This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Heterogeneously Perceived Incentives in Dynamic Environments:
Rationalization, Robustness and Unique Selectionsthanks: August 12, 2025. Contact: evan.piermont@rhul.ac.uk (Piermont), p.zuazogarin@hse.ru (Zuazo-Garin)

Evan Piermont Royal Holloway–University of London     Peio Zuazo-Garin HSE University–ICEF
Abstract

In dynamic settings each economic agent’s choices can be revealing of her private information. This elicitation via the rationalization of observable behavior depends each agent’s perception of which payoff-relevant contingencies other agents persistently deem as impossible. We formalize the potential heterogeneity of these perceptions as disagreements at higher-orders about the set of payoff states of a dynamic game. We find that apparently negligible disagreements greatly affect how agents interpret information and assess the optimality of subsequent behavior: When knowledge of the state space is only ‘almost common’, strategic uncertainty may be greater when choices are rationalized than when they are not— forward and backward induction predictions, respectively, and while backward induction predictions are robust to small disagreements about the state space, forward induction predictions are not. We also prove that forward induction predictions always admit unique selections à la Weinstein and Yildiz (2007) (also for spaces not satisfying richness) and backward induction predictions do not.

Keywords: Dynamic Games, Incomplete Information, Rationality, Robustness, Model Misspecification, Forward and Backward Induction, Common Knowledge, Unique Selections
JEL Classification: C72, D82, D83

Introduction

On October 26th, 1597, 13 vessels commanded by Joseon Kingdom admiral Yi Sun-sin stood in front of a Japanese fleet of over 120 warships in the Myeongnyang Strait, on the southwestern coast of the current Republic of Korea. Aware of the disparity of forces, Yi planned the location of the battle carefully and Myeongnyang was chosen due to its narrowness, which would preclude his small fleet to be vastly outnumbered once attacked, and its unique tidal conditions, with currents that strongly flow first to the North and abruptly switch southward in three-hour intervals. With both fleets at Myeongnyang, the favorable current and their formidable numerical superiority led the Japanese to attack. Yi’s army stood in the Northern end of the strait, where the northward flow was already calmer. Together with the narrowness of the strait, this enabled the Joseon fleet to repeal the attack until the direction of the currents switched. The unexpected reversion left the Japanese warships unable to maneuver, drifting backward and colliding among themselves, and the initial violent impulse of the new current in the Northern end helped the Joseon vessels strike their shaky enemy and sink 30 warships, eventually leading to the retreat of the Japanese fleet and the collapse of their campaign.

The tidal configuration of the strait, that is, the state (of, literally, nature), was a key determinant of the outcome of the conflict. Joseon’s commanders were aware of it, Japan’s were not; Joseon’s army based their strategic decisions on their knowledge of the state, Japan’s army based their strategic decisions persistently ignoring the state: They did not conceive the possibility of such an disadvantageous contingency, not even after observing that their enemy, who they knew to be familiar with the terrain and led by a commander of extraordinary skill, decided to fight in so extremely unbalanced conditions. The history relies thus on two elements every economist is familiar with. On the one hand, (nonstrategic) contingencies surrounding the interaction can affect outcomes, making economic agents eager to anticipate them; on the other, agents can update their initial assessments about these contingencies by eliciting others’ private information via the rationalization of observed behavior. For example, the gains from bargaining depend on the valuation parties assign to each agreement, and the initial sequence of offers is often aimed at outguessing the valuation of the other party and hiding one’s own valuation; the optimality of the price a financial trader offers depends on the risk-costs that govern other traders’ supply and demand distributions, and the trader can make inferences about these costs by making sense of others’ trading patterns.111See Kennan (2001) and Sannikov and Skrzypack (2016), respectively. Like at Myeongnyang, the success of rationalization relies on correctly assessing the possible contingencies others have in mind. The possible heterogeneity is incommensurable, but standard economic modeling systematically assumes that, despite different initial probabilistic assessments of each contingency, agents commonly agree on the set of all the possible contingencies, so that any state, even if initially ignored, can be considered upon observed behavior.

This paper introduces a general, tractable theory of persistent disagreements about the possible payoff-relevant, nonstrategic contingencies involved in a dynamic interaction. In it, we show that these disagreements, no matter how apparently negligible, can have a drastic impact on the way information is inferred from observed behavior and, in consequence, can revert our insights about some key features of standard economic predictions. Our findings cast doubts on the methodological validity of equilibrium refinements based on forward induction arguments (i.e., the idea that observed behavior is informative of private information and future plans): Arbitrarily small discrepancies regarding the set of possible contingencies can lead to an erratic interpretation of behavior that results in ‘anything goes’. Predictions based on forward induction are extremely sensitive to modeling details, to the extent that arbitrarily small misspecifications about what agents agree on can lead to unexpected predictions ignored in the benchmark model. This lack of robustness vanishes if, instead, predictions are based on backward induction; that is, if agents are assumed to render observed erratic behavior as a mistake and not as a product of unexpected private information and a signal of future intentions. Finally, we present an contagion argument à la global games that allows for creating unique selections for every forward induction outcome.

More specifically, our first contribution is methodological and consists in formalizing the idea that the set of payoff states of a dynamic game might not be commonly known. This is attained by endowing each player with a subjective payoff structure, which is a hierarchical construction whose first component is the set of states considered by the player, whose second component is the set of states the player understands her opponents to consider, and so on.

Unlike initial belief-hierarchies about the set of states, these subjective payoff structures are assumed to remain invariant in the player’s mind as the game unfolds, and place limits to how the player’s initial belief-hierarchy can be updated in response to observations throughout the game. We refer to a pair consisting of a subjective payoff structure and an initial belief-hierarchy as a subjective model, and to the profiles of subjective models in which players share common knowledge of a set of states, as a standard model.222Thus, in a sense, we reinterpret the set of states as a persistent belief instead of an objective, material object. This way, a model captures the dichotomy between the beliefs that can be updated, the initial ones, and the ones that will never be abandoned. In this respect, our definition of a subjective payoff structure goes hand by hand with the idea of Δ\Delta-restrictions à la Battigalli and Siniscalchi (2003), with the novelty that these restrictions may be not commonly known. In this sense, and in line with the ‘Wilson doctrine’, our approach consists in tearing apart common knowledge assumptions and increasing the level of subjectivity in the formalization of a game:333As recalled by Chung and Ely (2007), Wilson (1987) writes: “Game theory has a great advantage in explicitly analyzing the consequences of trading rules that presumably are really common knowledge; it is deficient to the extent it assumes other features to be common knowledge, such as one agent’s probability assessment about another’s preferences or information. […] I foresee the progress of game theory as depending on successive reduction in the base of common knowledge required to conduct useful analyses of practical problems. Only by repeated weakening of common knowledge assumptions will the theory approximate reality.” While we maintain common knowledge of the extensive form of the game (i.e., the rules: Players, turns, available actions, information feedback, etc.), the payoff structure (i.e, the information regarding preferences) is completely subjective, possibly representing delusion, and envisioned as a collection of element of a universal construction which is trivially commonly known, namely, the set of all possible utility functions. We would like to be emphatic about how natural this separation of rules and stakes is: While the possible outcomes of an economic interaction can be objectively described, there is nothing intrinsically objective in the notion of utility; the allocation of resources that results from each combination of reported preferences can be commonly known, the market shares that correspond to each different combination of pricing strategies can be commonly known, etc. But each individual agent’s preference over material outcomes is a personal trait of hers. In consequence, it seems reasonable to remove the set of states, or possible utilities, from the objective (in the sense of ‘commonly believed’) part of the game.444Of course, the subjectivization of payoff-functions is standard practice since Harsanyi (1967–1968); the formalization in this paper goes beyond that: The state space being subjectively perceived implies that the ways in which beliefs can be updated is subjective as well. By doing so, our approach allows for great heterogeneity regarding players’ interactive perception, to the extent that the standard case of a commonly known set of states turns out to be a knife-edge situation, and important established facts and understandings concluded from this kind of modeling, extremely fragile.

The first of such understandings is the insight that the outcomes of forward induction reasoning, usually formalized by extensive form rationalizability (Pearce, 1984, Battigalli, 1997), refine those of backward induction. This is known to be the case for games with complete information, both for the case of perfect information (Battigalli, 1997, Heifetz and Perea, 2015 and Perea, 2018a) and imperfect information (Chen and Micali, 2013, Perea, 2018c and Catonini, 2020). Whether an analogous result holds for the case of incomplete information remains an open question (backward induction being formalized here as an interim version of Penta’s (2011; 2015) backward rationalizability); however, our counterexamples establish that the answer is negative when the set of payoff-states of the dynamic game is not commonly known, no matter how small players’ discrepancies are.555While we did not attempt to prove this generalization, we conjecture that the refinement result should hold for dynamic games with commonly known sets of payoff-states. In our counterexamples for the case of not commonly known state spaces, it is precisely the lack of common knowledge what seems the key driving force. The key idea that triggers the observation is that when the state space is not commonly known, it can sometimes be possible for a player to choose rationally an action the other player cannot rationalize (the action may be undominated given the state space the first player considers, but not in the state space the first player thinks the second player considers). Thus, the second player may entertain any arbitrary conjecture about the future behavior of the first one, and this is something the first player can rationally exploit. The observation has some intrinsic interest. While the literature on strategic communication teaches that agents can have incentives to act is if they had less information than they do,666Dye (1985), for instance. in standard game-theoretic models it cannot be that a player rationally pretends not to be rational: If she did so, her opponents would be able to find ways to rationalize her behavior. Subjective payoff structures circumvent this impossibility in a natural manner.777The strategy of playing dumb is a commonplace in popular culture; as Suetonius (121) writes about Emperor Claudius: “He did not even keep quiet about his own stupidity, but in certain brief speeches he declared that he had purposely feigned it under [Caligula], because otherwise he could not have escaped alive and attained his present station. But he convinced no one, and within a short time a book was published, the title of which was ‘The Elevation of Fools and its thesis, that no one feigned folly.” The italics are ours.

The second understanding that collapses is the robustness of rationalizability to small misspecifications of players’ perception of payoff structures. Dekel, Fudenberg and Morris (2007) show that in static settings the predictions of interim (correlated) rationalizability are robust to small misspecifications of players’ higher-order beliefs about the payoff-states: Predictions excluded in the benchmark model do not to arise in slightly misspecified models (i.e., the solution concept is upper hemicontinuous on players’ belief-hierarchies). We partially extend this insight by proving in Proposition 2 that the correspondence that describes extensive form rationalizability is upper hemicontinuous on initial belief-hierarchies.888While we take for granted that this result is ‘folk knowledge’, to the best of our knowledge this is the first available proof of the result. However, the conclusion that extensive form rationalizability is robust in a general sense would be erroneous. Our examples show that the solution concept is not robust to misspecifications of models. It is possible to perturb a dynamic game with a commonly known set of payoff-states and unique extensive form rationalizable outcome so that, for every arbitrarily close by model, extensive form rationalizable outcomes are multiple. These observations shed important light on two aspects of the role of the state space in rationalization:

  • First, Penta (2012) and Chen (2012) show that in dynamic games with sufficiently rich set of payoff-states (i.e., state spaces in which for every strategy of every player there exists some payoff-state for which the strategy is strictly dominant), no refinement of interim rationalizability is robust to perturbations of initial beliefs.999Penta’s (2012) result is slightly stronger. By allowing for payoff-states about which players may have persistent private information, his result pertains not plain vanilla interim rationalizability, but interim sequential rationalizability, a solution concept that refines interim rationalizability by employing sequential rationality as a notion of optimal behavior, instead of plain rationality. For expositional purposes, throughout the paper we refer solely to ‘interim rationalizability’ and bundle Penta’s (2012) and Chen (2012) work together: The moral of both papers is that no refinement of the predictions given by solution concepts characterizing the behavioral implications of rationality and (only) initial common belief in rationality is robust. Our Proposition 2 shows that extensive form rationalizability is robust in this sense. Since extensive form rationalizability refines interim rationalizability, these results together imply that in games where the state space is rich, interim rationalizability and extensive form rationalizability coincide (Corollary 1). The intuition is obvious. If for every strategy there is some state that makes that strategy dominant then every observed behavior can be rationalized; thus, there is nothing rationalization helps to falsify. In consequence, while richness assumptions are innocuous in static games (specifying an interim belief makes every state ignored by the belief strategically irrelevant), our results highlight their problematic nature in dynamic environments.

  • Second, the failure of upper hemicontinuity of extensive form rationalizability on subjective payoff structures reveals the high sensitivity that rationalization processes display w.r.t common knowledge assumptions. On the one hand, as soon as the knife-edge supposition that players commonly agree about the state space is abandoned, the rationalization a player makes might be erroneous, or a player may fail to rationalize some observed behavior which was indeed rational. In consequence, despite the conceptual appeal of forward induction, the conditions for it to be applied successfully seem to crucially rely on knife-edge assumptions that are tremendously demanding, and hard to meet in practice.

We end our analysis of robustness to misspecifications of models by noting that there is a standard solution concept which is robust. Proposition 3 shows that the individual strategies given by backward induction, formalized here as an interim version of Penta’s (2011; 2015) backward rationalizability, are upper hemicontinuous in the whole space of subjective models. The key difference with extensive form rationalizability lies in the fact that, in the latter, the restrictions on beliefs are placed only on the histories that are reachable given the subjective payoff structure. Since this set of histories is not, in general, lower hemicontinuous on the subjective payoff structure, upper hemicontinuity of extensive form rationalizability collapses (that restrictions are satisfied along a sequence cannot guarantee that they are also satisfied in the limit). On the contrary, backward induction reasoning places restrictions on beliefs at every history, and thus, in a way, independent of the subjective payoff structure. We end by noting that, since backward rationalizability also refines interim rationalizability, in dynamic games satisfying richness backward rationalizability, extensive form rationalizability and interim rationalizability all coincide (Corollary 2), what further reinforces our critique of richness assumptions in dynamic games.

Finally, we present a partial structure theorem for extensive form rationalizability in the line of the seminal result by Weinstein and Yildiz (2007). Theorem 1 shows that for every standard model, every outcome consistent with extensive form rationalizability is the unique one also consistent with extensive form rationalizability for a sequence of profiles of perturbed subjective models. Notably, the theorem does not require the state space of the benchmark standard model to satisfy richness: Even if it does not, it can be approximated by profiles of subjective payoff structures that do satisfy richness at increasingly higher-orders (a player’s space may not be rich, but she may think that her opponents’ are, or may think that her opponents think that their opponents’ are, etc.; indeed, we show in Proposition 1 that this notion of higher-order richness is generic in the space of all subjective payoff structures). Thus, our structure theorem applies to virtually all finite dynamic games with incomplete information, unlike those by Penta (2012) and Chen (2012), that are restricted to games whose state space satisfies richness (and in which, as seen above, rationalization has no bite). However, while its proof is technically challenging, the conceptual consequences of Theorem 1 remain unclear. On the one hand, unlike Weinstein and Yildiz (2007), Penta (2012) and Chen (2012) we cannot claim that the predictions of the solution concept we study are generically unique: Our counterexamples document that a standard model can be approximated by a sequence of profiles of perturbed subjective models in which multiplicity of extensive form rationalizable outcomes is robust. On the other hand, it is not clear which relevant solution concept, if any, is a refinement of extensive form rationalizability, so that it is not clear how the theorem could be applied to deny the robustness of such hypothetical solution concept.

Nonetheless, the result suggests two possible lines of research. First, it leaves the characterization of the sharpest predictions robust to misspecifications of models as an open question. While the predictions of backwards rationalizability are robust, we document via a counterexample that they do not always admit unique selections. We conjecture then that, if a robust solution concept admitting unique selections exists, it must consist in some hybrid reasoning procedure combining elements of both backward and extensive-form rationalizability. Second, the proof of the theorem relies on a contagion argument that elaborates the classical one in global games and involves an explicit interplay between initial assessments, persistent information and the interpretation of observed behavior. It remains then to be explored whether such an interplay, slightly more complex than the standard one in dynamic global games, provides a rationale for the arousal of endogenous coordination.


The rest of the paper is structured as follows. Section 2 formalizes the basic elements for the analysis of dynamic games with incomplete information. Section 3 presents a tool to formalize players’ disagreements of arbitrary high-order about the set of payoff-states of the game, and extends the definition of extensive form rationalizability and backward rationalizability to account for this feature. Section 4 presents the main findings of the paper. Section 5 ends with some literature review and new directions for research. All the proofs are relegated to the appendix.

Preliminaries

The formalization of dynamic games in Section 2.1 is standard (see Penta, 2012), except for a small addition in Section 2.1.3 concerning the representation of the utility functions and the information thereof players entertain. Section 2.2 reviews the formalization of initial types (or belief-hierarchies) and is completely standard (see Harsanyi, 1967–1968, or Mertens and Zamir, 1985).

Dynamic games

A dynamic game consists in a list Γ,Υ\langle\Gamma,\Upsilon\rangle where (i)(i) Γ\Gamma, an extensive form, specifies the set of players, admissible sequences of choices, turns and past choices observed by each player at each turn, and (ii)(ii) Υ\Upsilon, a payoff structure, formalizes players’ possible preferences over the outcomes of the game and the information players may have about each others’ preferences.

Extensive forms

An extensive form consists in a list Γ=I,(Ai)iI,H,Z\Gamma=\left\langle I,(A_{i})_{i\in I},H,Z\right\rangle where II is s finite set of players, and:

  • For each player ii, AiA_{i} is a finite set of actions. A history represents the unfolding of the game and consists in a finite sequence of possibly simultaneous choices, i.e., on a finite sequence of elements from {h0}A\{h^{0}\}\cup A, where A:=JIAJA:=\bigcup_{J\subseteq I}A_{J} and AJ:=iJAiA_{J}:=\prod_{i\in J}A_{i} for any JIJ\subseteq I. We say that history hh^{\prime} follows history hh, denoted by hhh\preceq h^{\prime}, if hh^{\prime} obtains from adding finitely many possibly simultaneous choices to hh.101010That is, when there exists some (an)nNA(a^{n})_{n\leq N}\subseteq A such that h=(h;(an)nN)h^{\prime}=(h;(a^{n})_{n\leq N}).

  • HH and ZZ are finite and disjoint sets of histories such that (HZ,)(H\cup Z,\preceq) is a rooted and oriented tree with terminal nodes ZZ. Symbol h0h^{0} denotes the ex ante stage of the game, i.e., the root of the tree, and histories in HH and ZZ are referred to as partial and terminal, respectively. For any player ii and partial history hh, let Ai(h)A_{i}\left(h\right) denote the set of actions available to ii at hh. Player ii is active at hh if Ai(h)A_{i}(h) is nonempty; let HiH_{i} denote the set of these histories. We assume that: (i)(i) a player is never the only active one twice in a row, and (ii)(ii) whenever a player is active, at least two actions are available to her.

In this context, the set of player ii’s strategies is Si:=hHiAi(h)S_{i}:=\prod_{h\in H_{i}}A_{i}(h) and, as usual, the set of strategy profiles is denoted by S:=iISiS:=\prod_{i\in I}S_{i} and the set of player ii’s opponents strategies, by Si:=jiSjS_{-i}:=\prod_{j\neq i}S_{j}. Obviously, for each partial history hh each strategy ss induces a unique terminal history, z(s|h)z(s|h). Finally, let Si(h)S_{i}(h) and Si(h)S_{-i}(h) denote, respectively, the set of player ii’s strategies and the set of ii’s opponents’ strategies that reach partial history hh, and Hi(si)H_{i}(s_{i}), the set of player ii’s histories that can be reached when she chooses strategy sis_{i}.111111To be precise, Si(h)={siSi|hz(si;si|h0) for some siSi}S_{i}(h)=\{s_{i}\in S_{i}|h\preceq z(s_{-i};s_{i}|h^{0})\textup{ for some }s_{-i}\in S_{-i}\} and Si(h)=jiSj(h)S_{-i}(h)=\prod_{j\neq i}S_{j}(h) on the one hand, and Hi(si):={hHi|siSi(h)}H_{i}(s_{i}):=\{h\in H_{i}|s_{i}\in S_{i}(h)\} on the other.

Payoff structures

A standard payoff structure (for extensive form Γ\Gamma) consists of a list Υ:=(Θ0,(Θi,ui)iI)\Upsilon:=(\Theta_{0},(\Theta_{i},u_{i})_{i\in I}), where:

  • Θ0\Theta_{0} is a compact and metrizable set of states of nature and, for each player ii, Θi\Theta_{i} is a compact and metrizable set of payoff types. We denote the set of payoff states by Θ:=Θ0×iIΘi\Theta:=\Theta_{0}\times\prod_{i\in I}\Theta_{i} and, for each player ii, the set of profiles of player ii’s opponents’ payoff types by Θi:=jiΘj\Theta_{-i}:=\prod_{j\neq i}\Theta_{j}.

  • For each player ii, ui:Z×Θu_{i}:Z\times\Theta\rightarrow\mathds{R} is player ii’s continuous utility function.121212Throughout the paper finite sets are endowed with the discrete topology, \mathds{R} is endowed with the Euclidean topology, product sets are endowed with the product topology and spaces consisting of closed sets, with the Hausdorff metric. For a given topological space XX, its corresponding set of measures over the Borel σ\sigma-algebra is endowed with the weak\ast topology. Unless stated otherwise, measurability always refers to Borel measurability.

The set of all standard payoff structures (for extensive form Γ\Gamma) is denoted by 𝒫\mathscr{P}.131313A more general formalization of standard payoff structures allows for the set of payoff states not being a Cartesian product. We note that the whole theory developed in the paper extends to that case by introducing the necessary obvious changes. Thus, within a dynamic game with incomplete information Γ,Υ\langle\Gamma,\Upsilon\rangle, each payoff state θΘ\theta\in\Theta fully characterizes the profile of players’ utility functions: (ui(,θ))iI(u_{i}(\,\cdot\,,\theta))_{i\in I}. Besides, each player ii is assumed to know her payoff type θi\theta_{i} and (possibly) face uncertainty about the rest of components of θ\theta. Hence, besides the description of preferences, each payoff state θ\theta also determines the information players have about the payoff states: In every moment of the game, a player ii with payoff type θi\theta_{i} knows that the true payoff state is some element in {θi}×Θ0×Θi\{\theta_{i}\}\times\Theta_{0}\times\Theta_{-i}; in consequence, {θi}×Θ0×Θi\{\theta_{i}\}\times\Theta_{0}\times\Theta_{-i} is, precisely, the set of payoff states that player ii can eventually deem as possible as the game unfolds.

Canonical representation and convergence of payoff structures

The questions we explore throughout the paper are related to the strategic impact of small relaxation of players’ common knowledge of an information structure Υ\Upsilon. We develop this point in detail in Section 3.1, but before introducing it we need to first clarify what is understood as a small, or perturbation, in a standard payoff structure. The fact that the sets of payoff states can vary vastly across different standard payoff structures and be of radically different nature presents some inconvenience for an analyst interested in assessing how close two different standard payoff structures are form each other. To circumvent this issue we introduce a canonical representation that allows for envisioning every standard payoff structure as part of an object directly derived from the extensive form:

Definition 1 (Canonical representation of payoff structures).

Let Γ\Gamma be an extensive form and Υ=(Θ0,(Θi,ui)iI)\Upsilon=(\Theta_{0},(\Theta_{i},u_{i})_{i\in I}), a payoff structure. Then, the canonical representation of Υ\Upsilon is the list 𝒞(Υ):=𝒞0(Υ)×iI𝒞i(Υ)\mathcal{C}(\Upsilon):=\mathcal{C}_{0}(\Upsilon)\times\prod_{i\in I}\mathcal{C}_{i}(\Upsilon), where:

  1. (i)(i)

    𝒞0(Υ)\mathcal{C}_{0}(\Upsilon) is the set of possible profiles of utility functions:

    𝒞0(Υ):={(ui(,θ))iI|θΘ}.\mathcal{C}_{0}(\Upsilon):=\{(u_{i}(\,\cdot\,,\theta))_{i\in I}|\,\theta\in\Theta\}.
  2. (ii)(ii)

    For each player ii, 𝒞i(Υ)\mathcal{C}_{i}(\Upsilon) is the set of profiles of utility functions that player ii can consider possible as the game unfolds:

    𝒞i(Υ):={{(ui(,(θ0,θi,θi)))iI|(θ0,θi)Θ0×Θi}|θiΘi}.\mathcal{C}_{i}(\Upsilon):=\left\{\left\{(u_{i}(\,\cdot\,,(\theta_{0},\theta_{i},\theta_{-i})))_{i\in I}\left|(\theta_{0},\theta_{-i})\in\Theta_{0}\times\Theta_{-i}\right.\right\}\left|\,\theta_{i}\in\Theta_{i}\right.\right\}.

The interpretation of each element υ=(υ0,(υi)iI)𝒞(Υ)\upsilon=(\upsilon_{0},(\upsilon_{i})_{i\in I})\in\mathcal{C}(\Upsilon) is similar to that of each payoff state θ\theta: A description of each player ii’s preferences and information. First, υ0\upsilon_{0} consists in a profile specifying a utility function (υ0)i:Z(\upsilon_{0})_{i}:Z\rightarrow\mathds{R}. Second, υi\upsilon_{i} can be interpreted as the set of profiles that can be conceived by a player ii already endowed with a payoff type as the game unfolds—anything outside of υi\upsilon_{i} will never be part of ii’s beliefs (even if necessary for rationalizing observed behavior).

Now, the main reason for introducing Definition 1 is that it allows for envisioning payoff structures as informational constraints over a canonical object directly derived from the extensive form, namely, the space of profiles of all conceivable utility functions, iIZ\prod_{i\in I}\mathds{R}^{Z}: Simply notice that, given Υ\Upsilon, 𝒞0(Υ)\mathcal{C}_{0}(\Upsilon) is a compact subset of iIZ\prod_{i\in I}\mathds{R}^{Z} and that for each iIi\in I, each υi𝒞i(Υ)\upsilon_{i}\in\mathcal{C}_{i}(\Upsilon) is a compact subset of 𝒞0(Υ)\mathcal{C}_{0}(\Upsilon)—so that 𝒞i(Υ)\mathcal{C}_{i}(\Upsilon) is a compact subset of compact subsets of 𝒞0(Υ)\mathcal{C}_{0}(\Upsilon). The notion of convergence for canonical representations of standard payoff is very natural: Say that (𝒞(Υn))n(\mathcal{C}(\Upsilon^{n}))_{n\in\mathds{N}} converges to 𝒞(Υ)\mathcal{C}(\Upsilon) if:

  • 1.

    Sequence (𝒞0(Υn))n(\mathcal{C}_{0}(\Upsilon^{n}))_{n\in\mathds{N}} converges to 𝒞0(Υ)\mathcal{C}_{0}(\Upsilon) in the Hausdorff metric for compact subsets of iIZ\prod_{i\in I}\mathds{R}^{Z}.

  • 2.

    For each player ii, sequence (𝒞i(Υn))n(\mathcal{C}_{i}(\Upsilon^{n}))_{n\in\mathds{N}} converges to 𝒞i(Υ)\mathcal{C}_{i}(\Upsilon) in the Hausdorff metric for compact subsets of compact subsets of iIZ\prod_{i\in I}\mathds{R}^{Z}

The use of canonical representations greatly simplifies (at a conceptual level) our analysis of robustness in Section 4; perturbing payoff structures will consist on simply perturbing profiles of sets instead of perturbing both sets and utility functions whose domain includes these perturbed sets:

Definition 2 (Convergence of payoff structures).

Let Γ\Gamma be an extensive form. Then, we consider that a sequence of standard payoff structures (Υn)n(\Upsilon^{n})_{n\in\mathds{N}} converges to a standard payoff structure Υ\Upsilon if (𝒞(Υn))n(\mathcal{C}(\Upsilon^{n}))_{n\in\mathds{N}} converges to 𝒞(Υ)\mathcal{C}(\Upsilon).

Initial beliefs

At the beginning of dynamic game Γ,Υ\langle\Gamma,\Upsilon\rangle each player ii is endowed with a payoff type and an initial belief hierarchy about the state of nature and their opponents’ payoff types, Θ0×Θi\Theta_{0}\times\Theta_{-i}. The payoff type is assumed to remain invariant as the game progresses, but the initial belief hierarchy may be updated in response to observed behavior. More formally, for each player ii we have a type ti=(θi,πi)t_{i}=(\theta_{i},\pi_{i}), where θiΘi\theta_{i}\in\Theta_{i} is a payoff type and πi=(πi,1,πi,2,,πi,k,)\pi_{i}=(\pi_{i,1},\pi_{i,2},\dots,\pi_{i,k},\dots) is a belief hierarchy à la Brandenburger and Dekel (1993), where πi,1\pi_{i,1} represents player ii’s first-order belief about Θ0×Θi\Theta_{0}\times\Theta_{-i}, πi,2\pi_{i,2} represents player isi^{\prime}s second-order belief about both Θ0×Θi\Theta_{0}\times\Theta_{-i} and ii’s opponents first-order beliefs, and so on. We assume that the belief hierarchy is coherent, that is, that higher order beliefs marginalize to lower order beliefs and that coherency is common belief.141414The details of the formalization are standard in the literature but, to be more precise, a belief hierarchy for player ii consists in a sequence (πi,k)k(\pi_{i,k})_{k\in\mathds{N}} where: πi,1\pi_{i,1} is an element of Xi,1:=Δ(Θ0×Θi)X_{i,1}:=\Delta(\Theta_{0}\times\Theta_{-i}) and for each kk\in\mathds{N} subsequence (πi,1,,πi,k,πi,k+1)(\pi_{i,1},\dots,\pi_{i,k},\pi_{i,k+1}) is an element of Xi,k+1:={(πi,1,,πi,k,πi,k+1)Xi,k×Δ(Θ0×Θi×jiXj,k)|margΘ0×Θi×jiXj,kπi,k+1=πi,k}X_{i,k+1}:=\{(\pi_{i,1}^{\prime},\dots,\pi_{i,k}^{\prime},\pi_{i,k+1}^{\prime})\in X_{i,k}\times\Delta(\Theta_{0}\times\Theta_{-i}\times\prod_{j\neq i}X_{j,k})\,|\,\textup{marg}_{\Theta_{0}\times\Theta_{-i}\times\prod_{j\neq i}X_{j,k}}\pi_{i,k+1}^{\prime}=\pi_{i,k}^{\prime}\}.

Let Πi(Θ)\Pi_{i}(\Theta) denote the set of all such belief hierarchies of player ii and Ti(Θ)T_{i}(\Theta), the set of all types. We know from Brandenburger and Dekel (1993) that there exists a homeomorphism τi:Πi(Θ)Δ(Θ0×Ti(Θ))\tau_{i}:\Pi_{i}(\Theta)\rightarrow\Delta\left(\Theta_{0}\times T_{-i}(\Theta)\right), where Ti(Θ)=jiTj(Θ)T_{-i}(\Theta)=\prod_{j\neq i}T_{j}(\Theta). Throughout the paper, for any player ii and type ti=(θi,πi)t_{i}=(\theta_{i},\pi_{i}) we write θi(ti)\theta_{i}(t_{i}) to represent type tit_{i}’s payoff type and πi(ti)=(πi,1(ti),πi,2(ti),,πi,k(ti),)\pi_{i}(t_{i})=(\pi_{i,1}(t_{i}),\pi_{i,2}(t_{i}),\dots,\pi_{i,k}(t_{i}),\dots) to represent type tit_{i}’s belief hierarchy.

Heterogeneously perceived incentives

In these section we introduce the main methodological novelty of the paper. We develop a formalism that allows for players to entertain arbitrary, different perceptions, also at the higher-orders, regarding the payoff structure of the game: Player ii might be deluded about which the payoff structure player jj considers it to be, or may think that player jj is deluded about which the payoff structure player kk considers it to be, and so on. We formalize the objects that allow for this heterogeneous subjective approach to the game in Section 3.1 and, in Section 3.2, we extend the definitions of extensive form rationalizability and backward rationalizability to the present set-up.

Subjective payoff structures and subjective models

Definitions

Next, we introduce the main new formal element of the paper. A subjective payoff structure is a hierarchical construction that represents how a player perceives the payoff structure of a game, how she perceives her opponents to perceive the payoff structure of the game, and so on. Thus, technically, a subjective payoff structure consists in a list in which the first component is a standard payoff structure (the one the player considers to be the right one), the second component is a map that assigns a subjective payoff structure to each opponent (the one the player considers each opponent to consider the right one), and so on. One consistency condition is required (see part (iii)(iii) in the definition below): If player ii considers that Θj\Theta_{j} are the possible payoff types of player jj, then she must consider that jj considers that Θj\Theta_{j} is (at least in terms of canonical representations) the set of her own possible payoff types—i.e., there is common knowledge in the event that players know their own set of payoff types.

Definition 3 (Subjective payoff structure).

Let Γ\Gamma be an extensive form. A subjective payoff structure for player ii is a sequence di=(di,k)kd_{i}=(d_{i,k})_{k\in\mathds{N}} such that di,1𝒫i1:=𝒫d_{i,1}\in\mathscr{P}_{i}^{1}:=\mathscr{P} and, for every k2k\geq 2, we have (di,1,,di,k)𝒫ik(d_{i,1},\dots,d_{i,k})\in\mathscr{P}_{i}^{k}, where:

𝒫ik\displaystyle\mathscr{P}_{i}^{k} :={(di,1,,di,k)|(i)(di,1,,di,k1)𝒫ik1(ii)di,k:I{i}ji𝒫jk1, with di,k(j)𝒫jk for every ji,(iii)𝒞j(di,1)=𝒞j(di,2(j)) for every ji}.\displaystyle:=\left\{(d_{i,1},\dots,d_{i,k})\left|\begin{tabular}[]{r l}$(i)$&$(d_{i,1},\dots,d_{i,k-1})\in\mathscr{P}_{i}^{k-1}$\\[12.91663pt] $(ii)$&$d_{i,k}:I\setminus\{i\}\rightarrow\displaystyle\bigcup_{j\neq i}\mathscr{P}_{j}^{k-1},\textup{ with }d_{i,k}(j)\in\mathscr{P}_{j}^{k}\textup{ for every }j\neq i$,\\[12.91663pt] $(iii)$&$\mathcal{C}_{j}(d_{i,1})=\mathcal{C}_{j}(d_{i,2}(j))\textup{ for every }j\neq i$\end{tabular}\right.\right\}.

Let 𝒫i\mathscr{P}_{i}^{\infty} denote the set of all subjective payoff structures for player ii.

Throughout the paper, with some abuse of notation, we will indistinctly use Υ\Upsilon to refer to a standard payoff structure or to the subjective payoff structure that represents common knowledge of Υ\Upsilon; accordingly, we will treat 𝒫\mathscr{P} as a subset of 𝒫i\mathscr{P}_{i}^{\infty}.151515That is, for player ii, diΥd_{i}^{\Upsilon} such that di,1Υ=Θd_{i,1}^{\Upsilon}=\Theta, di,2Υ(j)=dj,1Υd_{i,2}^{\Upsilon}(j)=d_{j,1}^{\Upsilon}, di,3Υ(j)=dj,2Υd_{i,3}^{\Upsilon}(j)=d_{j,2}^{\Upsilon}, and so on. For a given subjective payoff structure did_{i} we denote:

  • By dj|i:=(di,k(j))k2d_{j|i}:=(d_{i,k}(j))_{k\geq 2} and di|i:=(dj|i)jid_{-i|i}:=(d_{j|i})_{j\neq i}, the subjective payoff structure ascribed by did_{i} to player jij\neq i, and the list of subjective payoff structures ascribed by did_{i} to every player jij\neq i, respectively.

  • By (di,1)0(d_{i,1})_{0}, the set of states of nature that corresponds to the set of payoff states in standard payoff structure di,1d_{i,1}, and by (di,1)j(d_{i,1})_{j}, for each jIj\in I, the set of player jj’s payoff types in that same set.

Since subjective payoff structures put restrictions on which payoff states can be considered by the player, obviously, not every type tit_{i} is necessarily consistent with every subjective payoff structure did_{i}; in order for the pair (di,ti)(d_{i},t_{i}) be consistent, the payoff type of tit_{i} must be in (di,1)i(d_{i,1})_{i}, the support of its first-order belief must be included in the projection on (di,1)0×(di,1)i(d_{i,1})_{0}\times(d_{i,1})_{-i} of di,1d_{i,1}, and so on.161616Formally, we require that (1)(1) (di,1)i(d_{i,1})_{i} contains θi(ti)\theta_{i}(t_{i}); (2)(2) that Proj(di,1)0×(di,1)i(di,1)\textup{Proj}_{(d_{i,1})_{0}\times(d_{i,1})_{-i}}(d_{i,1}) contains the support of πi,1(ti)\pi_{i,1}(t_{i}); (3)(3) that for every tit_{-i} in the support of τi(πi(ti))\tau_{i}(\pi_{i}(t_{i})) and every jij\neq i, (1)(1) and (2)(2) hold w.r.t. dj|id_{j|i}; and so on. For each profile of subjective payoff structures dd, we let Ti(di)T_{i}(d_{i}) and Ti(di):=jiTj(dj)T_{-i}(d_{-i}):=\prod_{j\neq i}T_{j}(d_{j}) denote, respectively, the set of player ii’s types that are consistent with did_{i}, and the set of opponents types that are consistent with did_{-i}. Finally, let T(d):=Ti(di)×Ti(di)T(d):=T_{i}(d_{i})\times T_{-i}(d_{-i}). Consistent combinations of subjective payoff structures and types play a central role in our analysis, since they fully specify players’ perception of the incentives of the game:

Definition 4 (Subjective models, Standard models).

Let Γ\Gamma be an extensive form. Then:

  • (i)(i)

    A subjective model for player ii is a pair (di,ti)(d_{i},t_{i}) where did_{i} is a subjective payoff structure for player ii and tit_{i} is a type consistent with did_{i}. Let i\mathscr{M}_{i}^{\infty} denote the set of subjective models for player ii.

  • (ii)(ii)

    We say that (di,ti)(d_{i},t_{i}) is standard if did_{i} represents common knowledge of a standard payoff structure. Let i1\mathscr{M}_{i}^{1} denote the set of standard subjective models for player ii.

  • (iii)(iii)

    A standard model is a profile of standard subjective models (di,ti)iI(d_{i},t_{i})_{i\in I} where the standard payoff structure Θ\Theta associated to each did_{i}, di,1d_{i,1}, is the same for every player ii. Let 1\mathscr{M}^{1} denote the set of standard models.

Thus, a subjective model fully specifies the player’s perception of the payoff structure of the game; the subjective payoff structure describes the information about preferences the player will restrict her attention to, and the type, her initial probabilistic assessment about the payoff states. If the player perceives the set of states as commonly known, her perception is represented by a standard subjective model and, if every player agree on their perception of the set of payoff states and this fact is commonly known, players’ perceptions are represented by a standard model. We present now an example aimed at illustrating these notions.

Example

Consider the following extensive form:

112211A1A_{1}D1D_{1}aaddA2A_{2}D2D_{2}

The mechanics of the game are simple: There are two players (11 and 22) with two available actions each (to advance and to go down); if, at her turn, a player chooses to go down the game ends, if, instead, she chooses to advance it is then the other player’s turn. Actions are observed, player 11 chooses first and has at most two turns, and player 22 has at most only one turn. Now, by adding a standard payoff structure Θ(n,+)\Theta^{(n,+)}, we can have a standard dynamic game as the following one, in which for a given nn, utility functions are commonly known:

112211
2+1n2+\frac{1}{n}
0
0
0
2
-1
1
1
A1A_{1}D1D_{1}aaddA2A_{2}D2D_{2}

Now, if we want to introduce the possibility of discrepancies among players, we can assume that player 22’s subjective payoff structure is d2nd_{2}^{n}, representing common knowledge of Θ(n,+)\Theta^{(n,+)},171717I.e., such that d2,1n=Θ(n,+)d_{2,1}^{n}=\Theta^{(n,+)}, d2,2n(1)=Θ(n,+)d_{2,2}^{n}(1)=\Theta^{(n,+)}, d2,3n(1)(2)=Θ(n,+)d_{2,3}^{n}(1)(2)=\Theta^{(n,+)}, etc. and that player 11’s subjective structure is given by d1nd_{1}^{n}, with a first component representing the following payoff structure, Θ(n,)\Theta^{(n,-)}:

112211
21n2-\frac{1}{n}
0
0
0
2
-1
1
1
A1A_{1}D1D_{1}aaddA2A_{2}D2D_{2}

and such that d1nd_{1}^{n} considers that player 22 considers Θ(n,+)\Theta^{(n,+)} to be commonly known.181818I.e., such that d1,1n=Θ(n,)d_{1,1}^{n}=\Theta^{(n,-)} and d2|1n=d2nd_{2|1}^{n}=d_{2}^{n}. Thus, (d1,d2)(d_{1},d_{2}) represent a situation in which player 22 considers that Θ(n,+)\Theta^{(n,+)} is commonly known and player 11 knows this, but considers the payoff structure to be Θ(n,)\Theta^{(n,-)}. For each subjective payoff structure, there is a unique type for the player consistent with it: For player 22, the one that represents common belief in the payoffs given by the unique state in Θ(n,+)\Theta^{(n,+)}, t2nt_{2}^{n}, and for player 11, the one that represents probability 1 belief in the payoffs given by the unique state in Θ(n,)\Theta^{(n,-)} and in type t2nt_{2}^{n}, t1nt_{1}^{n}. Thus, M1n:=(d1n,t1n)M_{1}^{n}:=(d_{1}^{n},t_{1}^{n}) and M2n:=(d2n,t2n)M_{2}^{n}:=(d_{2}^{n},t_{2}^{n}) are the unique possible subjective models; the first one is not standard, the second one is. Thus, (M1n,M2n)(M_{1}^{n},M_{2}^{n}) is not a standard model.

Richness and higher-order richness

Richness assumptions play a central role in the literature of structure theorems and unique selections (see Weinstein and Yildiz, 2007, Penta, 2012 or Chen, 2012). The reason is that states in which a strategy is dominant are employed as ‘seeds’ to initiate a contagion argument that propagates through the hierarchy of beliefs: A player might not believe in one of those states, but if she believes that her opponent believes in one of then, then the player should choose a best response to the strategy her opponent considers to be dominant. Formally, richness is defined as follows:

Definition 5 (Richness, c.f. Penta, 2012).

Let Γ\Gamma be an extensive form and Υ\Upsilon, a standard payoff structure. We say that Υ\Upsilon is rich if for every payoff type θi\theta_{i} and every strategy sis_{i} there exists some payoff type θi(θi,si)Θi\theta_{i}(\theta_{i},s_{i})\in\Theta_{i} such that sis_{i} is conditionally dominant for player ii at every payoff state θ\theta^{\prime} with θi=θi(θi,si)\theta_{i}^{\prime}=\theta_{i}(\theta_{i},s_{i}) and such that (θ0,θi,θi)(\theta_{0},\theta_{i},\theta_{-i}) and (θ0,θi(θi,si),θi)(\theta_{0},\theta_{i}(\theta_{i},s_{i}),\theta_{-i}) are payoff equivalent for every jij\neq i for every pair (θ0,θi)(\theta_{0},\theta_{-i}).

Notice that in the case of standard payoff structures, the notion is rather demanding. It requires that it is common knowledge among players that no strategy is known not to be conditionally dominant. Or, in other words, that it is commonly known among players that every strategy can be found to be conditionally dominant for some appropriate belief about the payoff states. On the contrary, subjective payoff structures allow for greatly relaxing this common knowledge assumption while keeping its bite. Players may know that richness does not hold, but they may consider that their opponents consider it to hold; or they may consider that it does not hold, that their opponents consider it not to hold, but consider that their opponents consider their opponents to consider that it holds. Thus, it is possible to approximate common knowledge of absence of richness while still having the chance to exploit an infection argument: I could perceive the game as not being rich, but perceive that my opponents perceive it to be rich, or only perceive that they perceive that their opponents perceive that it is rich, and so on. In order to formalize the idea that richness holds at high orders but not low ones let us introduce the following definition:

Definition 6 (Higher-order richness).

Let Γ\Gamma be an extensive form and did_{i}, a subjective payoff structure for player ii. Then we say that did_{i} satisfies:

  • (i)(i)

    1st1^{\textup{st}}-order richness if di,1d_{i,1} is rich.

  • (ii)(ii)

    kthk^{\textup{th}}-order richness, for k2k\geq 2, if dj|id_{j|i} satisfies (k1)th(k-1)^{\textup{th}}-richness for every jij\neq i.

  • (iii)(iii)

    Higher-order richness if did_{i} satisfies kthk^{\textit{th}}-order richness for some k1k\geq 1.

Notably, it turns out that higher-order richness is a generic property within the space of subjective payoff structures; thus, while richness is a very demanding notion for games with a commonly known state space, higher-order richness is extremely mild for games with subjective payoff structures:

Proposition 1 (Genericity of higher-order richness).

Let Γ\Gamma be an extensive form. Then, the set of subjective payoff structures that satisfy higher-order richness is open and dense in 𝒫i\mathscr{P}_{i}^{\infty}.

Discussion: Relaxation of all common knowledge assumptions

We would like to be emphatic about how natural this separation of rules and stakes is: While the possible outcomes of an economic interaction can be objectively described, there is nothing intrinsically objective in the notion of utility; the allocation of resources that results from each combination of reported preferences can be commonly known, the market shares that correspond to each different combination of pricing strategies can be commonly known, etc. But each individual agent’s preference over material outcomes is a personal trait of hers. In consequence, it seems reasonable to remove the set of states, or possible utilities, from the objective (in the sense commonly known) part of the game

Subjective payoff structures allow for relaxing all common knowledge assumptions about payoffs in a simple, tractable way; the only common knowledge assumptions they entail is the trivial one: The set of utility functions players reason about consists in functions that map outcomes to utilities. Furthermore, not only do subjective payoff structures allow for approximating models not satisfying any richness assumption with models that satisfy richness at high orders, but they also allow for approximating models satisfying certain knowledge assumptions (say, the no information case, as in Weinstein and Yildiz, 2007, or Chen, 2012) with models that satisfy different knowledge assumptions at high orders (say, models with persistent private information, as in Penta, 2012). In this sense, subjective payoff structures allow for enhancing the subjective nature of individual strategic decisions and, following the ‘Wilson doctrine’, develop game-theoretic analyses that dispense with unrealistic, excessively demanding common knowledge assumptions.

To the best of our knowledge, the methods for reducing common knowledge assumptions closest to our approach build on Battigalli and Siniscalchi’s (2003) Δ\Delta-restrictions. Formally, a Δ\Delta-restrictions is a specification of profiles of first-order beliefs, Δ:=(Δi)iI\Delta:=(\Delta_{i})_{i\in I}, so that for each player ii we have a collection Δi:=(Δθi)θiΘi\Delta_{i}:=(\Delta_{\theta_{i}})_{\theta_{i}\in\Theta_{i}} where ΔθiΔHi{h0}(Si×Θ0×Θi)\Delta_{\theta_{i}}\subseteq\Delta^{H_{i}\cup\{h^{0}\}}(S_{-i}\times\Theta_{0}\times\Theta_{-i}) for every payoff type θi\theta_{i}. The difference between Δ\Delta-restrictions and profiles of subjective payoff structures is mainly twofold. On the one hand, Δ\Delta-restrictions allow for placing restrictions on the evolution of strategic uncertainty, not only payoff uncertainty. On the other, Δ\Delta-restrictions impose constraints over higher-order beliefs by assuming that Δ\Delta is commonly known among players. Thus, it is possible to envision subjective payoff structures as Δ\Delta-restrictions that (1)(1) do not place constrains on beliefs about opponents’ behavior, and (2)(2) include specifications of constraints on higher-order beliefs that do not necessarily entail nontrivial common knowledge assumptions.

The restrictions embodied by subjective payoff structures also provide a reduced from method of representing unawareness in games. The uncertainty about the payoff state could be thought of as arising from the resolution of some unmodeled contingencies regarding physical aspects of the world external to the choices made by the players—a state of nature θ0iIZ\theta_{0}\in\prod_{i\in I}\mathds{R}^{Z} being shorthand for the particular resolution of this uncertainty that induces such payoffs to each player. Thus, although a player may easily envision the entirety of the abstract space iIZ\prod_{i\in I}\mathds{R}^{Z}, in actuality he considers on those states relating to some contingency he is aware of.

Now, when a player is aware of a contingency but initially deems it probability zero, he may revise his beliefs by reasoning about his opponents behavior. This is not so when the player is unaware of the contingency, since observing his opponents actions will not (in general) enlighten him to new contingencies. Thus, the essential attribute of our formalism—persistent, heterogeneous belief about the set of payoff states—can arise as a natural consequence of asymmetric awareness.

Of course, an agent’s perception of other agents’ awareness must be expressible in terms of things he is himself aware of. This is the motivation of various extant restrictions in the awareness literature—closure under subformula in Fagin and Halpern (1988), the confinidness property in Heifetz, Meier and Schipper (2006), property F1 in Piermont (2019). The translation to our set up would be: if ii is aware that jj is aware of θ\theta then ii must be himself aware of θ\theta. This conceptual restriction can be made formal via the edict: (di,2(j))0(di,1)0(d_{i,2}(j))_{0}\subseteq(d_{i,1})_{0}. The set of payoff states that ii considers jj to consider is a subset of those he himself is aware of.

Solution concepts

Conjectures and sequential rationality

The evolution throughout the game of each player’s subjective beliefs is presented by conjectures. Formally, for each player ii endowed with a subjective model (di,ti)(d_{i},t_{i}), a conjecture consists in a list of beliefs μi=(μi(h))hHi{h0}\mu_{i}=(\mu_{i}(h))_{h\in H_{i}\cup\{h^{0}\}} satisfying that:

  • (i)(i)

    For every history hh, either initial or in which player ii is active, μi(h)\mu_{i}(h) is a probability measure on Si×(di,1)0×Ti(di|i)S_{-i}\times(d_{i,1})_{0}\times T_{-i}(d_{-i|i}) that assigns probability 11 to Si×(di,1)0×Ti(di|i)S_{-i}\times(d_{i,1})_{0}\times T_{-i}(d_{-i|i}).

  • (ii)(ii)

    Whenever possible, beliefs are updated following conditional probability.

  • (iii)(iii)

    The initial beliefs represented by μi\mu_{i} are consisted with those represented by tit_{i}:

    marg(di,1)0×Ti(di|i)μi(h0)=τi(πi(ti)).\textup{marg}_{(d_{i,1})_{0}\times T_{-i}(d_{-i|i})}\mu_{i}(h^{0})=\tau_{i}(\pi_{i}(t_{i})).

We denote the set of conjectures consistent with subjective model (di,ti)(d_{i},t_{i}) by C(di,ti)\textup{C}(d_{i},t_{i}).

Given a payoff type θi\theta_{i} and a history hHi{h0}h\in H_{i}\cup\{h^{0}\}, each conjecture μi\mu_{i} naturally induces a conditional expected payoff for each strategy sis_{i}:

Ui(μi,si|θi,h):=Si×Θ0×Θiui(z(si;si|h),(θ0,θi,θi))d(margSi×Θ0×Θiμi(h)).U_{i}\left(\mu_{i},s_{i}\left|\theta_{i},h\right.\right):=\int_{S_{-i}\times\Theta_{0}\times\Theta_{-i}}u_{i}\left(z\left(s_{-i};s_{i}\left|h\right.\right),(\theta_{0},\theta_{i},\theta_{-i})\right)\textup{d}(\textup{marg}_{S_{-i}\times\Theta_{0}\times\Theta_{-i}}\mu_{i}(h)).

Finally, given a payoff type θi\theta_{i} and a conjecture μi\mu_{i}, the notion of sequential rationality is then captured via the set of (sequential) best responses:191919Notice that, if we denote the set of all conjecture of player ii as ΔHi{h0}(Si×(di,1)0×Ti(di|i))\Delta^{H_{i}\cup\{h^{0}\}}(S_{-i}\times(d_{i,1})_{0}\times T_{-i}(d_{-i|i})), then ri:Θi×ΔHi{h0}(Si×(di,1)0×Ti(di|i))Sir_{i}:\Theta_{i}\times\Delta^{H_{i}\cup\{h^{0}\}}(S_{-i}\times(d_{i,1})_{0}\times T_{-i}(d_{-i|i}))\rightrightarrows S_{i} is upper hemicontinuous and nonempty-valued.

ri(θi,μi):={siSi|sihHi(si)argmaxsiSiUi(μi,si|θi,h)}.r_{i}\left(\theta_{i},\mu_{i}\right):=\left\{s_{i}\in S_{i}\left|s_{i}\in\bigcap_{h\in H_{i}(s_{i})}\textup{arg}\underset{s^{\prime}_{i}\in S_{i}}{\textup{max}}\,U_{i}\left(\mu_{i},s^{\prime}_{i}\left|\theta_{i},h\right.\right)\right.\right\}.

Extensive form rationalizability

Extensive form rationalizability formalizes the notion of forward induction; that is, the idea that while forming conjectures about opponents’ behavior, players take observed past choices into account. Specifically, as clarified by the epistemic analysis by Battigalli and Siniscalchi (2002), players rationalize observed behavior to the ‘highest possible degree’ of strategic sophistication the behavior is consistent with: Rationality; or rationality and belief whenever possible in opponents’ rationality; or rationality, belief whenever possible in opponents’ rationality and belief whenever possible in opponents’ belief whenever possible in their opponents’ rationality; etc. Thus, the main conceptual feature of the solution concept is that observed behavior is employed to infer other players’ private information and that conjectures about behavior are based on this inference. To be more specific, if a player entertains certain beliefs about the payoff type of an opponent, upon observing a choice that would have been irrational for that payoff type, if possible, she should update her beliefs about the payoff type and believe that the true type is one for which the observed behavior is rational.

Formally, the original notion due to Pearce (1984) and Battigalli (1997) is generalized, in our context, as a correspondence that maps each player ii’s subjective model (di,ti)(d_{i},t_{i}) to a certain subset of strategies Fi(di,ti)Si\textup{F}_{i}(d_{i},t_{i})\subseteq S_{i}:

Definition 7 (Extensive form rationalizability, c.f. Pearce, 1984, Battigalli, 1997).

Let Γ\Gamma be an extensive form. Player ii’s set of (interim) extensive form rationalizable strategies for subjective model (di,ti)(d_{i},t_{i}) is Fi(di,ti):=k0Fi,k(di,ti)\textup{F}_{i}(d_{i},t_{i}):=\bigcap_{k\geq 0}\textup{F}_{i,k}(d_{i},t_{i}) where Fi,0(di,ti)=Si\textup{F}_{i,0}(d_{i},t_{i})=S_{i} and Ci,0F(di,ti)=Ci(di,ti)\textup{C}_{i,0}^{F}(d_{i},t_{i})=\textup{C}_{i}(d_{i},t_{i}), and for each k0k\geq 0,

Si(h)×Ti(di|i)Graph(Fi,k(di|i,)),μi(h)[(di,1)0×M]=1 for some MGraph(Fi,k(di|i,))}.\displaystyle:=\left\{\mu_{i}\in\textup{C}_{i,k}^{F}(d_{i},t_{i})\left|\begin{tabular}[]{l}$\textup{$\forall h\in H_{i}\cup\{h^{0}\}$ s.t.}$\\[8.61108pt] $S_{-i}(h)\times T_{-i}(d_{-i|i})\cap\textup{Graph}\left(\textup{F}_{-i,k}(d_{-i|i},\,\cdot\,)\right)\neq\emptyset$,\\[6.45831pt] $\mu_{i}(h)[(d_{i,1})_{0}\times M]=1\textup{ for some }M\subseteq\textup{Graph}\left(F_{-i,k}(d_{-i|i},\,\cdot\,)\right)$\end{tabular}\right.\right\}.
Fi,k+1(di,ti)\displaystyle\textup{F}_{i,k+1}(d_{i},t_{i}) :={siSi|μiCi,kF(di,ti) s.t. siri(θi(ti),μi)},\displaystyle:=\left\{s_{i}\in S_{i}\left|\exists\mu_{i}\in\textup{C}_{i,k}^{F}(d_{i},t_{i})\textup{ s.t. }s_{i}\in r_{i}(\theta_{i}(t_{i}),\mu_{i})\right.\right\},
Ci,k+1F(di,ti)\displaystyle\textup{C}_{i,k+1}^{F}(d_{i},t_{i}) :={μiCi,kF(di,ti)| hHi{h0} s.t.

Backward rationalizability

Backward rationalizability (see Penta, 2011, 2015) extends the standard notion of backward induction to games with incomplete information. The reasoning process behind differs from that behind extensive form rationalizability seen above: Here, upon observing unexpected behavior, players do not necessarily try to find a rationale for said behavior (even at the cost of dropping their belief in ‘common belief in rationality’); instead, they can interpret observed behavior as a mistake and maintain their belief that ‘common belief in rationality’ holds.202020It is not conceptually accurate to talk about common belief in rationality in a dynamic game, not at every history this belief can be held; however, we keep this terminology in the present discussion for expositional simplicity. Thus, the difference between the extensive form rationalizability and backward rationalizability is one of epistemic priority (see Catonini, 2019): While the former dictates to update higher-order beliefs about payoff states in order to perform a rationalization to the highest possible degree, the latter permits to maintain the higher-order beliefs about payoff states, believe that unexpected behavior is uninformative and keep the faith on continuation behavior being consistent with rationality and common belief thereof.

Under subjective payoff structures, Penta’s (2011; 2015) original notion is generalized as a correspondence that maps each player ii’s subjective model (di,ti)(d_{i},t_{i}) to a certain subset of strategies Bi(di,ti)Si\textup{B}_{i}(d_{i},t_{i})\subseteq S_{i}. It is easy to check that the following Definition 8 below captures the idea, formalized in epistemic language by Perea (2014) and Battigalli and De Vito (2018),212121Unlike Penta (2011, 2015), and in order to make the comparison with extensive form rationalizability more immediate, we drop the requirement that a player has beliefs about her own strategy. Nonetheless, it is immediate that the version we employ captures the idea of ‘continuation behavior consistent with common belief in rationality’ as formalized by rationality and common belief in future rationality in Perea (2014) or common full belief in optimal planning and in belief in continuation consistency in Battigalli and De Vito (2018). that players believe that rationality will hold at every possible higher-order in the continuation game, even after moves inconsistent with it. First, define the set of profiles of player ii’s opponents’ strategies that are equivalent to sis_{-i} in the continuation game following history hh as:

[si]h:={siSi|ji,hHj s.t. hh,sj(h)=sj(h)}.[s_{-i}]_{h}:=\left\{s_{-i}^{\prime}\in S_{-i}\left|\forall j\neq i,\forall h\in H_{j}\textup{ s.t. }h\prec h^{\prime},s^{\prime}_{j}(h)=s_{j}(h)\right.\right\}.

Then, backward rationalizability is defined in our context as:

Definition 8 (Backward rationalizability, c.f. Penta, 2011).

Let Γ\Gamma be an extensive form. Player ii’s set of (interim) backward rationalizable strategies for subjective model (di,ti)(d_{i},t_{i}) is Bi(di,ti):=kBi,k(di,ti)\textup{B}_{i}(d_{i},t_{i}):=\bigcap_{k\in\mathds{N}}\textup{B}_{i,k}(d_{i},t_{i}) where Bi,0(di,ti):=Si\textup{B}_{i,0}(d_{i},t_{i}):=S_{i} and Ci,0B(di,ti):=Ci(di,ti)\textup{C}_{i,0}^{B}(d_{i},t_{i}):=\textup{C}_{i}(d_{i},t_{i}), and for each k0k\geq 0,

Bi,k+1(di,ti)\displaystyle\textup{B}_{i,k+1}(d_{i},t_{i}) :={siSi|μiCi,kB(di,ti) s.t. siri(θi(ti),μi)},\displaystyle:=\left\{s_{i}\in S_{i}\left|\exists\mu_{i}\in\textup{C}_{i,k}^{B}(d_{i},t_{i})\textup{ s.t. }s_{i}\in r_{i}(\theta_{i}(t_{i}),\mu_{i})\right.\right\},
Ci,k+1B(di,ti)\displaystyle\textup{C}_{i,k+1}^{B}(d_{i},t_{i}) :={μiCkB(di,ti)|hHi{h0},(θ0,si,ti)supp μi(h)[si]hBi,k(di|i,ti)}.\displaystyle:=\left\{\mu_{i}\in\textup{C}_{k}^{B}(d_{i},t_{i})\left|\begin{tabular}[]{l}$\textup{$\forall h\in H_{i}\cup\{h^{0}\}$,}$\\[8.61108pt] $(\theta_{0},s_{-i},t_{-i})\in\textup{supp }\mu_{i}(h)\Rightarrow[s_{-i}]_{h}\cap\textup{B}_{-i,k}(d_{-i|i},t_{-i})\neq\emptyset$\end{tabular}\right.\right\}.

Results

The application of the theoretical framework developed above will shed light on the impact on strategic behavior of players’ discrepancies about the state space. Section 4.1 shows that extensive form rationalizable outcomes are not always a refinement of those by backward induction, and that they are not robust to misspecifications of subjective models. Section 4.2 shows that extensive form rationalizability delivers predictions that are robust to misspecifications of initial beliefs (Proposition 2) and that backward rationalizability delivers predictions that are robust to misspecification of subjective models (Proposition 3); a direct consequence of these results is that, in dynamic games where the set of payoff states is rich, extensive form rationalizability, backward rationalizability and interim rationalizability all coincide (Corollaries 1 and 2).222222By interim rationalizability we refer to Dekel, Fudenberg and Morris’s (2007) interim correlated rationalizability applied to the normal-form representation of the dynamic game. In Penta (2012) interim rationalizability is defined using sequential rationality instead of ex ante rationality, and the solution concept is dubbed interim sequential rationalizability. In the games considered by Chen (2012), where players have no private information, interim sequential rationalizability and interim rationalizability coincide (see Penta (2012), p. 648). Finally, in Section 4.3 we prove that, by perturbing standard models, it is possible to create unique selection arguments à la Weinstein and Yildiz (2007) so that, for any standard model and any extensive form rationalizable outcome, the outcome is the unique prediction along the misspecified subjective models (Theorem 1), and we end by presenting an example that illustrates why such a unique selection argument is impossible for backward rationalizability.

Rationalization as a refinement criterion

Main counterexample

Consider again the extensive form studied in Section 3.1.2. Let us endow it with a standard payoff structure Θ={θ}\Theta=\{\theta\} such that the resulting dynamic game is:

112211
22
0
0
0
2
-1
1
1
A1A_{1}D1D_{1}aaddA2A_{2}D2D_{2}

Now, for each player i=1,2i=1,2 let di0d_{i}^{0} represent the standard subjective payoff structure associated to Θ\Theta. Clearly, there is a unique type for player ii consistent with Θ\Theta, namely the one representing common belief in θ\theta, which we denote by ti0t_{i}^{0}. Thus, there is a unique profile of subjective models, (M1,M2)(M_{1},M_{2}) where Mi=(di0,ti0)M_{i}=(d_{i}^{0},t_{i}^{0}) for both i=1,2i=1,2, which is, indeed, a standard model. The strategic analysis of this game yields the following conclusions:

  • The unique prediction consistent with extensive form rationalizability is D1D_{1} (outcomewise). This is easy to see. First, player 11 will never choose A2A_{2} at her second turn. Second, if player 22 finds herself at A1A_{1}, while she can be surprised that 11 took the risk of advancing instead of guaranteeing a payoff of 22, she need not abandon the belief that 1 is rational. Player 11 may have advanced, rationally, by believing that 22 does not expect her (i.e., 2 does not expect 1) to be rational in her second turn, and therefore believing that 22 will advance as well. Thus, since 22 maintains the belief that 11 is rational, she will expect 11 to choose D2D_{2} in her first second turn, and will therefore choose dd. Being able to forecast this, player 11 will choose D1D_{1} in her first turn.

  • The unique prediction consistent with backward rationalizability is D1D_{1} as well. This is immediate.

Thus, both solution concepts uniquely predict D1D_{1} is played. Suppose now that we perturb the game using a sequence consisting of the subjective models defined in Section 3.1.2, (M1n,M2n)n(M_{1}^{n},M_{2}^{n})_{n\in\mathds{N}}. For each nn\in\mathds{N}, the situation can be represented as follows:

P1P_{1}P2P_{2}P1P_{1}
21n2-\frac{1}{n}
0
·····
2+1n2+\frac{1}{n}
0
0
0
2
-1
1
1
A1A_{1}D1D_{1}aaddA2A_{2}D2D_{2}

Remember that here players commonly agree on the payoffs that correspond to terminal histories (A1,d)(A_{1},d), (A1,a,D2)(A_{1},a,D_{2}) and (A1,d,A2)(A_{1},d,A_{2}). However, in player 22’s mind they also commonly agree about the payoffs corresponding to D1D_{1} being (2+1/n,0)(2+1/n,0), but this is wrong: In player 11’s mind, the payoff that corresponds to D1D_{1} is (21/n,0)(2-1/n,0), but she knows that player 22 believes in the commonly agreement on (2+1/n,0)(2+1/n,0). Obviously, the larger nn, the closer we are to the game with commonly known payoff structure Θ\Theta. This is the sense in which we say that (M1n,M2n)n(M_{1}^{n},M_{2}^{n})_{n\in\mathds{N}} converges to (M1,M2)(M_{1},M_{2}). For each given nn, the corresponding strategic analysis yields the following conclusions:

  • The predictions consistent with extensive form rationalizability are multiple now: D1D_{1}, (A1,d)(A_{1},d) and (A1,a,D2)(A_{1},a,D_{2}). This is easy to see. Upon observing A1A_{1}, which in her mind is strictly dominated, player 22 may drop the belief that player 11 is rational, and may have any arbitrary belief about 1’s future behavior. She might believe that player 11 will be completely irrational and choose A1A_{1} in her second turn, or she might believe player 11 will be more sensible next time and choose D2D_{2}. In consequence, both aa and dd are rationalizable choices for player 22. As a result, (A1,D2)(A_{1},D_{2}), (D1,A2)(D_{1},A_{2}) and (D1,D2)(D_{1},D_{2}) are all rationalizable strategies for player 11, depending on what she expects player 22 to choose.

  • The unique prediction consistent with backward rationalizability is D1D_{1} again. Also again, this is immediate.

The example shows that extensive form rationalizability refining backward rationalizability crucially depends on players commonly agreeing on the set of possible payoff states. As arbitrarily small disagreements arise, both (A1,d)(A_{1},d) and (A1,a,D2)(A_{1},a,D_{2}) become consistent with extensive form rationalizability; however, none of them is consistent with backward rationalizability. Importantly, the example delivers an additional lesson: Extensive form rationalizable outcomes fail to be upper hemicontinuous, even in standard models; all D1D_{1}, (A1,d)(A_{1},d) and (A1,a,D2)(A_{1},a,D_{2}) are consistent with extensive form rationalizability along the sequence, but only D1D_{1} is so in the limit.

Conclusions

Section 4.1.1 illustrates the critical dependence of rationalization processes to apparently negligible modeling details of the state space. As a result, the intuition that forward induction (i.e., extensive form rationalizability) always provides refinements of backward induction (i.e., backward rationalizability) is revealed as both extremely fragile and dependent on agents perfectly agreeing on the possible payoff-relevant nonstrategic contingencies of the environment. That is, arbitrarily small discrepancies among players can lead to the collapse of usual refinement criteria based on the use of observed behavior in order to form conjectures about future behavior.

The reasons for the conceptual discontinuity are evident. There is an option that is rational for player 11 that, because of an arbitrarily small discrepancy on the perception of the payoff structure, cannot be rationalized by player 22. Hence, player 11 can try to play dumb by choosing to advance and obtain, depending on the reaction of 22, a better payoff than the one backward rationalizability allows for. In contrast to the standard environment with a commonly known payoff structure—where it is always impossible that a choice rendered as irrational by her opponents is potentially beneficial for a player—arbitrarily small disagreements about the payoff structure ca permit an agent to profit from seemingly irrational choices. In technical terms, the observation is the result of the interplay between two factors:

  1. 1.

    The correspondence that maps subjective payoff structures to histories that can be reached by extensive form rationalizable strategies of the opponents is not lower hemicontinuous. This is an obvious fact, if only because it is well-known that the best response correspondence is not lower hemicontinuous. It is also easily visible in the example. In the limit, initial action A1A_{1} is rational for player 1 and thus, player 22’s history A1A_{1} is reachable by strategies in F1,1(M1)\textup{F}_{1,1}(M_{1}). On the contrary, for every nn\in\mathds{N}, history A1A_{1} is not reachable by any strategy in F1,1(M1n)\textup{F}_{1,1}(M_{1}^{n}). To put it explicitly: A history that can be reached in the limit may not be reached along the perturbation.

  2. 2.

    Extensive form rationalizability puts certain restrictions on beliefs only at some histories: requiring that player 22 believes that player 11 is rational at every history that could have been reached if this was the case, but no constraint is placed in the remaining histories. This, together with the lack of lower hemicontinuity mentioned above leads to the collapse of upper hemicontinuity. To see it, let (μ2n)n(\mu_{2}^{n})_{n\in\mathds{N}} be a sequence of player 22’s conjectures such that, for each nn\in\mathds{N}, μ2n\mu_{2}^{n} assigns, at every history hH2{h0}h\in H_{2}\cup\{h^{0}\} reachable by some strategy in F1,1(M1n)\textup{F}_{1,1}(M_{1}^{n}), probability 11 to player 11 choosing a strategy in F1,1(M1)\textup{F}_{1,1}(M_{1}). Clearly, this constraint only applies in h=h0h=h^{0}. Hence, even if the sequence is convergent, there is no way to guarantee that its limit, μ2\mu_{2}, will assign, at history A1A_{1}, probability 1 to player 11 choosing a strategy in F1,1(M1)\textup{F}_{1,1}(M_{1}): The elements in the sequence do not satisfy any particular requirement at history A1A_{1}. In consequence, it is impossible to ensure that a sequence of conjectures that justifies the inclusion of strategy s2s_{2} in each F2,2(M2n)\textup{F}_{2,2}(M_{2}^{n}) converges (or has a subsequence that does so) to a conjecture that justifies the inclusion of s2s_{2} in F2,2(M2)\textup{F}_{2,2}(M_{2}).

Besides the above, the example delivers other two additional important takeaways. First, that extensive form rationalizability is not robust to misspecifications of models (in particular, of subjective payoff structures): Arbitrarily small discrepancies on the set of payoff states can lead to predictions that where ignored in the benchmark model ((A1,d)(A_{1},d) and (A1,a,D1)(A_{1},a,D_{1}) in the example). Second, that multiplicity of the extensive form rationalizable outcomes can be robust: It is easy to see that, in the example, for every nn\in\mathds{N}, outcomes D1D_{1}, (A1,d)(A_{1},d) and (A1,a,D1)(A_{1},a,D_{1}) are all consistent with extensive form rationalizability for every model MM^{\prime} close enough to (M1n,M2n)(M_{1}^{n},M_{2}^{n}). These two observations, further discussed in Sections 4.2 and 4.3 respectively, contradict some established facts of static games. On the one hand, since in static games interim beliefs are not updated, all the uncertainty regarding utility functions can be encapsulated by the type, and we know from Dekel, Fudenberg and Morris (2007) that interim rationalizability (the static version of extensive form rationalizability) is robust to misspecifications of types. On the other hand, the seminal analysis of Weinstein and Yildiz (2007) reveals that multiplicity of interim rationalizability is nowhere robust, there exists no type with a neighborhood within which interim rationalizability is always multiple.

Robustness to misspecifications

Extensive form rationalizability

The example above shows that extensive form rationalizability is not robust: Arbitrarily small misspecifications of subjects’ perceptions of the payoff structure can lead to outcomes ignored by the benchmark model.232323Specifically, while the original standard model only admits D1D_{1}, predictions (A1,a,D2)(A_{1},a,D_{2}), (A1,d)(A_{1},d) and D1D_{1} all are consistent with the perturbed models. Notice however, that the models employed in the example, both the benchmark one and the perturbations, are such the corresponding payoff structures only admit one type. Thus, the lack of robustness of extensive form rationalizability follows from its lack of upper hemicontinuity on subjective payoffs structures. In this section we show that extensive form rationalizability does pass some partial robustness test: For a fixed subjective payoff structure, it is upper hemicontinuous on initial types.

Proposition 2 (Robustness of extensive form rationalizability to misspecifications of initial beliefs).

Let Γ\Gamma be an extensive form. Then for any player ii and any subjective payoff structure did_{i}, the correspondence Fi(di,):Ti(di)Si\textup{F}_{i}(d_{i},\cdot\,):T_{i}(d_{i})\rightrightarrows S_{i} is upper hemicontinuous.

The conceptual relevance of the result is twofold. First, standard economic modeling assumes that the state space of the game is commonly known and thus, imperturbable. Proposition 2 guarantees that, in these settings, extensive form rationalizable predictions are robust (also at individual level, i.e., in terms of strategies, not only outcomes). Second, as we argue next, the preposition warns against the perils and unfitting consequences of richness assumption in dynamic games. We know from Penta (2012) and Chen’s (2012) work that under richness, no refinement of interim rationalizability is robust. Now, Proposition 2 establishes that extensive form rationalizability is upper hemicontinuous in the settings studied by Penta (2012) and Chen (2012), and we know that extensive form rationalizability is a refinement of interim rationalizability. Hence, the conclusion, formally stated in Corollary 1, is immediate: Under richness, interim rationalizability (both ex ante and sequential) and extensive form rationalizability coincide.242424The argument is so straightforward that we omit any proof Thus, richness assumptions seem unsuitable for dynamic games; by killing all the bite of rationalization, they render the use of dynamic modeling a mostly redundant tool. Let us end the discussion by formally presenting the equivalence:252525Battigalli and Siniscalchi (2007) provide a similar insight in a weaker version of the following result (Proposition 5 in p.177): For rich Θ\Theta and state of nature θ\theta extensive form rationalizability and interim rationalizability coincide for those types consistent with initial common belief of θ\theta.

Corollary 1 (Triviality of rationalization under richness I).

Let Γ,Υ\langle\Gamma,\Upsilon\rangle be a dynamic game. Then, if Υ\Upsilon is rich, extensive form rationalizability and interim rationalizability coincide.

Backward rationalizability

The example in Section 3 shows that the failure of upper hemicontinuity of extensive form rationalizability on subjective payoff structures is due to the interplay of: (1)(1) the set of histories reachable by extensive-from rationalizability not being lower hemicontinuous on subjective payoff structures, and (2)(2) the rationalization requirement intrinsic to extensive form rationalizability depending on the subjective payoff structure.262626See bullet points 1.1. and 2.2. in Section 4.1.2. Fact (1)(1) implies that beliefs satisfying (2)(2) along a sequence may not imply that the limit belief satisfies (2)(2), because in the limit, the number of histories at which rationalization should me made can have exploded. Now, the reasoning process that explains backward rationalizability imposes the same constraint at every history,272727Namely, that continuation play will follow according to what backward rationalizability of the previous order would have dictated. and thus, if it holds for beliefs in a sequence, it is always satisfied as well by the limit belief. As the following proposition shows, this intuition backward rationalizability is immune to the mechanism that causes extensive form not to be robust, proves to be correct:

Proposition 3 (Robustness of backward rationalizability to misspecifications of models).

Let Γ\Gamma be an extensive form. Then for any player ii, the correspondence Bi:iSi\textup{B}_{i}:\mathscr{M}_{i}^{\infty}\rightrightarrows S_{i} is upper hemicontinuous.

It is immediate then that the result below obtains as a corollary. Clearly, this reinforces the critique on the assumption of richness in dynamic modeling.

Corollary 2 (Triviality of rationalization under richness II).

Let Γ,Υ\langle\Gamma,\Upsilon\rangle be a dynamic game. Then, if Υ\Upsilon is rich, backward rationalizability, extensive form rationalizability and interim rationalizability coincide.

Once backward rationalizable predictions are confirmed to be robust, a question arises naturally: Does backward rationalizability characterize the strongest predictions that are robust to misspecifications of models? That is, is it true that no (nontrivial) refinement of backward rationalizability is robust? The standard argument for results of this kind in static games is that by Weinstein and Yildiz (2007). In those games every interim rationalizable prediction is uniquely selected by some perturbation of the type. In consequence, no refinement of interim rationalizability is robust.282828Suppose it is and take the prediction ignored by the hypothetical refinement. In the perturbation, this prediction should be the unique prediction made by the hypothetical refinement. If the latter was upper hemicontinuous, it would also include it in the limit. The example in Section 4.3.2 below shows that not every prediction of backward rationalizability admits a perturbation that uniquely selects it. In consequence, our analysis leaves two open questions: (i)(i) does backward rationalizability characterize the strongest predictions that are robust to misspecifications of models (which, if true, would require a proof different from the usual approach)? And (ii)(ii) in case the answer to the previous question is negative, which is the solution concept that does characterize the strongest predictions robust to misspecifications of models?

Unique selections

Theorem for extensive form rationalizability

The following theorem shows that, for any subjective model and any outcome consistent with extensive form rationalizability, it is possible to create a perturbation of the model that allows for the chosen outcome to be uniquely selected along the perturbed sequence—uniquely selected outcomewise, not in terms of strategies. Notice that the theorem does not require that the state space associated to the standard model is rich: It is possible to perturb the state space using profiles of subjective payoff structures whose associated subjective models satisfy higher-order richness, at increasingly higher order the more the state space is approximated. This allows for the usual infection argument to be easily constructed. In consequence, the theorem establishes that dynamic games not satisfying richness also admit unique selections:

Theorem 1 (Structure theorem for extensive form rationalizability).

Let Γ,Υ\langle\Gamma,\Upsilon\rangle be a dynamic game. Then, for any profile of finite types tT(Θ)t\in T(\Theta) and any strategy profile sF(Υ,t)s\in\textup{F}(\Upsilon,t) there exists a sequence of profiles of models (dn,tn)n(d^{n},t^{n})_{n\in\mathds{N}} such that:

  • (i)(i)

    (di,ti)n(d_{i},t_{i})_{n\in\mathds{N}} converges to (di,ti)(d_{i},t_{i}).

  • (ii)(ii)

    For every nn\in\mathds{N} and every strategy profile snF(dn,tn)s^{n}\in\textup{F}(d^{n},t^{n}), z(sn|h0)=z(s|h0)z(s^{n}|h^{0})=z(s|h^{0}).

The interest of the theorem lies on illustrating that extensive form rationalizability admits unique selections in settings in which rationalization adds bite, not only on those in which extensive-from rationalizability is necessarily equivalent to interim rationalizability (because richness of the state space is assumed). On top of that, while the perturbations required do entail higher-order richness assumptions, we know from Proposition 1 that this requirement is generically satisfied in the set of all subjective payoff structures. Thus, not only the consequence, but also the very demanding nature of standard richness assumptions is alleviated.

However, the conceptual implications beyond the previous observations are not clear. On the one hand, (perfect Bayesian) equilibrium outcomes do not, in general, refine those of extensive form rationalizability, so Theorem 1 does not allow for concluding that equilibrium predictions are not robust—as Weinstein and Yildiz’s (2007) structure theorem allows for in the static case. On the other hand, (i)(i) only standard models are perturbed in the theorem (not arbitrary profiles of subjective models), and (ii)(ii) extensive form rationalizability is not robust to perturbations of subjective models. In consequence it cannot be concluded either that uniqueness of extensive form rationalizable predictions is a generic phenomenon in the space of profiles of subjective models. Indeed, the contrary is illustrated in the example in Section 4.1: It is possible that a standard model is perturbed so that, along the sequence, multiplicity of extensive form rationalizable outcomes is robust. Our last example below shows that unique selection arguments do not exist, at least as a rule, for backward rationalizability.

Counterexample for backward rationalizability

The following example illustrates that, while backward rationalizability is robust to misspecifications of models—see Proposition 3 above, the questions of whether it is the strongest robust solution concept remains open: It is not a conclusion that follows from a standard unique selection argument. As exemplified below, there exist situations in which it is not possible to perturb standard models so that backward rationalizable outcomes are uniquely selected. In consequence, even if true, there is no obvious way to claim that no refinement of backward rationalizability delivers robust predictions.

Consider the following dynamic game with standard payoff structure consisting of states θ1\theta_{1} and θ2\theta_{2}:

112211
44
44
3
3
0
0
2
2
A1A_{1}D1D_{1}aaddA2A_{2}D2D_{2}State θ1\theta_{1}
112211
0
0
3
0
1
1
2
2
A1A_{1}D1D_{1}aaddA2A_{2}D2D_{2}State θ2\theta_{2}

The description of the payoff structure is completed by assuming that player 11 always knows the payoff state, that the second player never knows it, and that these two are common knowledge. The set of backward rationalizable strategies of player 22 for arbitrary type t2t_{2} is:

  • (A)

    {a}\{a\}, if t2t_{2} assigns positive probability to θ2\theta_{2}. At state θ2\theta_{2} strategy (A1,A2)(A_{1},A_{2}) is strictly dominant for player 1 (who, remember, knows that the state is θ1\theta_{1}). Thus, if player 22 observes that player 11 advances in her first round, she updates her beliefs by excluding the possibility of θ1\theta_{1} being the true state. In consequence, aa becomes her only rational choice.

  • (B)

    {a,d}\{a,d\}, if t2t_{2} assigns null probability to θ2\theta_{2}. In this case it is unexpected for player 22 to observe player 1 advance, D1D_{1} would have been strictly dominant if the state had been θ1\theta_{1}. Thus, player 22 needs to perform an update of her beliefs from scratch, what allows for the following two possibilities: Was choosing A1A_{1} a mistake and, effectively, the state is θ1\theta_{1}? Or where player 22’s beliefs wrong and the state is θ2\theta_{2}, what justifies player 11’s decision to advance? Both updating alternatives (and the corresponding mixed beliefs) are admissible according to backward rationalizable reasoning, and the corresponding only rational choices are dd and aa, respectively (and one or either of them, if the updated is a mixed belief).

Pick now standard model M:=((Θ,t1),(Θ,t2))M:=((\Theta,t_{1}),(\Theta,t_{2})), where (i)(i) Θ:={θ1,θ2}\Theta:=\{\theta_{1},\theta_{2}\} and (ii)(ii) t1t_{1} and t2t_{2} represent common belief in states θ2\theta_{2} and θ1\theta_{1}, respectively. We claim now that no perturbation of MM can lead to dd being uniquely selected for player 22 and thus, to outcome (A1,d)(A_{1},d) (which is backward rationalizable for MM) being uniquely selected. To see it notice first that if a standard subjective model is close enough to Θ\Theta then it is the union of two disjoint sets of states, Θ1\Theta^{1} and Θ2\Theta^{2}, such that it is strictly dominant for player 11 to choose D1D_{1} at every θ1Θ1\theta_{1}^{\prime}\in\Theta^{1} and to choose (A1,A2)(A_{1},A_{2}) at every θ2Θ2\theta_{2}^{\prime}\in\Theta^{2}. Accordingly, for any perturbation of (Θ,t2)(\Theta,t_{2}), the same argument as in (A) and (B) leads to the following conclusions: (A’) if in the perturbed model Θ2\Theta^{2} gets positive probability then aa is player 2’s unique backward rationalizable strategy, and (B’) if in the perturbed model Θ2\Theta^{2} gets zero probability then both aa and dd are backward rationalizable for player 22. As a result, there is no perturbation of MM in which aa is not backward rationalizable for player 22.

Related literature

Our approach bears some similarity with the notion of unawareness (see Fagin and Halpern, 1988; Modica and Rustichini, 1994), and in particular with the state-space or semantic approach to modeling awareness (see Dekel, Lipman and Rustichini, 1998; Heifetz, Meier and Schipper, 2006, 2008; Piermont, 2019), in which agent’s are endowed with a coarse understanding of the true space of uncertainty. In particular, these approaches often model uncertainty via a partial order of increasingly expressive state-spaces: agent’s residing in more expressive state-spaces can only reason about events in lower state-spaces.

Agents who are introspectively unaware, however, might reason that there exist contingencies they are unaware of, without, of course, knowing exactly what such contingencies entail (see Halpern and Rêgo, 2009; Halpern and Piermont, 2019). Subjective payoff structures can capture both naive and introspective unawareness by changing restrictions on the relation between did_{i} and dj|id_{j|i}: when did_{i} is required to contain dj|id_{j|i} then the agent ii is naively unaware, as he does not consider it possible that jj considers a contingency he does not.

Note, most game-theoretic analyses of unawareness—Heifetz, Meier and Schipper (2014), Perea (2018b) and Guarino (2020), for instance—have focused on unawareness w.r.t. actions or strategies, not payoff states.

Lack of common knowledge of the type structure has also been studied by Ziegler (2019), and alternative misspecifications of models, by Esponda and Pouzo (2016) and, in the context of macroeconomic models, Hansen and Sargent (2001) and Cho and Kassa (2017).

The study of disagreements about payoff structures follows the literature on how small changes in beliefs and information at high orders affects strategic behavior. Rubinstein’s (1989) email game documents that behavior under common knowledge or under almost common knowledge can vary drastically. Later, Weinstein and Yildiz (2007) show that these discontinuities of behavior are not an isolated phenomenon, but rather, a pervasive feature of games with incomplete information. Penta (2012) and Chen (2012) extend this observation to dynamic games. Ely and Pȩski (2011) and Ruiz G. (2018) characterize the types in which these discontinuities arise in static and dynamic settings, respectively. Penta and Zuazo-Garin (2021) study the strategic impact of the discontinuities corresponding to higher-order uncertainty about the observability of choices, not preferences. Within the literature of robust mechanism design, Oury and Tercieux (2012) and Chen, Mueller-Frank and Pai (2020) study which social choice functions are implementable is a way robust to small misspecifications of higher-order beliefs. To this respect, the new notion of continuity in this paper and our results on the continuity of different solution concepts suggest novel questions and techniques for the study of implementability in dynamic mechanism design.

The counterintuitive nature of behavior being so discontinuous on information has sprung a literature that argues that these discontinuities are an artifact of very specific formalization or unrealistic assumptions on behavior. An approach consists in varying the notions of ‘similarity’ of belief hierarchies or ‘approximation’, as studied by Dekel, Fudenberg and Morris (2006), Chen, Di Tillio, Faingold and Xiong (2010; 2017) or Morris, Shin and Yildiz (2016). Another approach consists in showing that these discontinuities vanish with the introduction of bounded rationality and, specifically, under arbitrarily small departures from the benchmark of rationality and common belief thereof, as in Strzalecki (2014), Heifetz and Kets (2018), Germano, Weinstein and Zuazo-Garin (2020), Murayama (2020) or Jimenez-Gomez (2019). On the contrary, the literature on global games, starting from Carlsson and van Damme (1993), has embraced the discontinuities of choices as an intrinsic feature of strategic behavior and leveraged on them to explain diverse economic phenomena such as currency crises (Morris and Shin, 1998), bank runs (Angeletos, Hellwig and Pavan, 2006; 2007), conflict (Baliga and Sjöström, 2012), overvaluation in financial markets (Han and Kyle, 2017) or disclosure policies for stress tests (Inostroza and Pavan, 2018). Perturbations of state spaces, together with the unique selection argument in Theorem 1, allow for exploring variations of dynamic global games in which: (i)(i) information about the fundamentals is exogenously obtained solely at the beginning of the interaction, and new one is inferred only through observed behavior, and (ii)(ii) agents can entertain different perceptions of which payoff-relevant contingencies can eventually take place.

Acknowledgments and disclaimers

A very preliminary version of this project was occasionally presented under the title “Rationalization and robustness in dynamic games with incomplete information.” Thanks are due to Pierpaolo Battigalli, Emiliano Catonini, Pierfrancesco Guarino, Jaromír Kovářík, Avi Lichtig, Antonio Penta and Gabriel Ziegler for insightful comments and valuable feedback, and to audiences at HSE University-International College of Economics and Finance and Università Bocconi. Zuazo-Garin acknowledges financial support from the Spanish Ministry of Economy and Competitiveness, from the Department of Education, Language Policy and Culture of the Basque Government (grants ECO2012-31346 and POS-2016-2-0003 and IT568-13, respectively), from the ERC Programme (ERC Grant 579424) and from the Russian Academic Excellence Project ‘5-100’. Zuazo-Garin also expresses his gratitude to Northwestern University and Kellogg School of Management (the Department of Economics, MEDS and the CMS-EMS, particularly) for immense hospitality. The authors are responsible for all errors.

Appendix A Proofs I: Propositions

Proposition 1

See 1

Proof.

For denseness, take any arbitrary standard payoff structure satisfying richness Υ\Upsilon. Then, for any subjective payoff structure did_{i} and any nn\in\mathds{N} define din(di)d_{i}^{n}(d_{i}) by setting di0(di):=Υd_{i}^{0}(d_{i}):=\Upsilon and, for any n1n\geq 1, di,1n(di):=di,1d_{i,1}^{n}(d_{i}):=d_{i,1} and di|in(di):=din1(di|i)d_{-i|i}^{n}(d_{i}):=d_{-i}^{n-1}(d_{-i|i}). This way, din(di)d_{i}^{n}(d_{i}) is a subjective payoff structure satisfying that di,kn(di)=di,kd_{i,k}^{n}(d_{i})=d_{i,k} for every knk\leq n. Thus, clearly, (din(di))n(d_{i}^{n}(d_{i}))_{n\in\mathds{N}} converges to did_{i}. Openness follows from a standard inductive argument: It is immediate that the set of subjective payoff structure satisfying 1st{}^{\textup{st}}-order richness is open; given that, it is immediate as well that the set of subjective payoff structure satisfying 2nd{}^{\textup{nd}}-order richness is open as well, and so on. ∎

Proposition 2

See 2

Proof.

We proceed by induction and verify that the correspondence Fi,k(di,):Ti(di)Si\textup{F}_{i,k}(d_{i},\,\cdot\,):T_{i}(d_{i})\rightrightarrows S_{i} is upper hemicontinuous for every k0k\geq 0. The claim holds trivially for the initial case (k=0k=0) so we can focus on the proof of the inductive step. Fix k0k\geq 0 such that the claim holds; we verify next that it also does for k+1k+1. Fix player ii, subjective payoff structure did_{i} and sequence of types (tin)n(t_{i}^{n})_{n\in\mathds{N}} converging to some type tit_{i}, and pick strategy sis_{i} such that siFi(di,tin)s_{i}\in\textup{F}_{i}(d_{i},t_{i}^{n}) for every nn\in\mathds{N}. For each nn\in\mathds{N}, pick conjecture μin\mu_{i}^{n} that justifies the inclusion of sis_{i} in Fi,k+1(di,tin)\textup{F}_{i,k+1}(d_{i},t_{i}^{n}). Let (μinm)m(\mu_{i}^{n_{m}})_{m\in\mathds{N}} be a convergent subsequence of (μin)n(\mu_{i}^{n})_{n\in\mathds{N}} with limit μi\mu_{i}. Notice then that for any k\ell\leq k and any history hh reachable by some profile of opponents’ strategies in Fi,(di|i,)\textup{F}_{-i,\ell}(d_{-i|i},\,\cdot\,) we have that:

μi(h)[(di,1(i))0×Graph(Fi,(di|i,))]\displaystyle\mu_{i}(h)\left[(d_{i,1}(i))_{0}\times\textup{Graph}\left(\textup{F}_{-i,\ell}(d_{-i|i},\,\cdot\,)\right)\right]\geq
limsupm\displaystyle\geq\underset{m\rightarrow\infty}{\textup{limsup}} μi(h)nm[(di,1)0×Graph(Fi,(di|i,))]=1,\displaystyle\,\mu_{i}(h)^{n_{m}}\left[(d_{i,1})_{0}\times\textup{Graph}\left(\textup{F}_{-i,\ell}(d_{-i|i},\,\cdot\,)\right)\right]=1,

because we know from the induction hypothesis that Fi,(di|i,)\textup{F}_{-i,\ell}(d_{-i|i},\cdot\,) has closed graph. This, together with the continuity of marginalization and the upper hemicontinuity of the best response correspondence allows for concluding that μi\mu_{i} is a conjecture that justifies the inclusion of sis_{i} in Fi,k+1(di,ti)\textup{F}_{i,k+1}(d_{i},t_{i}). ∎

Proposition 3

Auxiliary results

Lemma 1.

Let ΔC(Ω)\Delta^{C}(\Omega) be a set of conditional probability systems with (compact and metrizable) space of uncertainty Ω\Omega and family of conditioning events CC. Consider sequence (Ωn)n(\Omega^{n})_{n\in\mathds{N}} of closed subsets of Ω\Omega and and closed Ω0Ω\Omega^{0}\subseteq\Omega such that: (i)(i) (Ωn)n(\Omega^{n})_{n\in\mathds{N}} converges to Ω0\Omega^{0} in the Hausdorff metric, and (ii)(ii) for any n{0}n\in\mathds{N}\cup\{0\}, Ωn\Omega^{n} has non-empty intersection with every conditioning event cCc\in C. Then, the following two hold:

  • (i)(i)

    For any sequence of conditional probability systems (μn)n(\mu^{n})_{n\in\mathds{N}} with limit μ\mu and such that μnΔC(Ωn)\mu^{n}\in\Delta^{C}(\Omega^{n}) for every nn\in\mathds{N} it holds that μΔC(Ω0)\mu\in\Delta^{C}(\Omega^{0}).

  • (ii)(ii)

    Sequence (ΔC(Ωn))n(\Delta^{C}(\Omega^{n}))_{n\in\mathds{N}} converges to ΔC(Ω0)\Delta^{C}(\Omega^{0}) in the Hausdorff metric.

Proof.

For claim (i)(i) we only need to check that supp μ(c)cΩ0\textup{supp }\mu(c)\subseteq c\cap\Omega^{0} for any cCc\in C.292929That updating according conditional probability is respected in the limit is a well-established fact. Fix arbitrary cCc\in C and pick xsupp μ(c)x\in\textup{supp }\mu(c). Since the support correspondence is lower hemicontinuous,303030See Theorem 17.14 in Aliprantis and Border (2007), p. 563. we know that there exists a subsequence (μnm)m(\mu^{n_{m}})_{m\in\mathds{N}} and a sequence (xm)m(x^{m})_{m\in\mathds{N}} with limit xx such that xmsupp μnm(c)cΩnmx^{m}\in\textup{supp }\mu^{n_{m}}(c)\subseteq c\cap\Omega^{n_{m}} for any mm\in\mathds{N}. Then, pick metric dd that topologizes Ω\Omega, and let dHd_{H} denote its corresponding Hausdorff metric. We know that dH(cΩnm,cΩ)0d_{H}(c\cap\Omega^{n_{m}},c\cap\Omega)\rightarrow 0 and therefore, that d(cΩnm,cΩ0)0d(c\cap\Omega^{n_{m}},c\cap\Omega^{0})\rightarrow 0. Since we have that d(xnm,x)0d(x^{n_{m}},x)\rightarrow 0 and cΩ0c\cap\Omega^{0} is closed, we conclude that xcΩ0x\in c\cap\Omega^{0}.

For claim (ii)(ii), proceed by contradiction and suppose that there exists some ε>0\varepsilon>0 such that for any mm\in\mathds{N} there exists some nmnn_{m}\geq n such that dH(ΔC(Ωnm),ΔC(Ω0))>εd_{H}(\Delta^{C}(\Omega^{n_{m}}),\Delta^{C}(\Omega^{0}))>\varepsilon. Then, for any mm\in\mathds{N} there exists some μnmΔC(Ωnm)\mu^{n_{m}}\in\Delta^{C}(\Omega^{n_{m}}) such that d(μnm,μ)>εd(\mu^{n_{m}},\mu)>\varepsilon for any μΔC(Ω0)\mu\in\Delta^{C}(\Omega^{0}). Now, since ΔC(Ω)\Delta^{C}(\Omega) is compact, we know that there exists a convergent subsequence (μnmr)r(\mu^{n_{m_{r}}})_{r\in\mathds{N}} with limit μ0\mu^{0}. We know from (i)(i) that μ0ΔC(Ω0)\mu^{0}\in\Delta^{C}(\Omega^{0}), and thus, since d(μnmr,μ0)0d(\mu^{n_{m_{r}}},\mu^{0})\rightarrow 0, we reached a contradiction. ∎

Corollary 3.

If sequence of subjective payoff structures (din)n(d_{i}^{n})_{n\in\mathds{N}} converges to subjective payoff structure did_{i}, then, sequence of sets of opponents’ types (Ti(di|in))n(T_{-i}(d_{-i|i}^{n}))_{n\in\mathds{N}} converges to Ti(di|i)T_{-i}(d_{-i|i}) in the Hausdorff metric.

Proof.

We proceed by induction. Let Xi1(di):=Proj(di,1)0×(di,1)i(di,1)X_{i}^{1}(d_{i}^{\prime}):=\textup{Proj}_{(d^{\prime}_{i,1})_{0}\times(d^{\prime}_{i,1})_{-i}}(d^{\prime}_{i,1}) for any subjective payoff structure did_{i}^{\prime}. Obviously, Xi1(din)X_{i}^{1}(d_{i}^{n}) converges to Xi1(di)X_{i}^{1}(d_{i}) in the Hausdorff metric, and thus, we know from Lemma 1 that (Zi1(din))n(Z_{i}^{1}(d_{i}^{n}))_{n\in\mathds{N}}, where Zi1(din):=Δ(Xi1(din))Z_{i}^{1}(d_{i}^{n}):=\Delta(X_{i}^{1}(d_{i}^{n})) for any nn\in\mathds{N}, converges to Zi1(di):=Δ(Xi1(di))Z_{i}^{1}(d_{i}):=\Delta(X_{i}^{1}(d_{i})) in the Hausdorff metric. Now, set inductively Zik1(di)=Δ(Xik1(di))Z_{i}^{k-1}(d^{\prime}_{i})=\Delta(X_{i}^{k-1}(d^{\prime}_{i})) and Xik(di′′):=Xik1(di′′)×jiZjk1(dj|i)X_{i}^{k}(d^{\prime\prime}_{i}):=X_{i}^{k-1}(d^{\prime\prime}_{i})\times\prod_{j\neq i}Z_{j}^{k-1}(d^{\prime}_{j|i}) for every k0k\geq 0 and every did^{\prime}_{i} and di′′d^{\prime\prime}_{i}, and suppose that k1k\geq 1 is such that Xik(din)X_{i}^{k}(d_{i}^{n}) converges to Xik(di)X_{i}^{k}(d_{i}) and (Zjk(dj|in))n(Z_{j}^{k}(d_{j|i}^{n}))_{n\in\mathds{N}} converges to Zjk(dj|i)Z_{j}^{k}(d_{j|i}) for any jij\neq i (all of them in the Hausdorff metric). Then, obviously, Xik+1(din)X_{i}^{k+1}(d_{i}^{n}), defined in the obvious way, converges to Xik+1(di)X_{i}^{k+1}(d_{i}) in the Hausdorff metric, and therefore, we know from Lemma 1 that (Zik+1(din))n(Z_{i}^{k+1}(d_{i}^{n}))_{n\in\mathds{N}} converges to Zik+1(di)Z_{i}^{k+1}(d_{i}). Obviously, it follows that (kjiZjk(dj|in))n(\prod_{k\in\mathds{N}}\prod_{j\neq i}Z_{j}^{k}(d_{j|i}^{n}))_{n\in\mathds{N}} converges to kjiZjk(dj|i)\prod_{k\in\mathds{N}}\prod_{j\neq i}Z_{j}^{k}(d_{j|i}) and hence, that (Ti(di|in))n(T_{-i}(d_{-i|i}^{n}))_{n\in\mathds{N}} converges to Ti(di|i)T_{-i}(d_{-i|i}). ∎

Corollary 4.

The set of player ii’s subjective models is closed; i.e., if pair (di,ti)𝒫i×Ti(d_{i},t_{i})\in\mathscr{P}_{i}^{\infty}\times T_{i} is the limit of a sequence of subjective models (din,tin)n(d_{i}^{n},t_{i}^{n})_{n\in\mathds{N}}, then (di,ti)(d_{i},t_{i}) is a subjective model too.

Proof.

Just notice that we know from the previous corollary that ((di,1n)0×Ti(di|in))n((d_{i,1}^{n})_{0}\times T_{-i}(d_{-i|i}^{n}))_{n\in\mathds{N}} converges to (di,1)0×Ti(di|i)(d_{i,1})_{0}\times T_{-i}(d_{-i|i}), and thus, we know from Lemma 1 that, if τinΔ((di,1n)0×Ti(di|in))\tau_{i}^{n}\in\Delta((d_{i,1}^{n})_{0}\times T_{-i}(d_{-i|i}^{n})) for every nn\in\mathds{N}, and (τin)n(\tau_{i}^{n})_{n\in\mathds{N}} converges to τi\tau_{i}, then τiΔ((di,1)0×Ti(di|i))\tau_{i}\in\Delta((d_{i,1})_{0}\times T_{-i}(d_{-i|i})). ∎

Proof of the proposition

See 3

Proof.

We will proceed by induction and verify that the correspondence Bi,k:iSi\textup{B}_{i,k}:\mathscr{M}_{i}^{\infty}\rightrightarrows S_{i} is upper hemicontinuous for every k0k\geq 0. The claim holds trivially for the initial case so we can focus on the proof of the inductive step. Fix k0k\geq 0 such that the claim holds; we verify next that it also does for k+1k+1. Fix player ii, convergent sequence of subjective models (Min)n(M_{i}^{n})_{n\in\mathds{N}} with limit MiM_{i} and strategy sis_{i} such that siBi(Min)s_{i}\in\textup{B}_{i}(M_{i}^{n}) for every nn\in\mathds{N}. For each nn\in\mathds{N}, pick conjecture μin\mu_{i}^{n} that justifies the inclusion of sis_{i} in Fi,k+1(Min)\textup{F}_{i,k+1}(M_{i}^{n}). Let (μinm)m(\mu_{i}^{n_{m}})_{m\in\mathds{N}} be a convergent subsequence of (μin)n(\mu_{i}^{n})_{n\in\mathds{N}} with limit μi\mu_{i}. For each nn\in\mathds{N} let dind_{i}^{n} denote the subjective payoff structure associated to subjective model MinM_{i}^{n} and did_{i}, the one associated to MiM_{i}. Notice then that for any k\ell\leq k and any history hh we have that:

Xi,n(h):={(si,ti)Si×Ti(di|i)|[si]hBi,(di|in,ti)h0}\displaystyle X_{i,\ell}^{n}(h):=\left\{(s_{-i},t_{-i})\in S_{-i}\times T_{-i}(d_{-i|i})\left|[s_{-i}]_{h}\cap\textup{B}_{-i,\ell}(d_{-i|i}^{n},t_{-i})\neq h^{0}\right.\right\} =\displaystyle=
siSi([si]h×Ti(di|i))Graph\displaystyle\bigcup_{s_{-i}\in S_{-i}}\left([s_{-i}]_{h}\times T_{-i}(d_{-i|i})\right)\cap\textup{Graph} (Bi,(di|in,)),\displaystyle\left(\textup{B}_{-i,\ell}(d_{-i|i}^{n},\,\cdot\,)\right),

and thus, we know by finiteness of SiS_{-i} and the induction hypothesis that Xi,(h)X_{i,\ell}(h) is closed. Then, it holds that:

m,μinm(h)[(di,1nm)0×Xi,nm(h)]=1\displaystyle\forall m\in\mathds{N},\,\mu_{i}^{n_{m}}(h)\left[(d_{i,1}^{n_{m}})_{0}\times X_{i,\ell}^{n_{m}}(h)\right]=1 m,μinm(h)[rnm(di,1r)0×Xi,r(h)]=1\displaystyle\Rightarrow\forall m\in\mathds{N},\,\mu_{i}^{n_{m}}(h)\left[\bigcup_{r\geq n_{m}}(d_{i,1}^{r})_{0}\times X_{i,\ell}^{r}(h)\right]=1
m,μi(h)[cl(rnm(di,1nm)0×Xi,r(h))]=1\displaystyle\Rightarrow\forall m\in\mathds{N},\,\mu_{i}(h)\left[\textup{cl}\left(\bigcup_{r\geq n_{m}}(d_{i,1}^{n_{m}})_{0}\times X_{i,\ell}^{r}(h)\right)\right]=1
μi(h)[mcl(rnm(di,1r)0×Xi,r(h))]=1\displaystyle\Rightarrow\mu_{i}(h)\left[\bigcap_{m\in\mathds{N}}\textup{cl}\left(\bigcup_{r\geq n_{m}}(d_{i,1}^{r})_{0}\times X_{i,\ell}^{r}(h)\right)\right]=1
μi(h)[(di,1)0×Xi,(h)]=1,\displaystyle\Rightarrow\mu_{i}(h)\left[(d_{i,1})_{0}\times X_{i,\ell}(h)\right]=1,

where:

Xi,(h):={(si,ti)Si×Ti(di|i)|[si]hBi,(di|i,ti)h0}.X_{i,\ell}(h):=\left\{(s_{-i},t_{-i})\in S_{-i}\times T_{-i}(d_{-i|i})\left|[s_{-i}]_{h}\cap\textup{B}_{-i,\ell}(d_{-i|i},t_{-i})\neq h^{0}\right.\right\}.

Notice that the last inclusion holds by virtue of Xi,n(h)X_{i,\ell}^{n}(h) being upper hemicontinuous on nn because of the induction hypothesis. The above, together with the continuity of marginalization and the upper hemicontinuity of the best response correspondence allows for concluding that μi\mu_{i} is a conjecture that justifies the inclusion of sis_{i} in Bi,k+1(Mi)\textup{B}_{i,k+1}(M_{i}). ∎

Appendix B Proofs II: Theorem

Additional notation

Distinguished histories

The histories of player ii that might be reached when every opponent jJI{i}j\in J\subseteq I\setminus\{i\} plays according to some given correspondence Wj(dj,):Tj(dj)Sj\textup{W}_{j}(d_{j},\,\cdot\,):T_{j}(d_{j})\rightrightarrows S_{j} is of special interest and formalized as follows:

Hi(WJ,dJ):={hHi{h0}|jJ(Sj(h)×Tj(dj)Graph(Wj(dj,)))h0}.\textup{H}_{i}(\textup{W}_{J},d_{J}):=\left\{h\in H_{i}\cup\{h^{0}\}\left|\prod_{j\in J}\left(S_{j}(h)\times T_{j}(d_{j})\cap\textup{Graph}(\textup{W}_{j}(d_{j},\,\cdot\,)\right))\neq h^{0}\right.\right\}.

Another kind of histories that play an important role throughout the proof are those in which the player updates beliefs from scratch. For each player ii and conjecture μi\mu_{i}, these are the histories that are considered unlikely to be reached by every history preceding them:

Hi(μi):={hHi{h0}|margSiμi(h)[Si(h)]=0 for every hh}.\textup{H}_{i}(\mu_{i}):=\left\{h\in H_{i}\cup\{h^{0}\}\left|\textup{marg}_{S_{-i}}\mu_{i}(h^{\prime})\left[S_{-i}(h)\right]=0\textup{ for every }h^{\prime}\prec h\right.\right\}.

Notice that every conjecture μi\mu_{i} is fully described by the beliefs μi(h)\mu_{i}(h) where hHi(μi)h\in\textup{H}_{i}(\mu_{i}): The belief that corresponds to any other history is obtained via the chain rule.

Strict rationalizability

First, for each player ii and strategy sis_{i} let [si][s_{i}] denote the set of player ii’s strategies that are outcome-equivalent to sis_{i} –those strategies sis_{i}^{\prime} such that si(h)=si(h)s_{i}^{\prime}(h)=s_{i}(h) for every history hHi(si)h\in H_{i}(s_{i}). Then, player ii’s set of (extensive form) strictly rationalizable strategies for subjective model (di,ti)(d_{i},t_{i}) is Fi0(di,ti):=k0Fi,k0(di,ti)\textup{F}_{i}^{0}(d_{i},t_{i}):=\bigcap_{k\geq 0}\textup{F}_{i,k}^{0}(d_{i},t_{i}) where Fi,00(di,ti)=Si\textup{F}_{i,0}^{0}(d_{i},t_{i})=S_{i} and for each k0k\geq 0,

Fi,k+10(di,ti)\displaystyle\textup{F}_{i,k+1}^{0}(d_{i},t_{i}) :={siSi|μiCi,k0(di,ti) s.t. ri(θi(ti),μi)=[si]},\displaystyle:=\left\{s_{i}\in S_{i}\left|\exists\mu_{i}\in\textup{C}_{i,k}^{0}(d_{i},t_{i})\textup{ s.t. }r_{i}(\theta_{i}(t_{i}),\mu_{i})=[s_{i}]\right.\right\},
Ci,k+10(di,ti)\displaystyle\textup{C}_{i,k+1}^{0}(d_{i},t_{i}) :={μiCi,k0(di,ti)|hHi(Fi,k0,di|i),μi(h)[(di,1)0×M]=1 for some MGraph(Fi,k0(di|i,))}.\displaystyle:=\left\{\mu_{i}\in\textup{C}_{i,k}^{0}(d_{i},t_{i})\left|\begin{tabular}[]{l}$\textup{$\forall h\in\textup{H}_{i}(\textup{F}_{-i,k}^{0},d_{-i|i})$,}$\\[8.61108pt] $\mu_{i}(h)[(d_{i,1})_{0}\times M]=1\textup{ for some }M\subseteq\textup{Graph}\left(\textup{F}_{-i,k}^{0}(d_{-i|i},\,\cdot\,)\right)$\end{tabular}\right.\right\}.

Related to this, we also add the following two notational shorthands, convenient for lemmas 3 and 5 in the proof of the theorem:

  • For each subjective payoff structure did_{i} and k0k\geq 0 we denote by [si|di]k[s_{i}|d_{i}]_{k} the set of player ii’s strategies that are equivalent to strategy sis_{i} at every history consistent with the opponents’ playing some profile of strategies in Fi,k0(di|i,)\textup{F}_{-i,k}^{0}(d_{-i|i},\,\cdot\,), that is:

    [si|di]k={siSi|si(h)=si(h) for every hHi(Fi,k0,di|i)}.[s_{i}|d_{i}]_{k}=\left\{s_{i}^{\prime}\in S_{i}\left|s_{i}^{\prime}(h)=s_{i}(h)\textup{ for every }h\in\textup{H}_{i}(\textup{F}_{-i,k}^{0},d_{-i|i})\right.\right\}.
  • The types consistent with subjective payoff structure did_{i} that admit strategy sis_{i} as Fi,k0(di,)\textup{F}_{i,k}^{0}(d_{i},\,\cdot\,) via some conjecture that assigns, initially, positive probability to every history reachable by some profile of opponents’ strategies in Fi,k10(di|i,)\textup{F}_{-i,k-1}^{0}(d_{-i|i},\,\cdot\,) is formalized as:

    (i)μi justifies the inclusion of si in Fi,k0(di,ti),(ii)hHi(Fi,k10,di|i),margSiμi(h0)[Si(h)]>0}.T_{i,k}(d_{i},s_{i}):=\left\{t_{i}\in T_{i}(d_{i})\left|\begin{tabular}[]{r l}\lx@intercol$\exists\mu_{i}\in\Delta^{H_{i}\cup\{h^{0}\}}(S_{-i}\times(d_{i,1})_{0}\times T_{-i}(d_{-i|i}))${ such that:}\hfil\lx@intercol\\[8.61108pt] $(i)$&$\mu_{i}\textup{ justifies the inclusion of $s_{i}$ in }\textup{F}_{i,k}^{0}(d_{i},t_{i})$,\\[6.45831pt] $(ii)$&$\forall h\in\textup{H}_{i}(\textup{F}_{-i,k-1}^{0},d_{-i|i}),\,\textup{marg}_{S_{-i}}\mu_{i}(h^{0})[S_{-i}(h)]>0$\end{tabular}\right.\right\}.
    Ti,k(di,si):={tiTi(di)| μiΔHi{h0}(×S-i(di,1)0T-i(d-|ii)) such that:

Proof of the theorem

See 1

Proof.

Fix standard payoff structure Υ\Upsilon, profile of finite types tT(Θ)t\in T(\Theta) and strategy profile sF(Υ,t)s\in\textup{F}(\Upsilon,t). Then, we know from lemmas 2 and 3 below that for each kk\in\mathds{N}:

  • (A)

    There exist some mkm_{k}\in\mathds{N} and a sequence of standard models (Υm,tk,m)mmk(\Upsilon^{m},t^{k,m})_{m\geq m_{k}} converging to (Υ,t)(\Upsilon,t) such that, for any mmkm\geq m_{k} and any player ii, tik,mTi(Υm)t_{i}^{k,m}\in T_{i}(\Upsilon^{m}) is finite, siFi,k0(Υm,tik,m)s_{i}\in\textup{F}_{i,k}^{0}(\Upsilon^{m},t_{i}^{k,m}) and for every jij\neq i,

    Hj(Fi,k0,Υm)=Hj(Fi,k,Υm)=Hj(Fi,k,Υ).\textup{H}_{j}(\textup{F}_{i,k}^{0},\Upsilon^{m})=\textup{H}_{j}(\textup{F}_{i,k},\Upsilon^{m})=\textup{H}_{j}(\textup{F}_{i,k},\Upsilon).
  • (B)

    For any mm\in\mathds{N} there exists a sequence of profiles of finite types (tk,m,)(t^{k,m,\ell})_{\ell\in\mathds{N}} converging to tk,mt^{k,m} such that for each player ii, tik,m,Ti,k(Υm,si)t_{i}^{k,m,\ell}\in T_{i,k}(\Upsilon^{m},s_{i}) (as defined in Section B.1.2) for every \ell\in\mathds{N}.

Now, for each player ii and each nn\in\mathds{N} set t~in:=timn,mn,n\tilde{t}_{i}^{n}:=t_{i}^{m_{n},m_{n},n}, and notice that:

  • t~in\tilde{t}_{i}^{n} is finite and included in Ti,mn(Υmn,si)T_{i,m_{n}}(\Upsilon^{m_{n}},s_{i}).

  • Hi(Fi,mn0,Υmn)=Hi(Fi,mn,Υmn)=Hi(Fi,mn,Υ)\textup{H}_{i}(\textup{F}_{-i,m_{n}}^{0},\Upsilon^{m_{n}})=\textup{H}_{i}(\textup{F}_{-i,m_{n}},\Upsilon^{m_{n}})=\textup{H}_{i}(\textup{F}_{-i,m_{n}},\Upsilon).

We know then from Lemma 5 below that:

  • (C)

    For any ii there exists some subjective model (din,tin)(d_{i}^{n},t_{i}^{n}) such that: di,mnn=di,mnd_{i,m_{n}}^{n}=d_{i,m_{n}}^{\prime} for di=Υmnd_{i}=\Upsilon^{m_{n}}, πi,mn(tin)=πi,mn(t~in)\pi_{i,m_{n}}(t_{i}^{n})=\pi_{i,m_{n}}(\tilde{t}_{i}^{n}) and as (for [si|Υmn]mn1[s_{i}|\Upsilon^{m_{n}}]_{m_{n}-1} defined in Section B.1.2),

    Fi,mn+1(din,tin)[si|Υmn]mn1,\textup{F}_{i,m_{n}+1}(d_{i}^{n},t_{i}^{n})\subseteq[s_{i}|\Upsilon^{m_{n}}]_{m_{n}-1},

    and hence,

    Fi(din,tin)[si|Υmn]mn1.\textup{F}_{i}(d_{i}^{n},t_{i}^{n})\subseteq[s_{i}|\Upsilon^{m_{n}}]_{m_{n}-1}.

It follows from (A) that:

[si|Υmn]mn1\displaystyle[s_{i}|\Upsilon^{m_{n}}]_{m_{n}-1} ={siSi|si(h)=si(h) for any hHi(Fi,mn10,Υmn)}\displaystyle=\left\{s_{i}^{\prime}\in S_{i}\left|s_{i}^{\prime}(h)=s_{i}(h)\textup{ for any }h\in\textup{H}_{i}(\textup{F}_{-i,m_{n}-1}^{0},\Upsilon^{m_{n}})\right.\right\}
={siSi|si(h)=si(h) for any hHi(Fi,mn1,Υ)}\displaystyle=\left\{s_{i}^{\prime}\in S_{i}\left|s_{i}^{\prime}(h)=s_{i}(h)\textup{ for any }h\in\textup{H}_{i}(\textup{F}_{-i,m_{n}-1},\Upsilon)\right.\right\}
{siSi|si(h)=si(h) for any hHi(Fi,Υ)}.\displaystyle\subseteq\left\{s_{i}^{\prime}\in S_{i}\left|s_{i}^{\prime}(h)=s_{i}(h)\textup{ for any }h\in\textup{H}_{i}(\textup{F}_{-i},\Upsilon)\right.\right\}.

Thus, (dn,tn)n(d^{n},t^{n})_{n\in\mathds{N}} is a sequence of profiles of subjective models converging to (Υ,t)(\Upsilon,t) such that for any player ii and nn\in\mathds{N},

Fi(din,tin){siSi|si(h)=si(h) for any hHi(Fi,Υ)}.\textup{F}_{i}(d_{i}^{n},t_{i}^{n})\subseteq\left\{s_{i}^{\prime}\in S_{i}\left|s_{i}^{\prime}(h)=s_{i}(h)\textup{ for any }h\in\textup{H}_{i}(\textup{F}_{-i},\Upsilon)\right.\right\}.

Now, fix nn\in\mathds{N} and snF(dn,tn)s^{n}\in\textup{F}(d^{n},t^{n}). Let JIJ\subseteq I be the set of those players ii for which h0Hih^{0}\in H_{i}. Clearly, h0Hi(F,Υ)h^{0}\in\textup{H}_{i}(\textup{F},\Upsilon) and thus, it must hold that sin(h0)=si(h0)s_{i}^{n}(h^{0})=s_{i}(h^{0}). Then, obviously, for every player jj such that (h0,(sin(h0))iJ)Hj(h^{0},(s_{i}^{n}(h^{0}))_{i\in J})\in H_{j}, it holds that (h0,(sin(h0))iJ)Hj(F,Υ)(h^{0},(s_{i}^{n}(h^{0}))_{i\in J})\in\textup{H}_{j}(\textup{F},\Upsilon) as well. Hence, an easy inductive argument enables to conclude that z(sn|h0)=z(s|h0)z(s^{n}|h^{0})=z(s|h^{0}). ∎

Proof of Lemma 2

First perturbation of payoff structures

Let Υ\Upsilon be a standard payoff structure. We first construct a perturbation (Υn)n(\Upsilon^{n})_{n\in\mathds{N}} that will ensure that any strategy of any player that is extensive form rationalizable given Υ\Upsilon is also strictly rationalizable along the tail of the perturbation. We proceed in three steps:

  1. 1.

    Set Θ0n:=Θ0\Theta_{0}^{n}:=\Theta_{0}.

  2. 2.

    Fix nn\in\mathds{N} and define the following set for each player ii:

    Θin:={θiqi(si,θi)|siSi,θiΘi and qi([n,)){}},\Theta_{i}^{n}:=\left\{\theta_{i}^{q_{i}}(s_{i},\theta_{i})\left|s_{i}\in S_{i},\theta_{i}\in\Theta_{i}\textup{ and }q_{i}\in([n,\infty)\cap\mathds{Q})\cup\{\infty\}\right.\right\},
  3. 3.

    For each player ii, define utility function:

    uin(z,(θ0,(θjqj(sj,θj))jI)):={ui(z,θ)+1qiifqi andifz=z(si;si) for some siSi,ui(z,θ)otherwise,u_{i}^{n}(z,(\theta_{0},(\theta_{j}^{q_{j}}(s_{j},\theta_{j}))_{j\in I})):=\left\{\begin{tabular}[]{l l}$u_{i}(z,\theta)+\frac{1}{q_{i}}$&$\textup{if}\,\,\,q_{i}\neq\infty\textup{ and}$\\[2.15277pt] &$\textup{\color[rgb]{1,1,1}\definecolor[named]{pgfstrokecolor}{rgb}{1,1,1}\pgfsys@color@gray@stroke{1}\pgfsys@color@gray@fill{1}if}\,\,\,z=z(s^{\prime}_{-i};s_{i})\textup{ for some }s^{\prime}_{-i}\in S_{-i},$\\[4.30554pt] $u_{i}(z,\theta)$&$\textup{otherwise}$,\end{tabular}\right.

    for each terminal history zz and each (θ0,(θjqj(sj,θ¯j))jI)Θ0n×iIΘin(\theta_{0},(\theta_{j}^{q_{j}}(s_{j},\bar{\theta}_{j}))_{j\in I})\in\Theta_{0}^{n}\times\prod_{i\in I}\Theta_{i}^{n}.

Set Υ:=(Θ0n,(Θin,uin)iI)\Upsilon:=(\Theta_{0}^{n},(\Theta_{i}^{n},u_{i}^{n})_{i\in I}). Clearly, (Υn)n(\Upsilon^{n})_{n\in\mathds{N}} is a well-defined sequence of standard payoff structures converging to Υ\Upsilon. Now, let us make few comments to better motivate the need of this construction:

  • (A)

    For every player ii, every strategy sis_{i}, every payoff type θiΘi\theta_{i}\in\Theta_{i}, and every qiq_{i}\neq\infty, there exists a payoff type θiqi(si,θi)Θin\theta_{i}^{q_{i}}(s_{i},\theta_{i})\in\Theta_{i}^{n} that mimics the payoffs of θi\theta_{i} for player ii except for a slight improvement in the payoffs corresponding to the terminal nodes consistent with sis_{i} (thus breaking all possible ties in expected utility in favor of sis_{i}). This payoff types are the key elements we use in Lemma 2 below to break ties and move from extensive form rationalizability to strict rationalizability.

  • (B)

    With some abuse of terminology it can be considered that Θ\Theta is a subset of Θn\Theta^{n}. That the states of nature of the former are contained in the latter is immediate from the construction. Player ii’s payoff types in Θi\Theta_{i} are not per se contained in those in Θin\Theta_{i}^{n}; however, there is a natural identification between payoff types θi\theta_{i} and θi(,)\theta_{i}^{\infty}(\,\cdot\,,\,\cdot\,). For our proof of Lemma 7 below we will, with some abuse of notation, leverage on this fact.

From rationalizability to strict rationalizability

The proof of the following lemma (and also of Lemma 3) will depend on the concept of finite conjecture. Given standard payoff structure Υ\Upsilon we say that conjecture μi\mu_{i} is finite if for every history hHi{h0}h\in H_{i}\cup\{h^{0}\} it holds that the belief hierarchy on Θ\Theta induced by μi(h)\mu_{i}(h), πi(h):=τi(margΘ0×Ti(Θ)μi(h))\pi_{i}(h):=\tau_{i}(\textup{marg}_{\Theta_{0}\times T_{-i}(\Theta)}\mu_{i}(h)), is that of a finite type. Then, we have that:

Lemma 2.

Let Γ\Gamma be an extensive form and Υ\Upsilon, a standard payoff structure. Then, for any player ii and any k0k\geq 0 the following two hold:

  • (i)(i)

    For any finite type tiTi(Θ)t_{i}\in T_{i}(\Theta), and any strategy siFi,k(Υ,ti)s_{i}\in\textup{F}_{i,k}(\Upsilon,t_{i}) there exists a nki,1n_{k}^{i,1}\in\mathds{N} and a sequence of finite types (tik,n)nnki,1(t_{i}^{k,n})_{n\geq n_{k}^{i,1}} converging to tit_{i} such that tinTi(Θn)t_{i}^{n}\in T_{i}(\Theta^{n}) and siFi,k0(Υn,tin)s_{i}\in\textup{F}_{i,k}^{0}(\Upsilon^{n},t_{i}^{n}) for any nnki,1n\geq n_{k}^{i,1}.

  • (ii)(ii)

    There exists some nki,2n_{k}^{i,2}\in\mathds{N} such that, for any nnki,2n\geq n_{k}^{i,2}, Fi,k0(Υn,ti)Fi,k(Υn,ti)\textup{F}_{i,k}^{0}(\Upsilon^{n},t_{i})\subseteq\textup{F}_{i,k}(\Upsilon^{n},t_{i}) for every type tiTi(Θn)t_{i}\in T_{i}(\Theta^{n}) and Hj(Fi,k0,Υn)=Hj(Fi,k,Υn)\textup{H}_{j}(\textup{F}_{i,k}^{0},\Upsilon^{n})=\textup{H}_{j}(\textup{F}_{i,k},\Upsilon^{n}) for every jij\neq i.

Proof.

We proceed by induction on kk. The initial case (k=0)(k=0) hold trivially for both claims, so we can focus on the proofs of the inductive step. Suppose that k0k\geq 0 is such that the claims hold; we verify next that they also hold for k+1k+1. We prove each claim separately:

Claim (i)(i). Fix player ii, finite type tit_{i} strategy siFi,k+1(Υ,ti)s_{i}\in\textup{F}_{i,k+1}(\Upsilon,t_{i}) and finite conjecture μi\mu_{i} that justifies the inclusion of sis_{i} in Fi,k+1(Υ,ti)\textup{F}_{i,k+1}(\Upsilon,t_{i}) (whose existence is guaranteed by Lemma 6 below). Next we will base on each μi\mu_{i} to construct a conditional probability system for Si×Θ0n×Ti(Θn)S_{-i}\times\Theta_{0}^{n}\times T_{-i}(\Theta^{n}), μin\mu_{i}^{n}, that will ensure that sis_{i} is in Fi,k+10(Υn,tik+1,n)\textup{F}_{i,k+1}^{0}(\Upsilon^{n},t_{i}^{k+1,n}) for some finite type tik+1,nTi(Θn)t_{i}^{k+1,n}\in T_{i}(\Theta^{n}). To this end set first Xi1:=X_{i}^{-1}:=\emptyset and next, for each =0,,k1\ell=0,\dots,k-1, Xi:=Hi(Fi,k,Υ)X_{i}^{\ell}:=\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon). Now, fix =0,,k1\ell=0,\dots,k-1 and history hXiXi1h\in X_{i}^{\ell}\setminus X_{i}^{\ell-1}. We know from the induction hypothesis that (a)(a) there exists some nki,2n_{k-\ell}^{i,2}\in\mathds{N} such that Hi(Fi,k,Υn)=Hi(Fi,k0,Υn)\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon^{n})=\textup{H}_{i}(\textup{F}_{-i,k-\ell}^{0},\Upsilon^{n}) for every nnki,2n\geq n_{k-\ell}^{i,2} and that (b)(b) for any pair (si,ti)(s_{-i},t_{-i}) in the support of the marginal of μi(h)\mu_{i}(h) on Si×Ti(Θ)S_{-i}\times T_{-i}(\Theta) there exists a sequence of finite types (tik,n(si,ti))nnki,1(si,ti)(t_{-i}^{k-\ell,n}(s_{-i},t_{-i}))_{n\geq n_{k-\ell}^{i,1}(s_{-i},t_{-i})} converging to tit_{-i} and satisfying the conditions in claim (i)(i) for kk-\ell. Then, set (remember that the support of μ(h)\mu(h) is finite):

nhi,1:=max{max{ni,1(si,ti)|(si,ti)supp μ(h)},ni,2}n_{h}^{i,1}:=\textup{max}\left\{\textup{max}\left\{\left.n_{\ell}^{i,1}(s_{-i},t_{-i})\right|(s_{-i},t_{-i})\in\textup{supp }\mu(h)\right\},n_{\ell}^{i,2}\right\}

and define, for each nnhi,1n\geq n_{h}^{i,1},

μin(h)[E]:=μi(h)[{(si,θ0,ti)Si×Θ0×Ti(Θ)|(si,θ0,tik,n(si,ti))E}].\displaystyle\mu_{i}^{n}(h)[E]:=\mu_{i}(h)\left[\left\{(s_{-i},\theta_{0},t_{-i})\in S_{-i}\times\Theta_{0}\times T_{-i}(\Theta)\left|\begin{tabular}[]{r l}\lx@intercol$(s_{-i},\theta_{0},t_{-i}^{k-\ell,n}(s_{-i},t_{-i}))\in E$\hfil\lx@intercol \end{tabular}\right.\right\}\right].

Finiteness of μi\mu_{i} guarantees that μi(h)\mu_{i}(h) is a well-defined probability measure on Si×Θ0n×Ti(Θn)S_{-i}\times\Theta_{0}^{n}\times T_{-i}(\Theta^{n}). Set then nk+1i,1:=max{nhi,1|hXiXi1,=0,,k1}n_{k+1}^{i,1}:=\textup{max}\{n_{h}^{i,1}|h\in X_{i}^{\ell}\setminus X_{i}^{\ell-1},\ell=0,\dots,k-1\}.

Notice now that the following three properties hold:313131Remember that Hi(μi)\textup{H}_{i}(\mu_{i}) was defined in SectionB.1.1 as the set of histories in which conjecture μi\mu_{i} updates beliefs from scratch.

  • (1)

    For every pair of histories hHi(μi)Hi(Fi,k,Υ)h\in\textup{H}_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon) and hHih^{\prime}\in H_{i}, if the marginal on SiS_{-i} of μi(h)\mu_{i}(h) assigns positive probability to Si(h)S_{-i}(h^{\prime}), then hHi(Fi,k0,Υn)h^{\prime}\in\textup{H}_{i}(\textup{F}_{-i,k-\ell}^{0},\Upsilon^{n}). To see this fix history hHi(μi)Hi(Fi,k,Υ)h\in\textup{H}_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon) and notice first that:

    μin(h)\displaystyle\mu_{i}^{n}(h) [Θ0n×Graph(Fi,k0(Υn,))]=\displaystyle\left[\Theta_{0}^{n}\times\textup{Graph}\left(\textup{F}_{-i,k-\ell}^{0}(\Upsilon^{n},\,\cdot\,)\right)\right]=
    =μi(h)[Θ0×{(si,ti)Si×Ti(Θ)|siFi,k0(Υn,tin(si,ti))}]\displaystyle=\mu_{i}(h)\left[\Theta_{0}\times\left\{(s_{-i},t_{-i})\in S_{-i}\times T_{-i}(\Theta)\left|s_{-i}\in\textup{F}_{-i,k-\ell}^{0}(\Upsilon^{n},t_{i}^{n}(s_{-i},t_{-i}))\right\}\right.\right]
    =μi(h)[Θ0×Graph(Fi,k(Υ,))]=1.\displaystyle=\mu_{i}(h)\left[\Theta_{0}\times\textup{Graph}\left(\textup{F}_{-i,k-\ell}(\Upsilon,\,\cdot\,)\right)\right]=1.

    Thus, for every history hHih^{\prime}\in H_{i} such that the marginal of μin(h)\mu_{i}^{n}(h) on SiS_{-i} assigns positive probability to Si(h)S_{-i}(h^{\prime}) we necessarily have that:

    hHi(Fi,k,Υn)h\in\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon^{n})

    and thus the claim must hold.

  • (2)

    For every history hHi(Fi,k,Υ)h^{\prime}\in\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon) there exists some hHi(μi)Hi(Fi,k,Υ)h\in\textup{H}_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon) such that the marginal on SiS_{-i} of μin(h)\mu_{i}^{n}(h) assigns positive probability to Si(h)S_{-i}(h^{\prime}). To see this, fix history hHi(Fi,k,Υ)h^{\prime}\in\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon). If hHi(μi)h^{\prime}\in\textup{H}_{i}(\mu_{i}) then there is nothing needed to be proved. If hHi(μi)h^{\prime}\notin\textup{H}_{i}(\mu_{i}) then we know that there exists some hHi(μi)h\in\textup{H}_{i}(\mu_{i}) such that hhh\prec h^{\prime} and the marginal of μi(h)\mu_{i}(h) on SiS_{-i} puts positive probability on Si(h)S_{-i}(h^{\prime}). Obviously, since hhh\prec h^{\prime} and hHi(Fi,k,Υ)h^{\prime}\in\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon) we conclude that, indeed, hHi(μi)Hi(Fi,k,Υ)h\in\textup{H}_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon). The fact that the marginals on SiS_{-i} of μi(h)\mu_{i}(h) and μin(h)\mu_{i}^{n}(h) coincide ensures that the claim is correct.

  • (3)

    For every pair of histories hHi(μi)Hi(Fi,k,Υ)h\in\textup{H}_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon) and hHi(Fi,k,Υ)h^{\prime}\notin\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon) the marginal of μi(h)\mu_{i}(h) on assigns zero probability to Si(h)S_{-i}(h^{\prime}). To see it remember that we know from the induction hypothesis that Hi(Fi,k,Υ)Hi(Fi,k0,Υn)\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon)\subseteq\textup{H}_{i}(\textup{F}_{-i,k-\ell}^{0},\Upsilon^{n}). Then, it follows from property (1) above that for every hHi(Fi,k,Υ)h\in\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon), the marginal of μin(h)\mu_{i}^{n}(h) on SiS_{-i} assigns zero probability to Si(h)S_{-i}(h^{\prime}) for every history hHi(Fi,k,Υ)h^{\prime}\notin\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon).

We already defined μin(h)\mu_{i}^{n}(h) for histories in hHi(μi)Hi(Fi,k,Υ)h\in\textup{H}_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon). Properties (2)(2) and (3)(3) above allows for defining μin(h)\mu_{i}^{n}(h) exclusively for every history hHi(Fi,k,Υ)Hi(Fi,+1,Υ)h\in\textup{H}_{i}(\textup{F}_{-i,k-\ell},\Upsilon)\setminus\textup{H}_{i}(\textup{F}_{-i,\ell+1},\Upsilon) using conditional probability. Clearly, μin\mu_{i}^{n} is a well-defined conjecture. Set now type tin:=(θin(si,θi(ti)),πin)t_{i}^{n}:=(\theta_{i}^{n}(s_{i},\theta_{i}(t_{i})),\pi_{i}^{n}) (notice the use of comment (A) in Section B.3.1), where:

πin:=τi1(margΘ0n×Ti(Θn)μin(h0)).\pi_{i}^{n}:=\tau_{i}^{-1}\left(\textup{marg}_{\Theta_{0}^{n}\times T_{-i}(\Theta^{n})}\mu_{i}^{n}(h^{0})\right).

Then, we have that tint_{i}^{n} is a finite element of Ti(Θn)T_{i}(\Theta^{n}), and sequence (tin)nnk+1i,1(t_{i}^{n})_{n\geq n_{k+1}^{i,1}} converges to tit_{i}. Notice that, by construction, μin\mu_{i}^{n} is an element of Ci,k0(Υn,tin)\textup{C}_{i,k}^{0}(\Upsilon^{n},t_{i}^{n}), the induction hypotheses guarantee that at each history hHi(Fi,k0,Υn)h\in\textup{H}_{i}(\textup{F}_{-i,k}^{0},\Upsilon^{n}) belief μin(h)\mu_{i}^{n}(h) assigns probability 1 to the graph of Fi,k0(Υn,)\textup{F}_{-i,k}^{0}(\Upsilon^{n},\,\cdot\,) and that, for each =1,k1\ell=1,\dots k-1, at each history hHi(Fi,k0,Υn)Hi(Fi,k+10,Υn)h\in\textup{H}_{i}(\textup{F}_{-i,k-\ell}^{0},\Upsilon^{n})\setminus\textup{H}_{i}(\textup{F}_{-i,k-\ell+1}^{0},\Upsilon^{n}) belief μin(h)\mu_{i}^{n}(h) assigns probability 1 to the graph of Fi,k0(Υn,)\textup{F}_{-i,k-\ell}^{0}(\Upsilon^{n},\,\cdot\,). In addition, it is evident that ri(θi(tin),μin)=[si]r_{i}(\theta_{i}(t_{i}^{n}),\mu_{i}^{n})=[s_{i}]. Hence, we conclude that siFi,k+10(Υn,tin)s_{i}\in\textup{F}_{i,k+1}^{0}(\Upsilon^{n},t_{i}^{n}). \bigstar

Claim (ii)(ii). Fix player ii and set n¯k+1i,1:=max{nkj,2|=1,,k and ji}\bar{n}_{k+1}^{i,1}:=\textup{max}\{n_{k}^{j,2}|\ell=1,\dots,k\textup{ and }j\neq i\} where each nkj,2n_{k}^{j,2} verifies part (ii)(ii) of the induction hypothesis for player jij\neq i. Then, pick type tit_{i}, strategy siFi,k+10(Υn,ti)s_{i}\in\textup{F}_{i,k+1}^{0}(\Upsilon^{n},t_{i}) and conjecture μi\mu_{i} that justifies the inclusion of sis_{i} in Fi,k+10(Υn,ti)\textup{F}_{i,k+1}^{0}(\Upsilon^{n},t_{i}). We know from part (ii)(ii) of the induction hypothesis that Fi,k0(Υn,ti)Fi,k(Υ,ti)\textup{F}_{-i,k}^{0}(\Upsilon^{n},t_{-i})\subseteq\textup{F}_{-i,k}(\Upsilon,t_{-i}) and Hi(Fi,k0,Υn)=Hi(Fi,k,Υn)\textup{H}_{i}(\textup{F}_{-i,k}^{0},\Upsilon^{n})=\textup{H}_{i}(\textup{F}_{-i,k},\Upsilon^{n}) for every tiTi(Θn)t_{-i}\in T_{-i}(\Theta^{n}). Hence, it clearly follow that μi\mu_{i} justifies the inclusion of sis_{i} in Fi,k+1(Υn,ti)\textup{F}_{i,k+1}(\Upsilon^{n},t_{i}). Now, we divide the proof of the second claim in the following three-part cycle:

  • Hj(Fi,k+10,Υn)Hj(Fi,k+1,Υn)\textup{H}_{j}(\textup{F}_{i,k+1}^{0},\Upsilon^{n})\subseteq\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon^{n}) for any jij\neq i and any nn¯k+1i,1n\geq\bar{n}_{k+1}^{i,1}. This follows immediately from part (i)(i), just proved above.

  • There exists some n¯k+1i,2\bar{n}_{k+1}^{i,2}\in\mathds{N} such that Hj(Fi,k+1,Υn)Hj(Fi,k+1,Υ)\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon^{n})\subseteq\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon) for any player jij\neq i and any nm¯k+1n\geq\bar{m}_{k+1}. Pick nk+1in_{k+1}^{i}; according to part (iii)(iii) of Lemma 7 we have that Hj(Fi,k+1,Υn)Hj(Fi,k+1,Υ)\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon^{n})\subseteq\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon) for any nn¯k+12,in\geq\bar{n}_{k+1}^{2,i}.

  • There exists some n¯k+1i,3\bar{n}_{k+1}^{i,3}\in\mathds{N} such that Hj(Fi,k+1,Υ)Hj(Fi,k+10,Υn)\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon)\subseteq\textup{H}_{j}(\textup{F}_{i,k+1}^{0},\Upsilon^{n}) for any jij\neq i and any nn¯k+1i,3n\geq\bar{n}_{k+1}^{i,3}. Let Ti0(Θ)T_{i}^{0}(\Theta) denote the set of finite types in Ti(Θ)T_{i}(\Theta). Because of upper hemicontinuity of Fi,k+1(Υ,)\textup{F}_{i,k+1}(\Upsilon,\,\cdot\,) we know that every tiTi0(Θ)t_{i}\in T_{i}^{0}(\Theta) has some open neighborhood Ui(ti)U_{i}(t_{i}) such that:

    tiUi(ti){ti}×Fi,k+1(Υ,ti)Ui(ti)×Fi,k+1(Υ,ti).\bigcup_{t_{i}^{\prime}\in U_{i}(t_{i})}\{t_{i}^{\prime}\}\times\textup{F}_{i,k+1}(\Upsilon,t_{i}^{\prime})\subseteq U_{i}(t_{i})\times\textup{F}_{i,k+1}(\Upsilon,t_{i}).

    Then, because of denseness of Ti0(di)T_{i}^{0}(d_{i}) we have that:

    Graph(Fi,k+1(Υ,))\displaystyle\textup{Graph}\left(\textup{F}_{i,k+1}(\Upsilon,\,\cdot\,)\right) =tiTi0(Θ)tiUi(ti){ti}×Fi,k+1(Υ,ti)\displaystyle=\bigcup_{t_{i}\in T_{i}^{0}(\Theta)}\bigcup_{t_{i}^{\prime}\in U_{i}(t_{i})}\{t_{i}^{\prime}\}\times\textup{F}_{i,k+1}(\Upsilon,t_{i}^{\prime})
    =tiTi0(Θ)Ui(ti)×Fi,k+1(Υ,ti)\displaystyle=\bigcup_{t_{i}\in T_{i}^{0}(\Theta)}U_{i}(t_{i})\times\textup{F}_{i,k+1}(\Upsilon,t_{i})

    Thus, hHj(Fi,k+1,Υ)h\in\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon) if and only if there exists some finite tit_{i} such that siSi(h)Fi,k+1(Υ,ti)s_{i}\in S_{i}(h)\cap\textup{F}_{i,k+1}(\Upsilon,t_{i}). In consequence, by finiteness of HiH_{i} and claim (i)(i) there exists some n¯k+1i,3\bar{n}_{k+1}^{i,3} such that, for every nn¯k+1i,3n\geq\bar{n}_{k+1}^{i,3}, if hHj(Fi,k+1,Υ)h\in\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon), then hHj(Fi,k+10,Υn)h\in\textup{H}_{j}(\textup{F}_{i,k+1}^{0},\Upsilon^{n}).

Thus, by setting nk+1i,2:=max{n¯k+1i,1,n¯k+1i,2,n¯k+1i,3}n_{k+1}^{i,2}:=\textup{max}\{\bar{n}_{k+1}^{i,1},\bar{n}_{k+1}^{i,2},\bar{n}_{k+1}^{i,3}\} we conclude that for any jij\neq i and any nn¯k+1i,2n\geq\bar{n}_{k+1}^{i,2},

Hj(Fi,k+1,Υ)=Hj(Fi,k+1,Υn)=Hj(Fi,k+10,Υn),\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon)=\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon^{n})=\textup{H}_{j}(\textup{F}_{i,k+1}^{0},\Upsilon^{n}),

and hence, the proof is complete. ∎

Proof of Lemma 3

Lemma 3.

Let Γ\Gamma be an extensive form. Then, for any kk\in\mathds{N}, any player ii, any subjective payoff structure did_{i}, any finite type tiTi(di)t_{i}\in T_{i}(d_{i}) and any strategy siFi,k0(di,ti)s_{i}\in\textup{F}_{i,k}^{0}(d_{i},t_{i}) there exists a sequence of finite types (tin)n(t_{i}^{n})_{n\in\mathds{N}} in Ti,k(di,si)T_{i,k}(d_{i},s_{i}) converging to tit_{i}.

Proof.

Fix kk\in\mathds{N}, player ii, subjective payoff structure did_{i}, type tiTi(di)t_{i}\in T_{i}(d_{i}) and strategy siFi,k0(di,ti)s_{i}\in\textup{F}_{i,k}^{0}(d_{i},t_{i}), and pick finite conjecture μi\mu_{i} that justifies the inclusion of sis_{i} in Fi,k0(di,ti)\textup{F}_{i,k}^{0}(d_{i},t_{i}) (the existence of such a conjecture is guaranteed by Lemma 6 below). Based on Hi(μi)H_{i}(\mu_{i}) we define next:323232Remember that Hi(μi)\textup{H}_{i}(\mu_{i}) was defined in SectionB.1.1 as the set of histories in which conjecture μi\mu_{i} updates beliefs from scratch.

Hi0(μi):=Hi(μi)Hi(Fi,k10,di|i).H_{i}^{0}(\mu_{i}):=H_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,k-1}^{0},d_{-i|i}).

The interest of Hi0(μi)H_{i}^{0}(\mu_{i}) is that it will eventually allow for recovering conjecture μi\mu_{i} at those histories consistent Fi,k10(di|i,)\textup{F}_{-i,k-1}^{0}(d_{-i|i},\,\cdot\,). For each history hHi0(μi)h\in H_{i}^{0}(\mu_{i}) we are going to find a sequence of finitely-supported beliefs (μ^im(h))mΔ(Si(h)×di,1×Ti(di|i))(\hat{\mu}_{i}^{m}(h))_{m\in\mathds{N}}\subseteq\Delta(S_{-i}(h)\times d_{i,1}\times T_{-i}(d_{-i|i})) that converges to μi(h)\mu_{i}(h) and satisfies certain desirable properties described below. For each mm\in\mathds{N} define probability measure μm(h0)\mu^{m}(h^{0}) on Δ(Si×di,1×Ti(di|i))\Delta(S_{-i}\times d_{i,1}\times T_{-i}(d_{-i|i})) by setting:

μim(h0)=(11m)μi(h0)+(1m)hHi0(μi)(1|Hi0(μi)|)μi(h).\mu_{i}^{m}(h^{0})=\left(1-\frac{1}{m}\right)\cdot\mu_{i}(h^{0})+\left(\frac{1}{m}\right)\cdot\sum_{h\in H_{i}^{0}(\mu_{i})}\left(\frac{1}{|H_{i}^{0}(\mu_{i})|}\right)\cdot\mu_{i}(h).

Clearly, μim(h0)\mu_{i}^{m}(h^{0}) is well-defined and finitely-supported. Now, for each mm\in\mathds{N} denote:

Him={hHi{h0}|(margSiμim(h0))[Si(h)]>0},H_{i}^{m}=\left\{h\in H_{i}\cup\{h^{0}\}\left|\left(\textup{marg}_{S_{-i}}\mu_{i}^{m}(h^{0})\right)[S_{-i}(h)]>0\right.\right\},

and for any hHimh\in H_{i}^{m} define belief μim(h)Δ(Si×di,1×Ti(di|i))\mu_{i}^{m}(h)\in\Delta(S_{-i}\times d_{i,1}\times T_{-i}(d_{-i|i})) by extending μim(h0)\mu_{i}^{m}(h^{0}) via conditional probability. Finally, let conjecture μimΔHi{h0}(Si×di,1×Ti(di|i))\mu_{i}^{m}\in\Delta^{H_{i}\cup\{h^{0}\}}(S_{-i}\times d_{i,1}\times T_{-i}(d_{-i|i})) and type t^im=(θi(ti),πim)\hat{t}_{i}^{m}=(\theta_{i}(t_{i}),\pi_{i}^{m}) be respectively defined as follows:

μim(h):={μim(h)for every hHim,μi(h)otherwise, and πim:=τi1(marg(di,1)0×Ti(di|i)μim(h0)).\displaystyle\mu_{i}^{m}(h):=\left\{\begin{tabular}[]{l l}$\mu_{i}^{m}(h)$&$\textup{for every }h\in H_{i}^{m}$,\\ $\mu_{i}(h)$&$\textup{otherwise,}$\end{tabular}\right.\textup{\hskip 1.42271ptand\hskip 5.69046pt}\pi_{i}^{m}:=\tau_{i}^{-1}\left(\textup{marg}_{(d_{i,1})_{0}\times T_{-i}(d_{-i|i})}\mu_{i}^{m}(h^{0})\right).

We claim now that the following four properties are satisfied:

  • (a)(a)

    (t^im)m(\hat{t}_{i}^{m})_{m\in\mathds{N}} is a sequence of finite types that converges to tit_{i}. For finiteness simply notice that μim(h0)\mu_{i}^{m}(h^{0}) is finitely-supported for every mm\in\mathds{N}. Convergence follows from (μim(h0))m(\mu_{i}^{m}(h^{0}))_{m\in\mathds{N}} converging to μi(h0)\mu_{i}(h^{0}) and marginalization being continuous.

  • (b)(b)

    For any mm\in\mathds{N} the marginal on SiS_{-i} of μim(h0)\mu_{i}^{m}(h^{0}) assigns positive probability to Si(h)S_{-i}(h) for every history hHi(Fi,k10,di)h\in\textup{H}_{i}(\textup{F}_{-i,k-1}^{0},d_{i}). Fix such hh. If hHi0(μi)h\in H_{i}^{0}(\mu_{i}), then:

    (margSiμim(h0))[Si(h)]\displaystyle(\textup{marg}_{S_{-i}}\mu_{i}^{m}(h^{0}))\left[S_{-i}(h)\right] (1m)(1|Hi0(μi)|)(margSiμi(h))[Si(h)]\displaystyle\geq\left(\frac{1}{m}\right)\cdot\left(\frac{1}{|H_{i}^{0}(\mu_{i})|}\right)\cdot(\textup{marg}_{S_{-i}}\mu_{i}(h))\left[S_{-i}(h)\right]
    =(1m)(1|Hi0(μi)|)>0.\displaystyle=\left(\frac{1}{m}\right)\cdot\left(\frac{1}{|H_{i}^{0}(\mu_{i})|}\right)>0.

    If hHi0(μi)h\notin H_{i}^{0}(\mu_{i}), then there must exist some history hhh^{\prime}\prec h such that hHi0(μi)h^{\prime}\in H_{i}^{0}(\mu_{i}) and (margSiμim(h))[Si(h)]>0(\textup{marg}_{S_{-i}}\mu_{i}^{m}(h^{\prime}))\left[S_{-i}(h)\right]>0, and thus:

    (margSiμim(h0))[Si(h)]\displaystyle(\textup{marg}_{S_{-i}}\mu_{i}^{m}(h^{0}))\left[S_{-i}(h)\right] (1m)(1|Hi0(μi)|)(margSiμi(h))[Si(h)]>0.\displaystyle\geq\left(\frac{1}{m}\right)\cdot\left(\frac{1}{|H_{i}^{0}(\mu_{i})|}\right)\cdot(\textup{marg}_{S_{-i}}\mu_{i}(h^{\prime}))\left[S_{-i}(h)\right]>0.
  • (c)(c)

    For any mm\in\mathds{N}, μim\mu_{i}^{m} is a well-defined conditional probability system consistent with did_{i} and t^im\hat{t}_{i}^{m}. Consistency with did_{i} and timt_{i}^{m} holds by construction; thus, all we need to check is that μim\mu_{i}^{m} does not violate the chain rule. This is easy to see by simply noticing that for any pair of different histories h,hHi0(μi)h,h^{\prime}\in H_{i}^{0}(\mu_{i}) the marginal on SiS_{-i} of μi(h)\mu_{i}(h) puts zero probability on Si(h)S_{-i}(h^{\prime}). This implies that for any h,hHimh,h^{\prime}\in H_{i}^{m} there are no inconsistency issues arising from belief update.333333Because if h,hHimh,h^{\prime}\in H_{i}^{m} then there exist h¯,h¯Hi0(μi)\bar{h},\bar{h^{\prime}}\in H_{i}^{0}(\mu_{i}) such that μi(h¯)\mu_{i}(\bar{h}) induces μim(h)\mu_{i}^{m}(h) and μi(h¯)\mu_{i}(\bar{h}^{\prime}) induces μim(h)\mu_{i}^{m}(h^{\prime}). Since the marginal on SiS_{-i} of μi(h¯)\mu_{i}(\bar{h}) (resp. μi(h¯)\mu_{i}(\bar{h}^{\prime})) puts zero probability on Si(h¯)S_{-i}(\bar{h^{\prime}}) (resp. Si(h¯)S_{-i}(\bar{h})), it follows that the marginal on SiS_{-i} of μim(h)\mu_{i}^{m}(h) (resp. μim(h)\mu_{i}^{m}(h^{\prime})) puts zero probability on Si(h)S_{-i}(h^{\prime}) (resp. Si(h)S_{-i}(h)). Clearly, there are no problems either for any pair h,hHimh,h^{\prime}\notin H_{i}^{m} (due to μi\mu_{i} being a conditional probability system). Since for any hHimh\in H_{i}^{m} the marginal on SiS_{-i} of μi(h)\mu_{i}(h) puts zero probability on Si(h)S_{-i}(h^{\prime}) for any hHimh^{\prime}\notin H_{i}^{m}, it follows that pairs hHimh\in H_{i}^{m} and hHimh^{\prime}\notin H_{i}^{m} are not problematic either.

    In addition it holds by construction that:

    supp μim(h0)Graph(Fi,k10(di|i,)).\textup{supp }\mu_{i}^{m}(h^{0})\subseteq\textup{Graph}\left(\textup{F}_{-i,k-1}^{0}(d_{-i|i},\,\cdot\,)\right).

    Thus we have that if hHi(Fi,k10,di|i)h\in\textup{H}_{i}(\textup{F}_{-i,k-1}^{0},d_{-i|i}) then:

    supp μim(h)supp μim(h0)Graph(Fi,k10(di|i,)).\textup{supp }\mu_{i}^{m}(h)\subseteq\textup{supp }\mu_{i}^{m}(h^{0})\subseteq\textup{Graph}\left(\textup{F}_{-i,k-1}^{0}(d_{-i|i},\,\cdot\,)\right).

    Hence belief in appropriate behavior is maintained whenever possible to do so. If hHi(Fi,k10,di|i)h\notin\textup{H}_{i}(\textup{F}_{-i,k-1}^{0},d_{-i|i}) then:

    supp μim(h)=supp μi(h),\textup{supp }\mu_{i}^{m}(h)=\textup{supp }\mu_{i}(h),

    what guarantees belief in appropriate continuation behavior.

  • (d)(d)

    There exists some m0m_{0}\in\mathds{N} such that ri(θi(ti),μim)=[si]r_{i}(\theta_{i}(t_{i}),\mu_{i}^{m})=[s_{i}] for any mm0m\geq m_{0}. This follows trivially from finiteness of SiS_{i}, continuity of conditional expected payoffs and the fact that (μim)m(\mu_{i}^{m})_{m\in\mathds{N}} converges to μi\mu_{i}.

This way, we conclude that siFi,k0(di,tim)s_{i}\in\textup{F}_{i,k}^{0}(d_{i},t_{i}^{m}) for every mm0m\geq m_{0}. Hence, if for each nn\in\mathds{N} we relabel as tin:=t^in+m0t_{i}^{n}:=\hat{t}_{i}^{n+m_{0}}, (tin)n(t_{i}^{n})_{n\in\mathds{N}} is the sequence we are looking for. ∎

Proof of Lemma 5

Second perturbation of payoff structures

Fix some standard payoff structure Υ\Upsilon. For each k{0}k\in\mathds{N}\cup\{0\}, each player ii and each let dikd_{i}^{k} denote the subjective payoff structure constructed recursively by setting:

  • (0)(0)

    di0(Θ):=Υd_{i}^{0}(\Theta):=\Upsilon^{*}, where Υ\Upsilon^{*} is some arbitrary standard payoff structure satisfying richness.

  • (1)(1)

    di1(Υ)d_{i}^{1}(\Upsilon) is defined by setting di,11(Υ):=Θd_{i,1}^{1}(\Upsilon):=\Theta and di|i1:=di0(Υ)d_{-i|i}^{1}:=d_{-i}^{0}(\Upsilon).

  • (k)(k)

    dik(Υ)d_{i}^{k}(\Upsilon) is defined by setting di,1k(Υ):=Θd_{i,1}^{k}(\Upsilon):=\Theta and di|ik(Υ):=dik1(Υ)d_{-i|i}^{k}(\Upsilon):=d_{-i}^{k-1}(\Upsilon)

Obviously, each dik(Υ)d_{i}^{k}(\Upsilon) is a subjective payoff structure satisfying higher-order richness, and sequence (dik(Υ))k(d_{i}^{k}(\Upsilon))_{k\in\mathds{N}} converges to Υ\Upsilon.

From strictness to uniqueness: Initial step

Lemma 4.

Let Γ\Gamma be an extensive form and Υ\Upsilon, a standard payoff structure. Then, for any player ii, any finite type tiTi(Θ)t_{i}\in T_{i}(\Theta), any strategy siFi,10(Υ,ti)s_{i}\in\textup{F}_{i,1}^{0}(\Upsilon,t_{i}) such that tiTi(Υ,si)t_{i}\in T_{i}(\Upsilon,s_{i}), there exists a finite type ti1t_{i}^{1} such that:

  • (i)(i)

    ti1Ti(di1(Υ))t_{i}^{1}\in T_{i}(d_{i}^{1}(\Upsilon)).

  • (ii)(ii)

    θi(ti1)=θi(ti)\theta_{i}(t_{i}^{1})=\theta_{i}(t_{i}).

  • (iii)(iii)

    Fi,2(di1(Υ),ti1)[si|Υ]0\textup{F}_{i,2}(d_{i}^{1}(\Upsilon),t_{i}^{1})\subseteq[s_{i}|\Upsilon]_{0}.

Proof.

For notational convenience, set di0:=di0(Θ)d_{i}^{0}:=d_{i}^{0}(\Theta) and di1:=di1(Θ)d_{i}^{1}:=d_{i}^{1}(\Theta). We are going to show first that for any player ii, any finite type tit_{i} and any strategy sis_{i} there exists a finite type ti0t_{i}^{0} such that:

  • (i)(i)

    Fi,1(di0,ti0)={si}\textup{F}_{i,1}(d_{i}^{0},t_{i}^{0})=\{s_{i}\}.

  • (ii)(ii)

    θi(ti0)\theta_{i}(t_{i}^{0}) and θi(ti)\theta_{i}(t_{i}) are payoff-equivalent for player ii’s opponents.

To see it fix player ii, finite type tit_{i} and strategy sis_{i} and pick player ii’s payoff type θi(ti,si)\theta_{i}(t_{i},s_{i}) that makes sis_{i} conditionally dominant and is payoff-equivalent to θi(ti)\theta_{i}(t_{i}) for player ii’s opponents.343434Richness of Υ\Upsilon^{*} allows for such a selection. Then, set ti0:=(θi(ti,si),πi(ti))t_{i}^{0}:=(\theta_{i}(t_{i},s_{i}),\pi_{i}(t_{i})). Obviously, ti0t_{i}^{0} is finite. Pick now arbitrary conjecture μi\mu_{i} consistent with subjective model (di0,ti0)(d_{i}^{0},t_{i}^{0}). By definition, for any history hHi{h0}h\in H_{i}\cup\{h^{0}\} and any strategy sis_{i}^{\prime} such that si(h)si(h)s_{i}^{\prime}(h)\neq s_{i}(h):

Ui(si,μi|θi(ti0),h)>Ui(si,μi|θi(ti0),h).U_{i}(s_{i},\mu_{i}|\theta_{i}(t_{i}^{0}),h)>U_{i}(s^{\prime}_{i},\mu_{i}|\theta_{i}(t_{i}^{0}),h).

Hence, ri(θi(ti0),μi)={si}r_{i}(\theta_{i}(t_{i}^{0}),\mu_{i})=\{s_{i}\} and thus, we conclude that Fi,1(di0,ti0)={si}\textup{F}_{i,1}(d_{i}^{0},t_{i}^{0})=\{s_{i}\}. In addition, as said above, θi(ti0)\theta_{i}(t_{i}^{0}) is by construction payoff-equivalent to θi(ti)\theta_{i}(t_{i}) for player ii’s opponents.

Now, to prove the claim of the lemma fix player ii, finite type tiTi(Θ)t_{i}\in T_{i}(\Theta) and strategy sis_{i} such that siFi,10(Υ,ti)s_{i}\in\textup{F}_{i,1}^{0}(\Upsilon,t_{i}) and tiTi,1(Υ,si)t_{i}\in T_{i,1}(\Upsilon,s_{i}). We know then that there exists some conjecture μ¯i\bar{\mu}_{i} consistent with Θ\Theta and tit_{i} and such that: (a)(a) ri(θi(ti),μ¯i)={s¯i}r_{i}(\theta_{i}(t_{i}),\bar{\mu}_{i})=\{\bar{s}_{i}\} and (b)(b) (margSiμ¯i(h0))[Si(h)]>0(\textup{marg}_{S_{-i}}\bar{\mu}_{i}(h^{0}))[S_{-i}(h)]>0 for every history hHi(Fi,00,Υ)h\in\textup{H}_{i}(\textup{F}_{-i,0}^{0},\Upsilon). Remember now that we just proved above that for any pair (si,ti)(s_{-i},t_{-i}) in the support of the marginal of μ¯i(h0)\bar{\mu}_{i}(h^{0}) on Si×Ti(di0)S_{-i}\times T_{-i}(d_{-i}^{0}) there exists some finite type ti0(si,ti)t_{-i}^{0}(s_{-i},t_{-i}) such that for any player ji,j\neq i,

  • Fj,1(dj0,tj0(sj,tj))={sj}\textup{F}_{j,1}(d_{j}^{0},t_{j}^{0}(s_{j},t_{j}))=\{s_{j}\}.

  • θj(tj0(sj,tj))\theta_{j}(t_{j}^{0}(s_{j},t_{j})) is payoff-equivalent to θj(tj)\theta_{j}(t_{j}) for player jj’s opponents.

Define then μ¯i1\bar{\mu}_{i}^{1} as follows:

μ¯i1[E]:=μ¯i(h0)[{(si,θ0,ti)Si×Θ0×Ti(Θ)|(si,θ0,ti0(si,ti))E}],\bar{\mu}_{i}^{1}[E]:=\bar{\mu}_{i}(h^{0})\left[\left\{(s_{-i},\theta_{0},t_{-i})\in S_{-i}\times\Theta_{0}\times T_{-i}(\Theta)\left|(s_{-i},\theta_{0},t_{-i}^{0}(s_{-i},t_{-i}))\in E\right.\right\}\right],

for any measurable ESi×(di,11)0×Ti(di0)E\subseteq S_{-i}\times(d_{i,1}^{1})_{0}\times T_{-i}(d_{-i}^{0}). Since SiS_{-i} and tit_{i} are finite it is immediate that μ¯i1\bar{\mu}_{i}^{1} is a well-defined probability measure in Δ(Si×(di,11)0×Ti(di0))\Delta(S_{-i}\times(d_{i,1}^{1})_{0}\times T_{-i}(d_{-i}^{0})). Define next belief hierarchy:

πi1:=τi1(marg(di,11)0×Ti(di0)μ¯i1),\pi_{i}^{1}:=\tau_{i}^{-1}\left(\textup{marg}_{(d_{i,1}^{1})_{0}\times T_{-i}(d_{-i}^{0})}\bar{\mu}_{i}^{1}\right),

and set ti1:=(θi(ti),πi1)t_{i}^{1}:=(\theta_{i}(t_{i}),\pi_{i}^{1}). Clearly, ti1t_{i}^{1} is finite and satisfies that θi(ti1)=θi(ti)\theta_{i}(t_{i}^{1})=\theta_{i}(t_{i}). Note in addition that the supports of the marginals on Θ0\Theta_{0} of μ¯i(h0)\bar{\mu}_{i}(h^{0}) and μ¯i1\bar{\mu}_{i}^{1} coincide, and thus, that the support of the marginal on Θ0\Theta_{0} of τi(πi(ti1))\tau_{i}(\pi_{i}(t_{i}^{1})) is contained in (di,11)0(d_{i,1}^{1})_{0}. It follows then that ti1Ti(di1)t_{i}^{1}\in T_{i}(d_{i}^{1}). Let’s finish verifying that:

Fi,2(di1,ti1)[s¯i|Υ]0.\textup{F}_{i,2}(d_{i}^{1},t_{i}^{1})\subseteq[\bar{s}_{i}|\Upsilon]_{0}.

To see it, pick arbitrary siFi,2(di1,ti1)s_{i}\in\textup{F}_{i,2}(d_{i}^{1},t_{i}^{1}) and conjecture μi\mu_{i} that justifies the inclusion of sis_{i} in Fi,2(di1,ti1)\textup{F}_{i,2}(d_{i}^{1},t_{i}^{1}). Now, for each pair (si,θi)(s_{-i},\theta_{-i}) denote:

Ti0(si,θi):={ti0(si,ti)|tiTi(Θ) and θi(ti0(si,ti))=θi},T_{-i}^{0}(s_{-i},\theta_{-i}):=\left\{t_{-i}^{0}(s_{-i},t_{-i})\left|t_{-i}\in T_{-i}(\Theta)\textup{ and }\theta_{-i}(t_{-i}^{0}(s_{-i},t_{-i}))=\theta_{-i}\right.\right\},

and notice that for any (si,θi,θ0)(s_{-i},\theta_{-i},\theta_{0}) we have that:353535We provide further clarification of the following development. The first two equalities follow from the construction of ti1t_{i}^{1}, from μi\mu_{i} initially assigning probability 11 to the graph of Fi,1(di0,)\textup{F}_{-i,1}(d_{-i}^{0},\,\cdot\,) and the fact that Fi,1(di0,ti0(si,ti))={si}\textup{F}_{-i,1}(d_{-i}^{0},t_{-i}^{0}(s_{-i},t_{-i}))=\{s_{-i}\}. The third one follows from the construction of μ¯i1\bar{\mu}_{i}^{1}. The fourth is obvious. The fifth equality follows, again, from the construction of μ¯1\bar{\mu}^{1}. The last one follows from the fact that if ti0(si,ti)=ti0(si,ti)t_{-i}^{0}(s_{-i}^{\prime},t_{-i})=t_{-i}^{0}(s_{-i},t_{-i}) then si=sis_{-i}^{\prime}=s_{-i}, because Fi,1(di0,ti0(si,ti))={si}\textup{F}_{-i,1}(d_{-i}^{0},t_{-i}^{0}(s_{-i},t_{-i}))=\{s_{-i}\}.

μi(h0)\displaystyle\mu_{i}(h^{0}) [{(si,θi,θ0)}×Πi(Θ)]=\displaystyle\left[\{(s_{-i},\theta_{-i},\theta_{0})\}\times\Pi_{-i}(\Theta)\right]=
=μi(h0)[{(si,θ0)}×Ti0(si,θi)]\displaystyle=\mu_{i}(h^{0})\left[\{(s_{-i},\theta_{0})\}\times T_{-i}^{0}(s_{-i},\theta_{-i})\right]
=μi(h0)[Si×{θ0}×Ti0(si,θi)]\displaystyle=\mu_{i}(h^{0})\left[S_{-i}\times\{\theta_{0}\}\times T_{-i}^{0}(s_{-i},\theta_{-i})\right]
=μ¯i1[Si×{θ0}×Ti0(si,θi)]\displaystyle=\bar{\mu}_{i}^{1}\left[S_{-i}\times\{\theta_{0}\}\times T_{-i}^{0}(s_{-i},\theta_{-i})\right]
=siSiμ¯i1[{(si,,θ0)}×Ti0(si,θi)]\displaystyle=\sum_{s_{-i}^{\prime}\in S_{-i}}\bar{\mu}_{i}^{1}\left[\{(s_{-i}^{\prime},,\theta_{0})\}\times T_{-i}^{0}(s_{-i},\theta_{-i})\right]
=siSiμ¯i(h0)[{(si,θ0)}×{tiTi(Θ)|ti0(si,ti)Ti0(si,θi)}]\displaystyle=\sum_{s_{-i}^{\prime}\in S_{-i}}\bar{\mu}_{i}(h^{0})\left[\{(s_{-i}^{\prime},\theta_{0})\}\times\left\{t_{-i}\in T_{-i}(\Theta)\left|t_{-i}^{0}(s_{-i}^{\prime},t_{-i})\in T_{-i}^{0}(s_{-i},\theta_{-i})\right.\right\}\right]
=μ¯i(h0)[{(si,θ0)}×{tiTi(Θ)|ti0(si,ti)Ti0(si,θi)}].\displaystyle=\bar{\mu}_{i}(h^{0})\left[\{(s_{-i},\theta_{0})\}\times\left\{t_{-i}\in T_{-i}(\Theta)\left|t_{-i}^{0}(s_{-i},t_{-i})\in T_{-i}^{0}(s_{-i},\theta_{-i})\right.\right\}\right].

Now, notice two things. First, we know that for any (si,ti)(s_{-i},t_{-i}) in the support of the marginal of μ¯i(h0)\bar{\mu}_{i}(h^{0}) on Si×Ti(Θ)S_{-i}\times T_{-i}(\Theta) θi(ti)\theta_{-i}(t_{-i}) and θi(ti0(si,ti))\theta_{-i}(t_{-i}^{0}(s_{-i},t_{-i})) are payoff-equivalent for player ii; thus, it follows from the above that μi(h0)\mu_{i}(h^{0}) and μ¯i(h0)\bar{\mu}_{i}(h^{0}) induce the same conditional expected payoffs at h0h^{0}. Second, the marginal of μ¯i(h0)\bar{\mu}_{i}(h^{0}) on SiS_{-i} assigns positive probability to Si(h)S_{-i}(h) for every hHi{h0}h\in H_{i}\cup\{h^{0}\}; in consequence, it follows from the above that μi\mu_{i} and μ¯i\bar{\mu}_{i} induce the same conditional expected payoffs at hh for every history hHi{h0}h\in H_{i}\cup\{h^{0}\}. Thus, we conclude that ri(θi(ti1),μi)={s¯i}r_{i}(\theta_{i}(t_{i}^{1}),\mu_{i})=\{\bar{s}_{i}\} and, thus, that si[s¯i|Υ]0s_{i}\in[\bar{s}_{i}|\Upsilon]_{0}. ∎

From strictness to uniqueness: Inductive step

Lemma 5.

Let Γ\Gamma be an extensive form. Then for any k1k\geq 1, any player ii, any standard payoff structure Υ\Upsilon such that Hj(Fi,k+1,Υ)Hj(Fi,k+10,Υ)\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon)\subseteq\textup{H}_{j}(\textup{F}_{i,k+1}^{0},\Upsilon) for every jij\neq i, any finite tiTi(Θ)t_{i}\in T_{i}(\Theta) and any strategy siFi,k0(Υ,ti)s_{i}\in\textup{F}_{i,k}^{0}(\Upsilon,t_{i}) such that tiTi(Υ,si)t_{i}\in T_{i}(\Upsilon,s_{i}) there exists a finite type tikt_{i}^{k} such that:

  • (i)(i)

    tikTi(dik(Υ))t_{i}^{k}\in T_{i}(d_{i}^{k}(\Upsilon)).

  • (ii)(ii)

    θi(tik)=θi(ti)\theta_{i}(t_{i}^{k})=\theta_{i}(t_{i}) and πi,k(tik)=πi,k(ti)\pi_{i,k}(t_{i}^{k})=\pi_{i,k}(t_{i}).

  • (iii)(iii)

    Fi,k+1(dik(Υ),tik)[si|Υ]k1\textup{F}_{i,k+1}(d_{i}^{k}(\Upsilon),t_{i}^{k})\subseteq[s_{i}|\Upsilon]_{k-1}.

Proof.

We proceed by induction on kk. The initial case (k=0k=0) is covered by Lemma 4 (the second part of property (ii)(ii) holds by vacuity) so we can focus on the proof of the inductive step. Suppose that k0k\geq 0 is such that the claims hold. We verify next that they also holds for k+1k+1. Fix first standard payoff structure Υ\Upsilon such that Hj(Fi,k+1,Υ)Hj(Fi,k+10,Υ)\textup{H}_{j}(\textup{F}_{i,k+1},\Upsilon)\subseteq\textup{H}_{j}(\textup{F}_{i,k+1}^{0},\Upsilon) for every jij\neq i and set, for notational convenience, dik+1:=dik+1(Υ)d_{i}^{k+1}:=d_{i}^{k+1}(\Upsilon); let us proceed now in two steps:

1. Parts (i)(i) and (ii)(ii): Construction of the required finite type

Fix player ii, finite tiTi(Θ)t_{i}\in T_{i}(\Theta) and strategy s¯i\bar{s}_{i} such that s¯iFi0(Υ,ti)\bar{s}_{i}\in\textup{F}_{i}^{0}(\Upsilon,t_{i}) and tiTi(Υ,s¯i)t_{i}\in T_{i}(\Upsilon,\bar{s}_{i}). Pick conjecture μ¯i\bar{\mu}_{i} that justifies the inclusion of s¯i\bar{s}_{i} in Fi,k+10(Υ,ti)\textup{F}_{i,k+1}^{0}(\Upsilon,t_{i}) and the inclusion of tit_{i} in Ti(Υ,s¯i)T_{i}(\Upsilon,\bar{s}_{i}). Now, we know from the induction hypothesis that for any pair (si,ti)(s_{-i},t_{-i}) in the support of the marginal of μ¯i(h0)\bar{\mu}_{i}(h^{0}) on Si×Ti(Θ)S_{-i}\times T_{-i}(\Theta) there exists some finite type tik(si,ti)t_{-i}^{k}(s_{-i},t_{-i}) such that for every player jij\neq i,

  • tjkTj(djk)=Tj(dj|ik+1)t_{j}^{k}\in T_{j}(d_{j}^{k})=T_{j}(d_{j|i}^{k+1}).

  • θjk=θj(tj)\theta_{j}^{k}=\theta_{j}(t_{j}) and πj,kk=πj,k(tj)\pi_{j,k}^{k}=\pi_{j,k}(t_{j}) for θjk=θj(tjk(sj,tj))\theta_{j}^{k}=\theta_{j}(t_{j}^{k}(s_{j},t_{j})) and πjk=πj(tjk(sj,tj))\pi_{j}^{k}=\pi_{j}(t_{j}^{k}(s_{j},t_{j})).

  • Fj,k+1(dj|ik+1,tjk(sj,tj))=Fj,k+1(djk,tjk(sj,tj))[sj|Υ]k1\textup{F}_{j,k+1}(d_{j|i}^{k+1},t_{j}^{k}(s_{j},t_{j}))=\textup{F}_{j,k+1}(d_{j}^{k},t_{j}^{k}(s_{j},t_{j}))\subseteq[s_{j}|\Upsilon]_{k-1}.

Define then probability measure μ¯ik+1Δ(Si×(di,1k+1)0×Ti(dik))\bar{\mu}_{i}^{k+1}\in\Delta(S_{-i}\times(d_{i,1}^{k+1})_{0}\times T_{-i}(d_{-i}^{k})) by setting:

μ¯ik+1[E]:=μ¯i(h0)[{(si,θ0,ti)Si×(di,1k+1)0×Ti(Θ)|(si,θ0,tik(si,ti))E}],\bar{\mu}_{i}^{k+1}[E]:=\bar{\mu}_{i}(h^{0})\left[\left\{(s_{-i},\theta_{0},t_{-i})\in S_{-i}\times(d_{i,1}^{k+1})_{0}\times T_{-i}(\Theta)\left|(s_{-i},\theta_{0},t_{-i}^{k}(s_{-i},t_{-i}))\in E\right.\right\}\right],

for any measurable ESi×(di,1k+1)0×Ti(dik)E\subseteq S_{-i}\times(d_{i,1}^{k+1})_{0}\times T_{-i}(d_{-i}^{k}). Finiteness of tit_{i} guarantees that μ¯ik+1\bar{\mu}_{i}^{k+1} is well-defined. Obviously, μ¯ik+1\bar{\mu}_{i}^{k+1} does not necessarily induce a whole conditional probability system via conditional probability (its marginal on SiS_{-i} may not have full-support); however, if we set μ¯ik+1(h0):=μ¯ik+1\bar{\mu}_{i}^{k+1}(h^{0}):=\bar{\mu}_{i}^{k+1} conditional probability enables to obtain a family of beliefs {μ¯ik+1(h)|hHi(Fi0,Υ)}\{\bar{\mu}_{i}^{k+1}(h)|h\in\textup{H}_{i}(\textup{F}_{-i}^{0},\Upsilon)\}. Notice that for any measurable ESi×Θ0×ΘiE\subseteq S_{-i}\times\Theta_{0}\times\Theta_{-i} it holds that:

μ¯ik+1(h)[E×Πi(dik)]=μ¯i(h)[E×Πi(Θ)],\bar{\mu}_{i}^{k+1}(h)[E\times\Pi_{-i}(d_{-i}^{k})]=\bar{\mu}_{i}(h)[E\times\Pi_{-i}(\Theta)],

for every history hHi(Fi,k0,Υ)h\in\textup{H}_{i}(\textup{F}_{-i,k}^{0},\Upsilon). Thus, for any such history hh we can define the conditional expected payoff induced by belief μ¯ik+1(h)\bar{\mu}_{i}^{k+1}(h) and any strategy sis_{i} for payoff type θi(ti)\theta_{i}(t_{i}), which, with some abuse of notation, we represent by Ui(si,μ¯ik+1|θi(ti),h)U_{i}(s_{i},\bar{\mu}_{i}^{k+1}|\theta_{i}(t_{i}),h), and, clearly, satisfies that:

Ui(si,μ¯ik+1|θi(ti),h)=Ui(si,μ¯i|θi(ti),h).U_{i}(s_{i},\bar{\mu}_{i}^{k+1}|\theta_{i}(t_{i}),h)=U_{i}(s_{i},\bar{\mu}_{i}|\theta_{i}(t_{i}),h). (1)

Define now belief hierarchy:

πik+1:=τi1(marg(di,1k+1)0×Ti(dik)μ¯ik+1),\pi_{i}^{k+1}:=\tau_{i}^{-1}\left(\textup{marg}_{(d_{i,1}^{k+1})_{0}\times T_{-i}(d_{-i}^{k})}\bar{\mu}_{i}^{k+1}\right),

and set tik+1:=(θi(ti),πik+1)t_{i}^{k+1}:=(\theta_{i}(t_{i}),\pi_{i}^{k+1}). Clearly, tik+1t_{i}^{k+1} is finite and has the same payoff type and kthk^{\textup{th}}-order beliefs as tit_{i}. Thus, it follows that tik+1Ti(dik+1)t_{i}^{k+1}\in T_{i}(d_{i}^{k+1}).363636Finiteness is immediate. To see that lower-order beliefs are maintained, simply notice that the probability assigned by μ¯i(h0)\bar{\mu}_{i}(h^{0}) to (si,θ0,ti)(s_{-i},\theta_{0},t_{-i}) is assigned by μ¯ik+1(h0)\bar{\mu}_{i}^{k+1}(h^{0}) to (si,θ0,tik(si,ti))(s_{-i},\theta_{0},t_{-i}^{k}(s_{-i},t_{-i})), being the (k1)th(k-1)^{\textup{th}}-order beliefs of tik(si,ti)t_{-i}^{k}(s_{-i},t_{-i}) exactly those of πi(ti)\pi_{-i}(t_{-i}). Finally, that tik+1Ti(dik+1)t_{i}^{k+1}\in T_{i}(d_{i}^{k+1}) follows from the fact that τi(πi(tik+1))\tau_{i}(\pi_{i}(t_{i}^{k+1})) only assigns positive probability to pairs (θ0,tik(si,ti))Θ0×Ti(dik)(\theta_{0},t_{-i}^{k}(s_{-i},t_{-i}))\in\Theta_{0}\times T_{-i}(d_{-i}^{k}) where (si,ti)(s_{-i},t_{-i}) is in the graph of Fi,k0(Θ,)\textup{F}_{-i,k}^{0}(\Theta,\,\cdot\,).

2. Part (iii)(iii): Verification of convenient behavior

It remains to be checked that:

Fi,k+2(dik+1,tik+1)[s¯i|Υ]k.\textup{F}_{i,k+2}(d_{i}^{k+1},t_{i}^{k+1})\subseteq[\bar{s}_{i}|\Upsilon]_{k}.

To see it fix strategy siFi,k+2(dik+1,tik+1)s_{i}\in\textup{F}_{i,k+2}(d_{i}^{k+1},t_{i}^{k+1}) and conjecture μi\mu_{i} that justifies the inclusion of sis_{i} in Fi,k+2(dik+1,tik+1)\textup{F}_{i,k+2}(d_{i}^{k+1},t_{i}^{k+1}). We want to verify that si[s¯i|Υ]ks_{i}\in[\bar{s}_{i}|\Upsilon]_{k}. The process is slightly arduous, so let us first present a road map describing the six intermediate steps required:

  • S1.

    For any (si,θi,θ0)Si×Θ0×Θi(s_{-i},\theta_{-i},\theta_{0})\in S_{-i}\times\Theta_{0}\times\Theta_{-i} we have that:

    μi(h0)[[si|Υ]k1×{(θ0,θi)}\displaystyle\mu_{i}(h^{0})[[s_{-i}|\Upsilon]_{k-1}\times\{(\theta_{0},\theta_{-i})\} ×Πi(dik)]=\displaystyle\times\Pi_{-i}(d_{-i}^{k})]=
    =μ¯ik+1(h0)[[si|Υ]k1×{(θ0,θi)}×Πi(Θ)].\displaystyle=\bar{\mu}_{i}^{k+1}(h^{0})[[s_{-i}|\Upsilon]_{k-1}\times\{(\theta_{0},\theta_{-i})\}\times\Pi_{-i}(\Theta)].

    In general, μi(h0)\mu_{i}(h^{0}) and μ¯ik+1(h0)\bar{\mu}_{i}^{k+1}(h^{0}) may induce different marginals on Si×Θ0×ΘiS_{-i}\times\Theta_{0}\times\Theta_{-i} because the types tit_{-i}^{\prime} to which tik+1t_{i}^{k+1} assigns positive probability do not necessarily have a unique strategy in Fi,k+1(dik,ti)\textup{F}_{-i,k+1}(d_{-i}^{k},t_{-i}^{\prime}). However, the marginals both beliefs induce expect ii’s opponents to behave analogously while each opponent jij\neq i finds herself in a history reachable by some strategy profile in Fk10(Υ,)\textup{F}_{k-1}^{0}(\Upsilon,\,\cdot\,).

  • S2.

    Furthermore, for any (si,θ0,θi)Si×Θ0×Θi(s_{-i},\theta_{0},\theta_{-i})\in S_{-i}\times\Theta_{0}\times\Theta_{-i} and any hHi(Fk0,Υ)h\in\textup{H}_{i}(\textup{F}_{k}^{0},\Upsilon) we have that:

    μi(h)[[si|Υ]k1×{(θ0,θi)}\displaystyle\mu_{i}(h)[[s_{-i}|\Upsilon]_{k-1}\times\{(\theta_{0},\theta_{-i})\} ×Πi(dik)]=\displaystyle\times\Pi_{-i}(d_{-i}^{k})]=
    =μ¯ik+1(h)[[si|Υ]k1×{(θ0,θi)}×Πi(Θ)].\displaystyle=\bar{\mu}_{i}^{k+1}(h)[[s_{-i}|\Upsilon]_{k-1}\times\{(\theta_{0},\theta_{-i})\}\times\Pi_{-i}(\Theta)].

    That is, despite both beliefs inducing possibly different marginals on Si×Θ0×ΘiS_{-i}\times\Theta_{0}\times\Theta_{-i} at different stages of the game, the observation in S1 generalizes from the initial history to every history that is consistent players having chosen some strategy profile in Fk10(Υ,)\textup{F}_{k-1}^{0}(\Upsilon,\,\cdot\,).

  • S3.

    For any pair (ti,si)(t_{-i},s_{-i}) in the graph of Fi,k0(Υ,)\textup{F}_{-i,k}^{0}(\Upsilon,\,\cdot\,) and any strategy si[si|Υ]k1s^{\prime}_{-i}\in[s_{-i}|\Upsilon]_{k-1} we have the following outcome equivalence:

    z((si;s¯i)|h0)=z((si;s¯i)|h0).z((s_{-i};\bar{s}_{i})|h^{0})=z((s_{-i}^{\prime};\bar{s}_{i})|h^{0}).
  • S4.

    For any pair (ti,si)(t_{-i},s_{-i}) in the graph of Fi,k0(Υ,)\textup{F}_{-i,k}^{0}(\Upsilon,\,\cdot\,) and any strategy si[si|Υ]k1s^{\prime}_{-i}\in[s_{-i}|\Upsilon]_{k-1} we have the following outcome equivalence:

    z((si;si)|h0)=z((si;si)|h0).z((s_{-i};s_{i})|h^{0})=z((s_{-i}^{\prime};s_{i})|h^{0}).
  • S5.

    For any strategy si{s¯i,si}s^{\prime}_{i}\in\{\bar{s}_{i},s_{i}\} and any history hHi(Fk0,Υ)h\in\textup{H}_{i}(\textup{F}_{k}^{0},\Upsilon):

    Ui(si,μi|θi(ti),h)=Ui(si,μ¯ik+1|θi(ti),h).U_{i}(s^{\prime}_{i},\mu_{i}|\theta_{i}(t_{i}),h)=U_{i}(s^{\prime}_{i},\bar{\mu}_{i}^{k+1}|\theta_{i}(t_{i}),h).

    At every history consistent players having chosen a strategy profile in Fi,k(Υ,)\textup{F}_{-i,k}(\Upsilon,\,\cdot\,) player ii’s type θi(ti)\theta_{i}(t_{i})’s expected utilities under μi\mu_{i} and μ¯ik+1\bar{\mu}_{i}^{k+1} coincide either when playing s¯i\bar{s}_{i} or sis_{i}. Notice that this is a direct consequence of the outcome equivalences in S3 and S4, and of S2, where we saw that μi\mu_{i} and μ¯ik+1\bar{\mu}_{i}^{k+1} induce beliefs that expect analogous behavior while each player jij\neq i finds herself in a history she finds reachable when opponents play according to Fj,k10(Υ,)\textup{F}_{-j,k-1}^{0}(\Upsilon,\,\cdot\,).

  • S6.

    For any history hHi(Fk0,Υ)h\in\textup{H}_{i}(\textup{F}_{k}^{0},\Upsilon), si(h)=s¯i(h)s_{i}(h)=\bar{s}_{i}(h). This follows from S5 and the fact that sis_{i} is optimal given μi\mu_{i} and s¯i\bar{s}_{i} is optimal given μ¯i\bar{\mu}_{i}, which in turn, induces an expected payoff similar enough to that of μ¯ik+1\bar{\mu}_{i}^{k+1}.

Once these six steps are verified part (iii)(iii) the lemma will follow immediately. Let’s proceed step by step:

  • S1.

    First, for every SiSiS_{-i}^{\prime}\subseteq S_{-i} and θiΘi\theta_{-i}\in\Theta_{-i} denote:

    Tik(Si,θi):=siSi{tik(si,ti)|tiTi(Θ) and θi(tik(si,ti))=θi}.T_{-i}^{k}(S_{-i}^{\prime},\theta_{-i}):=\bigcup_{s_{-i}\in S_{-i}^{\prime}}\left\{t_{-i}^{k}(s_{-i},t_{-i})\left|t_{-i}\in T_{-i}(\Theta)\textup{ and }\theta_{-i}(t_{-i}^{k}(s_{-i},t_{-i}))=\theta_{-i}\right.\right\}.

    Then, notice that for any triple (si,θ0,θi)Si×Θ0×Θi(s_{-i},\theta_{0},\theta_{-i})\in S_{-i}\times\Theta_{0}\times\Theta_{-i} we have that:

    μi(h0)[[si|Υ]k1×{(θ0,θi)}×Πi(Θ)]=\displaystyle\mu_{i}(h^{0})\left[[s_{-i}|\Upsilon]_{k-1}\times\{(\theta_{0},\theta_{-i})\}\times\Pi_{-i}(\Theta)\right]=
    =μi(h0)[[si|Υ]k1×{θ0}×Tik([si|Υ]k1,θi)]\displaystyle=\mu_{i}(h^{0})\left[[s_{-i}|\Upsilon]_{k-1}\times\{\theta_{0}\}\times T_{-i}^{k}([s_{-i}|\Upsilon]_{k-1},\theta_{-i})\right]
    =μi(h0)[Si×{θ0}×Tik([si|Υ]k1,θi)]\displaystyle=\mu_{i}(h^{0})\left[S_{-i}\times\{\theta_{0}\}\times T_{-i}^{k}([s_{-i}|\Upsilon]_{k-1},\theta_{-i})\right]
    =μ¯ik+1[Si×{θ0}×Tik([si|Υ]k1,θi)]\displaystyle=\bar{\mu}_{i}^{k+1}\left[S_{-i}\times\{\theta_{0}\}\times T_{-i}^{k}([s_{-i}|\Upsilon]_{k-1},\theta_{-i})\right]
    =si′′Siμ¯ik+1[{(si′′,,θ0)}×Tik([si|Υ]k1,θi)]\displaystyle=\sum_{s_{-i}^{\prime\prime}\in S_{-i}}\bar{\mu}_{i}^{k+1}\left[\{(s_{-i}^{\prime\prime},,\theta_{0})\}\times T_{-i}^{k}([s_{-i}|\Upsilon]_{k-1},\theta_{-i})\right]
    =si′′Siμ¯i(h0)[{(si′′,θ0)}×{tiTi(dik)|tik(si′′,ti)Tik([si|Υ]k1,θi)}]\displaystyle=\sum_{s_{-i}^{\prime\prime}\in S_{-i}}\bar{\mu}_{i}(h^{0})\left[\{(s_{-i}^{\prime\prime},\theta_{0})\}\times\left\{t_{-i}\in T_{-i}(d_{-i}^{k})\left|t_{-i}^{k}(s_{-i}^{\prime\prime},t_{-i})\in T_{-i}^{k}([s_{-i}|\Upsilon]_{k-1},\theta_{-i})\right.\right\}\right]
    =si′′[si|Θ]k1μ¯i(h0)[{(si′′,θ0)}×{tiTi(dik)|tik(si′′,ti)Tik([si|Υ]k1,θi)}]\displaystyle=\sum_{s_{-i}^{\prime\prime}\in[s_{-i}|\Theta]_{k-1}}\bar{\mu}_{i}(h^{0})\left[\{(s_{-i}^{\prime\prime},\theta_{0})\}\times\left\{t_{-i}\in T_{-i}(d_{-i}^{k})\left|t_{-i}^{k}(s_{-i}^{\prime\prime},t_{-i})\in T_{-i}^{k}([s_{-i}|\Upsilon]_{k-1},\theta_{-i})\right.\right\}\right]
    =si′′[si|Υ]k1μ¯i(h0)[{(si′′,θ0,θi)}×Πi(dik)]\displaystyle=\sum_{s_{-i}^{\prime\prime}\in[s_{-i}|\Upsilon]_{k-1}}\bar{\mu}_{i}(h^{0})\left[\{(s_{-i}^{\prime\prime},\theta_{0},\theta_{-i})\}\times\Pi_{-i}(d_{-i}^{k})\right]
    =μ¯i(h0)[[si|Υ]k1×{(θ0,θi)}×Πi(dik)].\displaystyle=\bar{\mu}_{i}(h^{0})\left[[s_{-i}|\Upsilon]_{k-1}\times\{(\theta_{0},\theta_{-i})\}\times\Pi_{-i}(d_{-i}^{k})\right].
  • S2.

    Fix history hHi(Fk0,Υ)h\in\textup{H}_{i}(\textup{F}_{k}^{0},\Upsilon) and siSi(h)s_{-i}\in S_{-i}(h), and pick some si[si|Υ]k1s_{-i}^{\prime}\in[s_{-i}|\Upsilon]_{k-1}. It is easy to see that siSi(h)s_{-i}^{\prime}\in S_{-i}(h): for any player jij\neq i and any history hHj{h0}h^{\prime}\in H_{j}\cup\{h^{0}\} that precedes or coincides with hh we have that hHj(Fk0,Υ)h^{\prime}\in\textup{H}_{j}(\textup{F}_{k}^{0},\Upsilon) and thus, it holds that sj(h)=sj(h)s^{\prime}_{j}(h^{\prime})=s_{j}(h^{\prime}). Hence, [si|Υ]k1Si(h)[s_{-i}|\Upsilon]_{k-1}\subseteq S_{-i}(h) and, in consequence, there must exist some si1,,siMSi(h)s_{-i}^{1},\dots,s_{-i}^{M}\in S_{-i}(h) such that the family {[si1|Υ]k1,,[siM|Υ]k1}\{[s_{-i}^{1}|\Upsilon]_{k-1},\dots,[s_{-i}^{M}|\Upsilon]_{k-1}\} is a partition of Si(h)S_{-i}(h). Then, it follows from S1 that:

    μi(h0)[Si(h)×Θ0×Ti(dik)]\displaystyle\mu_{i}(h^{0})[S_{-i}(h)\times\Theta_{0}\times T_{-i}(d_{-i}^{k})] =m=1Mμi(h0)[[sim|Υ]k1×Θ0×Ti(dik)]\displaystyle=\sum_{m=1}^{M}\mu_{i}(h^{0})[[s_{-i}^{m}|\Upsilon]_{k-1}\times\Theta_{0}\times T_{-i}(d_{-i}^{k})]
    =m=1Mμ¯ik+1(h0)[[sim|Υ]k1×Θ0×Ti(dik)]\displaystyle=\sum_{m=1}^{M}\bar{\mu}_{i}^{k+1}(h^{0})[[s_{-i}^{m}|\Upsilon]_{k-1}\times\Theta_{0}\times T_{-i}(d_{-i}^{k})]
    =μ¯ik+1(h0)[Si(h)×Θ0×Ti(dik)].\displaystyle=\bar{\mu}_{i}^{k+1}(h^{0})[S_{-i}(h)\times\Theta_{0}\times T_{-i}(d_{-i}^{k})].

    Remember now that μ¯ik+1(h0)\bar{\mu}_{i}^{k+1}(h^{0}) puts positive probability on Si(h)×Θ0×Ti(dik)S_{-i}(h)\times\Theta_{0}\times T_{-i}(d_{-i}^{k}) and thus, it follows that for any (si,θ0,θi)(s_{-i},\theta_{0},\theta_{-i}):

    μ¯ik+1(h)[[si|Υ]k1×\displaystyle\bar{\mu}_{i}^{k+1}(h)[[s_{-i}|\Upsilon]_{k-1}\times {(θ0,θi)}×Πi(dik)]=\displaystyle\{(\theta_{0},\theta_{-i})\}\times\Pi_{-i}(d_{-i}^{k})]=
    =μ¯ik+1(h0)[(Si(h)[si|Υ]k1)×{(θ0,θi)}×Πi(dik)]μ¯ik+1(h0)[Si(h)×Θ0×Ti(dik)]\displaystyle=\frac{\bar{\mu}_{i}^{k+1}(h^{0})[(S_{-i}(h)\cap[s_{-i}|\Upsilon]_{k-1})\times\{(\theta_{0},\theta_{-i})\}\times\Pi_{-i}(d_{-i}^{k})]}{\bar{\mu}_{i}^{k+1}(h^{0})[S_{-i}(h)\times\Theta_{0}\times T_{-i}(d_{-i}^{k})]}
    =μ¯i(h0)[(Si(h)[si|Υ]k1)×{(θ0,θi)}×Πi(Θ)]μ¯i(h0)[Si(h)×Θ0×Ti(Θ)]\displaystyle=\frac{\bar{\mu}_{i}(h^{0})[(S_{-i}(h)\cap[s_{-i}|\Upsilon]_{k-1})\times\{(\theta_{0},\theta_{-i})\}\times\Pi_{-i}(\Theta)]}{\bar{\mu}_{i}(h^{0})[S_{-i}(h)\times\Theta_{0}\times T_{-i}(\Theta)]}
    =μ¯i(h)[[si|Υ]k1×{(θ0,θi)}×Πi(Θ)].\displaystyle=\bar{\mu}_{i}(h)[[s_{-i}|\Upsilon]_{k-1}\times\{(\theta_{0},\theta_{-i})\}\times\Pi_{-i}(\Theta)].
  • S3.

    Fix pair (si,ti)(s_{-i},t_{-i}) in the graph of Fi,k0(Υ,)\textup{F}_{-i,k}^{0}(\Upsilon,\,\cdot\,) and si[si|Υ]k1s^{\prime}_{-i}\in[s_{-i}|\Upsilon]_{k-1}. Next, pick player jij\neq i such that there exists some history hHjh\in H_{j} that precedes z((si;s¯i)|h0)z((s_{-i};\bar{s}_{i})|h^{0}) and is reached by both sjs_{j} and sjs_{j}^{\prime}.373737 I.e., such that hz((si;s¯i)|h0)h\prec z((s_{-i};\bar{s}_{i})|h^{0}) and hHj(sj)Hj(sj)h\in H_{j}(s_{j})\cap H_{j}(s_{j}^{\prime}). Obviously, if there exists some hh satisfying the first requirement, there exists some history that satisfies both. We do not have to worry about players jj that lack such history hh: they do not make any choice along the path leading to z((si;s¯i)|h0)z((s_{-i};\bar{s}_{i})|h^{0}). To see that hHi(Fj,k10,Υ)h\in\textup{H}_{i}(\textup{F}_{j,k-1}^{0},\Upsilon) simply notice that: (a)(a) sF,k0(Υ,t)s_{\ell}\in\textup{F}_{\ell,k}^{0}(\Upsilon,t_{\ell}) for every player i,j\ell\neq i,j and (b)(b) s¯iFi,k+10(Υ,ti)\bar{s}_{i}\in\textup{F}_{i,k+1}^{0}(\Upsilon,t_{i}). Thus, we have that:

    hHj(Fi,k+10,Υ)i,jHj(F,k0,Υ)jHj(F,k10,Υ)=Hj(Fj,k10,Υ)h\in\textup{H}_{j}(\textup{F}_{i,k+1}^{0},\Upsilon)\cap\bigcap_{\ell\neq i,j}\textup{H}_{j}(\textup{F}_{\ell,k}^{0},\Upsilon)\subseteq\bigcap_{\ell\neq j}\textup{H}_{j}(\textup{F}_{\ell,k-1}^{0},\Upsilon)=\textup{H}_{j}(\textup{F}_{-j,k-1}^{0},\Upsilon)

    Then, we know by definition of [sj|Υ]k1[s_{j}|\Upsilon]_{k-1}, that sj(h)=sj(h)s_{j}(h)=s_{j}^{\prime}(h). This lets us conclude that z((si;s¯i)|h0)=z((si;s¯i)|h0)z((s_{-i};\bar{s}_{i})|h^{0})=z((s_{-i}^{\prime};\bar{s}_{i})|h^{0}) for every si[si|Υ]k1s_{-i}^{\prime}\in[s_{-i}|\Upsilon]_{k-1}.

  • S4.

    Fix pair (si,ti)(s_{-i},t_{-i}) in the graph of Fi,k0(Υ,)\textup{F}_{-i,k}^{0}(\Upsilon,\,\cdot\,) and si[si|Υ]k1s^{\prime}_{-i}\in[s_{-i}|\Upsilon]_{k-1}. Next, pick player jij\neq i such that there exists some history hHjh\in H_{j} that precedes z((si;si)|h0)z((s_{-i};s_{i})|h^{0}) and is reached by both sjs_{j} and sjs_{j}^{\prime}.383838See Footnote 37; the same logic applies here, mutatis mutandi. We verify first that hHj(Fj,k10,Υ)h\in\textup{H}_{j}(\textup{F}_{-j,k-1}^{0},\Upsilon). To begin, remember that we have that, for player i,j\ell\neq i,j, sF,k0(Υ,t)s_{\ell}\in\textup{F}_{\ell,k}^{0}(\Upsilon,t_{\ell}). Thus:

    hi,jHj(F,k,Υ).i,jHj(F,k1,Υ).h\in\bigcap_{\ell\neq i,j}\textup{H}_{j}(\textup{F}_{\ell,k},\Upsilon).\subseteq\bigcap_{\ell\neq i,j}\textup{H}_{j}(\textup{F}_{\ell,k-1},\Upsilon).

    Next, since siFi,k+2(dik+1,tik+1)s_{i}\in\textup{F}_{i,k+2}(d_{i}^{k+1},t_{i}^{k+1}), in particular, we know that siFi,k+1(dik+1,tik+1)s_{i}\in\textup{F}_{i,k+1}(d_{i}^{k+1},t_{i}^{k+1}). In consequence, since di,k+1k+1=di,k+1d_{i,k+1}^{k+1}=d_{i,k+1} and Fi,k+1\textup{F}_{i,k+1} only depends on the (k+1)th(k+1)^{\textup{th}}-order specification of the directory and the type, there must exist some finite type tiTi(Θ)t_{i}^{\prime}\in T_{i}(\Theta) such that siFi,k+1(Υ,ti)s_{i}\in\textup{F}_{i,k+1}(\Upsilon,t_{i}^{\prime}).393939The possibility of finding finite tit_{i} follows from tik+1t_{i}^{k+1} itself being finite. Thus, we have that:

    hHj(Fi,k1,dik+1)Hj(Fi,k1,Υ)Hj(Fi,k10,Υ)h\in\textup{H}_{j}(\textup{F}_{i,k-1},d_{i}^{k+1})\subseteq\textup{H}_{j}(\textup{F}_{i,k-1},\Upsilon)\subseteq\textup{H}_{j}(\textup{F}_{i,k-1}^{0},\Upsilon)

    the last inclusion being one of the conditions on Υ\Upsilon in the statement of the lemma. Hence we conclude that:

    hjHj(F,k10,Υ)=Hj(Fj,k10,Υ),h\in\bigcap_{\ell\neq j}\textup{H}_{j}(\textup{F}_{\ell,k-1}^{0},\Upsilon)=\textup{H}_{j}(\textup{F}_{-j,k-1}^{0},\Upsilon),

    and thus, by definition of [sj|Υ]k1[s_{j}|\Upsilon]_{k-1}, that sj(h)=sj(h)s_{j}(h)=s_{j}^{\prime}(h). Obviously, the latter implies that z((si;si)|h0)=z((si;si)|h0)z((s_{-i};s_{i})|h^{0})=z((s_{-i}^{\prime};s_{i})|h^{0}) for every si[si|Υ]k1s_{-i}^{\prime}\in[s_{-i}|\Upsilon]_{k-1}.

  • S5.

    Fix history hHi(Fk0,Υ)h\in\textup{H}_{i}(\textup{F}_{k}^{0},\Upsilon). Next pick strategies si1,,siMSi(h)s_{-i}^{1},\dots,s_{-i}^{M}\in S_{-i}(h) such that family {[si1|Υ]k1,[siM|Υ]k1}\{[s_{-i}^{1}|\Upsilon]_{k-1},\dots[s_{-i}^{M}|\Upsilon]_{k-1}\} is a partition of Si(h)S_{-i}(h). Based on the latter, set Si~(h):={[si1|Υ]k1,,[siM|Υ]k1}\widetilde{S_{i}}(h):=\{[s_{-i}^{1}|\Upsilon]_{k-1},\dots,[s_{-i}^{M}|\Upsilon]_{k-1}\} and define the two measures on Si~(h)×Θ0×Θi\widetilde{S_{i}}(h)\times\Theta_{0}\times\Theta_{-i} induced by setting, for any (s~i,θ0,θi)=([si|Υ]k1,θ0,θi)S~i(h)×Θ0×Θi(\tilde{s}_{-i},\theta_{0},\theta_{-i})=([s_{-i}|\Upsilon]_{k-1},\theta_{0},\theta_{-i})\in\widetilde{S}_{-i}(h)\times\Theta_{0}\times\Theta_{-i},

    μ~ik+1[(s~i,θ0,θi)]\displaystyle\widetilde{\mu}_{i}^{k+1}[(\tilde{s}_{-i},\theta_{0},\theta_{-i})] :=(margSi×Θ0×Θiμ¯ik+1(h))[s~i×{θ0,θi}],\displaystyle:=(\textup{marg}_{S_{-i}\times\Theta_{0}\times\Theta_{-i}}\bar{\mu}_{i}^{k+1}(h))[\tilde{s}_{-i}\times\{\theta_{0},\theta_{-i}\}],
    μ~i[(s~i,θ0,θi)]\displaystyle\widetilde{\mu}_{i}[(\tilde{s}_{-i},\theta_{0},\theta_{-i})] :=(margSi×Θ0×Θiμi(h))[s~i×{(θ0,θi)}].\displaystyle:=(\textup{marg}_{S_{-i}\times\Theta_{0}\times\Theta_{-i}}\mu_{i}(h))[\tilde{s}_{-i}\times\{(\theta_{0},\theta_{-i})\}].

    Notice that we know from S2 that μ~i=μ~ik+1\widetilde{\mu}_{i}=\widetilde{\mu}_{i}^{k+1}. It follows then that for any strategy si{s¯i,si}s^{\prime}_{i}\in\{\bar{s}_{i},s_{i}\}:

    Ui\displaystyle U_{i} (μi,si|θi(ti),h)=\displaystyle(\mu_{i},s^{\prime}_{i}|\theta_{i}(t_{i}),h)=
    =Si(h)×Θ0×Θiui(z(si;si),((θ0,θi);θi(ti)))d(margSi×Θ0×Θiμi(h))\displaystyle=\int_{S_{-i}(h)\times\Theta_{0}\times\Theta_{-i}}u_{i}(z(s_{-i};s^{\prime}_{i}),((\theta_{0},\theta_{i});\theta_{i}(t_{i})))\textup{d}(\textup{marg}_{S_{-i}\times\Theta_{0}\times\Theta_{-i}}\mu_{i}(h))
    =S~i(h)×Θ0×Θiui(z(s~i;si),((θ0,θi);θi(ti)))dμ~i\displaystyle=\int_{\widetilde{S}_{-i}(h)\times\Theta_{0}\times\Theta_{-i}}u_{i}(z(\tilde{s}_{-i};s^{\prime}_{i}),((\theta_{0},\theta_{-i});\theta_{i}(t_{i})))\textup{d}\widetilde{\mu}_{i}
    =S~i(h)×Θ0×Θiui(z(s~i;si),((θ0,θi);θi(ti)))dμ~ik+1\displaystyle=\int_{\widetilde{S}_{-i}(h)\times\Theta_{0}\times\Theta_{-i}}u_{i}(z(\tilde{s}_{-i};s^{\prime}_{i}),((\theta_{0},\theta_{-i});\theta_{i}(t_{i})))\textup{d}\widetilde{\mu}_{i}^{k+1}
    =Si(h)×Θ0×Θiui(z(si;si),((θ0,θi);θi(ti)))d(margSi×Θ0×Θiμ¯ik+1(h))\displaystyle=\int_{S_{-i}(h)\times\Theta_{0}\times\Theta_{-i}}u_{i}(z(s_{-i};s^{\prime}_{i}),((\theta_{0},\theta_{-i});\theta_{i}(t_{i})))\textup{d}(\textup{marg}_{S_{-i}\times\Theta_{0}\times\Theta_{-i}}\bar{\mu}_{i}^{k+1}(h))
    =Ui(μ¯ik+1,si|θi(ti),h),\displaystyle=U_{i}(\bar{\mu}_{i}^{k+1},s^{\prime}_{i}|\theta_{i}(t_{i}),h),

    being the second and fourth equalities consequences of the outcome equivalences in S4 and S5 (the abuse of notation z(s~i;si)=z(si;si)z(\tilde{s}_{-i};s_{i}^{\prime})=z(s_{-i};s_{i}^{\prime}) is is innocuous due to S4).

  • S6.

    Proceed by contradiction and suppose that there exists some history hHi(Fk0,Υ)h\in\textup{H}_{i}(\textup{F}_{k}^{0},\Upsilon) such that si(h)s¯i(h)s_{i}(h)\neq\bar{s}_{i}(h). Pick then strategy s^i\hat{s}_{i} that maximizes Ui(,μ¯i|θi(ti),h)U_{i}(\,\cdot\,,\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}) at every history hhh^{\prime}\succ h and define new strategy si0s_{i}^{0} as follows:

    si0(h):={s^i(h)if hh,si(h)h=h,s¯i(h)otherwise.s_{i}^{0}(h^{\prime}):=\left\{\begin{tabular}[]{l l}$\hat{s}_{i}(h^{\prime})$&$\textup{if }h^{\prime}\succ h$,\\ $s_{i}(h^{\prime})$&$h^{\prime}=h$,\\ $\bar{s}_{i}(h^{\prime})$&$\textup{otherwise.}$\end{tabular}\right.

    Let’s check next that si0ri(θi(ti),μ¯i)s_{i}^{0}\in r_{i}(\theta_{i}(t_{i}),\bar{\mu}_{i}). To do it conceive HiH_{i} as the disjoint union of the following four components:

    {hHi{h0}|hh}{h}{hHi{h0}|hh}{hHi{h0}|hh and hh}.\{h^{\prime}\in H_{i}\cup\{h^{0}\}|h^{\prime}\succ h\}\cup\{h\}\cup\{h^{\prime}\in H_{i}\cup\{h^{0}\}|h^{\prime}\prec h\}\cup\{h^{\prime}\in H_{i}\cup\{h^{0}\}|h\npreceq h^{\prime}\textup{ and }h^{\prime}\npreceq h\}.

    We distinguish then four cases:

    • \bullet

      hhh^{\prime}\succ h. The construction of si0s_{i}^{0} ensures that Ui(si0,μ¯i|θi(ti),h)=Ui(s^i,μ¯i|θi(ti),h)U_{i}(s_{i}^{0},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime})=U_{i}(\hat{s}_{i},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}); thus si0s_{i}^{0} must be a maximizer of Ui(,μ¯i|θi(ti),h)U_{i}(\,\cdot\,,\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}).

    • \bullet

      h=hh^{\prime}=h. The above, together with si0(h)=si(h)s_{i}^{0}(h^{\prime})=s_{i}(h^{\prime}), implies that:404040For the first inequality notice that si0s_{i}^{0} and sis_{i} are identical at h=hh^{\prime}=h, and that si0s_{i}^{0} is a maximizer of Ui(,μ¯i|θi(ti),h)U_{i}(\,\cdot\,,\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}) for every h′′hh^{\prime\prime}\succ h^{\prime}. The equality follows from (1) above.

      Ui(si0,μ¯i|θi(ti),h)Ui(si,μ¯i|θi(ti),h)=Ui(si,μ¯ik+1|θi(ti),h).U_{i}(s_{i}^{0},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime})\geq U_{i}(s_{i},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime})\\ =U_{i}(s_{i},\bar{\mu}_{i}^{k+1}|\theta_{i}(t_{i}),h^{\prime}).

      Now, we also have that:414141The first two equalities follow from S5 and the last, from equation (1) above. The inequality is a consequence of sis_{i} being a best response to μi\mu_{i} given θi(ti)\theta_{i}(t_{i}).

      Ui(si,μ¯ik+1|θi(ti),h)\displaystyle U_{i}(s_{i},\bar{\mu}_{i}^{k+1}|\theta_{i}(t_{i}),h^{\prime}) =Ui(si,μi|θi(ti),h)\displaystyle=U_{i}(s_{i},\mu_{i}|\theta_{i}(t_{i}),h^{\prime})
      Ui(s¯i,μi|θi(ti),h)\displaystyle\geq U_{i}(\bar{s}_{i},\mu_{i}|\theta_{i}(t_{i}),h^{\prime})
      =Ui(s¯i,μ¯ik+1|θi(ti),h)\displaystyle=U_{i}(\bar{s}_{i},\bar{\mu}_{i}^{k+1}|\theta_{i}(t_{i}),h^{\prime})
      =Ui(s¯i,μ¯i|θi(ti),h),\displaystyle=U_{i}(\bar{s}_{i},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}),

      and remember that s¯i\bar{s}_{i} is a maximizer of Ui(,μ¯i|θi(ti),h)U_{i}(\,\cdot\,,\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}). Thus, it must necessarily hold that Ui(si0,μ¯i|θi(ti),h)=Ui(s¯i,μ¯i|θi(ti),h)U_{i}(s_{i}^{0},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime})=U_{i}(\bar{s}_{i},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}). Hence, we conclude that si0s_{i}^{0} is a maximizer of Ui(,μ¯i|θi(ti),h)U_{i}(\,\cdot\,,\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}).

    • \bullet

      hhh^{\prime}\prec h. Notice that the construction of si0s_{i}^{0} on the one hand, and the facts that Ui(si0,μ¯i|θi(ti),h)=Ui(s¯i,μ¯i|θi(ti),h)U_{i}(s_{i}^{0},\bar{\mu}_{i}|\theta_{i}(t_{i}),h)=U_{i}(\bar{s}_{i},\bar{\mu}_{i}|\theta_{i}(t_{i}),h) and si0s_{i}^{0} and s¯i\bar{s}_{i} are equivalent at any history preceding hh, on the other, imply that Ui(si0,μ¯i|θi(ti),h)=Ui(s¯i,μ¯i|θi(ti),h)U_{i}(s_{i}^{0},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime})=U_{i}(\bar{s}_{i},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}). Thus, since s¯i\bar{s}_{i} maximizes Ui(,μ¯i|θi(ti),h)U_{i}(\,\cdot\,,\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}) it follows that si0s_{i}^{0} must maximize Ui(,μ¯i|θi(ti),h)U_{i}(\,\cdot\,,\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}) as well.

    • \bullet

      For any other hh^{\prime}, clearly, we have that:

      Ui(si0,μ¯i|θi(ti),h)=Ui(s¯i,μ¯i|θi(ti),h).U_{i}(s_{i}^{0},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime})=U_{i}(\bar{s}_{i},\bar{\mu}_{i}|\theta_{i}(t_{i}),h^{\prime}).

    We have then reached the following contradictory conclusions: si0[s¯i|Υ]ks_{i}^{0}\notin[\bar{s}_{i}|\Upsilon]_{k} and si0ri(θi(ti),μ¯i)=[s¯i|Υ]ks_{i}^{0}\in r_{i}(\theta_{i}(t_{i}),\bar{\mu}_{i})=[\bar{s}_{i}|\Upsilon]_{k}.

Hence, if sis_{i} is a best response to μi\mu_{i} for θi(ti)\theta_{i}(t_{i}) it must hold that si(h)=s¯i(h)s_{i}(h)=\bar{s}_{i}(h) for every hHi(Fk0,Υ)h\in\textup{H}_{i}(\textup{F}_{k}^{0},\Upsilon), that is,

Fi,k+2(dik+1,tik+1)[s¯i|Υ]k,\textup{F}_{i,k+2}(d_{i}^{k+1},t_{i}^{k+1})\subseteq[\bar{s}_{i}|\Upsilon]_{k},

and in consequence, the proof is finally complete. ∎

Proofs of the auxiliary results

Lemma 6: Availability of finite conjectures

Lemma 6.

Let Γ\Gamma be an extensive form and Υ\Upsilon, a standard payoff structure. Then for any k0k\geq 0, any player ii, any finite type tiTi(Θ)t_{i}\in T_{i}(\Theta) and every strategy siFi,k(Υ,ti)s_{i}\in\textup{F}_{i,k}(\Upsilon,t_{i}) there exists a finite conjecture μi\mu_{i} that justifies the inclusion of sis_{i} in Fi,k(Υ,ti)\textup{F}_{i,k}(\Upsilon,t_{i}).

Proof.

Fix k0k\geq 0, player ii, finite type Ti(Θ)T_{i}(\Theta) and strategy siFi,k+1(Υ,ti)s_{i}\in\textup{F}_{i,k+1}(\Upsilon,t_{i}) and pick conjecture μi\mu_{i} that justifies the inclusion of sis_{i} in Fi,k+1(Υ,ti)\textup{F}_{i,k+1}(\Upsilon,t_{i}). For convenience, let us denote Xi:=Θ0×Ti(Θ)X_{i}:=\Theta_{0}\times T_{-i}(\Theta). Since XiX_{i} is separable, we ca pick countable (xim)m(x_{i}^{m})_{m\in\mathds{N}} where {xim}m\{x_{i}^{m}\}_{m\in\mathds{N}} is dense in XiX_{i}.

Now, for each m,nm,n\in\mathds{N} let B(xim,1/n)B(x_{i}^{m},1/n) denote the ball of radius 1/n1/n around ximx_{i}^{m}. We know because of the upper hemicontinuity of Fi,(Υ,)\textup{F}_{-i,\ell}(\Upsilon,\,\cdot\,) for each =0,1,,k\ell=0,1,\dots,k that for each n,mn,m\in\mathds{N} there exists some open set Wi,n,mB(xim,1/n)W_{i}^{\ell,n,m}\subseteq B(x_{i}^{m},1/n) such that Fi,(Υ,ti)Fi,(Υ,tim)\textup{F}_{-i,\ell}(\Upsilon,t_{-i})\subseteq\textup{F}_{-i,\ell}(\Upsilon,t_{-i}^{m}) for every tit_{-i} is in the projection of Wi,n,mW_{i}^{\ell,n,m} on Ti(Θ)T_{-i}(\Theta), timt_{-i}^{m} being the projection on Ti(Θ)T_{-i}(\Theta) of ximx_{i}^{m}. Furthermore, for each =0,1,,k\ell=0,1,\dots,k and nn\in\mathds{N}, the family {Wi,n,m}m\{W_{i}^{\ell,n,m}\}_{m\in\mathds{N}} is an open cover of XiX_{i}, which, due to XiX_{i} being compact, we can assume as finite: {Wi,n,m}m=1M,n\{W_{i}^{\ell,n,m}\}_{m=1}^{M_{\ell,n}}.

Now, for each =0,1,,k\ell=0,1,\dots,k and nn\in\mathds{N} set:

Vi,n,m:=Wi,n,mr=1m1Wi,n,r.V_{i}^{\ell,n,m}:=W_{i}^{\ell,n,m}\setminus\bigcup_{r=1}^{m-1}W_{i}^{\ell,n,r}.

Notice that for each =0,1,,k\ell=0,1,\dots,k and nn\in\mathds{N} family {Vi,n,m}m=1M,n\{V_{i}^{\ell,n,m}\}_{m=1}^{M_{\ell,n}} is a partition of XiX_{i} consisting of measurable sets, contained in a ball of radius 1/n1/n. Since the set of finite types is dense in XiX_{i} for each =0,1,,k\ell=0,1,\dots,k and nn\in\mathds{N}, there exists some list (yi,n,m)m=1M,n(y_{i}^{\ell,n,m})_{m=1}^{M_{\ell,n}} such that, for each m=1,,M,nm=1,\dots,M_{\ell,n}, the projection on Ti(Θ)T_{-i}(\Theta) of yi,n,my_{i}^{\ell,n,m} is finite and yi,n,mVi,n,my_{i}^{\ell,n,m}\in V_{i}^{\ell,n,m}.

We turn now back to μi\mu_{i}. Let us denote Hi,k(μi):=Hi(μi)Hi(Fi,k,Θ)\textup{H}_{i,k}(\mu_{i}):=\textup{H}_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,k},\Theta) and then, define recursive, for each =0,1,,k1\ell=0,1,\dots,k-1, set:

Hi,(μi):=(Hi(μi)Hi(Fi,,Υ))Hi,k(μi).\textup{H}_{i,\ell}(\mu_{i}):=\left(\textup{H}_{i}(\mu_{i})\cap\textup{H}_{i}(\textup{F}_{-i,\ell},\Upsilon)\right)\setminus\textup{H}_{i,k}(\mu_{i}).

Next, we will construct a conditional probability system μin\mu_{i}^{n} for each nn\in\mathds{N}. First, for each each =0,1,,k\ell=0,1,\dots,k and each hHi,(μi){h0}h\in\textup{H}_{i,\ell}(\mu_{i})\setminus\{h^{0}\} set:

μin(h)[(si,yi,m)]:=μi(h)[{si}×Vi,n,m],\mu_{i}^{n}(h)[(s_{-i},y_{i}^{\ell,m})]:=\mu_{i}(h)[\{s_{-i}\}\times V_{i}^{\ell,n,m}],

for every siSis_{-i}\in S_{-i} and every m=1,,M,nm=1,\dots,M_{\ell,n}. Second, set μin(h0):=μi(h0)\mu_{i}^{n}(h^{0}):=\mu_{i}(h^{0}). Finally, for each hHi(μi)h\notin\textup{H}_{i}(\mu_{i}) define μin(h)\mu_{i}^{n}(h) using the chain rule. Notice that the marginals on SiS_{-i} of μi(h)\mu_{i}(h) and each μin(h)\mu_{i}^{n}(h) coincide for every history hh, and this guarantees that μin\mu_{i}^{n} is (or, has been) well-defined.

Now, notice also that, for each =0,1,,k\ell=0,1,\dots,k, μin(h)\mu_{i}^{n}(h) assigns probability one to the graph of Fi,(Υ,)\textup{F}_{-i,\ell}(\Upsilon,\,\cdot\,). Obviously, every μim\mu_{i}^{m} is consistent with type tit_{i}, and sequence (μin)n(\mu_{i}^{n})_{n\in\mathds{N}} converges to μi\mu_{i}. Thus, the upper hemicontinuity of ri(θi(ti),)r_{i}(\theta_{i}(t_{i}),\,\cdot\,) ensures the existence of some NN\in\mathds{N} such that siri(θi(ti),μin)s_{i}\in r_{i}(\theta_{i}(t_{i}),\mu_{i}^{n}) for every nNn\geq N. Hence, every μin\mu_{i}^{n} where nNn\geq N is a finite conjecture that justifies the inclusion of sis_{i} in Fi,k+1(Υ,ti)\textup{F}_{i,k+1}(\Upsilon,t_{i})

Lemma 7: A partial upper hemicontinuity result

Lemma 7.

Let Γ\Gamma be an extensive form and Υ\Upsilon, a standard payoff structure. Then for any k0k\geq 0, any player ii there exists some nkin_{k}^{i}\in\mathds{N} such that the following three hold:

  • (i)(i)

    For every type tiTi(Θ)t_{i}\in T_{i}(\Theta), Fi,k(Υ,ti)Fi,k(Υn,ti)Fi,k(Υn+,ti)\textup{F}_{i,k}(\Upsilon,t_{i})\subseteq\textup{F}_{i,k}(\Upsilon^{n},t_{i})\subseteq\textup{F}_{i,k}(\Upsilon^{n+\ell},t_{i}) for every nnkin\geq n_{k}^{i} and every \ell\in\mathds{N}.

  • (ii)(ii)

    For every type tiTi(Θ)t_{i}\in T_{i}(\Theta), nnkiFi,k(Υn,ti)Fi,k(Υ,ti)\bigcap_{n\geq n_{k}^{i}}\textup{F}_{i,k}(\Upsilon^{n},t_{i})\subseteq\textup{F}_{i,k}(\Upsilon,t_{i}).

  • (iii)(iii)

    For every player jij\neq i, Hi(Fj,k,Υn)=Hi(Fj,k,Υ)\textup{H}_{i}(\textup{F}_{j,k},\Upsilon^{n})=\textup{H}_{i}(\textup{F}_{j,k},\Upsilon) for every nnkin\geq n_{k}^{i}.

Proof.

We proceed by induction on kk. The claims hold trivially for the initial case (k=0k=0) so we can focus on the proof of the inductive step. Suppose that the claims hold for kk; we verify now that they also hold for k+1k+1:

Claim (i)(i). Fix player ii and set nk+1i,1:=max{nj|=0,,k and jI}n_{k+1}^{i,1}:=\textup{max}\{n_{\ell}^{j}|\ell=0,\dots,k\textup{ and }j\in I\}. We know from part (i)(i) of the induction hypothesis that for every nnk+1i,1n\geq n_{k+1}^{i,1} and every \ell\in\mathds{N},

Fi,k(Υ,ti)Fi,k(Υn,ti)Fi,k(Υn+,ti),\textup{F}_{i,k}(\Upsilon,t_{i})\subseteq\textup{F}_{i,k}(\Upsilon^{n},t_{i})\subseteq\textup{F}_{i,k}(\Upsilon^{n+\ell},t_{i}),

for every type tiTi(Θ)t_{i}\in T_{i}(\Theta) and,

Graph(Fi,k(Υ,))Graph(Fi,k(Υn,))Graph(Fi,k(Υn+,)).\textup{Graph}\left(\textup{F}_{-i,k}(\Upsilon,\,\cdot\,)\right)\subseteq\textup{Graph}\left(\textup{F}_{-i,k}(\Upsilon^{n},\,\cdot\,)\right)\subseteq\textup{Graph}\left(\textup{F}_{-i,k}(\Upsilon^{n+\ell},\,\cdot\,)\right).

We also know from part (iii)(iii) that,

Hi(Fi,k,Υn+)=Hi(Fi,k,Υn)=Hi(Fi,k,Υ).\textup{H}_{i}(\textup{F}_{-i,k},\Upsilon^{n+\ell})=\textup{H}_{i}(\textup{F}_{-i,k},\Upsilon^{n})=\textup{H}_{i}(\textup{F}_{-i,k},\Upsilon).

Fix now nnk+1i,1n\geq n_{k+1}^{i,1}, \ell\in\mathds{N}, type tiTi(Θ)t_{i}\in T_{i}(\Theta), strategy siFi,k+1(Υ,ti)s_{i}\in\textup{F}_{i,k+1}(\Upsilon,t_{i}) and conjecture μi\mu_{i} that justifies the inclusion of sis_{i} in Fi,k+1(Υ,ti)\textup{F}_{i,k+1}(\Upsilon,t_{i}). It follows from the above that siFi,k(Υn,ti)s_{i}\in\textup{F}_{i,k}(\Upsilon^{n},t_{i}) and that μi\mu_{i} is a conjecture that justifies the inclusion of sis_{i} in Fi,k+1(Υn,ti)\textup{F}_{i,k+1}(\Upsilon^{n},t_{i}). Fix now sinFi,k+1(Υn,ti)s_{i}^{n}\in\textup{F}_{i,k+1}(\Upsilon^{n},t_{i}) and conjecture μin\mu_{i}^{n} that satisfies the inclusion of sins_{i}^{n} in Fi,k+1(Υn,ti)\textup{F}_{i,k+1}(\Upsilon^{n},t_{i}). Again, it clearly follows from the above that siFi,k(Υn+,ti)s_{i}\in\textup{F}_{i,k}(\Upsilon^{n+\ell},t_{i}) and that μin\mu_{i}^{n} is a conjecture that satisfies the inclusion of sins_{i}^{n} in Fi,k+1(Υn+,ti)\textup{F}_{i,k+1}(\Upsilon^{n+\ell},t_{i}) (see the inclusion property in comment (B) in Section B.3.1). \bigstar

Claim (ii)(ii). Fix player ii and set again nk+1i,1:=max{nkj|=0,,k and jI}n_{k+1}^{i,1}:=\textup{max}\{n_{k}^{j}|\ell=0,\dots,k\textup{ and }j\in I\}. Fix type tiTi(Θ)t_{i}\in T_{i}(\Theta) and strategy sinnk+1i,2Fi,k+1(Υn,ti)s_{i}\in\bigcap_{n\geq n_{k+1}^{i,2}}\textup{F}_{i,k+1}(\Upsilon^{n},t_{i}), and take convergent sequence of conjectures (μim)mnk+1i,2(\mu_{i}^{m})_{m\geq n_{k+1}^{i,2}} such that, for each mnk+1i,2m\geq n_{k+1}^{i,2}, μim\mu_{i}^{m} justifies the inclusion of sis_{i} in Fi,k+1(Υnm,ti)\textup{F}_{i,k+1}(\Upsilon^{n_{m}},t_{i}) (being nmmn_{m}\geq m). Let μi\mu_{i} denote the limit of this sequence. Notice that we know from part (ii)(ii) of the induction hypothesis that:

sinnk+1i,2Fi,k(Υn,ti)Fi,k(Υ,ti).s_{i}\in\bigcap_{n\geq n_{k+1}^{i,2}}\textup{F}_{i,k}(\Upsilon^{n},t_{i})\subseteq\textup{F}_{i,k}(\Upsilon,t_{i}).

Now, set =0,,k\ell=0,\dots,k and pick history hHi(Fi,,Υ)h\in\textup{H}_{i}(\textup{F}_{-i,\ell},\Upsilon). Notice that:424242The first implication follows from parts (i)(i) and (iii)(iii) of the induction hypothesis. The second is a consequence of the graph of Fi,(Υnm,)\textup{F}_{-i,\ell}(\Upsilon^{n_{m}},\,\cdot\,) being closed. The third and the fourth inclusions are obvious and the last one follows from part (ii)(ii) the induction hypothesis.

mnk+1i,2,μim\displaystyle\forall m\geq n_{k+1}^{i,2},\,\mu_{i}^{m} (h)[Θ0nm×Graph(Fi,(Υnm,))]=1\displaystyle(h)[\Theta_{0}^{n_{m}}\times\textup{Graph}\left(\textup{F}_{-i,\ell}\left(\Upsilon^{n_{m}},\,\cdot\,\right)\right)]=1\Longrightarrow
\displaystyle\Longrightarrow mnk+1i,2,r0,μim+r(h)[Θ0nm×Graph(Fi,(Υnm,))]=1\displaystyle\,\forall m\geq n_{k+1}^{i,2},\forall r\geq 0,\,\mu_{i}^{m+r}(h)\left[\Theta_{0}^{n_{m}}\times\textup{Graph}\left(\textup{F}_{-i,\ell}\left(\Upsilon^{n_{m}},\,\cdot\,\right)\right)\right]=1
\displaystyle\Longrightarrow mnk+1i,2,limrμim+r(h)[Θ0nm×Graph(Fi,(Υnm,))]=1\displaystyle\,\forall m\geq n_{k+1}^{i,2},\,\underset{r\rightarrow\infty}{\textup{lim}}\mu_{i}^{m+r}(h)\left[\Theta_{0}^{n_{m}}\times\textup{Graph}\left(\textup{F}_{-i,\ell}\left(\Upsilon^{n_{m}},\,\cdot\,\right)\right)\right]=1
\displaystyle\Longrightarrow mnk+1i,2,μi(h)[Θ0×Graph(Fi,(Υnm,))]=1\displaystyle\,\forall m\geq n_{k+1}^{i,2},\,\mu_{i}(h)\left[\Theta_{0}\times\textup{Graph}\left(\textup{F}_{-i,\ell}\left(\Upsilon^{n_{m}},\,\cdot\,\right)\right)\right]=1
\displaystyle\Longrightarrow μi(h)[Θ0×mnk+1i,2Graph(Fi,(Υnm,))]=1\displaystyle\,\mu_{i}(h)\left[\Theta_{0}\times\bigcap_{m\geq n_{k+1}^{i,2}}\textup{Graph}\left(\textup{F}_{-i,\ell}\left(\Upsilon^{n_{m}},\,\cdot\,\right)\right)\right]=1
\displaystyle\Longrightarrow μi(h)[Θ0×Graph(Fi,(Υ,))]=1.\displaystyle\,\mu_{i}(h)\left[\Theta_{0}\times\textup{Graph}\left(\textup{F}_{-i,\ell}\left(\Upsilon,\,\cdot\,\right)\right)\right]=1.

Given the above, upper hemicontinuity of the best response operator ensures that μi\mu_{i} is a conjecture that justifies the inclusion of sis_{i} in Fi,k+1(Υ,ti)\textup{F}_{i,k+1}(\Upsilon,t_{i}). \bigstar

Claim (iii)(iii). Fix players ii and jij\neq i. It follow from part (i)(i) that Hi(Fj,k+1,Υ)Hi(Fj,k+1,Υn)\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon)\subseteq\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon^{n}) for every nnk+1j,1n\geq n_{k+1}^{j,1} (the latter has been defined in the proofs of the previous two claims). To prove the reverse inclusion simply notice that, since Hi{h0}H_{i}\cup\{h^{0}\} is finite, it follows from part (ii)(ii) that there must exist some nk+1j,n_{k+1}^{j,*}\in\mathds{N} such that:434343To better see it suppose by contradiction that for any mm\in\mathds{N} there exists some nmnn_{m}\geq n such that hHi(Fj,k+1,Υnm)h\in\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon^{n_{m}}). Then, we know from part (i)(i) that there exists some m¯\bar{m} such that hHi(Fj,k+1,Υm)h\in\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon^{m}) for every mm¯m\geq\bar{m} and thus, it follows from part (ii)(ii) that hHi(Fj,k+1,Υ).h\in\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon).

hHi(Fj,k+1,Υ)hHi(Fj,k+1,Υn)h\notin\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon)\Longrightarrow h\notin\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon^{n})

for every nnk+1j,n\geq n_{k+1}^{j,*}. Hence, it follows that Hi(Fj,k+1,Υn)=Hi(Fj,k+1,Υ)\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon^{n})=\textup{H}_{i}(\textup{F}_{j,k+1},\Upsilon) for every nnk+12:=max{max{nk+1j,1,nk+1j,}|jI}n\geq n_{k+1}^{2}:=\textup{max}\{\textup{max}\{n_{k+1}^{j,1},n_{k+1}^{j,*}\}|j\in I\}. \bigstar

Thus, to finish the proof simply set nk+1i:=max{nk+1i,1,nk+1i,2}n_{k+1}^{i}:=\textup{max}\{n_{k+1}^{i,1},n_{k+1}^{i,2}\}. ∎

References