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

Markovian Persuasion

Ehud Lehrer Tel Aviv University, Tel Aviv 69978, Israel. [email protected]  and  Dimitry Shaiderman Tel Aviv University, Tel Aviv 69978, Israel. [email protected]
Abstract.

In the classical Bayesian persuasion model an informed player and an uninformed one engage in a static interaction. The informed player, the sender, knows the state of nature, while the uninformed one, the receiver, does not. The informed player partially shares his private information with the receiver and the latter then, based on her belief about the state, takes an action. This action determines, together with the state of nature, the utility of both players. We consider a dynamic Bayesian persuasion situation where the state of nature evolves according to a Markovian law. In this repeated persuasion model an optimal disclosure strategy of the sender should, at any period, balance between getting high stage payoff and future implications on the receivers’ beliefs. We discuss optimal strategies under different discount factors and characterize when the asymptotic value achieves the maximal value possible.

The first author acknowledges the support of grants ISF 591/21 and DFG KA 5609/1-1.

Date:
JEL Classification: D72, D82, D83, K40, M31
Keywords: Markovian persuasion, Dynamic Bayesian persuasion, Markov chain, Asymptotic value, Absorbing set, Homothety.

1. Introduction

The literature devoted to Bayesian persuasion studies optimal policies by which an informed agent, the sender, discloses information to an uninformed agent, the receiver. Kamenika and Gentzkow [16] present a case where a prosecutor, who is fully informed about the state of nature, i.e., whether the suspect is guilty or innocent, wishes to persuade a judge, e.g to convict the suspect of a crime. This is a static scenario: upon the prosecutor’s disclosure, the judge takes a decision and the game is over.

In this paper we study a dynamic model where the interaction between the sender, who has a commitment power, and the receiver evolves over time. At each period nature randomly chooses a state and the sender gets informed about it. The decision of how much information to disclose to the receiver at each stage is the choice of the sender. The latter publicly announces an information provision policy, to which he is committed throughout. The receiver knows the details of this policy and thus, when she receives a signal from the sender she updates her belief about the true state. She then takes an action that is based solely on her posterior belief. This action, together with the realized state, determines not only her own payoff, but also that of the sender. In short, in this dynamic signaling interaction, the sender is in charge of the informational aspects of the interaction (talking), while the receiver is in charge of doing. Neither may affect the evolution of the realized states, which is entirely exogenous.

We assume that the evolution of states follows a Markov chain. The dynamic of the receiver’s belief is governed by both the disclosure policy of the sender and a Markov transition matrix. After observing the signal sent by the sender, the receiver updates her belief according to Bayes’ law. Due to the Markovian transition law, this belief shifts to another belief. Since all terms are known to both players, the nature of the shift from prior belief to a posterior one is also known to both.

When committing to an information provision policy, the sender takes into account the resulting posterior belief of the receiver, which has two effects. The first is on the action taken by the receiver and directly on his own stage payoff. In a dynamic setting, as discussed here, it has also an effect on the belief over future states, and consequently on future payoffs. The posterior resulting from a Bayesian updating in one period is then shifted by the Markov transition matrix and becomes the initial belief in the next one. Balancing between these two, potentially contradicting effects, is the dynamic programming problem faced by the sender.

In this paper we study the long-run optimal values that the sender may achieve in an irreducible and aperiodic Markov chain.111In a companion paper we examine, based on the current study, the general case. This value depends on his discount factor. As in Aumann and Maschler [6] and in Kamenika and Gentzkow [16], the optimal signaling strategy follows a splitting scheme of beliefs in accordance with the concavification of a certain function222That is, the lowest concave function above a given function.. While in a static model this function is the payoff function of the sender, here, this function takes account of the underlying dynamics and combines present and future payoffs.

The first set of results deal with a patient sender. It turns out that the stationary distribution of the Markov chain plays a central role. The reason is that in case the sender keeps mute, the receiver’s beliefs become converge, very fast,333At an exponential rate (see e.g., Theorem 4.9 in [20]). to the stationary distribution.

We show that as the sender becomes increasingly patient,444The method of treating patient players in not new to the literature of dynamic interactions. Fudenberg and Maskin [11] introduced a Folk Theorem characterizing the set of equilibria in a repeated game played by patient players. Typically in these kind of games, it is impossible to provide a description of the value or of the set of equilibria that correspond to a specific discount factor more informative than that given by Bellman equation (see Abreu et. al. [1]). the optimal values converges to a certain level, called the asymptotic value. Moreover, this value cannot exceed the optimal level obtained in the static case when the prior belief is the stationary distribution. A natural question is under what conditions this static optimal value can be obtained in the dynamic model. Our main theorem provides a full characterization. In order to describe it we introduce the notion of an absorbing set.

A set CC of prior distributions is said to be absorbing, if any distribution pp in CC evolves by the Markov law to another distribution which is within the convex boundaries of CC. The stationary distribution, for instance, is absorbing as it shifts to itself. The main theorem roughly asserts that the asymptotic value is as high as it can get (i.e., the static value at the stationary distribution), precisely when there is an absorbing set by which the optimal static value is defined (as a weighted average). The intuition behind this result is the following. If any set AA by means of which the optimal static value is defined is non-absorbing, then regardless of the policy the sender uses, the posterior beliefs of the receiver would exit the boundaries of AA infinitely often, at times having positive limit frequency. Upon each such exit, the sender’s payoff would take a hit which he would not be able to compensate for in subsequent time periods. Therefore, the dream scenario for a patient sender is to find a set AA by means of which the optimal static value is defined, so that he could confine the posteriors beliefs of the receiver within the convex boundaries of AA throughout the entire game.

The second type of results is non-asymptotic in nature. For a certain region around the stationary distribution we provide a closed form expression for the values corresponding to any level of patience. In the case where this region includes a neighborhood of the stationary distribution, we give asymptotically effective555That is, bounds whose difference is arbitrarily small for a patient enough sender. two-sided bounds for the values corresponding to any level of patience. Moreover, the effectiveness of those bounds depends on the geometry and size of the described neighborhood.

The closest paper to the present one is that of Renault et. al. [22]. It deals with a specific type of Markov chains, homothetic ones, and with a utility function which gives a fixed amount to the sender in a certain region and zero in its complement. Renault et. al. [22] show that in this model the optimal information provision policy is a greedy one (starting from a certain random time period on). Namely, the sender’s greedy policy instructs him to maximize his current stage payoff at any time period, ignoring future effects of this policy. Also closely related to our work are the works of Ely [9] and Farhadi and Teneketzis [12], which deal with a case where there are two states, whose evolution is described by a Markov chain with one absorbing state. In these two works, the receiver is interested in detecting the jump to the absorbing state, whereas the sender seeks to prolong the time to detection of such a jump.

In a broader context, our paper can be ascribed to a vast growing literature on dynamic information design problems, specifically with the analysis of situations in which the sender is required to convey signals sequentially to a receiver, which he may base on his additional private information. Without dwelling upon the exact details of their models, whose frameworks cover diverse methodological approaches, we refer the reader to Mailath and Samuelson [19], Honryo [14], Phelan [21], Lingenbrink and Iyer [18], Wiseman [26], Athey and Bagwell [4], Arieli and Babichenko [2], Arieli et. al. [3], Dziuda and Gradwohl [8], Escobar and Toikka [10], Ganglmair and Tarantino [13], and Augenblick and Bodoh-Creed [5] for an acquaintance with the literature.

The paper is organized as follows. Section 2 presents the model and an example by which we explain the notations, new concepts and the main results. The asymptotic results as well as the main theorem are provided in Section 3. Results related to non-asymptotic values are given in Section 4. Section 5 provides a characterization of homothetic matrices in terms of the asymptotic value. The proofs are given in Section 6.


2. The Model

Let K={1,,k}K=\{1,...,k\} be a finite set of states. Assume that (Xn)n1(X_{n})_{n\geq 1} is an irreducible and aperiodic Markov chain over KK with prior probability pΔ(K)p\in\Delta(K) and a transition rule given by the stochastic matrix MM. We assume that (Xn)n1(X_{n})_{n\geq 1} are defined on the probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}).

A sender is an agent who is informed at each time period nn of the realized value xnx_{n} of XnX_{n}. Upon obtaining this information a sender is prescribed to send a signal sns_{n}, from a finite set of signals SS with cardinality at least kk.666This assumption is in place to make sure the sender can disclose (xn)n(x_{n})_{n}.

A receiver is an agent who, at any time period nn, is instructed to make a decision bnb_{n} from a set of possible set of decisions BB, assumed to be a compact metric space. This decision may take into account the first nn signals s1,,sns_{1},...,s_{n} of the sender.

The payoffs of the sender and the receiver at time period nn are given by the utilities v(xn,bn)v(x_{n},b_{n}) and w(xn,bn)w(x_{n},b_{n}), respectively, so that they depend solely on the realized state xnx_{n} and the decision bnb_{n}. Both the sender and the receiver discount their payoffs by a factor λ[0,1)\lambda\in[0,1). We denote this game by Γλ(p)\Gamma_{\lambda}(p). As in the models of Renault et. al. [22], Ely [9], and Farhadi and Teneketzis [12] the receiver receives information only through the sender.

A signaling strategy σ\sigma of the sender in Γλ(p)\Gamma_{\lambda}(p) is described by a sequence of stage strategies (σn)(\sigma_{n}), where each σn\sigma_{n} is a mapping σn:(K×S)n1×KΔ(S).\sigma_{n}:(K\times S)^{n-1}\times K\to\Delta(S). Thus, the signal sns_{n}, sent by the sender at time nn is distributed by the lottery σn\sigma_{n}, which may depend on all past states x1,,xn1x_{1},...,x_{n-1} and past signals s1,,sns_{1},...,s_{n} together with the current state xnx_{n}. Let Σ\Sigma be the space of all signaling strategies.

A standard assumption in many Bayesian persuasion models is that of commitment by the sender. That is, we assume that the sender commits to a signaling strategy σ\sigma at the start of the game Γλ(p)\Gamma_{\lambda}(p), and makes it known to the receiver. The commitment assumption enables the receiver to update her beliefs on the distribution of states (Xn)(X_{n}) based on the signals (sn)(s_{n}) he receives from the sender. Formally, by Kolmogorov’s Extension Theorem, each signaling strategy σ\sigma together with (Xn)n1(X_{n})_{n\geq 1} induce a unique probability measure p,σ\mathbb{P}_{p,\sigma} on the space 𝒴=(K×S)\mathcal{Y}=(K\times S)^{\mathbb{N}}, determined by the laws

(2.1) p,σ(x1,s1,,xn,sn)=(p(x1)i=1n1Mxi,xi+1)×(i=1nσi(x1,s1,,xi1,si1,xi)(si)).\mathbb{P}_{p,\sigma}(x_{1},s_{1},...,x_{n},s_{n})=\left(p(x_{1})\prod_{i=1}^{n-1}M_{x_{i},x_{i+1}}\right)\times\\ \left(\prod_{i=1}^{n}\sigma_{i}(x_{1},s_{1},...,x_{i-1},s_{i-1},x_{i})(s_{i})\right).

Thus, the posterior probability pnp_{n}^{\ell} the receiver assigns to the event {Xn=}\{X_{n}=\ell\}, given the signals s1,,sns_{1},...,s_{n} and the strategy σ\sigma, is given by the formula

(2) pn=p,σ(Xn=|s1,,sn).p_{n}^{\ell}=\mathbb{P}_{p,\sigma}\left(X_{n}=\ell\,|\,s_{1},...,s_{n}\right).

Set pn=(pn)Kp_{n}=(p_{n}^{\ell})_{\ell\in K}. A second key assumption of our model is that the receiver’s decision at any time period nn depends only on pnp_{n}. Such an assumption includes the natural situation in which the receiver seeks to maximize her expected payoff based on his current belief (e.g., Renault et. al. (2015)). Denote by θ:Δ(K)B\theta:\Delta(K)\to B the decision policy of the receiver, that is, the mapping which depicts the decision of the receiver as a function of his belief. The last assumption of our model is that the function u:Δ(K)u:\Delta(K)\to\mathbb{R} defined by u(q)=Kqv(,θ(q))u(q)=\sum_{\ell\in K}q^{\ell}v(\ell,\theta(q)) is continuous. To summarize, our assumptions imply that the signaling strategy σ\sigma of the sender determines his payoff at any time period nn. Moreover, the total expected payoff to the sender in Γλ(p)\Gamma_{\lambda}(p) under the signaling strategy σ\sigma can now be written as

(3) γλ(p,σ):=𝔼p,σ[(1λ)n=1λn1u(pn)],\gamma_{\lambda}(p,\sigma):=\mathbb{E}_{p,\sigma}\left[(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}u(p_{n})\right],

where 𝔼p,σ\mathbb{E}_{p,\sigma} is the expectation w.r.t. p,σ\mathbb{P}_{p,\sigma}. The value of the game Γλ(p)\Gamma_{\lambda}(p) is vλ(p)=supσΣγλ(p,σ)v_{\lambda}(p)=\sup_{\sigma\in\Sigma}\gamma_{\lambda}(p,\sigma).

Example 1.

Assume that KK consists of two states, HH and LL. Each probability measure (p,1p)(p,1-p) over KK is determined by a parameter p[0,1]p\in[0,1], corresponding to the distribution mass assigned to the state HH by the respective probability measure. Suppose that the receiver’s set of decisions BB is [0,1][0,1] and that when his belief is (p,1p)(p,1-p) his decision policy θ\theta chooses pp.

As for the sender, when the decision made by the receiver is pp and the state is HH, his utility is 22 minus 33 times the distance between pp and 12\frac{1}{2}. That is, v(H,p)=23|p12|v(H,p)=2-3|p-\frac{1}{2}|. In words, when HH is the realized state, the closer the decision of the receiver to 12\frac{1}{2}, the higher the payoff of the sender. However, when the decision made by the receiver is pp and the state is LL, his utility is p/10p/10. That is, v(L,p)=p/10v(L,p)=p/10. When LL is the realized state, the closer the decision of the receiver to 11, the higher the payoff of the sender.

We obtain that u((p,1p))=pv(H,p)+(1p)v(L,p)=p(23|p12|)+(1p)p/10u((p,1-p))=pv(H,p)+(1-p)v(L,p)=p(2-3|p-\frac{1}{2}|)+(1-p)p/10. The graph of uu, which for the sake of convenience is plotted as a function of pp, can be found on the left panel of Figure 1. Thus, uu is convex on the interval [0,0.5][0,0.5] and concave on the interval [0.5,1][0.5,1]. This graph illustrates the short-term strategic incentives of the sender. Indeed, the minimal possible stage payoff, 0, occurs when the sender reveals that the realized state is LL (corresponding to p=0p=0). A revelation of the state HH (corresponding to p=1p=1) would result in a payoff of 0.50.5. The maximal payoff for the sender, 1.0451.045, is reached for p=0.581p=0.581. As explained below (see Example 3), for that pp the optimal signaling strategy would instruct the sender to not reveal any information regarding the realized state.

In a dynamic model the amount of information revealed by the sender is a result of an interplay between the one-shot payoff, uu, and the transition law governing the evolution of future states. The tension between these two factors is discussed in the rest of this paper.

Refer to caption
Figure 1. The graphs of uu and (Cavu\text{Cav}\,u)

3. The Main Theorem

3.1. The Existence of Asymptotic Value.

To state our first result we need to introduce some notations. First, let πM\pi_{M} be the unique stationary distribution of MM. Second, for any function g:Δ(K)g:\Delta(K)\to\mathbb{R}, define the function (Cavg)(\text{Cav}\,g) by

(Cavg)(q):=inf{h(q):h:Δ(K)concave,hg},qΔ(K).(\text{Cav}\,g)(q):=\inf\{h(q)\,:\,h:\Delta(K)\to\mathbb{R}\,\,\text{concave},\,h\geq g\},\ \ \ \ \forall q\in\Delta(K).

To showcase the Cav operator in action, we show the graph of (Cavu\text{Cav}\,u), where uu is that given in Example 1, on the right-hand panel of Figure 1.

Our first result reveals that the effect of pΔ(K)p\in\Delta(K) on the value vλ(p)v_{\lambda}(p) of a sufficiently patient sender, i.e., with λ\lambda close to 11, is negligible compared to the effects of uu and MM. Moreover, as the patience level λ\lambda gets closer to 11 the sequences vλ(p)v_{\lambda}(p) converge for any pp. Formally, this result is stated as follows.

Theorem 1.

There exists a scalar vv_{\infty}\in\mathbb{R}, v(Cavu)(πM)v_{\infty}\leq(\emph{Cav}\,u)(\pi_{M}), such that for every ε>0\varepsilon>0 and every pΔ(K)p\in\Delta(K) there exists 0<δ<10<\delta<1 such that

(4) |vλ(p)v|<ε,λ>δ.|v_{\lambda}(p)-v_{\infty}|<\varepsilon,\ \ \ \ \ \forall\lambda>\delta.

As it turns out, the upper bound on vv_{\infty} described in Theorem 1 is tight. In our next section we shall give a geometric criteria on the achievement of this upper bound. To do so we begin by introducing and studying the notion of MM-absorbing sets.

3.2. MM-absorbing sets.

Definition 1.

A non-empty set CΔ(K)C\subseteq\Delta(K) is said to be MM-absorbing if 777conv(C)\text{conv}(C) is the convex hull of the set CC. qMconv(C)qM\in\emph{conv}(C) for every qCq\in C.

The intuition behind this choice of terminology is the following. Since qqMq\mapsto qM is a linear operator, if CC is MM-absorbing, so in conv(C)\text{conv}(C). However for conv(C)\text{conv}(C), MM-absorption exactly describes the situation in which the image of conv(C)\text{conv}(C) under MM lies inside (is absorbed in) conv(C)\text{conv}(C). In a dual fashion, if CΔ(K)C\subseteq\Delta(K) is a closed convex MM-absorbing set, then since by Krein-Milman Theorem888ext(C)\text{ext}(C) is the set of the extreme points of CC: those points in CC that cannot be expressed as a convex combination of two distinct points in CC. conv(ext(C))=C\text{conv}(\text{ext}(C))=C, we get that ext(C)\text{ext}(C) is also MM-absorbing. This implies, in particular, that since Δ(K)\Delta(K) is MM-absorbing, so is the set of its extreme points (i.e., the set of all mass-point distributions). Lastly, note that since conv(C1)conv(C2)conv(C1C2)\text{conv}(C_{1})\cup\text{conv}(C_{2})\subseteq\text{conv}(C_{1}\cup C_{2}), if C1C_{1} and C2C_{2} are both MM-absorbing, then so is C1C2C_{1}\cup C_{2}.


Example 2.

[Example 1, continued] Suppose that the transition rule underlying the interaction in Example 1 is given by

(5) M=( 0.10.90.60.4).M=\begin{pmatrix}\ 0.1&0.9\\ 0.6&0.4\end{pmatrix}.

Note that the stationary distribution πM\pi_{M} of MM equals (0.4,0.6)(0.4,0.6) (which corresponds to p=0.4p=0.4 in Figure 1) and the pair ((0.4,0.6),(Cavu)((0.4,0.6))((0.4,0.6),(\emph{Cav}\,u)((0.4,0.6)) is on the straight segment of the graph of (Cavu)(\emph{Cav}\,u) (see the right panel of Figure 1).

Moreover, the set containing the points (0,1)(0,1) and (0.5,0.5)(0.5,0.5) (presented in Figure 1 as the points 0 and 0.50.5) is not MM-absorbing. The reason is that (0,1)(0,1) is mapped by MM to (0,1)M=(0.6,0.4)(0,1)M=(0.6,0.4), which is not in the convex hull of (0,1)(0,1) and (0.5,0.5)(0.5,0.5).


In order to enhance the intuition about MM-absorbing sets, assume that at any point qΔ(K)q\in\Delta(K) a football is passed in a straight line to the point qMqM. The orbit generated by the football at qq is the union of line segments n1[qMn1,qMn]\bigcup_{n\geq 1}[qM^{n-1},qM^{n}], where [x,y]={αx+(1α)y: 0α1}[x,y]=\{\alpha x+(1-\alpha)y:\,0\leq\alpha\leq 1\}. When the set CC is MM-absorbing, the orbit generated by the football whose starting point is qCq\in C never exists of conv(C)\text{conv}(C).

We proceed with some basic examples of MM-absorbing sets. The simplest are the singleton {πM}\{\pi_{M}\} and the entire set, Δ(K)\Delta(K). To describe additional examples, consider the 1\ell_{1}-,2\ell_{2}-, and \ell_{\infty}-norms on Δ(K)\Delta(K), denoted by q1:=K|q|\|q\|_{1}:=\sum_{\ell\in K}|q^{\ell}|, q2:=K(q)2\|q\|_{2}:=\sqrt{\sum_{\ell\in K}(q^{\ell})^{2}} and q:=maxK|q|\|q\|_{\infty}:=\max_{\ell\in K}|q^{\ell}|, respectively, for qΔ(K)q\in\Delta(K). Denote by Mi\|M\|_{i} the operator norm999 Mi=maxxi=1xMi\|M\|_{i}=\max_{\|x\|_{i}=1}\|xM\|_{i}. of MM w.r.t.  the i\ell_{i}-norm, i{1,2,}i\in\{1,2,\infty\}.

For every i{1,2,}i\in\{1,2,\infty\} we have,

(6) qMπMi=qMπMMiMiqπMi.\|qM-\pi_{M}\|_{i}=\|qM-\pi_{M}M\|_{i}\leq\|M\|_{i}\|q-\pi_{M}\|_{i}.

It is known that M\|M\|_{\infty} equals the maximum 1\ell_{1}-norm of a row of MM (see, e.g., Example 5.6.5 on p.345 in [15]), and therefore M=1.\|M\|_{\infty}=1. Also,101010 This follows from the fact that M2\|M\|_{2} equals the maximum singular value of MM (see e.g., Example 5.6.6 on p.346 in [15]). M2=M=1\|M\|_{2}=\|M\|_{\infty}=1. Thus, in view of Eq. (6), any ball (either open or closed) w.r.t. to the 2\ell_{2} or \ell_{\infty}-norms centered at πM\pi_{M} is MM-absorbing. Moreover, if MM is doubly stochastic it is known that111111This follows from the fact that M1\|M\|_{1} equals the maximum 1\ell_{1}-norm of a column of MM (e.g., Example 5.6.5 on p. 344–345 in [15])). M1=1\|M\|_{1}=1 and therefore in that case any ball (either open or closed) w.r.t. the 1\ell_{1}-norm, centered at πM\pi_{M}, is also MM-absorbing. See Figure 2.


Refer to caption
Figure 2. Absorbing sets.

In all the examples above the MM-absorbing sets contain πM\pi_{M}. This is not a coincidence. Indeed, for every MM-absorbing set CC, the image of conv(C)\text{conv}(C) under the linear map MM is also contained in conv(C)\text{conv}(C). Therefore, by Brouwer’s fixed-point theorem, MM possesses a fixed point in121212clconv(C)\text{cl}\,\text{conv}(C) is the closure of conv(C)\text{conv}(C). clconv(C)\text{cl}\,\text{conv}(C). As the only fixed point of MM is πM\pi_{M}, we deduce that πMclconv(C)\pi_{M}\in\text{cl}\,\text{conv}(C) for every MM-absorbing set CC.

We end the discussion on MM-absorbing sets with the following proposition whose content and proof (delegated to Section 6) may enhance the intuition about absorbing sets.

Proposition 1.

Let CC be an MM-absorbing set. Then, CC contains a countable MM-absorbing set.


3.3. The Main Theorem.

To state our main result we begin with a review of some basic concepts from the theory of concave functions. First, for each g:Δ(K)g:\Delta(K)\to\mathbb{R} let Graph[g]:={(q,g(q)):qΔ(K)}\text{Graph}[g]:=\{(q,g(q)):\,q\in\Delta(K)\}. Since (Cavu\text{Cav}\,u) is a concave function, Graph[(Cavu)]\text{Graph}[(\text{Cav}\,u)] can be supported at (πM,(Cavu)(πM))(\pi_{M},(\text{Cav}\,u)(\pi_{M})) by a hyperplane. We may parametrize each such supporting hyperplane by a point in k\mathbb{R}^{k} as follows; first, for every zkz\in\mathbb{R}^{k} define fz:kf_{z}:\mathbb{R}^{k}\to\mathbb{R} by fz(x):=(Cavu)(πM)+z,qπMf_{z}(x):=(\text{Cav}\,u)(\pi_{M})+\langle z,q-\pi_{M}\rangle. Second, set

Λ:={zk:(Cavu)(q)fz(q),qΔ(K)}.\Lambda:=\{z\in\mathbb{R}^{k}\,:\,(\text{Cav}\,u)(q)\leq f_{z}(q),\,\forall q\in\Delta(K)\}.

As fz(πM)=(Cavu)(πM)f_{z}(\pi_{M})=(\text{Cav}\,u)(\pi_{M}) for every zz, the set Λ\Lambda corresponds to all supporting hyperplanes of Graph[(Cavu)]\text{Graph}[(\text{Cav}\,u)] at (πM,(Cavu)(πM))(\pi_{M},(\text{Cav}\,u)(\pi_{M})). For every zΛz\in\Lambda let

Az:={qΔ(K):u(q)=fz(q)}.A_{z}:=\{q\in\Delta(K)\,:\,u(q)=f_{z}(q)\}.

The set AzA_{z} can thus be interpreted as the projection to the first kk coordinates of the intersection of Graph[u]\text{Graph}[u] with the supporting hyperplane to Graph[(Cavu)]\text{Graph}[(\text{Cav}\,u)] at (πM,(Cavu)(πM))(\pi_{M},(\text{Cav}\,u)(\pi_{M})) parametrized by zz. A visualization of AzA_{z} when k=3k=3 is given in Figure 3.

Refer to caption
Figure 3. A visualization of AzA_{z} for zΛz\in\Lambda.
Proposition 2.

We have the following:

  • (i)

    If AzA_{z} contains an MM-absorbing set for some zΛz\in\Lambda, then v=(Cavu)(πM)v_{\infty}=(\emph{Cav}\,u)(\pi_{M}).

  • (ii)

    If v=(Cavu)(πM)v_{\infty}=(\emph{Cav}\,u)(\pi_{M}) then for every zΛz\in\Lambda, AzA_{z} contains a countable MM-absorbing set.

Why MM-absorbing sets contained in the AzA_{z}’s are of importance? This has to do with the control the sender has on the receiver’s beliefs. Indeed, once a belief is in the convex hull of an MM-absorbing subset CAzC\subseteq A_{z}, zΛz\in\Lambda, its shift under MM, which describes the evolution of the posterior in one time period, also lies in conv(C)\text{conv}(C). At this point in time the sender may send messages that would induce posteriors within CC, and in particular is AzA_{z}. As (Cavu)(\text{Cav}\,u) is an affine function on conv(Az)\text{conv}(A_{z}) (see Lemma 4 in Section 6), the weighted average of the values of (Cavu)(\text{Cav}\,u) evaluated at these posteriors, is equal to the value of (Cavu)(\text{Cav}\,u) evaluated πM\pi_{M}.

In the main theorem we summarize the results of Theorem 1 and Proposition 2. This theorem characterizes when patient senders can obtain a value close to the maximum possible, the upper bound stated in Theorem 1.

Theorem 2.

The following statements are equivalent:

  • (i)

    For every ε>0\varepsilon>0 and pΔ(K)p\in\Delta(K) there exists 0<δ<10<\delta<1 such that

    (7) |vλ(p)(Cavu)(πM)|<ε,λ>δ.|v_{\lambda}(p)-(\emph{Cav}\,u)(\pi_{M})|<\varepsilon,\ \ \ \ \ \forall\lambda>\delta.
  • (ii)

    There exists zΛz\in\Lambda such that AzA_{z} contains an MM-absorbing set.

  • (iii)

    For every zΛz\in\Lambda, AzA_{z} contains a countable MM-absorbing set.

Note that this characterization is stated in terms of the primitives of the model: uu and MM. Moreover, it unravels the sensitivity of a patient sender to the interrelationships between uu and MM, as the sets AzA_{z}, zΛz\in\Lambda, are fully determined by the former, whereas MM-absorbing sets are clearly determined by the latter. An interesting question that arises naturally in this context is how sensitive is a patient sender to such interrelationships. Assume, for instance, that AzA_{z} does not contain an MM-absorbing set for some zΛz\in\Lambda. Can one quantify the difference between (Cavu)(πM)(\text{Cav}\,u)(\pi_{M}) and vv_{\infty} in terms of uu and MM?

Example 3.

[Example 3 continued] Consider again the matrix MM in Eq. (5) and recall that πM\pi_{M} equals (0.4,0.6)(0.4,0.6). For the function uu given in Example 1 we have that AzA_{z} consisting of the two points (0,1)(0,1) and (0.5,0.5)(0.5,0.5) in Δ(K)\Delta(K) (corresponding to the points 0 and 0.50.5 in Figure 1) for every supporting hyperplane zΛz\in\Lambda. Since this set is not MM-absorbing, we conclude by Theorem 2 that for sufficiently patient sender the value is strictly less than (Cavu)(πM)(\emph{Cav}\,u)(\pi_{M}). However, if

(8) M=( 1/21/21/65/6),M=\begin{pmatrix}\ 1/2&1/2\\ 1/6&5/6\end{pmatrix},

then πM=(0.25,0.75)\pi_{M}=(0.25,0.75). Therefore as Az={(0,1),(0.5,0.5)}A_{z}=\{(0,1),(0.5,0.5)\} identifies with the set of extreme points of the ball of radius 0.250.25 w.r.t. the \ell_{\infty}-norm around πM\pi_{M}, we deduce that AzA_{z} is MM-absorbing for every zΛz\in\Lambda. Thus, Theorem 2 ensures that the value vλ(p)v_{\lambda}(p) of a sufficiently patient sender (i.e., when λ{\lambda} is close to 11) is close to (Cavu)(πM)=0.512(\emph{Cav}\,u)(\pi_{M})=0.512.

Under both transition matrices, the maximal payoff possible is obtained at p=0.581p=0.581. This point is located in the region where uu is equal to (Cavu)(\emph{Cav}\,u). We shall prove in Lemma 1 in Section 6 that at this point the sender has no incentive to alter the prior belief of the receiver. He therefore reveals no information to the latter. Such a result holds for any continuous uu and any point pp on which uu agrees with (Cavu)(\emph{Cav}\,u).

We end the section with a mathematical theorem on MM-absorbing sets, which is a by product of our main economical result.


4. Additional Results: A Non-Asymptotic Approach and a Strong Law

4.1. The value vλv_{\lambda} for every λ\lambda.

As it turns out, the case where v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}) encompasses information about the behavior of vλv_{\lambda} across all λ[0,1)\lambda\in[0,1). To showcase this, we begin by recalling that as any union of MM-absorbing sets is also MM-absorbing, by Proposition 2, if v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}), one may associate with each zΛz\in\Lambda a maximal (w.r.t. inclusion) MM-absorbing set BzAzB_{z}\subseteq A_{z}. Set D:=zΛconv(Bz)D:=\bigcup_{z\in\Lambda}\text{conv}(B_{z}) (see Figure 4).


Refer to caption
Figure 4. A visualization of DD.

In our next theorem we give an exact formula for vλ(p)v_{\lambda}(p) for all pDp\in D and all λ[0,1)\lambda\in[0,1) in the case where v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}). Therefore, the case v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}) is not of a mere asymptotic nature, but rather unravels the full information regarding vλv_{\lambda} for any discount factor λ\lambda on the domain DΔ(K)D\subseteq\Delta(K). However, outside of DD, the exact behavior of vλv_{\lambda}, even in the case v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}), remains an open problem.

Theorem 3.

Assume that v=(Cavu)(πM)v_{\infty}=(\emph{Cav}\,u)(\pi_{M}). Then, if pDp\in D we have

(9) vλ(p)=(Cavu)((1λ)p(IdkλM)1),λ[0,1),v_{\lambda}(p)=(\emph{Cav}\,u)\left((1-\lambda)p\,(\emph{Id}_{k}-\lambda M)^{-1}\right),\ \ \forall\lambda\in[0,1),

where Idk\emph{Id}_{k} is the k×kk\times k dimensional identity matrix.

Let us now give some intuition regarding the formula given in Theorem 3. First, recall that the sum of an infinite geometric series with common ratio r[0,1)r\in[0,1) and scale factor 11 equals 1/(1r)1/(1-r). As MM is stochastic, we will argue in the proof of Theorem 3 that we apply a matrix version of the formula for the sum of infinite geometric series so that

(1λ)p(IdkλM)1=(1λ)n=1λn1pMn1.(1-\lambda)p\,(\text{Id}_{k}-\lambda M)^{-1}=(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}pM^{n-1}.

Moreover, as we shall see in the proof section (e.g., Eq. (13)), for any signaling strategy σ\sigma and prior belief pΔ(K)p\in\Delta(K) we have that 𝔼p,σpn=pMn1\mathbb{E}_{p,\sigma}p_{n}=pM^{n-1}. Roughly speaking, the proof of Theorem 3 is divided into two parts. In the first, we come up with a signaling strategy σ\sigma for which the stage payoffs achieve their maximal possible value, shown to be equal to (Cavu\text{Cav}\,u)(pMn1)(pM^{n-1}) for n1n\geq 1. The second part connects (1λ)n=1λn1(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}\,(Cavu\text{Cav}\,u)(pMn1)(pM^{n-1}) with the formula given in (9). Such an argument is valid on DD, because (Cavu\text{Cav}\,u) is an affine function on each conv(Bz)\text{conv}(B_{z}), zΛz\in\Lambda.

As it turns out, the topological structure of the set DD around the point πM\pi_{M} may be utilized to derive two-sided bounds on vλv_{\lambda} for every discount factor λdk\lambda\in dk. Formally, consider the case where πM\pi_{M} is an interior point of DD (in the topological space Δ(K)\Delta(K) equipped with the 2\ell_{2}-norm). Thus, the Convergence Theorem for Markov Chains (e.g., Theorem 4.9 in [20]) guarantees that for every pΔ(K)p\in\Delta(K) there exists some finite time NN so that pMNDpM^{N}\in D. By employing bounds on the mixing rates of irreducible and aperiodic Markov chains we obtain the following theorem.

Theorem 4.

Assume that v=(Cavu)(πM)v_{\infty}=(\emph{Cav}\,u)(\pi_{M}) and that πM\pi_{M} has a full support (i.e.,131313int(D)(D) is the interior of DD. πMint(D)\pi_{M}\in\emph{int}(D)). Then, there exists a finite positive integer NN such that for every pΔ(K)p\in\Delta(K) and every λ[0,1)\lambda\in[0,1) we have the following lower and upper bounds:

(4.1) (1λ)n=1Nλn1u(pMn1)vλ(p)λ(p)(1λ)n=1Nλn1(Cavu)(pMn1),(1-\lambda)\sum_{n=1}^{N}\lambda^{n-1}u(pM^{n-1})\leq v_{\lambda}(p)-\mathcal{I}_{\lambda}(p)\\ \leq(1-\lambda)\sum_{n=1}^{N}\lambda^{n-1}(\emph{Cav}\,u)(pM^{n-1}),

where

λ(p):=λN(Cavu)((1λ)pMN(IdkλM)1).\mathcal{I}_{\lambda}(p):=\lambda^{N}(\emph{Cav}\,u)\left((1-\lambda)pM^{N}\,(\emph{Id}_{k}-\lambda M)^{-1}\right).

Moreover, NN can be chosen to be of the form clog2rD1c\lceil\log_{2}r_{D}^{-1}\rceil, where c>0c>0 is an integer constant (which may depend on MM) and rD:=sup{r>0:B1(πM,r)D}r_{D}:=\sup\{r>0:\,B_{\ell_{1}}(\pi_{M},r)\subseteq D\}, where B1(πM,r)B_{\ell_{1}}(\pi_{M},r) is the ball of radius rr w.r.t. the 1\ell_{1}-norm centered at πM\pi_{M}.

To illustrate the patience level required to make the bounds in Theorem 4 effective. we provide a numerical example. Assume that rD=210r_{D}=2^{-10}, so that DD contains a very small neighborhood of πM\pi_{M}. Hence, to attain the value in the left-hand side of Eq. (4.1) the sender can simply avoid sharing any private information along the first 10c10\,c time periods. As we shall see in the proof of Theorem 4, after those 10c10\,c time periods, the belief of the receiver will enter the region DD. As soon as this takes place, we will exploit in the proof of Theorem 4 the recursive structure of Γλ(p)\Gamma_{\lambda}(p) as well as the result of Theorem 3, to prescribe to the sender a signaling strategy that will guarantee him λ(p)\mathcal{I}_{\lambda}(p) in the remaining time periods of the game. Finally, if we choose the discount factor λ\lambda so that 1λ10c=(1λ)n=110cλn1<ε1-\lambda^{10c}=(1-\lambda)\sum_{n=1}^{10c}\lambda^{n-1}<\varepsilon, for some small ε\varepsilon, the lower bound and upper bound on vλv_{\lambda} in Eq. (4.1) will defer by at most ε×maxΔ(K)|u|\varepsilon\times\max_{\Delta(K)}|u|.

4.2. A Strong Law.

The next result is concerned with convergence in the strong sense, namely with the p,σ\mathbb{P}_{p,\sigma}-almost-surely behavior of the random variable limλ1(1λ)n=1λn1u(pn)\lim_{\lambda\to 1^{-}}(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}u(p_{n}) for a specific prior pp and a strategy σ\sigma. The weak-type of convergence employed so far is concerned with the limit of the expectations. In contrast, the next theorem deals with the almost-surely convergence of the payoff, corresponding to a sender interested in the payoffs he actually obtains. These are the payoffs he truly gets along realized paths, not just their expected values.

As it turns out, finite MM-absorbing sets can be used to deduce a strong law for the distribution of limλ1(1λ)n=1λn1u(pn)\lim_{\lambda\to 1^{-}}(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}u(p_{n}) when v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}).

Theorem 5.

Assume that AzA_{z} contains a finite MM-absorbing set CC for some zΛz\in\Lambda. Then, there exist an MM-absorbing subset QCQ\subseteq C and a strategy σΣ\sigma\in\Sigma such that for every pconv(Q)p\in\emph{conv}(Q) it holds that

(11) limλ1(1λ)n=1λn1u(pn)=(Cavu)(πM),p,σ-a.s.\lim_{\lambda\to 1^{-}}(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}u(p_{n})=(\emph{Cav}\,u)(\pi_{M}),\ \ \ \mathbb{P}_{p,\sigma}\emph{-}\emph{a.s.}

Moreover, if πMint(Q)\pi_{M}\in\emph{int}(Q), then (11) holds for every pΔ(K)p\in\Delta(K).

In words, under the assumptions of Theorem 5, for almost every infinite sequences of realizations (x1,x2,)(x_{1},x_{2},...) of the Markov chain (Xn)n1(X_{n})_{n\geq 1}, if the sender is patient enough he can guarantee himself by following σ\sigma a payoff close to (Cavu)(πM)(\text{Cav}\,u)(\pi_{M}).


5. When the Main Theorem Holds for every uu: Homothety.

The previous results shed light on the connection between MM and uu. Specifically, the main theorem characterizes when v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}) in terms of MM and uu. A natural question arises as to when this result holds for a fixed MM and for every uu. To answer this question, we need to introduce the notion of a homothety.

Definition 3.

A linear map, ψ:kk\psi:\mathbb{R}^{k}\to\mathbb{R}^{k} is said to be homothety with respect to the pair (v,β)k×[0,1)(v,\beta)\in\mathbb{R}^{k}\times[0,1) if ψ\psi maps each point xkx\in\mathbb{R}^{k} into the point βx+(1β)v\beta x+(1-\beta)v. The point vv is called the center and β\beta is called the ratio.

It is clear that when ψ\psi is a homothety with respect to (v,β)(v,\beta), the point vv is a fixed point of ψ\psi. Moreover, ψ\psi reduces the distance from any point xx to vv by a factor of β\beta. When vint(Δ(K))v\in\text{int}(\Delta(K)), or equivalently141414For any measure μ\mu defined on a finite probability space, we denote by supp(μ)\text{supp}(\mu) its support, i.e., the set of elements to which μ\mu assigns a positive probability. |supp(v)|=k|\text{supp}(v)|=k, and β[0,1)\beta\in[0,1), the matrix MψM^{\psi} defined by the homothety ψ\psi is an irreducible aperiodic stochastic matrix. In particular, the stationary distribution of MM is vv. Therefore, we shall say that an irreducible aperiodic stochastic matrix MM is a homothety if the mapping ϕ:xxM\phi:x\mapsto xM is a homothety of k\mathbb{R}^{k} with center πM\pi_{M} and ration β\beta, for some β[0,1)\beta\in[0,1).

When a stochastic matrix is a homothety, the transition from one state to another follows this law: each state stays unchanged with probability 1β1-\beta and moves according to the distribution vv to other states with probability β\beta.

We proceed by describing an interesting class of MM-absorbing sets of a homothety MM. A set EΔ(K)E\subseteq\Delta(K) is said to be star shaped around pΔ(K)p\in\Delta(K) if [p,q]E[p,q]\subseteq E for every qEq\in E. In words, assume that an observer is located at the point pΔ(K)p\in\Delta(K). Then EE is star shaped around pp if the line of sight, [p,q][p,q], to any point qEq\in E lies entirely in EE. Assume now that EE is star shaped around πM\pi_{M}. If MM is a homothety, then qM[πM,q)Econv(E)qM\in[\pi_{M},q)\subseteq E\subseteq\text{conv}(E) for every qEq\in E. Hence, when MM is a homothety, every star shaped set around any πM\pi_{M} is MM-absorbing.

Let MM be irreducible and aperiodic. Our next result gives a characterization of when MM is homothety in terms of vv_{\infty}. To make the result transparent, note that, by Theorem 1, vv_{\infty} is constant on Δ(K)\Delta(K), and as such is simply a function of uu and MM. In our new characterization we let uu vary over the space of all continuous functions defined on Δ(K)\Delta(K).Therefore vv_{\infty} also varies accordingly.

Theorem 6.

MM is a homothety if and only if v=(Cavu)(πM)v_{\infty}=(\emph{Cav}\,u)(\pi_{M}) for every continuous function uu.


6. Proofs

We start this section by reviewing the notion of a split, a cornerstone in the field of Bayesian persuasion. This can be described informally as follows: Given a lottery XX over KK with law pΔ(K)p\in\Delta(K), to which extent can the sender manipulate (split) pp using his signals. The answer, given by Blackwell (1951) and Aumann and Maschler (1995), is that for every choice of distributions q1,,q|S|Δ(K)q_{1},...,q_{|S|}\in\Delta(K) and convex weights (αi)i=1|S|Δ(S)(\alpha_{i})_{i=1}^{|S|}\in\Delta(S) s.t. i=1|S|αiqi=p\sum_{i=1}^{|S|}\alpha_{i}q_{i}=p, the sender can correlate his lottery over signals, YY, with the lottery XX, so that on the event that siSs^{i}\in S is chosen (having marginal probability αi\alpha_{i}) the posterior belief over states becomes qiq_{i}. This lottery YY will obey the rule

(12) (Y=si|X=)=αiqip,i=1,,|S|,K.\mathbb{P}(Y=s^{i}\,|\,X=\ell)=\frac{\alpha_{i}q_{i}^{\ell}}{p^{\ell}},\ \ \forall i=1,...,|S|,\forall\,\,\ell\in K.

Thus, if we denote by

SSp={{(qi,αi)}i=1|S|:qiΔ(K)i,(αi)i=1|S|Δ(S),s.t.i=1|S|αiqi=p},\SS_{p}=\bigg{\{}\{(q_{i},\alpha_{i})\}_{i=1}^{|S|}\,:\,q_{i}\in\Delta(K)\,\,\forall i,\,(\alpha_{i})_{i=1}^{|S|}\in\Delta(S),\,\text{s.t.}\sum_{i=1}^{|S|}\alpha_{i}q_{i}=p\bigg{\}},

the set of all possible splits in pΔ(K)p\in\Delta(K), then the signaling strategy σ\sigma can be interpreted as follows: At the first stage of Γλ(p)\Gamma_{\lambda}(p), σ\sigma specifies which element of SSp\SS_{p} the sender should pick. Then, for each time period n2n\geq 2, given that pn1=qp_{n-1}=q for some qΔ(K)q\in\Delta(K), σn\sigma_{n} prescribes which element of SSqM\SS_{qM} the sender should pick. Indeed, prior to getting the signal sns_{n}, the receiver updates her belief on the distribution of XnX_{n}, by taking the multiplication of her posterior belief qq on the distribution of Xn1X_{n-1} with the transition matrix MM, resulting in the belief qMqM. It thus follows that the sequence (pn)(p_{n}) satisfies the following important distributional law:

(13) 𝔼p,σ(pn+1|pn)=pnM,n1.\mathbb{E}_{p,\sigma}\,(p_{n+1}\,|\,p_{n})=p_{n}M,\ \ \ \ \forall n\geq 1.

Consequently, 𝔼p,σpn+1=(𝔼p,σp1)Mn=pMn\mathbb{E}_{p,\sigma}\,p_{n+1}=(\mathbb{E}_{p,\sigma}p_{1})M^{n}=pM^{n} for every n1n\geq 1. In particular if p=πMp=\pi_{M}, then 𝔼πM,σpn=πM\mathbb{E}_{\pi_{M},\sigma}\,p_{n}=\pi_{M} for every n1n\geq 1.

In light of this fresh view of σ\sigma, standard arguments (e.g., Theorem 2.20 in [25]) lead to the following ‘Bellman type’ recursive formula for vλ(p)v_{\lambda}(p):

(14) vλ(p)=sup{(qi,αi)}iSSp{(1λ)i=1|S|αiu(qi)+λi=1|S|αivλ(qiM)}.v_{\lambda}(p)=\sup_{\{(q_{i},\alpha_{i})\}_{i}\in\SS_{p}}\bigg{\{}(1-\lambda)\sum_{i=1}^{|S|}\alpha_{i}u(q_{i})+\lambda\sum_{i=1}^{|S|}\alpha_{i}v_{\lambda}(q_{i}M)\bigg{\}}.

Consider the operator ϕ:Δ(K)Δ(K)\phi:\Delta(K)\to\Delta(K) defined by ϕ(q)=qM\phi(q)=qM. Since |S|k|S|\geq k, Carathéodory’s Theorem (see, e.g., Corollary 17.1.5 in [24]) implies that the expression on the right-hand side of Eq. (14) equals (Cav{(1λ)u+λvλϕ})(p)(\text{Cav}\,\{(1-\lambda)u+\lambda v_{\lambda}\circ\phi\})(p). Thus, in conjunction with Eq. (14) we obtain the following key relation:

(15) vλ(p)=(Cav{(1λ)u+λ(vλϕ)})(p).v_{\lambda}(p)=(\text{Cav}\,\{(1-\lambda)u+\lambda\,(v_{\lambda}\circ\phi)\})(p).

In particular, this shows that the function vλ:Δ(K)v_{\lambda}:\Delta(K)\to\mathbb{R} is concave for every λ\lambda. As ϕ\phi is linear, vλϕv_{\lambda}\circ\phi is also concave. Then, by the definition of Cav we infer from Eq. (15) the following two-sided inequality:

(16) (1λ)minΔ(K)u+λ(vλϕ)(p)vλ(p)(1λ)(Cavu)(p)+λ(vλϕ)(p).(1-\lambda)\min_{\Delta(K)}u+\lambda(v_{\lambda}\circ\phi)(p)\leq v_{\lambda}(p)\leq(1-\lambda)(\text{Cav}\,u)(p)+\lambda(v_{\lambda}\circ\phi)(p).

Since the sender can always decide to not reveal any information at pp, i.e., to choose the split {(qi,αi)}iSSp\{(q_{i},\alpha_{i})\}_{i}\in\SS_{p}, where qi=pq_{i}=p for all i=1,,|S|i=1,...,|S|, and thereafter play optimally in the game Γλ(pM)\Gamma_{\lambda}(pM), we also have that vλ(p)(1λ)u(p)+λ(vλϕ)(p)v_{\lambda}(p)\geq(1-\lambda)u(p)+\lambda(v_{\lambda}\circ\phi)(p). The latter combined with the right hand side of Eq. (16) gives the following result:

Lemma 1.

Assume that pΔ(K)p\in\Delta(K) satisfies u(p)=(Cavu)(p)u(p)=(\emph{Cav}\,u)(p). Then, for any λ[0,1)\lambda\in[0,1), the optimal signaling strategy σλ\sigma_{\lambda} in Γλ(p)\Gamma_{\lambda}(p), would instruct the sender to not reveal any information at pp.

Set v¯(p):=lim infλ1vλ(p)\underline{v}(p):=\liminf_{\lambda\to 1^{-}}v_{\lambda}(p) and v¯(p):=lim supλ1vλ(p)\bar{v}(p):=\limsup_{\lambda\to 1^{-}}v_{\lambda}(p). Then, by (16), v¯(p)=v¯(pM)\underline{v}(p)=\underline{v}(pM) and v¯(p)=v¯(pM)\bar{v}(p)=\bar{v}(pM) for every pΔ(K)p\in\Delta(K). As our Markov chain is irreducible and aperiodic, the Convergence Theorem for Markov chains ensures that pMnπMpM^{n}\to\pi_{M} as nn\to\infty. A combination of the above two arguments yields the following proposition.

Proposition 3.

The functions v¯\underline{v} and v¯\bar{v} are constant on Δ(K)\Delta(K).

It follows from Proposition 3 that on our way to proving Theorem 1 we need to show that v¯=v¯\underline{v}=\bar{v}. This requires classical tools and techniques from the field of repeated games with incomplete information [23]. We begin by introducing, for every NN\in\mathbb{N} and pΔ(K)p\in\Delta(K), the NN-stage game ΓN(p)\Gamma_{N}(p) over the strategy space Σ\Sigma whose payoff is given by the formula

(17) γN(p,σ)=𝔼p,σ(1Nn=1Nu(pn)).\gamma_{N}(p,\sigma)=\mathbb{E}_{p,\sigma}\left(\frac{1}{N}\sum_{n=1}^{N}u(p_{n})\right).

The value of ΓN(p)\Gamma_{N}(p) will be denoted by vN(p)v_{N}(p). Standard continuity and compactness based arguments (see, e.g., Theorem 2.14 on p.15 in [25]) show that vN(p)=maxσΣγN(p,σ).v_{N}(p)=\max_{\sigma\in\Sigma}\gamma_{N}(p,\sigma).

The following proposition established a number of fundamental properties of vN(p)v_{N}(p) which will play an important role in our future proofs.

Proposition 4.

We have the following:

  • (i)

    vN:Δ(K)v_{N}:\Delta(K)\to\mathbb{R} is concave for every NN\in\mathbb{N}.

  • (ii)

    For every NN\in\mathbb{N}, the function vN:Δ(K)v_{N}:\Delta(K)\to\mathbb{R} is Lipschitz (w.r.t. the 1\ell^{1}-norm) with constant u:=maxΔ(K)|u|\|u\|_{\infty}:=\max_{\Delta(K)}|u| for every NN\in\mathbb{N}.

  • (iii)

    The sequence {NvN(πM)}N\{Nv_{N}(\pi_{M})\}_{N} is sub-additive.

  • (iv)

    The sequence {vN(πM)}N\{v_{N}(\pi_{M})\}_{N} converges.

  • (v)

    The sequence {vbN(πM)}N\{v_{b^{N}}(\pi_{M})\}_{N} is non-increasing for every bb\in\mathbb{N}.

Proof of Proposition 4.

The proof of (i) uses the following classical neat argument. Let q1,q2Δ(K)q_{1},q_{2}\in\Delta(K) and α(0,1)\alpha\in(0,1) such that p=αq1+(1α)q2p=\alpha q_{1}+(1-\alpha)q_{2}. Prior to the start of ΓN(p)\Gamma_{N}(p) the sender writes a computer program ZZ which draws either the digit 11 or 22 according to a random lottery. This lottery, denoted also ZZ, is dependent of X1X_{1}, and independent of (Xn)n2(X_{n})_{n\geq 2}. The conditional laws are given as follows:

(18) (Z=1|X1=)=1(Z=2|X1=)=αq1p,K.\mathbb{P}(Z=1\,|\,X_{1}=\ell)=1-\mathbb{P}(Z=2\,|\,X_{1}=\ell)=\frac{\alpha q_{1}^{\ell}}{p^{\ell}},\ \ \forall\ell\in K.

Then, at the first stage the sender sends a messenger151515The messenger cannot be the receiver. to collect on his behalf the realized value of X1X_{1}, type it into his program ZZ, and then tell him the output of the program. If the outcome of ZZ is 11 (22) the sender plays his optimal strategy in Γ(q1)\Gamma(q_{1}) (G(q2)G(q_{2})). We now argue that this strategy would guarantee him αvN(q1)+(1α)vN(q2)\alpha v_{N}(q_{1})+(1-\alpha)v_{N}(q_{2}) in ΓN(p)\Gamma_{N}(p). Indeed, the latter follows by Bayes’ law, which implies that the outcome of the program is 11 (resp., 22) with probability α\alpha (resp., 1α1-\alpha) and that the posterior distribution of X1X_{1} is q1q_{1} (resp., q2q_{2}). The last argument required to depict the situation in which the sender faces ΓN(q1)\Gamma_{N}(q_{1}) (resp., ΓN(q2)\Gamma_{N}(q_{2})) based on the message of the messenger is that after hearing the message (11 or 22), the sender must ask the messenger to inform him of the value of X1X_{1} which was realized!.

To prove (ii), observe that by conditioning on the outcome of X1X_{1} one has that

(19) γN(p,σ)=KpγN(δ,σ),\gamma_{N}(p,\sigma)=\sum_{\ell\in K}p^{\ell}\gamma_{N}(\delta^{\ell},\sigma),

for every σΣ\sigma\in\Sigma, where δ\delta^{\ell} is the Dirac measure supported on K\ell\in K. Therefore, by the triangle inequality, for any p,qΔ(K)p,q\in\Delta(K) and any σΣ\sigma\in\Sigma,

(20) |γN(p,σ)γN(q,σ)|uK|pq|.|\gamma_{N}(p,\sigma)-\gamma_{N}(q,\sigma)|\leq\|u\|_{\infty}\sum_{\ell\in K}|p^{\ell}-q^{\ell}|.

As ΓN(p)\Gamma_{N}(p) can be viewed as a finite game in extensive form, it admits a normal-form game description, and thus (ii) follows from the following basic inequality: for any two zero-sum matrix games AA and BB of equal dimensions it holds that |val(A)val(B)|AB|\text{val}(A)-\text{val}(B)|\leq\|A-B\|_{\infty}.

We move on to (iii). For every σΣ\sigma\in\Sigma and every N,LN,L\in\mathbb{N}, the (N+L)(N+L)’th stage game payoff γN+L(πM,σ)\gamma_{N+L}(\pi_{M},\sigma) equals

(21) NN+L𝔼πM,σ[1Nn=1Nu(pn)]+LN+L𝔼πM,σ[1Ln=N+1N+Lu(pn)].\frac{N}{N+L}\,\mathbb{E}_{\pi_{M},\sigma}\left[\frac{1}{N}\sum_{n=1}^{N}u(p_{n})\right]+\frac{L}{N+L}\,\mathbb{E}_{\pi_{M},\sigma}\left[\frac{1}{L}\sum_{n=N+1}^{N+L}u(p_{n})\right].

As the belief of the receiver at the start of the (N+1)(N+1)’st time period, prior to obtaining the signal sN+1s_{N+1}, equals pNMp_{N}M, we may bound the latter from above by

(22) NN+LvN(πM)+LN+L𝔼πM,σ[vL(pNM)].\frac{N}{N+L}v_{N}(\pi_{M})+\frac{L}{N+L}\mathbb{E}_{\pi_{M},\sigma}\left[v_{L}(p_{N}M)\right].

By Jensen’s inequality applied for the functions {vN}\{v_{N}\}, which are known to be concave by (i), in conjunction with the fact that at the initial belief πM\pi_{M}, all the steps of the sequence (pn)(p_{n}) have expectation πM\pi_{M}, we obtain that the latter is at most

(23) NN+LvN(πM)+LN+LvL(𝔼πM,σ(pNM))=NvN(πM)+LvL(πM).\frac{N}{N+L}v_{N}(\pi_{M})+\frac{L}{N+L}v_{L}(\mathbb{E}_{\pi_{M},\sigma}(p_{N}M))=Nv_{N}(\pi_{M})+Lv_{L}(\pi_{M}).

Since this holds for every σ\sigma, we conclude that(N+L)vN+L(πM)NvN(πM)+LvL(πM)(N+L)v_{N+L}(\pi_{M})\leq Nv_{N}(\pi_{M})+Lv_{L}(\pi_{M}), which completes the proof of (iii).

The proof of (iv) can be deduced directly from (iii) based on the following basic result from analysis; if {an}\{a_{n}\} is a sub-additive sequence, then {an/n}\{a_{n}/n\} converges. Lastly, by a repeated use of (iii) we obtain that if LL divides N,N, then

(24) NvN(πM)(NL1)vNL1(πM)+LvL(πM)NL(LvL(πM)).Nv_{N}(\pi_{M})\leq(\frac{N}{L}-1)v_{\frac{N}{L}-1}(\pi_{M})+Lv_{L}(\pi_{M})\leq\cdots\leq\frac{N}{L}(Lv_{L}(\pi_{M})).

Therefore, vN(πM)vL(πM)v_{N}(\pi_{M})\leq v_{L}(\pi_{M}), which is sufficient to prove (v). ∎

We are now in position to show that v¯=v¯\underline{v}=\bar{v}. We split the proof into two steps, the first of which is.

Lemma 2.

limNvN(πM)v¯\lim_{N\to\infty}v_{N}(\pi_{M})\leq\underline{v}.

Proof of Lemma 2.

Let ε>0\varepsilon>0. Define Tε:=min{n1:qMnB1(πM,ε),qΔ(K)}T_{\varepsilon}:=\min\{n\geq 1:\,qM^{n}\in B_{\ell_{1}}(\pi_{M},\varepsilon),\ \,\forall q\in\Delta(K)\}. A well-known bound on the mixing rates of MM (see, e.g., Eq. (4.34) in Section 4.5 of [20]) states that there exists a constant cc (which may depend on MM) such that Tεclog2ε1T_{\varepsilon}\leq c\lceil\log_{2}\varepsilon^{-1}\rceil.

Consider now the following signaling strategy σΣ\sigma\in\Sigma. In the first NεN_{\varepsilon} time periods of the game Γλ(πM)\Gamma_{\lambda}(\pi_{M}), the sender plays optimally in ΓNε(πM)\Gamma_{N_{\varepsilon}}(\pi_{M}), where NεN_{\varepsilon} satisfies (i) Nε(clog2ε1)/εN_{\varepsilon}\geq(c\lceil\log_{2}\varepsilon^{-1}\rceil)/\varepsilon and (ii) vNε(πM)limNvN(πM)εv_{N_{\varepsilon}}(\pi_{M})\geq\lim_{N\to\infty}v_{N}(\pi_{M})-\varepsilon. Then, starting from the (Nε+1)(N_{\varepsilon}+1)’st stage σ\sigma, asks the sender to ignore all information obtained along the next TεT_{\varepsilon} time periods, and send the same signal s0Ss_{0}\in S during these times. Starting from the (Nε+Tε+1)(N_{\varepsilon}+T_{\varepsilon}+1)’st stage, by following σ\sigma, the sender will play an optimal strategy in ΓN(pNε+TεM)\Gamma_{N}(p_{N_{\varepsilon}+T_{\varepsilon}}M) along the next NεN_{\varepsilon} time periods, and subsequently play s0s_{0} at each of the next TεT_{\varepsilon} time periods, thereby ignoring all information collected at times Nε+Tε+1,,2Nε+2TεN_{\varepsilon}+T_{\varepsilon}+1,...,2N_{\varepsilon}+2T_{\varepsilon}.

The extension of the definition of σ\sigma to all time periods n1n\geq 1 is done inductively. Let (bm)m=1(b_{m})_{m=1}^{\infty} be defined by bm:=(m1)Nε+(m1)Tεb_{m}:=(m-1)N_{\varepsilon}+(m-1)T_{\varepsilon}. For each m1m\geq 1, the strategy σ\sigma instructs the sender to play an optimal strategy in ΓNε(pbmM)\Gamma_{N_{\varepsilon}}(p_{b_{m}}M) starting from the (bm+1)(b_{m}+1)’st time period for NεN_{\varepsilon} periods, followed by sending the fixed signal s0Ss_{0}\in S at times bm+Nε+1,,bm+1b_{m}+N_{\varepsilon}+1,...,b_{m+1}.

By the definition of TεT_{\varepsilon}, we have that pbmM=pbmTεMTε+1B1(πM,ε)p_{b_{m}}M=p_{b_{m}-T_{\varepsilon}}M^{T_{\varepsilon}+1}\in B_{\ell_{1}}(\pi_{M},\varepsilon) for every m2m\geq 2. Hence, by the definition of σ\sigma, item (ii) of Proposition 4, and the choice of NεN_{\varepsilon} we obtain that

(25) 1Nεn=bm+1bm+Nε𝔼πM,σ(u(pn))=𝔼πM,σ(vNε(pbmM))𝔼πM,σ(vNε(πM)uε)limNvN(πM)(u+1)ε,m2.\begin{split}\frac{1}{N_{\varepsilon}}\sum_{n=b_{m}+1}^{b_{m}+N_{\varepsilon}}\mathbb{E}_{\pi_{M},\sigma}(u(p_{n}))&=\mathbb{E}_{\pi_{M},\sigma}(v_{N_{\varepsilon}}(p_{b_{m}}M))\\ &\geq\mathbb{E}_{\pi_{M},\sigma}(v_{N_{\varepsilon}}(\pi_{M})-\|u\|_{\infty}\varepsilon)\\ &\geq\lim_{N\to\infty}v_{N}(\pi_{M})-(\|u\|_{\infty}+1)\varepsilon,\ \ \ \ \ \forall m\geq 2.\end{split}

As Nε/Tε1/εN_{\varepsilon}/T_{\varepsilon}\geq 1/\varepsilon, using Eq. (25) we obtain that

(26) lim infN1Nn=1N𝔼πM,σ(u(pn))limNvN(πM)O(ε).\liminf_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\mathbb{E}_{\pi_{M},\sigma}(u(p_{n}))\geq\lim_{N\to\infty}v_{N}(\pi_{M})-O(\varepsilon).

Combining this with a Tauberian theorem (see, e.g., Theorem 3.1  in [25]) and Proposition 3 we conclude that

(27) v¯=lim infλ1vλ(πM)lim infλ1γλ(πM,σ)limNvN(πM)O(ε).\underline{v}=\liminf_{\lambda\to 1^{-}}v_{\lambda}(\pi_{M})\geq\liminf_{\lambda\to 1^{-}}\gamma_{\lambda}(\pi_{M},\sigma)\geq\lim_{N\to\infty}v_{N}(\pi_{M})-O(\varepsilon).

Since ε>0\varepsilon>0 was arbitrary all along, the proof is complete. ∎

The second step towards showing that v¯=v¯\bar{v}=\underline{v} is given in the following lemma.

Lemma 3.

v¯limNvN(πM)\bar{v}\leq\lim_{N\to\infty}v_{N}(\pi_{M}).

Proof of Lemma 3..

For each λ[0,1)\lambda\in[0,1), let σλΣ\sigma_{\lambda}\in\Sigma be such that161616Such sλs_{\lambda} exists by standard arguments of continuity and compactness, e.g. Theorem 2.14 on p.15 in [25] vλ(πM)=γλ(πM,σλ).v_{\lambda}(\pi_{M})=\gamma_{\lambda}(\pi_{M},\sigma_{\lambda}). Since uu is bounded on Δ(K)\Delta(K), we may apply the following decomposition (see, e.g., Eq. (1) in [17]) of the discounted average into a convex combination of finite averages:

(28) γλ(πM,σλ)=(1λ)2n=0(n+1)λnγn(πM,σλ).\gamma_{\lambda}(\pi_{M},\sigma_{\lambda})=(1-\lambda)^{2}\sum_{n=0}^{\infty}(n+1)\lambda^{n}\gamma_{n}(\pi_{M},\sigma_{\lambda}).

Now fix ε>0\varepsilon>0. By item (iv) of Proposition 4, there exists N0N_{0}\in\mathbb{N} such that

(29) γn(πM,σλ)limNvN(πM)+ε,n>N0,λ[0,1).\gamma_{n}(\pi_{M},\sigma_{\lambda})\leq\lim_{N\to\infty}v_{N}(\pi_{M})+\varepsilon,\ \ \ \forall n>N_{0},\,\,\,\forall\lambda\in[0,1).

Next, as |γn(πM,σλ)|u|\gamma_{n}(\pi_{M},\sigma_{\lambda})|\leq\|u\|_{\infty} for every nn and every λ\lambda, we may take λ0[0,1)\lambda_{0}\in[0,1) for which

(30) (1λ)2n=0N0(n+1)λnγn(πM,σλ)ε,λ>λ0.(1-\lambda)^{2}\sum_{n=0}^{N_{0}}(n+1)\lambda^{n}\gamma_{n}(\pi_{M},\sigma_{\lambda})\leq\varepsilon,\ \ \ \forall\lambda>\lambda_{0}.

Thus, by combining Eqs. (28), (29), and (30) we conclude that vλ(πM)limNvN(πM)+2εv_{\lambda}(\pi_{M})\leq\lim_{N\to\infty}v_{N}(\pi_{M})+2\varepsilon for every λ>λ0\lambda>\lambda_{0}, thus proving the lemma. ∎


Combining Lemmas 2 and 3 we obtain the string of inequalities

limNvN(πM)v¯v¯limNvN(πM).\lim_{N\to\infty}v_{N}(\pi_{M})\leq\underline{v}\leq\bar{v}\leq\lim_{N\to\infty}v_{N}(\pi_{M}).

Hence, v¯=v¯\bar{v}=\underline{v}. Thus, the proof of Theorem 1 will be concluded if we show the following simple claim to be true.

Claim 1.

For every λ[0,1)\lambda\in[0,1) and every N1N\geq 1 we have that vλ(πM)<(Cavu)(πM)v_{\lambda}(\pi_{M})<(\emph{Cav}\,u)(\pi_{M}) and vN(πM)<(Cavu)(πM)v_{N}(\pi_{M})<(\emph{Cav}\,u)(\pi_{M}).

Proof of Claim 1..

The claim follows easily from the fact that for any σΣ\sigma\in\Sigma, we have, by Jensen’s inequality, that the expected payoff at any time period nn satisfies

(31) 𝔼πM,su(pn)𝔼πM,s(Cavu)(pn)(Cavu)(𝔼πM,spn)=(Cavu)(πM).\mathbb{E}_{\pi_{M},s}u(p_{n})\leq\mathbb{E}_{\pi_{M},s}(\text{Cav}\,u)(p_{n})\leq(\text{Cav}\,u)(\mathbb{E}_{\pi_{M},s}\,p_{n})=(\text{Cav}\,u)(\pi_{M}).

As the first step toward the proof of Proposition 2 and Theorems 3 and 4 we prove the following basic lemma.

Lemma 4.

For every zΛz\in\Lambda we have:

  • (i)

    Cavu(q)=fz(q)\emph{Cav}\,u(q)=f_{z}(q) for every qconv(Az)q\in\emph{conv}(A_{z}).

  • (ii)

    Let (αi)i=1m(\alpha_{i})_{i=1}^{m}, αi>0\alpha_{i}>0 for every ii, iαi=1\sum_{i}\alpha_{i}=1 and (qi)i=1mΔ(K)(q_{i})_{i=1}^{m}\in\Delta(K) such that iαiqi=πM\sum_{i}\alpha_{i}q_{i}=\pi_{M}. If iαiu(qi)=(Cavu)(πM)\sum_{i}\alpha_{i}u(q_{i})=(\emph{Cav}\,u)(\pi_{M}), then qiAzq_{i}\in A_{z} for every ii.

Proof of Lemma 4..

Let qconv(Az)q\in\text{conv}(A_{z}). Take (qi)Az(q_{i})\in A_{z} and convex weights (αi)(\alpha_{i}) such that q=iαiqiq=\sum_{i}\alpha_{i}q_{i}. Since Cavu\text{Cav}\,u is concave and qiAzq_{i}\in A_{z}, we have

(31) (Cavu)(q)iαi(Cavu)(qi)iαiu(qi)=iαifz(qi)=fz(q),(\text{Cav}\,u)(q)\geq\sum_{i}\alpha_{i}(\text{Cav}\,u)(q_{i})\geq\sum_{i}\alpha_{i}u(q_{i})=\sum_{i}\alpha_{i}f_{z}(q_{i})=f_{z}(q),

where the last equality follows from the fact that fzf_{z} is affine. Since by the definition of Λ\Lambda we have (Cavu)(q)fz(q)(\text{Cav}\,u)(q)\leq f_{z}(q) for every qΔ(K)q\in\Delta(K), we have shown (i). For (ii) assume that there exists qi0Azq_{i_{0}}\notin A_{z}. Then u(qi0)<fz(qi0)u(q_{i_{0}})<f_{z}(q_{i_{0}}), and since αi0>0\alpha_{i_{0}}>0 and zΛz\in\Lambda, we have

(32) (Cavu)(πM)=iαiu(qi)<iαifz(qi)=fz(πM)=(Cavu)(πM).(\text{Cav}\,u)(\pi_{M})=\sum_{i}\alpha_{i}u(q_{i})<\sum_{i}\alpha_{i}f_{z}(q_{i})=f_{z}(\pi_{M})=(\text{Cav}\,u)(\pi_{M}).

We reached a contradiction. ∎

Next, let us now prove the following proposition, unveiling the special advantages of MM-absorbing subsets of AzA_{z} for zΛz\in\Lambda.

Proposition 5.

Assume that for some zΛz\in\Lambda, the set AzA_{z} admits an MM-absorbing subset CC. Then,

(33) vλ(p)=(Cavu)(λp(IdkλM)1),λ[0,1),v_{\lambda}(p)=(\emph{Cav}\,u)\left(\lambda p\,(\emph{Id}_{k}-\lambda M)^{-1}\right),\ \ \forall\lambda\in[0,1),

for every pconv(C)p\in\emph{conv}(C), where Idk\emph{Id}_{k} is the kk-dimensional identity matrix.

Proof of Proposition 5..

To show Eq. (33) we will show that a two-sided inequality holds. First, by applying the Jensen’s inequality in same fashion as in Eq. (31), we get that for every σΣ\sigma\in\Sigma and every λ[0,1)\lambda\in[0,1) it holds that

(6.1) γλ(p,σ)(1λ)n=1λn1(Cavu)(𝔼p,σpn)=(1λ)n=1λn1(Cavu)(pMn1).\gamma_{\lambda}(p,\sigma)\leq(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}(\text{Cav}\,u)(\mathbb{E}_{p,\sigma}\,p_{n})\\ =(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}(\text{Cav}\,u)(pM^{n-1}).

Since pconv(C)p\in\text{conv}(C) and the set conv(C)\text{conv}(C) is MM-absorbing, pMnconv(C)pM^{n}\in\text{conv}(C) for all n1n\geq 1. Hence, as conv(C)conv(Az)\text{conv}(C)\subseteq\text{conv}(A_{z}), item (i) of Lemma 4 implies that

(35) (1λ)n=1λn1(Cavu)(pMn1)=(1λ)n=1λn1fz(pMn1)=fz((1λ)n=1λn1pMn1)=(Cavu)((1λ)n=1λn1pMn1)\begin{split}(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}(\text{Cav}\,u)(pM^{n-1})&=(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}f_{z}(pM^{n-1})\\ &=f_{z}\left((1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}pM^{n-1}\right)\\ &=(\text{Cav}\,u)\left((1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}pM^{n-1}\right)\end{split}

where the last equlity follows from item (i) of Lemma 4 as well. Moreover, since MM is a stochastic matrix, we have the following well-know formula (e.g., Theorem 2.29 in [25]):

(36) (1λ)n=1λn1pMn1=λp(Idk(1λ)M)1,λ[0,1).(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}pM^{n-1}=\lambda p\,(\text{Id}_{k}-(1-\lambda)M)^{-1},\ \ \forall\lambda\in[0,1).

A combination of Eqs. (6.1), (35), and (36) yields that vλ(p)v_{\lambda}(p) is at most (Cavu)((1λ)p(IdkλM)1)(\text{Cav}\,u)\left((1-\lambda)p\,(\text{Id}_{k}-\lambda M)^{-1}\right) for every pconv(C)p\in\text{conv}(C) and every λ[0,1)\lambda\in[0,1). Let us now show that the inverse direction holds as well. We start by defining for each qΔ(K)q\in\Delta(K) the set SSqCSSq\SS_{q}^{C}\subseteq\SS_{q} by

(37) SSqC:={{(qi,αi)}i=1|S|:qiCi=1,,|S|}\SS_{q}^{C}:=\{\{(q_{i},\alpha_{i})\}_{i=1}^{|S|}\,:\,q_{i}\in C\,\,\,\,\forall i=1,...,|S|\}

Since |S|k|S|\geq k, Carathéodory’s Theorem shows that SSqC\SS_{q}^{C}\neq\emptyset whenever qconv(C)q\in\text{conv}(C). Consider now the strategy σC\sigma^{C} defined as follows: at each n1n\geq 1, if pn1=qconv(C)p_{n-1}=q\in\text{conv}(C), σC\sigma^{C} will chose an element in SSqMC\SS_{qM}^{C}; otherwise, if pn1=qΔ(K)conv(C)p_{n-1}=q\in\Delta(K)\setminus\text{conv}(C), then σC\sigma^{C} will chose some element in SSq\SS_{q}. As pconv(C)p\in\text{conv}(C), and conv(C)\text{conv}(C) is MM-absorbing, we have that under the strategy σC\sigma^{C}, supp(pn)C\text{supp}(p_{n})\subseteq C for every n1n\geq 1. Indeed, we show this by induction on nn. For n=1n=1, since pconv(C)p\in\text{conv}(C), supp(p1)C\text{supp}(p_{1})\subseteq C by the definition of σz\sigma^{z}. Assume now that supp(pn)C\text{supp}(p_{n})\subseteq C for some n1n\geq 1. Since CC is MM-absorbing, supp(pnM)conv(C)\text{supp}(p_{n}M)\subseteq\text{conv}(C), and thus by the definition of σC\sigma^{C} we see that supp(pn+1)C\text{supp}(p_{n+1})\subseteq C as well. The latter, coupled with CAzC\subseteq A_{z} implies that the discounted payoff under σC\sigma^{C} can be computed as follows:

(38) γλ(p,σC)=(1λ)n=1λn1𝔼p,σC[u(pn)]=(1λ)n=1λn1𝔼p,σC[fz(pn)]=(1λ)n=1λn1fz(𝔼p,σCpn)=(1λ)n=1λn1fz(pMn1)=(1λ)n=1λn1(Cavu)(pMn1),\begin{split}\gamma_{\lambda}(p,\sigma^{C})&=(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}\mathbb{E}_{p,\sigma^{C}}[u(p_{n})]\\ &=(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}\mathbb{E}_{p,\sigma^{C}}\,[f_{z}(p_{n})]\\ &=(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}f_{z}\left(\mathbb{E}_{p,\sigma^{C}}\,p_{n}\right)\\ &=(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}f_{z}\left(pM^{n-1}\right)\\ &=(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}(\text{Cav}\,u)(pM^{n-1}),\end{split}

where we note that the third equality holds because fzf_{z} is affine, and the last equality is a consequence of item (i) of Lemma 4. Hence, by Eqs. (35) and (36) we have that γλ(p,σC)\gamma_{\lambda}(p,\sigma^{C}) equals (Cavu)((1λ)p(IdkλM)1)(\text{Cav}\,u)\left((1-\lambda)p\,(\text{Id}_{k}-\lambda M)^{-1}\right), thus showing the converse inequality. ∎

We proceed with the proof of Proposition 2.

Proof of Proposition 2..

Assume that CC is an MM-absorbing subset of AzA_{z} where zΛz\in\Lambda. Let qconv(C)q\in\text{conv}(C). First, by Proposition 5 we have

vλ(q)=(Cavu)((1λ)q(IdkλM)1),λ[0,1).v_{\lambda}(q)=(\text{Cav}\,u)\left((1-\lambda)q\,(\text{Id}_{k}-\lambda M)^{-1}\right),\ \ \forall\lambda\in[0,1).

Next, since qMnπMqM^{n}\to\pi_{M} as nn\to\infty, we have that (1/n)=1nqM1πM(1/n)\sum_{\ell=1}^{n}qM^{\ell-1}\to\pi_{M} as nn\to\infty. Thus, by a Tauberian Theorem (see, e.g., Theorem 3.1. in [25]) we obtain that (1λ)n=1λn1qMn1πM(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}qM^{n-1}\to\pi_{M} as λ1\lambda\to 1^{-}. Hence, from the identity given in (36) we obtain that (1λ)q(IdkλM)1πM(1-\lambda)q\,(\text{Id}_{k}-\lambda M)^{-1}\to\pi_{M} as λ1\lambda\to 1^{-}. Moreover, since (Cavu)(\text{Cav}\,u) is concave on Δ(K)\Delta(K) it is also continuous at πM\pi_{M} (see, e.g., Theorem 10.1 in [24]). A combination of the above arguments with Theorem 1 yields

(6.2) v=limλ1vλ(q)=limλ1(Cavu)((1λ)q(IdkλM)1)=(Cavu)(πM),v_{\infty}=\lim_{\lambda\to 1^{-}}v_{\lambda}(q)=\lim_{\lambda\to 1^{-}}(\text{Cav}\,u)\left((1-\lambda)q\,(\text{Id}_{k}-\lambda M)^{-1}\right)\\ =(\text{Cav}\,u)(\pi_{M}),

thus proving item (i) of Proposition 2.

Let us continue with the proof of item (ii). Assume that v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}). We have (i) limNvN(πM)=(Cavu)(πM)\lim_{N\to\infty}v_{N}(\pi_{M})=(\text{Cav}\,u)(\pi_{M}), (ii) vN(πM)(Cavu)(πM)v_{N}(\pi_{M})\leq(\text{Cav}\,u)(\pi_{M}) for every NN (by Claim 1) and (iii) {vbN(πM)}N\{v_{b^{N}}(\pi_{M})\}_{N} is non-increasing for every bb\in\mathbb{N} (by item (v) of Proposition 4). A combination of (i), (ii), and (iii) shows that vN(πM)=(Cavu)(πM)v_{N}(\pi_{M})=(\text{Cav}\,u)(\pi_{M}) for every NN. Let σN\sigma^{N} be an optimal strategy in ΓN(πM)\Gamma_{N}(\pi_{M}). Denote by (pnN)n(p_{n}^{N})_{n} the sequence of posteriors induced by σλ\sigma^{\lambda} and the prior probability πM\pi_{M}. By Jensen’s inequality, 𝔼πM,σN[u(pnN)](Cavu)(πM)\mathbb{E}_{\pi_{M},\sigma^{N}}[u(p^{N}_{n})]\leq(\text{Cav}\,u)(\pi_{M}) for every nn. Hence, as γN(πM,σN)=(Cavu)(πM)\gamma_{N}(\pi_{M},\sigma^{N})=(\text{Cav}\,u)(\pi_{M}), we obtain that 𝔼πM,σN[u(pnN)]=(Cavu)(πM)\mathbb{E}_{\pi_{M},\sigma^{N}}[u(p^{N}_{n})]=(\text{Cav}\,u)(\pi_{M}) for every n=1,,Nn=1,...,N. Fix λ(0,1)\lambda\in(0,1). We see that

(40) γλ(πM,σN)(1λ)n=1Nλn1(Cavu)(πM)λNu\gamma_{\lambda}(\pi_{M},\sigma^{N})\geq(1-\lambda)\sum_{n=1}^{N}\lambda^{n-1}(\text{Cav}\,u)(\pi_{M})-\lambda^{N}\|u\|_{\infty}

for every N1N\geq 1. Letting NN\to\infty we get vλ(πM)(Cavu)(πM)v_{\lambda}(\pi_{M})\geq(\text{Cav}\,u)(\pi_{M}). Since by Claim 1 the inverse inequality holds as well, we deduce that vλ(πM)=(Cavu)(πM)v_{\lambda}(\pi_{M})=(\text{Cav}\,u)(\pi_{M}). Therefore, there exists a strategy σλ\sigma^{\lambda} such that γλ(πM,σλ)=(Cavu)(πM)\gamma_{\lambda}(\pi_{M},\sigma^{\lambda})=(\text{Cav}\,u)(\pi_{M}). Denote the sequence of posteriors induced by σλ\sigma^{\lambda} and the prior probability πM\pi_{M} by (pnλ)n(p^{\lambda}_{n})_{n}. By Jensen’s inequality, 𝔼πM,σλ[u(pnλ)](Cavu)(πM)\mathbb{E}_{\pi_{M},\sigma^{\lambda}}\,[u(p^{\lambda}_{n})]\leq(\text{Cav}\,u)(\pi_{M}), and we therefore obtain that 𝔼πM,σλ[u(pnλ)]=(Cavu)(πM)\mathbb{E}_{\pi_{M},\sigma^{\lambda}}[u(p^{\lambda}_{n})]=(\text{Cav}\,u)(\pi_{M}) for every nn. Moreover, as supp(pnλ)\text{supp}(p_{n}^{\lambda}) is finite and 𝔼πM,σλpnλ=πM\mathbb{E}_{\pi_{M},\sigma^{\lambda}}\,p^{\lambda}_{n}=\pi_{M} for every nn, item (ii) of Lemma 4 implies that supp(pnλ)Az\text{supp}(p_{n}^{\lambda})\subseteq A_{z} for every zΛz\in\Lambda and every nn. Set C:=n1supp(pnλ)C:=\bigcup_{n\geq 1}\text{supp}(p_{n}^{\lambda}). We have CAzC\subseteq A_{z} for every zΛz\in\Lambda. Moreover, as the set of signals SS of the receiver is finite, we have that CC is the countable union of finite sets, and thus is countable.

We claim that CC is MM-absorbing. Indeed, if qCq\in C, then there exists some nn such that pnλ=qp_{n}^{\lambda}=q with positive probability. Since 𝔼πM,σλ(pn+1|pnλ=q)=qM\mathbb{E}_{\pi_{M},\sigma^{\lambda}}(p_{n+1}\,|\,p_{n}^{\lambda}=q)=qM, we obtain that qMconv(supp(pn+1λ))conv(C)qM\in\text{conv}(\text{supp}(p_{n+1}^{\lambda}))\subseteq\text{conv}(C). To summarize, CC is a countable MM-absorbing subset of AzA_{z} for every zΛz\in\Lambda, as desired. ∎

We may now deduce the proofs of Theorems 3 and 4 as follows.

Proofs of Theorems 3 and 4..

By Proposition 2, the sets BzB_{z}, zΛz\in\Lambda, are well defined (i.e., non-empty). As they are MM-absorbing, we can apply Proposition 5 to any point pconv(Bz)p\in\text{conv}(B_{z}) to get the result of Theorem 3.

As for the proof of Theorem 4, let E:=B1(πM,r)E:=B_{\ell_{1}}(\pi_{M},r) where r<rDr<r_{D} satisfies log2r1=log2rD1\lceil\log_{2}r^{-1}\rceil=\lceil\log_{2}r_{D}^{-1}\rceil. By the definition of rDr_{D} we have EDE\subseteq D (see, in the statement of the theorem). By employing a classical bound on the mixing rates of MM (e.g., Eq. (4.34) in Section 4.5 in [20]) we may take NN of the form clog2r1=clog2rDc\lceil\log_{2}r^{-1}\rceil=c\lceil\log_{2}r_{D}\rceil (where cc is a positive integer which may depend on MM) so that pMNEpM^{N}\in E for every pΔ(K)p\in\Delta(K). Thus if the sender does not use his private information along the first NN time periods, and then plays optimally starting from the (N+1)(N+1)’st time period, he guarantees (1λ)n=1λn1u(pMn1)+λNvλ(pMN)(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}u(pM^{n-1})+\lambda^{N}v_{\lambda}(pM^{N}). As pMNDpM^{N}\in D, Theorem 3 ensures that λNvλ(pMN)=λ(p)\lambda^{N}v_{\lambda}(pM^{N})=\mathcal{I}_{\lambda}(p), thus proving the left hand side of Eq. (4.1). On the other hand, by using Jensen inequality for both Cavu\text{Cav}\,u and vλv_{\lambda}, we see that for every σΣ\sigma\in\Sigma and pΔ(K)p\in\Delta(K)

γλ(p,σ)(1λ)n=1λn1Cavu(pMn1)+λNvλ(pMN),\gamma_{\lambda}(p,\sigma)\leq(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}\text{Cav}\,u(pM^{n-1})+\lambda^{N}v_{\lambda}(pM^{N}),

and thus, since λNvλ(pMN)=λ(p)\lambda^{N}v_{\lambda}(pM^{N})=\mathcal{I}_{\lambda}(p) by Theorem 3, we obtain the right-hand side of Eq. (4.1). ∎

We move on with the proof of Theorem 5.

Proof of Theorem 5..

Let C={q1,,qr}C=\{q_{1},...,q_{r}\}. Since CC is MM-absorbing, we can assign for each i=1,,ri=1,...,r a distribution αiΔ(C)\alpha^{i}\in\Delta(C) so that qiM=j=1rαjiqjq_{i}M=\sum_{j=1}^{r}\alpha^{i}_{j}q_{j}. Moreover, by Carathéodory’s Theorem, we can choose αi\alpha^{i} so that |supp(αi)|k|\text{supp}(\alpha^{i})|\leq k for every ii. Define the r×rr\times r matrix WW by Wi,j:=αjiW_{i,j}:=\alpha^{i}_{j}; WW is a stochastic matrix. Also, define the r×kr\times k dimensional matrix PP by Pi,:=qiP_{i,\ell}:=q_{i}^{\ell}, where i{1,,r}i\in\{1,...,r\} and K\ell\in K. We have the following algebraic relation:

(41) PM=WP.PM=WP.

Let \mathcal{R} be a communicating class of WW whose states are recurrent (see, e.g., Lemma 1.26 in [20]). Denote by WW_{\mathcal{R}} the restriction of the matrix WW to the set of states \ell\in\mathcal{R}. The ||×|||\mathcal{R}|\times|\mathcal{R}| matrix WW_{\mathcal{R}} is clearly stochastic. Moreover, since \mathcal{R} is a communication class, WW_{\mathcal{R}} is also irreducible. Next, let us denote by PP_{\mathcal{R}} the k×||k\times|\mathcal{R}| matrix with entries (P)i,:=qi(P_{\mathcal{R}})_{i,\ell}:=q_{i}^{\ell}, where ii\in\mathcal{R} and K\ell\in K. It follows from Eq. (41) that

(42) PM=WP.P_{\mathcal{R}}M=W_{\mathcal{R}}P_{\mathcal{R}}.

As WW_{\mathcal{R}} is stochastic, Eq. (42) implies that the set Q={qi}iQ=\{q_{i}\}_{i\in\mathcal{R}} is MM-absorbing.

Consider the following strategy σΣ\sigma\in\Sigma. If pconv(Q)p\in\text{conv}(Q), split pp into kk elements of QQ according to some prior μpΔ(Q)\mu_{p}\in\Delta(Q). Otherwise, ignore private information and send the fixed signal s0Ss_{0}\in S. Assume that pn=qp_{n}=q for some n1n\geq 1. If qconv(Q)q\notin\text{conv}(Q), the sender ignores his private information and sends the fixed signal s0Ss_{0}\in S. Next, if qconv(Q)Qq\in\text{conv}(Q)\setminus Q, σ\sigma instructs the sender to choose a split from 𝒮qMQ\mathcal{S}_{qM}^{Q} (see (37) for the definition of 𝒮qMQ\mathcal{S}_{qM}^{Q}). Finally, if q=qjQq=q^{j}\in Q, σ\sigma instructs the sender to split qMqM into {(qi,wij)}i\{(q_{i},w^{j}_{i})\}_{i\in\mathcal{R}}, where (w1j,,w||j)(w^{j}_{1},...,w^{j}_{|\mathcal{R}|}) is the jj’th row of WW_{\mathcal{R}}. Note that such a split is available to the sender because each row in WW contains at most kk non-zero elements, and k|S|k\leq|S|.

It follows from the definition of σ\sigma that for each pconv(Q)p\in\text{conv}(Q), the sequence of posteriors (pn)(p_{n}) follows a Markov chain over the state space QQ with initial probability μp\mu_{p} and transition rule given by the stochastic matrix WW_{\mathcal{R}}. Since the latter is irreducible, we may employ the Ergodic Theorem for Markov chains (e.g., Theorem C.1 in [20]) to obtain that

(43) limN1Nn=1Nu(pn)=i=1||νiu(qi),p,σ-a.s.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}u(p_{n})=\sum_{i=1}^{|\mathcal{R}|}\nu_{i}u(q_{i}),\ \ \ \mathbb{P}_{p,\sigma}\emph{-}\text{a.s.}

where ν=(νi)i=1||\nu=(\nu_{i})_{i=1}^{|\mathcal{R}|} is the unique stationary distribution of WW_{\mathcal{R}}. Furthermore, since QAzQ\subseteq A_{z}, we have that

(44) i=1||νiu(qi)=i=1||νifz(qi)=fz(i=1||νiqi)=(Cavu)(i=1||νiqi),\sum_{i=1}^{|\mathcal{R}|}\nu_{i}u(q_{i})=\sum_{i=1}^{|\mathcal{R}|}\nu_{i}f_{z}(q_{i})=f_{z}\left(\sum_{i=1}^{|\mathcal{R}|}\nu_{i}q_{i}\right)=(\text{Cav}\,u)(\sum_{i=1}^{|\mathcal{R}|}\nu_{i}q_{i}),

where the last equality follows from item (i) of Lemma 4. Next, by multiplying by ν\nu from the left both sides of Eq. (42) we obtain

(45) νPM=νWP=νP,\nu P_{\mathcal{R}}M=\nu W_{\mathcal{R}}P_{\mathcal{R}}=\nu P_{\mathcal{R}},

which together with the uniqueness of πM\pi_{M} implies that νP=πM\nu P_{\mathcal{R}}=\pi_{M}. However, as πM=νP=i=1||νiqi\pi_{M}=\nu P_{\mathcal{R}}=\sum_{i=1}^{|\mathcal{R}|}\nu_{i}q_{i}, we deduce that

(46) limN1Nn=1Nu(pn)=(Cavu)(πM),p,σ-a.s.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}u(p_{n})=(\text{Cav}\,u)(\pi_{M}),\ \ \ \mathbb{P}_{p,\sigma}\emph{-}\text{a.s.}

Therefore, by a Tauberian theorem (see, e.g., Theorem 3.1. in [25]),

(47) limλ1(1λ)n=1λn1u(pn)=(Cavu)(πM),p,σ-a.s.\lim_{\lambda\to 1^{-}}(1-\lambda)\sum_{n=1}^{\infty}\lambda^{n-1}u(p_{n})=(\text{Cav}\,u)(\pi_{M}),\ \ \ \mathbb{P}_{p,\sigma}\emph{-}\text{a.s.}

Finally, assume that πMint(conv(Q))\pi_{M}\in\text{int}(\text{conv}(Q)). Then, since MM is irreducible and aperiodic, for some finite time period NN for every pΔ(K)p\in\Delta(K). Then, by the definition of σ\sigma, we see that (46) holds for every pΔ(K)p\in\Delta(K) and thus so does (47) ∎

The proof of Proposition 1 may enhance the intuition about absorbing sets.

Proof of Proposition 1.

By Carathéodory’s Theorem (see, e.g., Corollary 17.1.5 in [24]) we may assign with each qCq\in C kk distributions w1(q),,wk(q)Cw_{1}(q),...,w_{k}(q)\in C such that qMconv({w1(q),,wk(q)}qM\in\text{conv}(\{w_{1}(q),...,w_{k}(q)\}. Define a correspondence ξ:C2C\xi:C\to 2^{C} by ξ(q)={w1(q),,wn(q)}\xi(q)=\{w_{1}(q),...,w_{n}(q)\}. In particular, qMξ(q)qM\in\xi(q). The set 𝒜(q):=n=1ξn1(q)\mathcal{A}(q):=\bigcup_{n=1^{\infty}}\xi^{n-1}(q) is MM-absorbing for every qCq\in C, where ξn1\xi^{n-1} is the composition of ξ\xi with itself n1n-1 times. Indeed, let w𝒜(q)w\in\mathcal{A}(q), and let n1n\geq 1 such that wξn1(q)w\in\xi^{n-1}(q). By the definition of ξ\xi we have that wMconv(ξ(w))conv(ξn(q))conv(𝒜(q))wM\in\text{conv}(\xi(w))\subseteq\text{conv}(\xi^{n}(q))\subseteq\text{conv}(\mathcal{A}(q)) as desired. ∎

We end the proof section with the proof of Theorem 6.

Proof of Theorem 6..

Suppose that MM is a homothety and let us fix uC(Δ(K))u\in C(\Delta(K)). As uu is continuous, Carathéodory’s Theorem (see, e.g., Corollary 17.1.5 in [24]) implies that there exist q1,,qmΔ(K)q_{1},...,q_{m}\in\Delta(K), mkm\leq k, and positive convex weights (αi)i=1m(\alpha_{i})_{i=1}^{m} such that πM=i=1mαiqi\pi_{M}=\sum_{i=1}^{m}\alpha_{i}q_{i} and (Cavu)(πM)=i=1mαiu(qi)(\text{Cav}\,u)(\pi_{M})=\sum_{i=1}^{m}\alpha_{i}u(q_{i}). Hence, by item (ii) of Lemma 4, qiAzq_{i}\in A_{z} for every ii and every zΛz\in\Lambda. Therefore, πMconv(Az)\pi_{M}\in\text{conv}(A_{z}) for every zΛz\in\Lambda. The latter coupled with the convexity of conv(Az)\text{conv}(A_{z}) implies that conv(Az)\text{conv}(A_{z}) is star shaped around πM\pi_{M}. Hence, since MM is a homothety we get that conv(Az)\text{conv}(A_{z}) is MM-absorbing for any zΛz\in\Lambda. By the definition of an MM-absorbing set, AzA_{z} must also be MM-absorbing for any zΛz\in\Lambda. By Proposition 2, we deduce that v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}), proving the first direction of Theorem 6.

Suppose now that v=(Cavu)(πM)v_{\infty}=(\text{Cav}\,u)(\pi_{M}) for every uC(Δ(K))u\in C(\Delta(K)). For each iKi\in K, let eiΔ(K)e_{i}\in\Delta(K) be the Dirac measure concentrated on the ii’th coordinate of k\mathbb{R}^{k}. Fix iKi\in K and consider for each n1n\geq 1 the vector ein=πM+(πMei)/ne_{i}^{n}=\pi_{M}+(\pi_{M}-e_{i})/n. Clearly, πM[ei,ein]\pi_{M}\in[e_{i},e_{i}^{n}] for all nn. Next, as πMint(Δ(K))\pi_{M}\in\text{int}(\Delta(K)), there exists NiN_{i} such that einΔ(K)e_{i}^{n}\in\Delta(K) for every nNin\geq N_{i}. For each nNin\geq N_{i} we take an element ui,nC(Δ(K))u_{i,n}\in C(\Delta(K)) satisfying ui,n(q)=1u_{i,n}(q)=1 for q{ei,ein}q\in\{e_{i},e_{i}^{n}\} and ui,n(q)<1u_{i,n}(q)<1 for qΔ(K){ei,ein}q\in\Delta(K)\setminus\{e_{i},e_{i}^{n}\}. We have that (Cavu)(πM)=1(\text{Cav}\,u)(\pi_{M})=1. Moreover, by the definition of ui,nu_{i,n}, 0Λ0\in\Lambda (for Λ\Lambda corresponding to u=ui,nu=u_{i,n}), because the hyperplane f0(x)=(Cavu)(πM)=1f_{0}(x)=(\text{Cav}\,u)(\pi_{M})=1 for all xkx\in\mathbb{R}^{k}, supports (Cavu)(\text{Cav}\,u) at πM\pi_{M}. As A0={ei,ein}A_{0}=\{e_{i},e_{i}^{n}\}, Proposition 2 shows that {ei,ein}\{e_{i},e_{i}^{n}\} contains an MM-absorbing subset. However, as MM has a unique stationary distribution πM{ei,ein}\pi_{M}\notin\{e_{i},e_{i}^{n}\} we get that neither {ei}\{e_{i}\} nor {ein}\{e_{i}^{n}\} is MM-absorbing. Therefore, {ei,ein}\{e_{i},e_{i}^{n}\} must be MM-absorbing for every nNin\geq N_{i}. In particular, eiM(ei,ein]e_{i}M\in(e_{i},e_{i}^{n}] for every nNin\geq N_{i}. Since einπMe_{i}^{n}\to\pi_{M} as nn\to\infty, we obtain that eiM(ei,πM]e_{i}M\in(e_{i},\pi_{M}]. As ii was arbitrary, we have shown that for each iKi\in K there exists βi[0,1)\beta_{i}\in[0,1) such that eiM=βiei+(1βi)πMe_{i}M=\beta_{i}e_{i}+(1-\beta_{i})\pi_{M}. Since qMqq\to Mq is a linear operator, to prove that MM is a homothety is suffices to show that βi=βj\beta_{i}=\beta_{j} for all ijKi\neq j\in K.

The proof now bifurcates according to the dimension of Δ(K)\Delta(K). First, let us assume that k=2k=2. Since MM is irreducible, there exists a unique α(0,1)\alpha\in(0,1) so that πM=αe1+(1α)e2\pi_{M}=\alpha e_{1}+(1-\alpha)e_{2}. We have

(48) πMM=(αe1+(1α)e2)M=α(e1M)+(1α)(e2M)=α(β1e1+(1β1)πM)+(1α)(β2e2+(1β2)πM).\begin{split}\pi_{M}M&=(\alpha e_{1}+(1-\alpha)e_{2})M\\ &=\alpha(e_{1}M)+(1-\alpha)(e_{2}M)\\ &=\alpha(\beta_{1}e_{1}+(1-\beta_{1})\pi_{M})+(1-\alpha)(\beta_{2}e_{2}+(1-\beta_{2})\pi_{M}).\end{split}

By plugging πM=αeq+(1α)e2\pi_{M}=\alpha e_{q}+(1-\alpha)e_{2} into the last expression in Eq. (48) and using simple algebraic manipulations, we get that the convex weight ρ\rho of e1e_{1} in the convex decomposition of πMM\pi_{M}M to e1e_{1} and e2e_{2} equals

αβ1+α2(1β1)+(1α)(1β2)α.\alpha\beta_{1}+\alpha^{2}(1-\beta_{1})+(1-\alpha)(1-\beta_{2})\alpha.

However, as πM=πMM\pi_{M}=\pi_{M}M, we must have that ρ=α\rho=\alpha. After some further simple algebraic manipulations, one gets that the equality ρ=α\rho=\alpha is equivalent to β1β2=α(β1β2)\beta_{1}-\beta_{2}=\alpha(\beta_{1}-\beta_{2}). As α(0,1)\alpha\in(0,1), we obtain that β1=β2\beta_{1}=\beta_{2}, thus proving that MM is a homothety whenever k=2k=2.

Next, let k3k\geq 3. Assume that βi<βj\beta_{i}<\beta_{j} for some ijKi\neq j\in K. Define v=(ei+ej)/2v=(e_{i}+e_{j})/2. Since |supp(v)|=2|\text{supp}(v)|=2, whereas |supp(πM)|=k3|\text{supp}(\pi_{M})|=k\geq 3, we have that πMv\pi_{M}\neq v. Consider for each nn\in\mathbb{N} the vector vn=πM+(πMv)/nv_{n}=\pi_{M}+(\pi_{M}-v)/n. Then πM[v,vn]\pi_{M}\in[v,v_{n}] for every nn. Moreover, since πMint(Δ(K))\pi_{M}\in\text{int}(\Delta(K)) there exists N0N_{0} such that for every nN0n\geq N_{0} we have vnΔ(K)v_{n}\in\Delta(K). As at the beginning of the proof we take for each nN0n\geq N_{0} an element unC(Δ(K))u_{n}\in C(\Delta(K)) satisfying un(q)=1u_{n}(q)=1 for q{v,vn}q\in\{v,v_{n}\} and un(q)<1u_{n}(q)<1 for qΔ(K){v,vn}q\in\Delta(K)\setminus\{v,v_{n}\}. Hence, by arguing as before for ei,eine_{i},e_{i}^{n} and uinu_{i}^{n}, only this time for v,vnv,v_{n} and unu_{n}, we obtain that vM(v,vn]vM\in(v,v_{n}] for every nN0n\geq N_{0}. As vnπMv_{n}\to\pi_{M} as nn\to\infty we see that vM(v,πM]vM\in(v,\pi_{M}]. On the other hand,

(49) vM=12(eiM+ejM)=12(βiei+(1βi)πM)+12(βjej+(1βj)πM)=(1βi2βj2)πM+βiv+12(βjβi)ej.\begin{split}vM&=\frac{1}{2}\,(e_{i}M+e_{j}M)\\ &=\frac{1}{2}\,(\beta_{i}e_{i}+(1-\beta_{i})\pi_{M})+\frac{1}{2}\,(\beta_{j}e_{j}+(1-\beta_{j})\pi_{M})\\ &=\left(1-\frac{\beta_{i}}{2}-\frac{\beta_{j}}{2}\right)\pi_{M}+\beta_{i}v+\frac{1}{2}\,(\beta_{j}-\beta_{i})e_{j}.\end{split}

As 0βi<βj<10\leq\beta_{i}<\beta_{j}<1, this implies that vMvM lies in the relative interior of the triangle defined by πM\pi_{M}, vv, and eje_{j}. This of course contradicts the fact that vM(v,πM]vM\in(v,\pi_{M}]. Hence, βi=βj\beta_{i}=\beta_{j} for every ijKi\neq j\in K, thus proving that MM is a homothety. ∎

References

  • [1] Abreu, D., Pearce, D., and Stacchetti, E. (1990), Toward a Theory of Discounted Repeated Games with Imperfect Monitoring. Econometrica, 58(5), 1041–1063.
  • [2] Arieli, I. and Babichenko, Y. (2019), Private Bayesian persuasion. Journal Economic Theory, 182 ,185–217.
  • [3] Arieli, I., Babichenko, Y., Smorodinsky, R. and Yamashita, T. (2020), Optimal Persuasion via Bi-Pooling. In:The Twenty-First ACM Conference on Economics and Computation.
  • [4] Athey, S. and Bagwell, K. (2008), Collusion with Persistent Cost Shocks. Econometrica, 76(3), 493-540.
  • [5] Augenblick, N. and Bodoh-Creed, A. (2018), To Reveal or not to Reveal: Privacy Preferences and Economic Frictions. Games and Economic Behavior, 110, pp. 318–329
  • [6] Aumann, R.J. and Maschler, M. (1995), Repeated games with incomplete information. With the collaboration of R. Stearns, Cambridge, MA: MIT Press.
  • [7] Blackwell, D. (1951), Comparison of Experiments. In: Proceedings of the Second Berkeley Symposium on Mathematical Statistics and Probability, 1950, pp. 93–102. University of California Press, Berkeley and Los Angeles, Calif.
  • [8] Dziuda, W. and Gradwohl, R. (2015), Achieving Cooperation Under Privacy Concerns. Am. Econ. J. Microecon., 7, 142–173
  • [9] Ely, J.C. (2017), Beeps. American Economic Review, 107(1): 31–53.
  • [10] Escobar, J.F. and Toikka, J. (2013), Efficiency in Games with Markovian Private Information. Econometrica, 81, 1887–1934.
  • [11] Fudenberg, D. and Maskin, E. (1986), The Folk Theorem in Repeated Games with Discounting or with Incomplete Information. Econometrica, 54(3), 533–554.
  • [12] Farhadi, F. and Teneketzis, D. (2021), Dynamic Information Design: A Simple Problem on Optimal Sequential Information Disclosure. Dyn Games Appl.
  • [13] Ganglmair, B. and Tarantino, E. (2014), Conversation With Secrets. Rand J. Econ., 45, 273–302.
  • [14] Honryo, T. (2018), Dynamic Persuasion. J. Econom. Theory, 178, 36–58.
  • [15] Horn, R.A. and Johnson, C.R. (2013), Matrix analysis. Second edition. Cambridge University Press, Cambridge.
  • [16] Kamenica, E. and Gentzkow, M. (2011), Bayesian Persuasion. American Economic Review, 101 (6): 2590–2615.
  • [17] Lehrer, E. and Sorin, S. (1992), A Uniform Tauberian Theorem in Dynamic Programming. Math. Oper. Res. 17, no. 2, 303–307.
  • [18] Lingenbrink, D. and Iyer, K. (2019) Optimal Signaling Mechanisms in Unobservable Queues with Strategic Customers. Oper. Res. 67, no. 5, 1397-1416.
  • [19] Mailath, G. and Samuelson, L. (2001), Who Wants a Good Reputation?. Rev. Econom. Stud., 68, no. 2, 415–441.
  • [20] Levin, D. A.; Peres, Y. (2017), Markov chains and mixing times, second edition. With contributions by Elizabeth L. Wilmer. With a chapter on “Coupling from the past” by James G. Propp and David B. Wilson. American Mathematical Society, Providence, RI.
  • [21] Phelan, C. (2006), Public Trust and Government Betrayal. J. Econom. Theory, 127(1), 27–43.
  • [22] Renault, J., Solan, E., and Vieille, N. (2017), Optimal Dynamic Information Provision. Games Econ. Behav., 104, 329–349.
  • [23] Renault, J. (2018), Repeated games with incomplete information. In: Meyers R. (eds) Encyclopedia of Complexity and Systems Science. Springer, Berlin, Heidelberg.
  • [24] Rockafellar, R. T. (1970), Convex analysis. Princeton University Press.
  • [25] Solan, E. (2021), A Course in Stochastic Game Theory. Cambridge University Press. To appear.
  • [26] Wiseman, T. (2008), Reputation and Impermanent Types. Games Econ. Behav., 62, 190–210.