Skip to main content
Log in

Analysis of Markovian Competitive Situations Using Nonatomic Games

  • Published:
Dynamic Games and Applications Aims and scope Submit manuscript

Abstract

We study discrete-time dynamic games with effectively identical players who possess private states that evolve randomly. Players in these games are concerned with their undiscounted sums of period-wise payoffs in the finite-horizon case and discounted sums of stationary period-wise payoffs in the infinite-horizon case. In the general semi-anonymous setting, the other players influence a particular player’s payoffs and state evolutions through the joint state–action distributions that they form. When dealing with large finite games, we find it profitable to exploit symmetric mixed equilibria of a reduced feedback type for the corresponding nonatomic games (NGs). These equilibria, when continuous in a certain probabilistic sense, can be used to achieve near-equilibrium performances when there are a large but finite number of players. We focus on the case where independently generated shocks drive random actions and state transitions. The NG equilibria we consider are random state-to-action maps that pay no attention to players’ external environments. Yet, they can be adopted for a variety of real situations where the knowledge about the other players can be incomplete. Results concerning finite horizons also form the basis of a link between an NG’s stationary equilibrium and good stationary profiles for large finite games.

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

References

  1. Adlakha S, Johari R (2013) Mean field equilibrium in dynamic games with complementarities. Oper Res 61:971–989

    MathSciNet  MATH  Google Scholar 

  2. Al-Najjar NI, Smorodinsky R (2001) Large nonanonymous repeated games. Games Econ Behav 37:26–39

    MathSciNet  MATH  Google Scholar 

  3. Al-Najjar NI (2008) Large games and the law of large numbers. Games Econ Behav 64:1–34

    MathSciNet  MATH  Google Scholar 

  4. Aumann RJ (1964a) Markets with a continuum of traders. Econometrica 32:39–50

    MathSciNet  MATH  Google Scholar 

  5. Aumann RJ (1964b) Mixed and behavior strategies in infinite extensive games. In: Dresher M, Shapley L, Tucker A (eds) Annals of mathematics studies, 52, Advances in game theory. Princeton University Press, Princeton, pp 627–650

    Google Scholar 

  6. Balbus L, Dziewulski P, Reffett K, Woźny L (2017) A qualitative theory of large games with strategic complementarities. Econ Theory (forthcoming)

  7. Balder EJ (1995) A unifying approach to existence of nash equilibria. Int J Game Theory 24:79–94

    MathSciNet  MATH  Google Scholar 

  8. Balder EJ (2002) A unifying pair of Cournot–Nash equilibrium existence results. J Econ Theory 102:437–470

    MathSciNet  MATH  Google Scholar 

  9. Banerjee A, Moll B (2010) Why does misallocation persist? Am Econ J Macroecon 2:189–206

    Google Scholar 

  10. Bensoussan A, Sung KCJ, Yam SCP, Yung SP (2014) Linear-quadratic mean field games. 2016. Linear-quadratic mean field games. J Optim Theory Appl 169:496–529

    MATH  Google Scholar 

  11. Bergin J, Bernhardt D (1995) Anonymous sequential games: existence and characterization of equilibria. Econ Theor 5:461–489

    MathSciNet  MATH  Google Scholar 

  12. Carmona G (2004) Nash equilibria of games with a continuum of players. Working Paper, Universidade Nova de Lisboa

  13. Carmona R, Lacker D (2015) A probabilistic weak formulation of mean field games and applications. Ann Appl Probab 25:1189–1231

    MathSciNet  MATH  Google Scholar 

  14. Ethier SN, Kurtz TG (1986) Markov processes: characterization and convergence. Wiley, New York

    MATH  Google Scholar 

  15. Fudenberg D, Levine DK (1988) Open-loop and closed-loop equilibria in dynamic games with many players. J Econ Theory 44:1–18

    MathSciNet  MATH  Google Scholar 

  16. Gabaix X, Lasry J-M, Lions P-L, Moll B (2016) The dynamics of inequality. Econometrica 84:2071–2111

    MathSciNet  MATH  Google Scholar 

  17. Green EJ (1980) Non-cooperative price taking in large dynamic markets. J Econ Theory 22:155–182

    MATH  Google Scholar 

  18. Green EJ (1984) Continuum and finite-player noncooperative models of competition. Econometrica 52:975–993

    MathSciNet  MATH  Google Scholar 

  19. Hart S, Hildenbrand W, Kohlberg E (1974) On equilibrium allocation as distributions on the commodity space. J Math Econ 1:159–167

    MathSciNet  MATH  Google Scholar 

  20. Hildenbrand W (1974) Core and equilibria of a large economy. Princeton University Press, Princeton

    MATH  Google Scholar 

  21. Hopenhayn HA (1992) Entry, exit, and firm dynamics in long run equilibrium. Econometrica 60:1127–1150

    MathSciNet  MATH  Google Scholar 

  22. Housman D (1988) Infinite player noncooperative games and the continuity of the Nash equilibrium correspondence. Math Oper Res 13:488–496

    MathSciNet  MATH  Google Scholar 

  23. Huang M, Caines PE, Malham PR (2003) Individual and Mass Behaviour in Large Population Stochastic Wireless Power Control Problems: Centralized and Nash equilibrium solutions. In: Proceedings of the 42nd IEEE conference on decision and control, Maui, Hawaii, pp 98–103

  24. Huang M, Malhame RP, Caines PE (2006) Large population stochastic dynamic games: closed-loop McKean–Vlasov systems and the Nash certainty equivalence principle. Commun Inf Syst 6:221–252

    MathSciNet  MATH  Google Scholar 

  25. Jovanovic B, Rosenthal RW (1988) Anonymous sequential games. J Math Econ 17:77–88

    MathSciNet  MATH  Google Scholar 

  26. Kalai E (2004) Large robust games. Econometrica 72:1631–1665

    MathSciNet  MATH  Google Scholar 

  27. Khan MA (1985) On Extensions of the Cournot-Nash Theorem. In: Aliprantis CD, Burkinshaw O, Rothman NJ (eds) Advances in equilibrium theory. Springer lecture notes in economics and mathematical systems, pp 79–106

  28. Khan MA, Rath KP, Sun Y (1997) On the existence of pure strategy equilibrium in games with a continuum of players. J Econ Theory 76:13–46

    MATH  Google Scholar 

  29. Lasry J-M, Lions P-L (2007) Mean field games. Jpn J Math 2:229–260

    MathSciNet  MATH  Google Scholar 

  30. Light B, Weintraub GY (2019) Mean field equilibrium: uniqueness, existence, and comparative statics. Working Paper, Stanford University, CA

  31. Mas-Colell A (1984) On a theorem of Schmeidler. J Math Econ 13:201–206

    MathSciNet  MATH  Google Scholar 

  32. Mertens JF, Parthasarathy T (1987) Equilibria for discounted stochastic games. CORE Discussion Paper No. 8750

  33. Parthasarathy KR (2005) Probability measures on metric spaces. AMS Chelsea Publishing, Providence

    MATH  Google Scholar 

  34. Podczeck K (2009) On purification of measure-valued maps. Econ Theor 38:399–418

    MathSciNet  MATH  Google Scholar 

  35. Podczeck K (2010) On existence of rich fubini extensions. Econ Theor 45:1–22

    MathSciNet  MATH  Google Scholar 

  36. Rath KP (1992) A direct proof of the existence of pure strategy equilibria in games with a continuum of players. Econ Theor 2:427–433

    MathSciNet  MATH  Google Scholar 

  37. Rath KP (1996) Existence and upper hemicontinuity of equilibrium distributions of anonymous games with discontinuous payoffs. J Math Econ 26:305–324

    MathSciNet  MATH  Google Scholar 

  38. Reny PJ, Perry M (2006) Toward a strategic foundation for rational expectations equilibrium. Econometrica 74:1231–1269

    MathSciNet  MATH  Google Scholar 

  39. Sabourian H (1990) Anonymous repeated games with a large number of players and random outcomes. J Econ Theory 51:92–110

    MathSciNet  MATH  Google Scholar 

  40. Schmeidler D (1973) Equilibrium points of nonatomic games. J Stat Phys 7:295–300

    MathSciNet  MATH  Google Scholar 

  41. Shapley LS (1953) Stochastic games. Proc Nat Acad Sci 39:1095–1100

    MathSciNet  MATH  Google Scholar 

  42. Solan E (1998) Discounted stochastic games. Math Oper Res 23:1010–1021

    MathSciNet  MATH  Google Scholar 

  43. Sun Y (2006) The exact law of large numbers via Fubini extension and characterization of insurable risks. J Econ Theory 126:31–69

    MathSciNet  MATH  Google Scholar 

  44. Weintraub GY, Benkard CL, van Roy B (2008) Markov perfect industry dynamics with many firms. Econometrica 76:1375–1411

    MathSciNet  MATH  Google Scholar 

  45. Weintraub GY, Benkard CL, van Roy B (2011) Industry dynamics: foundations for models with an infinite number of firms. J Econ Theory 146:1965–1994

    MathSciNet  MATH  Google Scholar 

  46. Wiszniewska-Matyszkiel A (2000) Existence of pure equilibria in games with nonatomic space of players. Topol Methods Nonlinear Anal 16:339–349

    MathSciNet  MATH  Google Scholar 

  47. Wiszniewska-Matyszkiel A (2002) Discrete time dynamic games with continuum of players, I: decomposable games. Int Game Theory Rev 4:331–342

    MathSciNet  MATH  Google Scholar 

  48. Wiszniewska-Matyszkiel A (2003) Discrete time dynamic games with continuum of players, II: semi-decomposable games. Int Game Theory Rev 5:27–40

    MathSciNet  MATH  Google Scholar 

  49. Wiszniewska-Matyszkiel A (2005) A dynamic game with continuum of players and its counterpart with finitely many players. In: Nowak AS, Szajowski K (eds) Advances in dynamic games. Annals of the international society of dynamic games, vol 7. Birkhäuser, Boston, pp 455–469

    MATH  Google Scholar 

  50. Wiszniewska-Matyszkiel A (2008) Common resources, optimality and taxes in dynamic games with increasing number of players. J Math Anal Appl 337:840–841

    MathSciNet  MATH  Google Scholar 

  51. Wiszniewska-Matyszkiel A (2014) Open and Closed Loop Nash Equilibria in Games with a Continuum of Players. J Optim Theory Appl 160:280–301

    MathSciNet  MATH  Google Scholar 

  52. Yang J (2017) A link between semi-anonymous sequential games and their large finite counterparts. International Journal of Game Theory 46:383–433

    MathSciNet  MATH  Google Scholar 

  53. Yang J (2020) Supplementary Material to the Paper “Analysis of Markovian Competitive Situations using Nonatomic Games”. Working Paper, Rutgers University

Download references

Acknowledgements

The author wishes to thank Professor Thomas G. Kurtz for referring him to Ethier and Kurtz [14], a source instrumental to the proof of Lemma 3. He would also like to thank an anonymous referee for pointing out critical errors concerning the definition of \(\mathcal{K}(A,B,\pi _B,C)\), as well as the proofs of Lemma 4 and Proposition 1. In addition, the author wishes to acknowledge the support of the National Science Foundation of China Grant 11871362.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jian Yang.

Additional information

Publisher's Note

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

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 31 KB)

Appendices

Technical Developments in Sect. 4

The Prohorov Metric: The Prohorov metric \(\rho _A\) for a metric space A is such that, for any distributions \(\pi ,\pi '\in \mathcal{P}(A)\),

$$\begin{aligned} \rho _A(\pi ,\pi ')=\inf \left( \epsilon >0\mid \pi '((A')^\epsilon )+\epsilon \ge \pi (A'),\; \text{ for } \text{ all } A'\in {\mathscr {B}}(A)\right) , \end{aligned}$$
(A.1)

where

$$\begin{aligned} (A')^\epsilon =\{a\in A\mid d_A(a,a')<\epsilon \text{ for } \text{ some } a'\in A'\}. \end{aligned}$$
(A.2)

The metric \(\rho _A\) is known to generate the weak topology for \(\mathcal{P}(A)\).

According to Parthasarathy [33] (Theorem II.7.1), the strong LLN applies to the empirical distribution under the weak topology, and hence under the Prohorov metric. In the following, we state its weak version.

Lemma 1

Given separable metric spaces A and B, suppose distribution \(\pi _A\in \mathcal{P}(A)\) and measurable mapping \(y\in \mathcal{M}(A,B)\). Then, for any \(\epsilon >0\), as long as n is large enough,

$$\begin{aligned} (\pi _A)^n\left( \left\{ a\equiv (a_m)_{m=1,\ldots ,n}\in A^n\mid \rho _B(\varepsilon (a) \cdot y^{-1},\pi _A \cdot y^{-1})<\epsilon \right\} \right) >1-\epsilon . \end{aligned}$$

For a separable metric space A, a point \(a\in A\), and an \((n-1)\)-point empirical distribution space \(\pi \in \mathcal{P}_{n-1}(A)\), we use \((a,\pi )_n\) to represent the member of \(\mathcal{P}_n(A)\) that has an additional 1/n weight on the point a, but with probability masses in \(\pi \) being reduced to \((n-1)/n\) times of their original values. For \(a\in A^n\) and \(m=1,\ldots ,n\), we have \((a_m,\varepsilon (a_{-m}))_n=\varepsilon (a)\). Concerning the Prohorov metric, we have also a simple but useful observation.

Lemma 2

Let A be a separable metric space. Then, for any \(n=2,3,\ldots \), \(a\in A\), and \(\pi \in \mathcal{P}_{n-1}(A)\),

$$\begin{aligned} \rho _A\left( (a,\pi )_n,\pi \right) \le \frac{1}{n}. \end{aligned}$$

Proof

Let \(A'\in {\mathscr {B}}(A)\) be chosen. If \(a\notin A'\), then

$$\begin{aligned} (a,\pi )_n(A')\le \pi (A')\le (a,\pi )_n(A')+\frac{1}{n}; \end{aligned}$$
(A.3)

if \(a\in A'\), then

$$\begin{aligned} (a,\pi )_n(A')-\frac{1}{n}\le \pi (A')\le (a,\pi )_n(A'). \end{aligned}$$
(A.4)

Hence, it is always true that

$$\begin{aligned} \mid (a,\pi )_n(A')-\pi (A')\mid \le \frac{1}{n}. \end{aligned}$$
(A.5)

In view of (A.1) and (A.2), we have

$$\begin{aligned} \rho _A\left( (a,\pi )_n,\pi \right) \le \frac{1}{n}. \end{aligned}$$
(A.6)

We have thus completed the proof. \(\square \)

The following result is important for showing the near-trajectory evolution of aggregate environments in large multi-period games. Among other things, it relies on Lemma 1.

Lemma 3

Given a separable metric space A and complete separable metric spaces B and C, suppose \(y_n\in \mathcal{M}(A^n,B^n)\) for every \(n\in {\mathbb {N}}\), \(\pi _A\in \mathcal{P}(A)\), \(\pi _B\in \mathcal{P}(B)\), and \(\pi _C\in \mathcal{P}(C)\). If

$$\begin{aligned} (\pi _A)^n\left( \{a\in A^n\mid \rho _B(\varepsilon (y_n(a)),\pi _B)<\epsilon \}\right) >1-\epsilon , \end{aligned}$$

for any \(\epsilon >0\) and any n large enough, then

$$\begin{aligned} (\pi _A\otimes \pi _C)^n\left( \{(a,c)\in (A\times C)^n\mid \rho _{B\times C}(\varepsilon (y_n(a),c),\pi _B\otimes \pi _C)<\epsilon \}\right) >1-\epsilon , \end{aligned}$$

for any \(\epsilon >0\) and any n large enough.

Proof

Suppose sequence \(\{\pi '_{B1},\pi '_{B2},\ldots \}\) weakly converges to the given probability measure \(\pi _B\), and sequence \(\{\pi '_{C1},\pi '_{C2},\ldots \}\) weakly converges to the given probability measure \(\pi _C\). We are to show that the sequence \(\{\pi '_{B1}\otimes \pi '_{C1},\pi '_{B2}\otimes \pi '_{C2},\ldots \}\) weakly converges to \(\pi _B\otimes \pi _C\).

Let F(B) denote the family of uniformly continuous real-valued functions on B with bounded support. Let F(C) be similarly defined for C. We certainly have

$$\begin{aligned} \left\{ \begin{array}{l} \lim _{k\rightarrow +\infty }\int _B f(b)\cdot \pi '_{Bk}(db)=\int _B f(b)\cdot \pi _B(db),\quad \forall f\in F(B),\\ \lim _{k\rightarrow +\infty }\int _C f(c)\cdot \pi '_{Ck}(dc)=\int _C f(c)\cdot \pi _C(dc),\quad \forall f\in F(C). \end{array}\right. \end{aligned}$$
(A.7)

Define F so that

$$\begin{aligned} \begin{array}{l} F=\{f\mid f(b,c)=f_B(b)\cdot f_C(c)\; \text{ for } \text{ any } (b,c)\in B\times C,\\ \;\;\;\;\;\;\;\;\;\;\;\; \text{ where } f_B\in F(B)\cup \{\mathbf{1}\} \text{ and } f_C\in F(C)\cup \{\mathbf{1}\}\}, \end{array} \end{aligned}$$
(A.8)

where \(\mathbf{1}\) stands for the function whose value is 1 everywhere. By (A.7) and (A.8),

$$\begin{aligned} \lim _{k\rightarrow +\infty }\int _{B\times C}f(b,c)\cdot (\pi '_{Bk}\otimes \pi '_{Ck})(d(b,c))=\int _{B\times C}f(b,c)\cdot (\pi _B\otimes \pi _C)(d(b,c)). \end{aligned}$$
(A.9)

According to Ethier and Kurtz [14] (Proposition III.4.4), F(B) and F(C) happen to be \(\mathcal{P}(B)\) and \(\mathcal{P}(C)\)’s convergence determining families, respectively. As B and C are complete, Ethier and Kurtz ([14], Proposition III.4.6, whose proof involves Prohorov’s Theorem, i.e., the equivalence between tightness and relative compactness of a collection of probability measures defined for complete separable metric spaces) further states that F as defined through (A.8) is convergence determining for \(\mathcal{P}(B\times C)\). Therefore, we have the desired weak convergence by (A.9).

Let \(\epsilon >0\) be given. In view of the above product-measure convergence and the equivalence between the weak topology and that induced by the Prohorov metric, there must be \(\delta _B>0\) and \(\delta _C>0\), such that \(\rho _B(\pi '_B,\pi _B)<\delta _B\) and \(\rho _C(\pi '_C,\pi _C)<\delta _C\) will imply

$$\begin{aligned} \rho _{B\times C}(\pi '_B\otimes \pi '_C,\pi _B\otimes \pi _C)<\epsilon . \end{aligned}$$
(A.10)

By (A.1) and the given hypothesis, there is \({\bar{n}}^1\in {\mathbb {N}}\), so that for \(n={\bar{n}}^1,{\bar{n}}^1+1,\ldots \),

$$\begin{aligned} (\pi _A)^n({\tilde{A}}_n)>1-\frac{\epsilon }{2}, \end{aligned}$$
(A.11)

where \({\tilde{A}}_n\) contains all \(a\in A^n\) such that

$$\begin{aligned} \rho _B(\varepsilon (y_n(a)),\pi _B)<\delta _B. \end{aligned}$$
(A.12)

By (A.1) and Lemma 1, on the other hand, there is \({\bar{n}}^2\in {\mathbb {N}}\), so that for \(n={\bar{n}}^2,\bar{n}^2+1,\ldots \),

$$\begin{aligned} (\pi _C)^n({\tilde{C}}_n)>1-\frac{\epsilon }{2}, \end{aligned}$$
(A.13)

where \({\tilde{C}}_n\) contains all \(c\in C^n\) such that

$$\begin{aligned} \rho _C(\varepsilon (c),\pi _C)<\delta _C. \end{aligned}$$
(A.14)

For any \(n={\bar{n}}^1\vee {\bar{n}}^2,{\bar{n}}^1\vee {\bar{n}}^2+1,\ldots \), let (ac) be an arbitrary member of \({\tilde{A}}_n\times {\tilde{C}}_n\). We have from (A.10), (A.12), and (A.14) that,

$$\begin{aligned} \rho _{B\times C}(\varepsilon (y_n(a),c),\pi _B\otimes \pi _C)<\epsilon . \end{aligned}$$
(A.15)

Noting the facilitating (ac) is but an arbitrary member of \({\tilde{A}}_n\times {\tilde{C}}_n\), we see that

$$\begin{aligned} \begin{array}{l} (\pi _A\otimes \pi _C)^n\left( \{(a,c)\in (A\times C)^n\mid \rho _{B\times C}(\varepsilon (y_n(a),c),\pi _B\otimes \pi _C)<\epsilon \}\right) \\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\ge (\pi _A)^n({\tilde{A}}_n)\times (\pi _C)^n({\tilde{C}}_n), \end{array} \end{aligned}$$
(A.16)

which by (A.11) and (A.13), is greater than \(1-\epsilon \). \(\square \)

Because the equivalence between tightness and relative compactness of a collection of probability measures is indirectly related to the proof of Lemma 3, we require B and C to be complete separable metric spaces.

Lemma 4

Given separable metric spaces A, B, C, and D, as well as distributions \(\pi _A\in \mathcal{P}(A)\), \(\pi _B\in \mathcal{P}(B)\), and \(\pi _C\in \mathcal{P}(C)\), suppose \(y_n\in \mathcal{M}(A^n,B^n)\) for every \(n\in {\mathbb {N}}\) and \(z\in \mathcal{K}(B,C,\pi _C,D)\). If

$$\begin{aligned} (\pi _A\otimes \pi _C)^n\left( \{a\in A^n,c\in C^n\mid \rho _{B\times C}(\varepsilon (y_n(a),c),\pi _B\otimes \pi _C)<\epsilon \}\right) >1-\epsilon , \end{aligned}$$

for any \(\epsilon >0\) and any n large enough, then

$$\begin{aligned} (\pi _A\otimes \pi _C)^n\left( \left\{ a\in A^n,c\in C^n\mid \rho _D(\varepsilon (y_n(a),c)\cdot z^{-1},(\pi _B\otimes \pi _C)\cdot z^{-1})<\epsilon \right\} \right) >1-\epsilon , \end{aligned}$$

for any \(\epsilon >0\) and any n large enough.

Proof

Let \(\epsilon >0\) be given. Since \(z\in \mathcal{K}(B,C,\pi _C,D)\), there exist \(C'\in {\mathscr {B}}(C)\) satisfying

$$\begin{aligned} \pi _C(C')>1-\frac{\epsilon }{2}, \end{aligned}$$
(A.17)

as well as

$$\begin{aligned} \delta \in (0,\epsilon /2], \end{aligned}$$
(A.18)

such that for any \(b,b'\in B\) and \(c,c'\in C'\) satisfying \(d_{B\times C}((b,c),(b',c'))<\delta \),

$$\begin{aligned} d_D(z(b,c),z(b',c'))<\epsilon . \end{aligned}$$
(A.19)

For any subset \(D'\) in \({\mathscr {B}}(D)\), we therefore have

$$\begin{aligned} (z^{-1}(D'))^\delta \cap (B\times C')\subseteq z^{-1}((D')^\epsilon ). \end{aligned}$$
(A.20)

This leads to \((z^{-1}(D'))^\delta \setminus (B\times (C \setminus C'))\subseteq z^{-1}((D')^\epsilon )\), and hence due to (A.17),

$$\begin{aligned} (\pi _B\otimes \pi _C)\left( z^{-1}((D')^\epsilon )\right) \ge (\pi _B\otimes \pi _C)\left( (z^{-1}(D'))^\delta \right) -\frac{\epsilon }{2}. \end{aligned}$$
(A.21)

On the other hand, by the hypothesis, we know for n large enough,

$$\begin{aligned} (\pi _A\otimes \pi _C)^n(E'_n)>1-\delta , \end{aligned}$$
(A.22)

where

$$\begin{aligned} E'_n=\{a\in A^n,c\in C^n\mid \rho _{B\times C}(\varepsilon (y_n(a),c),\pi _B\otimes \pi _C)<\delta \}\in {\mathscr {B}}^n(A\times C). \end{aligned}$$
(A.23)

By (A.23), for any \((a,c)\in E'_n\) and \(F'\in {\mathscr {B}}(B\times C)\),

$$\begin{aligned} (\pi _B\otimes \pi _C)((F')^\delta )\ge [\varepsilon (y_n(a),c)](F')-\delta . \end{aligned}$$
(A.24)

Combining the above, we have, for any \((a,c)\in E'_n\) and \(D'\in {\mathscr {B}}(D)\),

$$\begin{aligned} \begin{array}{l} [(\pi _B\otimes \pi _C)\cdot z^{-1}]((D')^\epsilon )=(\pi _B\otimes \pi _C)(z^{-1}((D')^\epsilon ))\\ \;\;\;\;\;\;\ge (\pi _B\otimes \pi _C)((z^{-1}(D'))^\delta )-\epsilon /2\ge [\varepsilon (y_n(a),c)](z^{-1}(D'))-\delta -\epsilon /2\\ \;\;\;\;\;\;\ge [\varepsilon (y_n(a),c)](z^{-1}(D'))-\epsilon =([\varepsilon (y_n(a),c)]\cdot z^{-1})(D')-\epsilon . \end{array} \end{aligned}$$
(A.25)

where the first inequality is due to (A.21), the second inequality is due to (A.24), and the third inequality is due to (A.18). That is, we have

$$\begin{aligned} \rho _D\left( \varepsilon (y_n(a),c)\cdot z^{-1},(\pi _B\otimes \pi _C)\cdot z^{-1}\right) \le \epsilon ,\quad \forall (a,c)\in E'_n. \end{aligned}$$
(A.26)

In view of (A.18) and (A.22), we have the desired result. \(\square \)

With Lemmas 1 to 4 ready, we can now prove Proposition 1 and then Theorem 1.

Proof of Proposition 1:

Let \(t=1,\ldots ,{\bar{t}}-1\) and \(x\in \mathcal{K}(S,G,\gamma ,X)\) be given. Define a map \(z\in \mathcal{M}(S\times G\times I,S)\), such that

$$\begin{aligned} z(s,g,i)=\theta _t\left( s,x(s,g),M(\sigma ,x),i\right) ,\quad \forall s\in S,g\in G,i\in I, \end{aligned}$$
(A.27)

where \(M(\sigma ,x)\) is given in (7). In view of (10) and (A.27), we have, for any \(S'\in {\mathscr {B}}(S)\),

$$\begin{aligned} \begin{array}{l} [T_t(x)\circ \sigma ](S')=\int _S\int _G\int _I\mathbf{1}(z(s,g,i)\in S')\cdot \iota (di)\cdot \gamma (dg)\cdot \sigma (ds)\\ \;\;\;\;\;\;=(\sigma \otimes \gamma \otimes \iota )(\{(s,g,i)\in S\times G\times I\mid z(s,g,i)\in S'\})=(\sigma \otimes \gamma \otimes \iota )(z^{-1}(S')). \end{array} \end{aligned}$$
(A.28)

For \(n\in {\mathbb {N}}\), \(g\equiv (g_m)_{m=1,\ldots ,n}\in G^n\), and \(i\equiv (i_m)_{m=1,\ldots ,n}\in I^n\), also define an operator \(T'_n(g,i)\) on \(\mathcal{P}_n(S)\) so that \(T'_n(g,i)\circ \varepsilon (s)=\varepsilon (s')\), where for \(m=1,2,\ldots ,n\),

$$\begin{aligned} s'_m=z(s_m,g_m,i_m)=\theta _t\left( s_m,x(s_m,g_m),M(\sigma ,x),i_m\right) . \end{aligned}$$
(A.29)

It is worth noting that (A.29) is different from the earlier (15). In view of (A.27) and (A.29), we have, for \(S'\in {\mathscr {B}}(S)\), that \([T'_n(g,i)\circ \varepsilon (s)](S')\) equals

$$\begin{aligned} \frac{1}{n}\cdot \sum _{m=1}^n\mathbf{1}\left( z(s_m,g_m,i_m)\in S'\right) =\varepsilon ((s_1,g_1,i_1),\ldots ,(s_n,g_n,i_n))\left( z^{-1}(S')\right) . \end{aligned}$$
(A.30)

Combining (A.28) and (A.30), we arrive to a key observation that

$$\begin{aligned} T_t(x)\circ \sigma =(\sigma \otimes \gamma \otimes \iota )\cdot z^{-1},\quad \text{ while } \quad T'_n(g,i)\circ \varepsilon (s)=\varepsilon (s,g,i)\cdot z^{-1}. \end{aligned}$$
(A.31)

In the rest of the proof, we first show the asymptotic closeness between \(T_t(x)\circ \sigma \) and \(T'_n(g,i)\circ \varepsilon (s_n(a))\), and then that between the latter and \(T_{nt}(x,g,i)\circ \varepsilon (s_n(a))\).

First, due to the hypothesis on the convergence of \(\varepsilon (s_n(a))\) to \(\sigma \), the completeness of the spaces S, G, and I and hence also the completeness of \(G\times I\), as well as Lemma 3,

$$\begin{aligned} (\pi \otimes \gamma \otimes \iota )^n(\{(a,g,i)\in (A\times G\times I)^n\mid \rho _{S\times G\times I}(\varepsilon (s_n(a),g,i),\sigma \otimes \gamma \otimes \iota )<\epsilon '\})>1-\epsilon ', \end{aligned}$$
(A.32)

for any \(\epsilon '>0\) and any n large enough. We then show z as defined in (A.27) is a member of \(\mathcal{K}(S,G\times I,\gamma \otimes \iota ,S)\) according to the definition around (17) using (S1) and \(x\in \mathcal{K}(S,G,\gamma ,X)\).

Fix any \(\epsilon >0\). By (S1), there exist \(\delta >0\) and \(I'\in {\mathscr {B}}(I)\) with \(\iota (I')>1-\epsilon /2\) such that

$$\begin{aligned} d_S(\theta _t(s,y,M(\sigma ,x),i),\theta _t(s',y',M(\sigma ,x),i'))<\epsilon , \end{aligned}$$
(A.33)

for any \((s,y),(s',y')\in S\times X\) and \(i,i'\in I'\) satisfying \(d_{S\times X\times I}((s,y,i),(s',y',i'))<\delta \). Since \(x\in \mathcal{K}(S,G,\gamma ,X)\), there exist \(\delta '\in (0,\delta /2]\) and \(G'\in {\mathscr {B}}(G)\) with \(\gamma (G')>1-\epsilon /2\) such that

$$\begin{aligned} d_X(x(s,g),x(s',g'))<\frac{\delta }{2}, \end{aligned}$$
(A.34)

for any \(s,s'\in S\) and \(g,g'\in G'\) satisfying \(d_{S\times G}((s,g),(s',g'))<\delta '\).

Now suppose \(s,s'\in S\) and \((g,i),(g',i')\in G'\times I'\) satisfy

$$\begin{aligned} d_{S\times G\times I}((s,g,i),(s',g',i'))<\delta '. \end{aligned}$$
(A.35)

By the first inequality of (3), \(d_{S\times G}((s,g),(s',g'))<\delta '\). This and the fact that \(g,g'\in G'\) would result in (A.34). Due to the first inequality of (3), another consequence of (A.35) is

$$\begin{aligned} d_{S\times I}((s,i),(s',i'))<\delta '\le \frac{\delta }{2}. \end{aligned}$$
(A.36)

With the second inequality of (3), we can conclude from (A.34) and (A.36) that

$$\begin{aligned} d_{S\times X\times I}((s,x(s,g),i),(s',x(s',g'),i'))<\delta . \end{aligned}$$
(A.37)

As \(i,i'\in I'\), we can see from (A.33) that

$$\begin{aligned} d_S(\theta _t(s,x(s,g),M(\sigma ,x),i),\theta _t(s',x(s',g') ,M(\sigma ,x),i'))<\epsilon . \end{aligned}$$
(A.38)

In addition, the measures of \(G'\) and \(I'\) would lead to

$$\begin{aligned} (\gamma \otimes \iota )(G'\times I')\ge 1-(1-\gamma (G'))-(1-\iota (I'))>1-\epsilon . \end{aligned}$$
(A.39)

Since \(\epsilon >0\) is arbitrary, (17), (A.35), (A.38), and (A.39) would together mean that z as defined through (A.27) is a member of \(\mathcal{K}(S,G\times I,\gamma \otimes \iota ,S)\).

By Lemma 4, this fact along with (A.32) will lead to the strict dominance of \(1-\epsilon '\) by

$$\begin{aligned} (\pi \otimes \gamma \otimes \iota )^n(\{(a,g,i)\in (A\times G\times I)^n\mid \rho _S(\varepsilon (s_n(a),g,i)\cdot z^{-1},(\sigma \otimes \gamma \otimes \iota )\cdot z^{-1})<\epsilon '\}), \end{aligned}$$
(A.40)

for any \(\epsilon '>0\) and any n large enough. By (A.31), this is equivalent to that, given \(\epsilon >0\), there exists \({\bar{n}}^1\in {\mathbb {N}}\) so that for any \(n={\bar{n}}^1,{\bar{n}}^1+1,\ldots \),

$$\begin{aligned} (\pi \otimes \gamma \otimes \iota )^n \left( {\tilde{A}}_n(\epsilon )\right) >1-\frac{\epsilon }{2}, \end{aligned}$$
(A.41)

where \({\tilde{A}}_n(\epsilon )\in {\mathscr {B}}^n(A\times G\times I)\) is equal to

$$\begin{aligned} \left\{ (a,g,i)\in (A\times G\times I)^n\mid \rho _S\left( T_t(x)\circ \sigma ,T'_n(g,i)\circ \varepsilon (s_n(a))\right) <\frac{\epsilon }{2}\right\} . \end{aligned}$$
(A.42)

Next, note that the only difference between \(T_{nt}(x,g,i)\circ \varepsilon (s_n(a))\) and \(T'_n(g,i)\circ \varepsilon (s_n(a))\) lies in that \(\varepsilon (s_{n,-m}(a),g_{-m})\) is used in the former as in (15), whereas \(\sigma \otimes \gamma \) is used in the latter as in (A.29). Here, \(s_{n,-m}(a)\) refers to the vector \((s_{n1}(a),\ldots ,s_{n,m-1}(a),s_{n,m+1}(a),\ldots ,s_{nn}(a))\). By (S2), there is \(\delta \in (0,\epsilon /4]\) and \(I'\in {\mathscr {B}}(I)\) with

$$\begin{aligned} \iota (I')>1-\frac{\epsilon }{4}, \end{aligned}$$
(A.43)

so that for any \((s,g,i)\in S\times G\times I'\) and any \(\mu '\in \mathcal{P}(S\times X)\) satisfying \(\rho _{S\times X}(M(\sigma ,x),\mu ')<\delta \),

$$\begin{aligned} d_S\left( \theta _t(s,x(s,g),M(\sigma ,x),i), \theta _t(s,x(s,g),\mu ',i)\right) <\frac{\epsilon }{2}. \end{aligned}$$
(A.44)

For each \(n\in {\mathbb {N}}\), define \(I'_n\) so that

$$\begin{aligned} I'_n=\left\{ i\equiv (i_m)_{m=1,\ldots ,n}\in I^n\mid \text{ more } \text{ than } \left( 1-\frac{\epsilon }{2}\right) \cdot n \text{ components } \text{ come } \text{ from } I'\right\} . \end{aligned}$$
(A.45)

Also important is that by (A.44) and (A.45), for any \(S'\in {\mathscr {B}}(S)\) and \(i\equiv (i_m)_{m=1,\ldots ,n}\in I'_n\),

$$\begin{aligned} \left[ T_{nt}(x,g,i)\circ \varepsilon (s_n(a))\right] \left( (S')^{\epsilon /2}\right) +\frac{\epsilon }{2}\ge \left[ T'_n(g,i)\circ \varepsilon (s_n(a))\right] (S'), \end{aligned}$$
(A.46)

whenever

$$\begin{aligned} \rho _{S\times X}\left( M(\sigma ,x),M_n(\varepsilon (s_{n,-m}(a)),x,g_{-m})\right) <\delta . \end{aligned}$$
(A.47)

It can be shown that \(I'_n\) will occupy a big chunk of \(I^n\) as measured by \(\iota ^n\) when n is large. Define a map q from I to \(\{0,1\}\) so that \(q(i)=1\) or 0 depending on whether or not \(i\in I'\). By (A.43), \(\iota \cdot q^{-1}\) is a Bernoulli distribution with \((\iota \cdot q^{-1})(\{1\})>1-\epsilon /4\). So by (A.45), \(I'_n\) contains all \(i\equiv (i_m)_{m=1,\ldots ,n}\in I^n\) that satisfy

$$\begin{aligned} \rho _{\{0,1\}}(\varepsilon (i)\cdot q^{-1},\iota \cdot q^{-1})<\frac{\epsilon }{4}. \end{aligned}$$
(A.48)

Therefore, by Lemma 1, there exits \({\bar{n}}^2\in {\mathbb {N}}\), so that for \(n={\bar{n}}^2,{\bar{n}}^2+1,\ldots \),

$$\begin{aligned} \iota ^n(I'_n)>1-\frac{\epsilon }{4}. \end{aligned}$$
(A.49)

We can also demonstrate that (A.47) will be highly likely when n is large. By Lemma 3 and the hypothesis on the convergence of \(\varepsilon (s_n(a))\) to \(\sigma \), we know \(\varepsilon (s_n(a),g)\) will converge to \(\sigma \otimes \gamma \) in probability. Due to Lemma 2, this conclusion applies to the sequence \(\varepsilon (s_{n,-m}(a),g_{-m})\) as well. The fact that \(x\in \mathcal{K}(S,G,\gamma ,X)\) certainly leads to \((\text{ prj }^{S\times G}_S,x)\in \mathcal{K}(S,G,\gamma ,S\times X)\). So by Lemma 4, there is \({\bar{n}}^3\in {\mathbb {N}}\), so that for \(n={\bar{n}}^3,{\bar{n}}^3+1,\ldots \),

$$\begin{aligned} (\pi ^n\otimes \gamma ^n)\left( {\tilde{B}}_n(\delta )\right) >1-\frac{\epsilon }{4}, \end{aligned}$$
(A.50)

where

$$\begin{aligned} {\tilde{B}}_n(\delta )=\{(a,g)\in A^n\times G^n\mid (A.47) \text{ is } \text{ true }\}\in {\mathscr {B}}^n(A\times G). \end{aligned}$$
(A.51)

Consider arbitrary \(n={\bar{n}}^1\vee {\bar{n}}^2\vee {\bar{n}}^3,\bar{n}^1\vee {\bar{n}}^2\vee {\bar{n}}^3+1,\ldots \), \((a,g,i)\in \tilde{A}_n(\epsilon )\cap ({\tilde{B}}_n(\delta )\times I'_n)\), and \(S'\in {\mathscr {B}}(S)\). By (A.1) and (A.42), we see that

$$\begin{aligned}{}[T'_n(g,i)\circ \varepsilon (s_n(a))]\left( (S')^{\epsilon /2}\right) +\frac{\epsilon }{2}\ge [T_t(x)\circ \sigma ](S'). \end{aligned}$$
(A.52)

Combining this with (A.46), (A.47), and (A.51), we obtain

$$\begin{aligned}{}[T_{nt}(x,g,i)\circ \varepsilon (s_n(a))]\left( (S')^\epsilon \right) +\epsilon \ge [T'_n(g,i)\circ \varepsilon (s_n(a))]\left( (S')^{\epsilon /2}\right) +\frac{\epsilon }{2}\ge [T_t(x)\circ \sigma ](S'). \end{aligned}$$
(A.53)

According to (A.1), this means

$$\begin{aligned} \rho _S\left( T_{nt}(x,g,i)\circ \varepsilon (s_n(a)),T_t(x)\circ \sigma \right) \le \epsilon . \end{aligned}$$
(A.54)

Therefore, for \(n\ge {\bar{n}}^1\vee {\bar{n}}^2\vee {\bar{n}}^3\),

$$\begin{aligned} \begin{array}{l} (\pi \otimes \gamma \otimes \iota )^n\left( \{(a,g,i)\in (A\times G\times I)^n\mid \rho _S(T_{nt}(x,g,i)\circ \varepsilon (s_n(a)),T_t(x)\circ \sigma )\le \epsilon \}\right) \\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\ge (\pi \otimes \gamma \otimes \iota )^n\left( \tilde{A}_n(\epsilon )\cap ({\tilde{B}}_n(\delta )\times I'_n)\right) , \end{array}\end{aligned}$$
(A.55)

whereas the latter is, in view of (A.41), (A.49), and (A.50), greater than \(1-\epsilon \). \(\square \)

Proof of Theorem 1:

We use induction to show that, for each \(\tau =0,1,\ldots ,{\bar{t}}-t+1\),

$$\begin{aligned} \left( \sigma _t\otimes \gamma ^\tau \otimes \iota ^\tau \right) ^n\left( \tilde{A}_{n\tau }(\epsilon )\right) >1-\frac{\epsilon }{{\bar{t}}-t+2}, \end{aligned}$$
(A.56)

for any \(\epsilon >0\) and n large enough, where \(\tilde{A}_{n\tau }(\epsilon )\in {\mathscr {B}}^n(S\times G^\tau \times I^\tau )\) is such that, for any \((s_t,g_{[t,t+\tau -1]},i_{[t,t+\tau -1]})\in {\tilde{A}}_{n\tau }(\epsilon )\),

$$\begin{aligned} \rho _S\left( T_{n,[t,t+\tau -1]}(x_{[t,t+\tau -1]},g_{[t,t+\tau -1]}, i_{[t,t+\tau -1]})\circ \varepsilon (s_t),T_{[t,t+\tau -1]}(x_{[t,t+\tau -1]})\circ \sigma _t\right) <\epsilon . \end{aligned}$$
(A.57)

Once the above is achieved, we can then define \({\tilde{A}}_n(\epsilon )\) required in the theorem by

$$\begin{aligned} {\tilde{A}}_n(\epsilon )=\bigcap _{\tau =0}^{{\bar{t}}-t+1}\left[ {\tilde{A}}_{n\tau }(\epsilon )\times G^{n\cdot ({\bar{t}}-t+1-\tau )}\times I^{n\cdot ({\bar{t}}-t+1-\tau )}\right] . \end{aligned}$$
(A.58)

By this and (A.56), we have \(\left( \sigma _t\otimes \gamma ^{{\bar{t}}-t+1}\otimes \iota ^{{\bar{t}}-t+1}\right) ^n\left( \tilde{A}_n(\epsilon )\right) \) greater than

$$\begin{aligned} \begin{array}{l} 1-\sum _{\tau =0}^{{\bar{t}}-t+1}\left[ 1-\left( \sigma _t\otimes \gamma ^{{\bar{t}}-t+1}\otimes \iota ^{{\bar{t}}-t+1}\right) \left( {\tilde{A}}_{n\tau }(\epsilon )\times G^{n\cdot ({\bar{t}}-t+1-\tau )}\times I^{n\cdot ({\bar{t}}-t+1-\tau )}\right) \right] \\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;=1-\sum _{\tau =0}^{{\bar{t}}-t+1}\left[ 1-(\sigma _t\otimes \gamma ^\tau \otimes \iota ^\tau )({\tilde{A}}_{n\tau }(\epsilon ))\right] \\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;>1-({\bar{t}}-t+2)\cdot \left[ 1-(1-\epsilon /({\bar{t}}-t+2))\right] =1-\epsilon , \end{array}\end{aligned}$$
(A.59)

for any \(\epsilon >0\) and n large enough.

Now we proceed with the induction process. First, note that \(T_{n,[t,t-1]}\circ \varepsilon (s_t)\) is merely \(\varepsilon (s_t)\) itself and \(T_{[t,t-1]}\circ \sigma _t\) is merely \(\sigma _t\) itself. Hence, we will have (A.56) for \(\tau =0\) for any \(\epsilon >0\) and n large enough just by Lemma 1. Then, for some \(\tau =1,2,\ldots ,{\bar{t}}-t+1\), suppose

$$\begin{aligned} \left( \sigma _t\otimes \gamma ^{\tau -1}\otimes \iota ^{\tau -1}\right) ^n\left( \tilde{A}_{n,\tau -1}(\epsilon )\right) >1-\frac{\epsilon }{{\bar{t}}-t+2}, \end{aligned}$$
(A.60)

for any \(\epsilon >0\) and n large enough. We may apply Proposition 1 to the above, while at the same time identifying \(S\times G^{\tau -1}\times I^{\tau -1}\) with A, \(\sigma _t\otimes \gamma ^{\tau -1}\otimes \iota ^{\tau -1}\) with \(\pi \), \(x_{t+\tau -1}\) with x, \(T_{n,[t,t+\tau -2]}(x_{[t,t+\tau -2]},g_{[t,t+\tau -2]},i_{[t,t+\tau -2]})\circ \varepsilon (s_t)\) with \(\varepsilon (s_n(a))\), and \(T_{[t,t+\tau -2]}(x_{[t,t+\tau -2]})\circ \sigma _t\) with \(\sigma \). This way, we will verify (A.56) for any \(\epsilon >0\) and n large enough. Therefore, the induction process can be completed. \(\square \)

Technical Developments in Sect. 5

Proof of Proposition 2:

Because payoffs are bounded, the value functions are bounded too. We then prove by induction on t. By (20), we know the result is true for \(t={\bar{t}}+1\). Suppose for some \(t={\bar{t}},{\bar{t}}-1,\ldots ,2\), we have the continuity of \(v_{t+1}(s_{t+1},\sigma _{t+1},x_{[t+1,{\bar{t}}]},x_{t+1})\) in \(s_{t+1}\). By this induction hypothesis, the distribution-wise uniform continuity of \(x_t\), (S1), (F1), and the boundedness of the value functions, we see the continuity of the right-hand side of (21) in \(s_t\). So, \(v_t(s_t,\sigma _t,x_{[t{\bar{t}}]},x_t)\) is continuous in \(s_t\), and we have completed our induction process. \(\square \)

Proof of Proposition 3:

We prove by induction on t. By (20) and (24), we know the result is true for \(t={\bar{t}}+1\). Suppose for some \(t={\bar{t}},{\bar{t}}-1,\ldots ,2\), we have the convergence of \(v_{n,t+1}(s_{t+1,1},\varepsilon (s^n_{t+1,-1}),x_{[t+1,{\bar{t}}]},x_{t+1})\) to \(v_{t+1}(s_{t+1,1},\sigma _{t+1},x_{[t+1,{\bar{t}}]},x_{t+1})\) at an \(s_{t+1,1}\)-independent rate when \(s_{t+1,-1}\equiv (s_{t+1,2},s_{t+1,3},\ldots )\) is sampled from \(\sigma _{t+1}\). Now, suppose \(s_{t,-1}\equiv (s_{t2},s_{t3},\ldots )\) is sampled from \(\sigma _t\). Let also \(g\equiv (g_1,g_2,\ldots )\) be generated through sampling on \((G,{\mathscr {B}}(G),\gamma )\) and \(i\equiv (i_1,i_2,\ldots )\) be generated through sampling on \((I,{\mathscr {B}}(I),\iota )\). In the remainder of the proof, let \(s^n_t\equiv (s_{t1},s_{t2},\ldots ,s_{tn})\) for any arbitrary \(s_{t1}\in S\), \(g^n\equiv (g_1,\ldots ,g_n)\) and \(i^n\equiv (i_1,\ldots ,i_n)\).

Due to Lemma 1, \(\varepsilon (s^n_{t,-1})\) will converge to \(\sigma _t\). By Lemma 2, \(\varepsilon (s^n_t)\) will converge to \(\sigma _t\) at an \(s_{t1}\)-independent rate. By Proposition 1, we know that \(T_{nt}(x_t,g^n,i^n)\circ \varepsilon (s^n_t)\) will converge to \(T_t(x_t)\circ \sigma _t\) in probability at an \(s_{t1}\)-independent rate, and by Lemma 2 again, so will \([T_{nt}(x_t,g^n,i^n)\circ \varepsilon (s^n_t)]_{-1}\) to \(T_t(x_t)\circ \sigma _t\). Now Lemma 3 will lead to the convergence in probability of \(\varepsilon (s^n_{t,-1},g^n_{-1})\) to \(\sigma _t\otimes \gamma \). Due to \(x_t\)’s distribution-wise uniform continuity, Lemma 4 will lead to the convergence in probability of \(M_n(\varepsilon (s^n_{t,-1}),x_t,g^n_{-1})\) to \(M(\sigma _t,x_t)\). Thus,

  1. \(\psi _t(s_{t1},x_t(s_{t1},g_1),M_n(\varepsilon (s^n_{t,-1}),x_t,g^n_{-1}))\) will converge to \(\psi _t(s_{t1},x_t(s_{t1},g_1), M(\sigma _t,x_t))\) in probability at an \(s_{t1}\)-independent rate due to (F2);

  2. \(v_{n,t+1}(\theta _t(s_{t1},x_t(s_{t1},g_1),M_n(\varepsilon (s^n_{t,-1}),x_t,g^n_{-1}),i_1),[T_{nt}(x_t,g^n,i^n)\circ \varepsilon (s^n_t)]_{-1}, x_{[t+1,{\bar{t}}]},x_{t+1})\) will converge to \(v_{t+1}(\theta _t(s_{t1},x_t(s_{t1},g_1),M_n(\varepsilon (s^n_{t,-1}),x_t,g^n_{-1})),i_1),T_t(x_t)\circ \sigma _t,x_{[t+1,{\bar{t}}]},x_{t+1})\) in probability at an \(s_{t1}\)-independent rate due to the induction hypothesis; the latter will in turn converge to \(v_{t+1}(\theta _t(s_{t1},x_t(s_{t1},g_1),M(\sigma _t,x_t),i_1),T(x_t)\circ \sigma _t,x_{[t+1,{\bar{t}}]},x_{t+1})\) in probability at an \(s_{t1}\)-independent rate due to (S2) and Proposition 2.

As per-period payoffs are bounded, all value functions are bounded. The above convergences will then lead to the convergence of the right-hand side of (25) to the right-hand side of (21) at an \(s_{t1}\)-independent rate. That is, \(v_{nt}(s_{t1},\varepsilon (s^n_{t,-1}),x_{[t{\bar{t}}]},x_t)\) will converge to \(v_t(s_{t1},\sigma _t,x_{[t{\bar{t}}]},x_t)\) at a rate independent of \(s_{t1}\). We have completed the induction process. \(\square \)

Proof of Theorem 2:

Let us consider subgames starting with some time \(t=1,2,\ldots ,{\bar{t}}\). For convenience, we let \(\sigma _t=T_{[1,t-1]}(x^*_{[1,t-1]})\circ \sigma _1\). Now let \(s_t\equiv (s_{t1},s_{t2},\ldots )\) be generated through sampling on \((S,{\mathscr {B}}(S),\sigma _t)\), \(g\equiv (g_1,g_2,\ldots )\) be generated through sampling on \((G,{\mathscr {B}}(G),\gamma )\), and \(i\equiv (i_1,i_2,\ldots )\) be generated through sampling on \((I,{\mathscr {B}}(I),\iota )\). In the remainder of the proof, we let \(s^n_t\equiv (s_{t1},\ldots ,s_{tn})\), \(s^n_{t,-1}\equiv (s_{t2},\ldots ,s_{tn})\), \(g^n\equiv (g_1,\ldots ,g_n)\), and \(i^n\equiv (i_1,\ldots ,i_n)\).

By Lemma 1 and Proposition 1, we know that \(\varepsilon (s^n_t)=\varepsilon (s_{t1},\ldots ,s_{tn})\) converges to \(\sigma _t\) in probability, and also that \(T_{nt}(x^*_t,g^n,i^n)\circ \varepsilon (s^n_t)\) converges to \(T_t(x^*_t)\circ \sigma _t\) in probability. Due to Lemma 2, \(\varepsilon (s^n_{t,-1})\) and \([T_{nt}(x^*_t,g^n,i^n)\circ \varepsilon (s^n_t)]_{-1}\) will have the same respective convergences. Also, Lemma 3 will lead to the convergence in probability of \(\varepsilon (s^n_{t,-1},g^n_{-1})\) to \(\sigma _t\otimes \gamma \). Due to \(x_t\)’s distribution-wise uniform continuity, Lemma 4 will lead to the convergence in probability of \(M_n(\varepsilon (s^n_{t,-1}),x_t,g^n_{-1})\) to \(M(\sigma _t,x_t)\). Then,

  1. \(\psi _t(s_{t1},y(s_{t1},g_1),M_n(\varepsilon (s^n_{t,-1}),x_t,g_{-1}))\) will converge to \(\psi _t(s_{t1},y(s_{t1},g_1), M(\sigma _t,x_t))\) in probability at a y-independent rate due to (F2);

  2. \(v_{n,t+1}(\theta _t(s_{t1},y(s_{t1},g_1), M_n(\varepsilon (s^n_{t,-1}),x_t,g^n_{-1}),i_1),[T_{nt}(x^*_t,g^n,i^n)\circ \varepsilon (s^n_t)]_{-1}, x^*_{[t+1,{\bar{t}}]},x^*_{t+1})\) will converge to \(v_{t+1}(\theta _t(s_{t1},y(s_{t1},g_1),M_n(\varepsilon (s^n_{t,-1}),x_t,g^n_{-1}),i_1),T_t(x^*_t)\circ \sigma _t, x^*_{[t+1,{\bar{t}}]}, x^*_{t+1})\) in probability at a y-independent rate due to Proposition 3, which, due to (S2) and Proposition 2, will converge to \(v_{t+1}(\theta _t(s_{t1},y(s_{t1},g_1),M(\sigma _t,x_t),i_1),T_t(x^*_t)\circ \sigma _t,x^*_{[t+1,{\bar{t}}]},x^*_{t+1})\) in probability at a y-independent rate.

As per-period payoffs are bounded, all value functions are bounded. By (21) and (25), the above convergences will then lead to the convergence of the left-hand side of (31) to the left-hand side of (27). At the same time, the right-hand side of (31) plus \(\epsilon \) will converge to the right-hand side of (27) due to the convergence of \(\varepsilon (s^n_{t,-1})\) to \(\sigma _t\), Proposition 3, and the uniform boundedness of the value functions. By (27), as long as n is large enough, (31) will be true for any \(\epsilon >0\) and \(y\in \mathcal{M}(S\times G,X)\). This would then lead to (32) due to Theorem 1 and the boundedness of payoff functions. \(\square \)

Technical Developments in Sect. 6

Value Functions for the Stationary Case: For \(t=0,1,\ldots \), we define \(v_t(s,\sigma ,x,y)\) as the total expected payoff a player can make from period 1 to t, when he starts period 1 with a state \(s\in S\) and a state-variable profile \(\sigma \), while all players keep on using the strategy x from period 1 to t with the exception of the current player in the very beginning, who deviates to \(y\in \mathcal{M}(S\times G,X)\) then. As a terminal condition, we have

$$\begin{aligned} v_0(s,\sigma ,x,y)=0. \end{aligned}$$
(C.1)

Due to the stationarity of the setting, we have, for \(t=1,2,\ldots \),

$$\begin{aligned} \begin{array}{ll} v_t(s,\sigma ,x,y)=\int _G[\psi (s,y(s,g),M(\sigma ,x))\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+\alpha \cdot \int _I v_{t-1}(\theta (s,y(s,g),M(\sigma ,x),i),\sigma ,x,x)\cdot \iota (di)]\cdot \gamma (dg). \end{array} \end{aligned}$$
(C.2)

This is much like (21); but with (39) being true, the last term in (C.2) actually appears simpler than its counterpart. Using (C.1) and (C.2), we can inductively show that

$$\begin{aligned} \mid v_{t+1}(s,\sigma ,x,y)-v_t(s,\sigma ,x,y)\mid \le \alpha ^t\cdot {{\bar{\psi }}},\quad t=0,1,\ldots . \end{aligned}$$
(C.3)

The sequence \(\{v_t(s,\sigma ,x,y)\}_{t=0,1,\ldots }\) is thus Cauchy with a limit point \(v_\infty (s,\sigma ,x,y)\). This \(v_\infty (s,\sigma ,x,y)\) can be understood as the infinite-horizon total discounted expected payoff a player can obtain by starting with state s and environment \(\sigma \), while all players adhere to the action plan x except for the current player in the beginning, who deviates to y then.

Now we move on to the n-player game \(\Gamma _n\) with the same stationary features provided by \(\psi \), \(\theta \), and \(\alpha \). The in-action environment experienced by player m will be \(M_n(\varepsilon (s_{-m}),x,g_{-m})\), as defined in (14), when the other players start with the state vector \(s_{-m}\equiv (s_l)_{l\ne m}\), all act according to some x strategy, and experience the pre-action shock vector \(g_{-m}\equiv (g_l)_{l\ne m}\). Given any strategy \(x\in \mathcal{M}(S\times G,X)\), pre-action shock vector \(g\equiv (g_m)_{m=1,\ldots ,n}\in G^n\), and post-action shock vector \(i\equiv (i_m)_{m=1,\ldots ,n}\in I^n\), we define \(T_n(x,g,i)\) as the operator on \(\mathcal{P}_n(S)\) that converts a period’s state-variable profile into that of the next period. Following the transient version (15), \(\varepsilon (s')=T_n(x,g,i)\circ \varepsilon (s)\) is such that

$$\begin{aligned} s'_m=\theta \left( s_m,x(s_m,g_m),M_n(\varepsilon (s_{-m}),x,g_{-m}), i_m\right) ,\quad \forall m=1,2,\ldots ,n. \end{aligned}$$
(C.4)

Let \(v_{nt}(s_1,\varepsilon (s_{-1}),x,y)\) be the total expected payoff player 1 can make from period 1 to t, when the player’s starting state is \(s_1\in S\), the other players’ initial states are given by the vector \(s_{-1}\equiv (s_m)_{m\ne 1}\), and all players adopt the strategy \(x\in \mathcal{M}(S\times G,X)\) with the exception of player 1, who adopts the strategy \(y\in \mathcal{M}(S\times G,X)\) in the beginning. We have

$$\begin{aligned} v_{n0}(s_1,\varepsilon (s_{-1}),x,y)=0. \end{aligned}$$
(C.5)

For \(t=1,2,\ldots \), similarly to (25), \(v_{nt}(s_1,\varepsilon (s_{-1}),x,y)\) is equal to

$$\begin{aligned} \begin{array}{l} \int _{G^n}\gamma ^n(dg)\times \{\psi \left( s_1,y(s_1,g_1), M_n(\varepsilon (s_{-1}),x,g_{-1})\right) +\alpha \cdot \int _{I^n}\iota ^n(di)\times \\ \;\;\;\;\;\;\;\;\;\;\;\;\times v_{n,t-1}\left( \theta (s_1,y(s_1,g_1),M_n (\varepsilon (s_{-1}),x,g_{-1}),i_1),[T_n(x,g,i)\circ \varepsilon (s)]_{-1},x,x\right) \}. \end{array} \end{aligned}$$
(C.6)

In (C.6), \([T_n(x,g,i)\circ \varepsilon (s)]_{-1}\) stands for \(\varepsilon (s'_{-1})\), while \(s'\) comes from \(\varepsilon (s')=T_n(x,g,i)\circ \varepsilon (s)\). Using (C.5) and (C.6), we can inductively show that

$$\begin{aligned} \mid v_{n,t+1}(s_1,\varepsilon (s_{-1}),x,y)-v_{nt}(s_1,\varepsilon (s_{-1}),x,y)\mid \le \alpha ^t\cdot {{\bar{\psi }}},\forall t=0,1,\ldots . \end{aligned}$$
(C.7)

Thus, the sequence \(\{v_{nt}(s_1,\varepsilon (s_{-1}),x,y)\}_{t=0,1,\ldots }\) is Cauchy with limit \(v_{n\infty }(s_1,\varepsilon (s_{-1}),x,y)\).

Proof of Theorem 3:

Let \(\epsilon >0\) be fixed. For \(t=1,2,\ldots \) satisfying \(t\ge \ln (6{{\bar{\psi }}}/(\epsilon \cdot (1-\alpha )))/\ln (1/\alpha )+1\), we have from (C.6) and (C.7),

$$\begin{aligned} \mid v_{n\infty }(s_1,\varepsilon (s_{-1}),x^*,y)-v_{nt} (s_1,\varepsilon (s_{-1}),x^*,y)\mid <\frac{\epsilon }{6}. \end{aligned}$$
(C.8)

Therefore, we need merely to select such a large t and show that, when n is large enough,

$$\begin{aligned} \int _{S^n} v_{n t}(s_1,\varepsilon (s_{-1}),x^*,x^*)\cdot (\sigma ^*)^n(ds)\ge \int _{S^n} v_{n t}(s_1,\varepsilon (s_{-1}),x^*,y)\cdot (\sigma ^*)^n(ds)-\frac{2\epsilon }{3}. \end{aligned}$$
(C.9)

For \(t=1,2,\ldots \), since \((x^*,\sigma ^*)\) forms an equilibrium for \(\Gamma \), we know (42) is true. This, as well as (C.2) and (C.3), lead to

$$\begin{aligned} \alpha ^{t-\tau }\cdot \left[ \int _S v_\tau (s,\sigma ^*,x^*,y)\cdot \sigma ^*(ds)-\int _S v_\tau (s,\sigma ^*,x^*,x^*)\cdot \sigma ^*(ds)\right] \le \frac{2\alpha ^{t-1}\cdot {{\bar{\psi }}}}{1-\alpha }\le \frac{\epsilon }{3}. \end{aligned}$$
(C.10)

for \(\tau =1,2,\ldots ,t\), \(g\in G\), \(s\in S\), and \(y\in \mathcal{M}(S\times G,X)\).

We associate entities here with those defined in Sect. 5 when \({\bar{t}}\) there is fixed at the t here. To signify the difference in the two notational systems, we add superscript “K” to symbols defined in the previous section. For instance, we write \(v^K_\tau \) for the \(v_\tau \) defined in that section, which has a different meaning than the \(v_\tau \) here. Now, our \(\alpha ^{t-\tau }\cdot v_\tau (s,\sigma ^*,x^*,y)\) can be understood as \(v^K_{t+1-\tau }(s,\sigma ^*,x',y)\), with \(x'\equiv (x'_{t+1-\tau },\ldots ,x'_t)\in (\mathcal{M}(S\times G,X))^\tau \) being such that \(x'_{t'}=x^*\) for \(t'=t+1-\tau ,\ldots ,t\). Due to the consistency of \(\sigma ^*\) with \(x^*\) through the definition (39), we can understand \(\sigma ^*\) as \(T^K_{[1,\tau -1]}(x'_{[1,\tau -1]})\circ \sigma ^K_1\), where \(x'_{[1,\tau -1]}\equiv (x'_1,\ldots ,x'_{\tau -1})\in (\mathcal{M}(S\times G,X))^{\tau -1}\) is such that \(x'_{t'}=x^*\) for \(t'=1,2,\ldots ,\tau -1\).

With these correspondences, (C.10) can be translated into something akin to (27), with the only difference being that \(-\epsilon /3\) should be added to all the right-hand sides. That is, we now know that the current \((x^*,\sigma ^*)\) offers an \((\epsilon /3)\)-Markov equilibrium for the nonatomic game \(\Gamma ^{K}(\sigma ^*)\) with \({\bar{t}}=t\), \(\theta ^K_\tau =\theta \), and \(\psi ^K_\tau =\alpha ^{\tau -1}\cdot \psi \). Even though Theorem 2 is nominally about going from a 0-equilibrium for the nonatomic game to \(\epsilon \)-equilibria for finite games, we can follow exactly the same logic used to prove it to go from an \((\epsilon /3)\)-equilibrium for the nonatomic game to \((2\epsilon /3)\)-equilibria for finite games.

Thus, from one of the theorem’s claims, we can conclude that, for n large enough and any \(y\in \mathcal{M}(S\times G,X)\),

$$\begin{aligned} \int _{S^n}\left( \sigma ^K_1\right) ^n(ds)\cdot v^K_{nt}\left( s_{1},\varepsilon (s_{-1}),x'_{[1t]},x'_1\right) \ge \int _{S^n}\left( \sigma ^K_1\right) ^n(ds)\cdot v^{K}_{nt}\left( s_{1},\varepsilon (s_{-1}),x'_{[1t]},y\right) -\frac{2\epsilon }{3}, \end{aligned}$$
(C.11)

where \(x'_{[1t]}\) is again to be understood as the strategy that takes action \(x^*(s,g)\) whenever the most immediate state–shock pair is (sg). But this translates into (C.9). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, J. Analysis of Markovian Competitive Situations Using Nonatomic Games. Dyn Games Appl 11, 184–216 (2021). https://doi.org/10.1007/s13235-020-00356-x

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13235-020-00356-x

Keywords

JEL Classification

Navigation