({\mathcal {X}}'))\subset C(\epsilon , \varLambda (\kappa ({\mathcal {X}}))) \subset B[\epsilon , A] \cup B[\epsilon , \varLambda (\kappa ({\mathcal {X}})){\setminus } A]\).

Note that \(\varLambda (\kappa ({\mathcal {X}})){\setminus } A\) is a compact set.Footnote 35 Let \(\delta =\min \{\delta _0, \delta _1\}\) and fix \(\mathcal {X}'\in \mathbb {X}\) with \(\tau (\mathcal {X},\mathcal {X}')<\delta \). If \(A'\cap B[\epsilon , A]^c\ne \emptyset \), then \(A' \cap B[\epsilon , \varLambda (\kappa ({\mathcal {X}})){\setminus } A]\ne \emptyset \) and \(A' \cap B[\epsilon , A]\ne \emptyset \). In addition, when \(\epsilon <\pi \), it follows that \(B[\epsilon , A] \bigcap B[\epsilon , \varLambda (\kappa ({\mathcal {X}})){\setminus } A]=\emptyset \). Since \(A\) and \(\varLambda (\kappa ({\mathcal {X}})){\setminus } A\) are compact sets, both \(B[\epsilon , A]\) and \(B[\epsilon , \varLambda (\kappa ({\mathcal {X}})){\setminus } A]\) are closed. Thus, we obtain a partition of the connected set \(A'\) into two non-empty and disjoint closed sets, \(A' \cap B[\epsilon , \varLambda (\kappa ({\mathcal {X}})){\setminus } A]\) and \(A' \cap B[\epsilon , A]\), which is a contradiction. Therefore, for any \(\mathcal {X}'\in \mathbb {X}\) with \(\tau (\mathcal {X},\mathcal {X}')<\delta \), we have that \(A'\subset B[\epsilon , A]\). \(\square \)

Lemma 2

Let \({\mathcal {G}}={\mathcal {G}}( (K_{t},\varGamma _{t},u_{t})_{t\in T_1 \cup T_2})\) be a generalized payoff secure and upper semicontinuous game. Then, \({\mathcal {G}}\) satisfies continuous security.

Proof

Given \((m,a)\notin \varLambda ({\mathcal {G}})\), generalized payoff security guarantees that, for any \(\epsilon >0\), there exists \((U^\epsilon , (\varphi ^\epsilon _t)_{t \in T_1\cup T_2}, \alpha ^\epsilon )\) satisfying item (i) of Definition 10. Thus, to guarantee that \({\mathcal {G}}\) is continuous secure, it is sufficient to prove that \((U^\epsilon , \alpha ^\epsilon )\) satisfies Definition 10 (ii) for some \(\epsilon >0\). Suppose, by contradiction, that for any \(n \in {\mathbb {N}}\), there is \((f_n,a_n)\in \widehat{\mathcal {F}}^1\times \widehat{\mathcal {F}}^2 \) satisfying,

  1. (a)

    \((m(f_n),a_n)\in U^{\frac{1}{n}}\),

  2. (b)

    \(f_n(t) \in \varGamma _t(m(f_n),a_n)\) for almost all \(t \in T_1\),

  3. (c)

    \(a_{n,t} \in \varGamma _t(m(f_n), a_{n,-t})\) for all \(t \in T_2\),

  4. (d)

    for almost all \(t \in T_1, u_t(f_n(t), m(f_n),a_n)\ge \alpha ^{\frac{1}{n}}(t)\),

  5. (e)

    for any \(t \in T_2, u_t(m(f_n), a_{n,t}, a_{n, -t})\ge \alpha ^{\frac{1}{n}}(t)\).

Since we can assume that \(\bigcap _n U^{\frac{1}{n}}=\{(m,a)\},\) it follows from (a) that \((m(f_n),a_n) \rightarrow _n (m,a)\). Conditions (b)-(c) guarantee, by using analogous arguments to those made in the proof of Theorem 1 (Claim A), that we can find a strategy profile \(f\in {\mathcal {F}}^1((K_t)_{t\in T_1})\) such that \(m=m(f)\) and there is a full measure set \(T_1'\subseteq T_1\) such that \(f(t)\in L_S(f_n(t)),\,\forall t \in T'_1 \).

In addition, as correspondences of admissible strategies have closed graph, it follows that (i) for almost all \(t \in T_1, f(t)\in \varGamma _t(m(f),a)\); (ii) for all \(k\in T_2, a_k \in \varGamma _k(m(f),a_{-k}).\)

Hence, as \((m,a)\notin \varLambda ({\mathcal {G}})\), there is a non-negligible set of agents that are sub-optimizing, i.e., there exists \(\delta >0\) such that either for a positive measure set \(T''_1\subseteq T_1\),

$$\begin{aligned} u_t(f(t), m, a) \, + \, \delta \, < \sup _{x \in \varGamma _t(m,a)} u_t(x,m,a),\quad \forall t \in T''_1, \end{aligned}$$

or for some \(t\in T_2\),

$$\begin{aligned} u_t(m, a_{t}, a_{-t}) \,+ \, \delta \, < \sup _{x \in \varGamma _t(m,a_{-t})} u_t(m,x,a_{-t}). \end{aligned}$$

This last condition implies that

$$\begin{aligned} \sum _{t \in T_2} u_t(m, a_{t}, a_{-t}) \,+ \, \delta \, < \sum _{t \in T_2} \,\sup _{x \in \varGamma _t(m,a_{-t})} u_t(m,x,a_{-t}). \end{aligned}$$

Since \({\mathcal {G}}\) is upper semicontinuous and \((m(f_n),a_n) \rightarrow _n (m,a)\), it follows from the definition of \(f\) that for \(n\in {\mathbb {N}}\) large enough, we have that either for all \(t\in T'_1 \cap T''_1\),

$$\begin{aligned} u_t(f_n(t), m(f_n), a_n) + \, 0.5\,\delta \, < \sup _{x \in \varGamma _t(m,a)} u_t(x,m,a), \end{aligned}$$
(1)

or

$$\begin{aligned} \sum _{t \in T_2}\,\,u_t(m(f_n), a_{n,t}, a_{n,-t}) + \, 0.5\, \delta \,< \sum _{t \in T_2}\,\,\sup _{x \in \varGamma _t(m,a_{-t})} u_t(m,x,a_{-t}). \end{aligned}$$

The later inequality implies that there is \(t \in T_2\) such that

$$\begin{aligned} u_t(m(f_n), a_{n,t}, a_{n,-t}) + \, \frac{0.5\, \delta }{\# T_2}\,< \sup _{x \in \varGamma _t(m,a_{-t})} u_t(m,x,a_{-t}). \end{aligned}$$
(2)

On the other hand, it follows from conditions (d)–(e) above and Definition 11(ii) that for \(n\in {\mathbb {N}}\) large enough, there exists \(T_n\subseteq T_1\) with \(\mu (T_n)\ge \mu (T_1)-\frac{1}{n}\) such that, for any \(t \in T_n\),

$$\begin{aligned} u_t(f_n(t), m(f_n), a_n) \ge \sup _{x \in \varGamma _t(m,a)} u_t(x,m,a) - \frac{1}{n}, \end{aligned}$$

and for every atomic player \(t\in T_2\),

$$\begin{aligned} u_t(m(f_n), a_{n,t}, a_{n,-t}) \ge \sup _{x \in \varGamma _t(m,a_{-t})} u_t(m,x,a_{-t})- \frac{1}{n}. \end{aligned}$$

Thus, \(\underline{\lim }_n u_t(f_n(t), m(f_n), a_n) \ge \sup _{x \in \varGamma _t(m,a)} u_t(x,m,a)\) for almost all non-atomic player \(t \in T_1\). Also, for each atomic player \(t\in T_2\), we have that \(\underline{\lim }_n u_t(m(f_n), a_{n,t}, a_{n,-t}) \ge \sup _{x \in \varGamma _t(m,a_{-t})} u_t(m,x,a_{-t})\). Hence, taking the lower limit in (1) and (2), we obtain a contradiction. \(\square \)

Proposition 2

\(({\mathbb {G}}_d, \rho )\) is a complete metric space.

Proof

Since \((K_t, \varGamma _t)_{t \in T_1 \cup T_2}\) does not change, \(({\mathbb {G}}_d, \rho )\) can be considered as a subset of the space of bounded functions \({\mathcal {B}}:={\mathcal {U}}(T_1 \times \widehat{K}\times \widehat{ M}\times \widehat{{\mathcal {F}}}^2) \times \prod _{t \in T_2} {\mathcal {U}}_t(\widehat{M}\times \widehat{{\mathcal {F} }}^2)\), where for any \(t\in T_2\), the set \({\mathcal {U}}_t(\widehat{M}\times \widehat{{\mathcal {F} }}^2)\) is the collection of bounded functions \((m,a_t,a_{-t}) \rightarrow u_t(m,a_t,a_{-t})\) which are quasi-concave on \(a_t\). Note that \(({\mathcal {B}}, \rho )\) is a complete metric space and, therefore, it is sufficient to ensure that \({\mathbb {G}}_d\) is a closed subset of \({\mathcal {B}}\).

Fix a sequence \(\{{\mathcal {G}}_n\}_{n \in {\mathbb {N}}} \subset {\mathbb {G}}_d\), with \({\mathcal {G}}_n = {\mathcal {G}}_n( (u^n_{t})_{t\in T_1 \cup T_2})\) for any \(n \in {\mathbb {N}}\), which converges to \(\overline{\mathcal {G}}={\overline{\mathcal {G}}} ( (\overline{u}_{t})_{t\in T_1 \cup T_2}) \in {\mathcal {B}}.\) We want to prove that \(\overline{\mathcal {G}} \in {\mathbb {G}}_d\).

Claim. \(\overline{\mathcal {G}}\) is generalized payoff secure.

Given \((m,a)\in \widehat{M}\times \widehat{\mathcal {F}}^2\) and \(\epsilon >0\), generalized payoff security of \({\mathcal {G}}_n\) at \(((m,a), 0.5\, \epsilon )\) implies that there exists \((U^{n} , (\varphi ^{n}_t)_{t \in T_1\cup T_2}, \alpha ^{n})\) satisfying the requirements of Definition 11.

Thus, for \(n\) large enough, for almost all \(t \in T_1\), for all \(k \in T_2\), and for every \((m',a')\in U^n\), we have that,

$$\begin{aligned} \overline{u}_t(x,m',a')&> \,\,u^n_t(x,m',a') - 0.25\, \epsilon&\ge \,\,\alpha ^{n}(t) - \,0.25\,\epsilon ,\,\,\,\,\,\forall x\in \varphi ^{n}_t(m',a');\\ \overline{u}_k(m',x, a'_{-t})&> \,\,u^n_k (m'. x, a'_{-t})- 0.25\, \epsilon&\ge \,\, \alpha ^{n}(k)-\,0.25\, \epsilon , \,\,\,\,\, \forall x\in \varphi ^{n}_k(m',a'). \end{aligned}$$

Furthermore, Definition 11(ii) ensures that, for \(n\) large enough,

$$\begin{aligned} \left( \alpha ^{n}(k)-\,0.25\,\epsilon \right) + \epsilon&= (\alpha ^{n}(k) + 0.5 \epsilon ) + 0,25\,\epsilon \\&\ge \sup _{x \in \varGamma _t(m,a_{-k})}\, u^n_t(m,x,a_{-k}) + 0.25\, \epsilon \\&> \sup _{x \in \varGamma _t(m,a_{-k})}\, \overline{u}_t(m,x,a_{-k}),\quad \forall k \in T_2. \end{aligned}$$
$$\begin{aligned} \mu \left\{ t \in T_1: \left( \alpha ^{n}(t)-\,0.25\,\epsilon \right) + \epsilon \ge \sup _{x \in \varGamma _t(m,a)}\, \overline{u}_t(x,m,a)\right\} \ge \mu (T_1)-0.5\,\epsilon . \end{aligned}$$

Therefore, taking \(n\) large enough and choosing \((U^{n} , (\varphi ^{n}_t)_{t \in T_1\cup T_2}, \alpha ^{n}-\,0.25\,\epsilon )\), we ensure that \(\overline{\mathcal {G}}\) is generalized payoff secure at \(((m,a),\epsilon )\). \(\square \)

It is a direct consequence of Carbonell-Nicolau (2010, Lemma 1, page 425) that \(\overline{\mathcal {G}}\) is upper semicontinuous and atomic players’ objective functions \((\overline{u}_t)_{t \in T_2}\) are quasi-concave. In addition, the same arguments made in the proof of Proposition 1 guarantee that, for every \((m,a)\in \widehat{M}\times \widehat{\mathcal {F}}^2\), the map \((t,x)\in T_1 \times \widehat{K} \rightarrow \overline{u}_t(x,m,a)\) is measurable. This concludes the proof. \(\square \)

Proof of Theorem 5

Since \(({\mathbb {G}}_d,\rho )\) is complete, it follows from the proofs of Theorems 1–4 that it is sufficient to ensure that \(\varLambda \) still has a closed graph when its domain is extended to \({\mathbb {G}}_d\).

Let \(\{(\mathcal {G}_n,(m_n,a_n))\}_{n \in {\mathbb {N}}} \subset \text{ Graph }(\varLambda )\) such that \((\mathcal {G}_n,(m_n,a_n))\rightarrow (\overline{\mathcal {G}},(\overline{m},\overline{a}))\), where \({\mathcal {G}}_n= {\mathcal {G}}_n((u^n_{t})_{t\in T_1 \cup T_2})\) and \(\overline{\mathcal {G}}= \overline{\mathcal {G}}((\overline{u}_{t})_{t\in T_1 \cup T_2})\in \mathbb {G}_d\). We want to prove that \((\overline{m},\overline{a}) \in \varLambda ({\overline{\mathcal {G}}})\).

Since \((m_n,a_n)\in \varLambda ({\mathcal {G}_n})\), there is \(f_n\in \widehat{\mathcal {F}}^1\) such that

  1. (a)

    the function \(g_n: T_1 \rightarrow {\mathbb {R}}^m\) given by \(g_n(t)=H(t, f_n(t))\) is measurable and \(m_n=m(f_n)\);

  2. (b)

    for almost all \(t \in T_1\) both \(f_n(t)\in \varGamma _t(m_n,a_n)\) and

    $$\begin{aligned} u_t^n(f_n(t), m_n,a_{n})=\sup \limits _{x\in \varGamma _t(m_n, a_n)}{u_t^n(x, m_n,a_{n})}. \end{aligned}$$

Claim A

There exists \(\overline{f}\in \widehat{\mathcal {F}}^1\) such that \(\overline{m}=\int _{T_1}{H(t,\overline{f}(t))d\mu }\).

Proof

By analogous arguments to those in the proof of Theorem 1 (Claim A), we can find a strategy profile \(\overline{f}\in \widehat{\mathcal {F}}^1\) and a full measure set \(T^*_1\subseteq T_1\) such that \(\overline{m}=m(\overline{f}), \overline{f}(t)\in L_S(f_n(t)),\,\forall t \in T^*_1\). \(\square \)

Claim B

For almost all \(t\in T_1, \overline{f}(t)\in \varGamma _t(\overline{m},\overline{a})\). In addition, for any \(t \in T_2, \overline{a}_t \in \varGamma _t(\overline{m},\overline{a}_{-t})\).

Proof

It follows from the proof of Claim A that there is a full measure set \(T^*_1\subseteq T_1\) such that \(\overline{f}(t)\in L_S(f_n(t)),\,\forall t \in T^*_1\). Thus, the closed graph property of correspondences of admissible strategies ensures that (i) for all \(t \in T^*_1, \overline{f}(t)\in \varGamma _t(\overline{m},\overline{a})\); and (ii) for all \(t\in T_2, \overline{a}_t \in \varGamma _t(\overline{m},\overline{a}_{-t}).\) \(\square \)

Claim C

The following properties hold

  1. (i)

    For almost all \(t \in T_1, \overline{f}(t) \in \mathop {\hbox {argmax}}\nolimits _{x \in \varGamma _t(\overline{m},\overline{a})}\,\,\overline{u}_t(x,\overline{m},\overline{a}).\)

  2. (ii)

    For any \(t\in T_2, \overline{a}_t\in \mathop {\hbox {argmax}}\nolimits _{x\in \varGamma _t(\overline{m},\overline{a}_{-t})}\overline{u}_t(\overline{m},x,\overline{a}_{-t}).\)

Proof

Suppose that at least one of the properties (i) and (ii) does not hold, i.e., there is a non-negligible set of agents that are suboptimizing. Since \({\overline{\mathcal {G}}}\) is upper semicontinuous, identical arguments to those made in the proof of Lemma 2 to obtain conditions (1) and (2) imply that there is \(\xi >0\) such that,\(^{36}\) for \(n\) large enough, either

$$\begin{aligned} \overline{u}_t(f_n(t), m_n, a_n) \, + \, \xi \, < \sup _{x \in \varGamma _t(\overline{m},\overline{a})} \overline{u}_t(x,\overline{m},\overline{a}),\quad \forall t \in T^{**}_1\subseteq T_1, \end{aligned}$$

where \(T^{**}_1\) is a positive measure set, or there exists \(t \in T_2\) such thatFootnote 36

$$\begin{aligned} \overline{u}_t(m_n, a_{n,t}, a_{n,-t}) \, + \, \xi \, < \sup _{x \in \varGamma _t(\overline{m},\overline{a}_{-t})} \overline{u}_t(\overline{m},x,\overline{a}_{-t}). \end{aligned}$$

Thus, as \(\rho ({\mathcal {G}}_n, \overline{\mathcal {G}})\rightarrow _n 0\), for \(n\) large enough at least one of the following conditions hold:

$$\begin{aligned}&\displaystyle u^n_t(f_n(t), m_n, a_n) \, + \, 0.5\,\xi <\sup _{x \in \varGamma _t(\overline{m},\overline{a})} \overline{u}_t(x,\overline{m},\overline{a}),\quad \forall t \in T^{*}_1 \cap T^{**}_1;&\end{aligned}$$
(3)
$$\begin{aligned}&\displaystyle u^n_t(m_n, a_{n,t}, a_{n,-t}) \, + \, 0.5\,\xi < \sup _{x \in \varGamma _t(\overline{m},\overline{a}_{-t})} \overline{u}_t(\overline{m},x,\overline{a}_{-t}).&\end{aligned}$$
(4)

On the other hand, since \(\overline{\mathcal {G}}\) is a generalized payoff secure large game, for every \(\epsilon >0\), there exists \((U^\epsilon , (\varphi ^\epsilon _t)_{t \in T_1\cup T_2}, \alpha ^\epsilon )\) satisfying Definition 11. In particular, as \((m_n,a_n)\rightarrow _n (\overline{m},\overline{a})\), there exists a set \(T_\epsilon \subseteq T_1\) with \(\mu (T_\epsilon )\ge \mu (T_1)-\epsilon \) such that, for any \(n\) large enough, we have \((m_n,a_n) \in U^\epsilon \) and the following properties hold for every \((t,k)\in T_\epsilon \times T_2\):

$$\begin{aligned} \sup _{x \in \varGamma _t(m_n,a_n)} \overline{u}_t (x,m_n,a_n)&\ge \sup _{x \in \varphi ^\epsilon _t(m_n,a_n)} \overline{u}_t (x,m_n,a_n)\nonumber \\&\ge \alpha ^\epsilon (t) \,\,\ge \,\, \sup _{x \in \varGamma _t(\overline{m},\overline{a})} \overline{u}_t(x,\overline{m},\overline{a}) - \epsilon ; \end{aligned}$$
(5)
$$\begin{aligned} \sup _{x \in \varGamma _k(m_n,a_{n,-k})} \overline{u}_k (m_n,x, a_{n,-k})&\ge \sup _{x \in \varphi ^\epsilon _k(m_n,a_{n,-k})} \overline{u}_k (m_n,x, a_{n,-k}) \nonumber \\&\ge \alpha ^\epsilon (k) \,\,\ge \,\, \sup _{x \in \varGamma _k(\overline{m},\overline{a}_{-k})} \overline{u}_k(\overline{m},x, \overline{a}_{-k}) - \epsilon . \end{aligned}$$
(6)

As objective functions are bounded, the uniform convergence of \({\mathcal {G}}_n\) to \(\overline{\mathcal {G}}\) ensures that, for \(n\) large enough and for each \((t,k) \in T^*_1\times T_2\),

$$\begin{aligned} u^n_t (f_n(t), m_n, a_n) + \epsilon&= \sup _{x \in \varGamma _t(m_n,a_n)} u^n_t (x,m_n,a_n) +\epsilon \nonumber \\&\ge \sup _{x \in \varGamma _t(m_n,a_n)} \overline{u}_t (x,m_n,a_n); \end{aligned}$$
(7)
$$\begin{aligned} u^n_k (m_n, a_{n,k}, a_{n,-k}) + \epsilon&= \sup _{x \in \varGamma _k(m_n,a_{n,-k})} u^n_k (m_n,x, a_{n,-k}) +\epsilon \nonumber \\&\ge \sup _{x \in \varGamma _k(m_n,a_{n,-k})} \overline{u}_k (m_n,x, a_{n,-k}). \end{aligned}$$
(8)

Therefore, for every \(\epsilon >0\) and for each \((t,k) \in (T_1^* \cap T_\epsilon ) \times T_2\), taking the lower limit as \(n\) goes to infinity on inequalities (7)–(8) it follows that,

$$\begin{aligned} \underline{\lim }_n\, u^n_t (f_n(t), m_n, a_n) + \epsilon&\ge \sup _{x \in \varGamma _t(\overline{m},\overline{a})} \overline{u}_t(x,\overline{m},\overline{a}) - \epsilon , \end{aligned}$$
(9)
$$\begin{aligned} \underline{\lim }_n\, u^n_k (m_n, a_{n,k}, a_{n,-k}) + \epsilon&\ge \sup _{x \in \varGamma _k(\overline{m},\overline{a}_{-k})} \overline{u}_k(\overline{m},x, \overline{a}_{-k}) - \epsilon . \end{aligned}$$
(10)

Taking the limit as \(\epsilon \) goes to zero on (910) and taking the lower limit as \(n\) goes to infinity on (34), we obtain a contradiction.

It follows from Claims A and C that \((\overline{m},\overline{a}) \in \varLambda ({\overline{\mathcal {G}}})\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Correa, S., Torres-Martínez, J.P. Essential equilibria of large generalized games. Econ Theory 57, 479–513 (2014). https://doi.org/10.1007/s00199-014-0821-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-014-0821-3

Keywords

JEL Classification

Navigation