1 Introduction

The theory of network games has become a booming field and attracted increasing attention from social and computer scientists.Footnote 1 One direction of research in this field is to model social interactions by letting two-player games be simultaneously played by players connected through network structures. In this paper, we study such a network extension of matching pennies.Footnote 2 There is generally no pure strategy Nash equilibrium (PNE for short) in this model, so we study its best-response dynamics, which are mimics of individuals’ adaptive behaviors and the limit cycles of which to be introduced soon could also be viewed as extensions of PNE.

Matching pennies is an asymmetric game. It can be intuitively interpreted as that the row player (called a conformist) preferring to coordinate with her opponent’s action, while the column player (called a rebel) preferring to anti-coordinate. Therefore, in a network extension of matching pennies, it is natural to assume that a pair of conformists play a coordination game, a pair of rebels play an anti-coordination game, and a conformist-rebel pair plays matching pennies. With this in mind, the model, to be referred to as the network matching pennies game (NMP for short), can be decomposed into three elementary base games, namely, matching pennies, a coordination game, and an anti-coordination game, among which matching pennies is the most essential one, because it is the only one that has both types of players. The NMP model is originally proposed by Jackson (2008, P.271) for the analysis of the phenomenon of fashion and subsequently analyzed by Cao et al. (2013), Cao and Yang (2014), and Zhang et al. (2018). However, the potential applications of NMP are not restricted to fashion. For example, the dynamics of NMP, as studied in this paper, may be viewed as learning processes of heterogeneous agents with different learning rules. Conforming and rebelling behaviors are also closely related to herding and contrarian behaviors, which are well recognized as being critical to understanding financial markets (Park and Sabourian 2011). In fact, there are at least three types of conforming/rebelling behaviors, namely, pure conformity/rebellion, instrumental conformity/rebellion, and informational conformity/rebellion, as detailed in Young (2001) and Zhang et al. (2018).

1.1 Results

We mainly focus on the simultaneous best response dynamics (simultaneous BRD for short). Since the simultaneous BRD is deterministic, and there are a total of finitely many states, it will eventually enter a cycle, which is usually called a limit cycle in the literature on dynamic systems. We are mainly concerned with the lengths of limit cycles (LLC for short) under simultaneous BRD. It is clear that a PNE must exist for the corresponding NMP when \(\hbox {LLC}=1\).

Dynamic analyses on general network structures are usually very complicated. For the particular model that we are interested in, even the static problem of deciding whether a PNE exists is computationally hard (Cao and Yang 2014). For this reason, we shall investigate two benchmark cases: the case with a single type of agents and the case without direct interaction among agents of the same type, and three special network structures: lines, rings or stars. Although these network structures seem simple, some of the analyses are nontrivial. In addition, they are frequently analyzed in the literature on social learning and cultural dynamics, a field that is closely related to this paper.Footnote 3

Our main findings are summarized as follows. (i) We show that \(\hbox {LLC}=1\) or 2 when all agents are of the same type (Theorem 1).Footnote 4 (ii) We generalize the simple fact that matching pennies has \(\hbox {LLC}=4\) to NMPs with all edges being conformist-rebel ones under certain additional conditions (Theorem 2). (iii) When the underlying network is a line or a star, all the possible LLCs for NMPs are 1, 2 and 4 (Theorems 3, 5). When the network is a ring, however, LLC can be as large as twice the size of the network (Remark 5), although in most cases the properties with rings are similar to those with lines (Theorem 4). (iv) If a network is sufficiently large, then regardless of the initial action profiles, NMP has \(\hbox {LLC}=1\) for almost all type configurations when the network is a line or a ring (Theorem 6(a)), and \(\hbox {LLC}=4\) for approximately one half of the type configurations when the network is a star (Theorem 6(b)).

In an appendix, we also investigate the sequential BRD. It turns out that when the underlying network is a line or a ring, NMP is an ordinal potential game, which implies that the sequential BRD converges to a PNE, if and only if the network is not completely heterophilic (i.e., not all edges are conformist-rebel ones) (Theorem 7). For star networks, we consider a particular sequential BRD, where at each time step unsatisfied agents are randomly selected to deviate with the probability that each unsatisfied agent is selected being positive. We show that when the underlying network is a star, the probability that the above sequential BRD converges is one if and only if at least one half of the peripheral agents are of the same type as the central agent, a condition that is equivalent to the existence of a PNE (Theorem 8).

1.2 Related work

The NMP model is originally proposed by Jackson (2008), who observed that when each agent has no fewer conformist neighbors than rebel neighbors, the existence of a PNE is guaranteed. Applying a partial potential analysis, Zhang et al. (2018) established an improvement of the above observation by showing that we only need to require each conformist have no fewer conformist neighbors than rebel neighbors to guarantee the existence of a PNE. The major difference between this paper and Zhang et al. (2018) is that the latter paper studies sequential BRD by transforming the discrete dynamic into a system of ordinary differential equations, whereas the present paper mainly focuses on simultaneous BRD, a dynamic that is much less understood in the literature than its sequential counterpart. In addition, the present paper also carries out a rigorous analysis, instead of an approximate analysis as in Zhang et al. (2018), of the sequential BRD. The major difference between this paper and Cao and Yang (2014) is that the latter focuses on computational issues. While we are concerned with LLC and convergence of BRDs, Cao et al. (2013) concentrated on the cooperation level problem in NMP through numerical simulations.

Since the coordination game and the anti-coordination game are both base games, NMP can also be viewed as a combination of the network coordination game and the network anti-coordination game. The literature on network coordination games is too vast for us to give a complete survey here. In particular, Apt et al. (2017) considered general coordination games on networks where agents typically have different action sets. Several other related useful results for this paper will be introduced when they are applied. The network anti-coordination game, in contrast, attracts considerably less attention. The most influential papers on this topic are Bramoullé et al. (2004) and Bramoullé (2007).Footnote 5 In the field of social physics, a closely related and extensively studied model is the minority game (also known as the El Farol Bar problem), a model that was proposed by Arthur (1994) and developed by Challet and Zhang (1997) with numerous follow-ups. We refer the reader to Szabó et al. (2014) for a recent study of the evolutionary matching pennies game on bipartite regular networks.

Our research falls into the booming field of network games. The NMP model is a mixture of a network game with strategic complements and a network game with strategic substitutes, the two most extensively studied classes of network game models. Indeed, it exhibits a simple form of strategic complement for conformists and a simple form of strategic substitute for rebels. In a recent paper by Bramoullé et al. (2014), mixtures of strategic complements and strategic substitutes are allowed, but action sets in their paper are continuous. In Hernandez et al. (2013), both strategic complement and strategic substitute are considered. However, for any specific values of parameters in their model, strategic complement and strategic substitute do not coexist. Other recent related works include Southwell and Cannings (2013), Ramazi et al. (2016), Haslegrave and Cannings (2017), and Shirado and Christakis (2017).Footnote 6

The rest of this paper is organized as follows. Section 2 formally introduces the NMP model. Section 3 deals with two benchmark cases for the simultaneous BRD. Section 4 considers three special network structures for the simultaneous BRD. Section 5 concludes the paper with several further discussions. Most of the proofs and the investigation of the sequential BRD are organized in an appendix.

2 The model

In the NMP model, there are two types of agents, conformists and rebels, located on a fixed network. Conformists like to match the majority action of her neighbors, while rebels prefer to match the minority. Each agent has two available actions, named 0 and 1, which can be interpreted as wearing white or black skirts, buying or not buying an iphone, etc.Footnote 7 According to their types, each pair of neighboring agents play one of the three base games, the payoffs of which are described in Fig. 1. We assume that each agent takes the same action in all of the base games she plays. The overall payoff of each agent is simply the sum of the payoffs that she receives in all the games she plays.

Fig. 1
figure 1

The three base games. (1) The coordination game (left): a conformist versus a conformist; (2) the anti-coordination game (middle): a rebel versus a rebel; (3) matching pennies (right): a conformist versus a rebel. Conformists are represented by circles, and rebels by triangles

Formally, we use \(\{C,R\}\) to denote the type set, where C stands for conformists and R for rebels. For each agent i, \(T_i\in \{C,R\}\) is her type. An NMP is represented by a triple \(G=(N,\mathrm{g},{\mathbf {T}})\), where

  • \(N=\{1,2,\ldots ,n\}\) is the set of agents;

  • \(\mathrm{g}\subseteq N\times N\) is the set of links (edges);

  • \({\mathbf {T}}=(T_1,T_2,\ldots ,T_n)\in \{C,R\}^n\) is the configuration of agents’ types.

The underlying network \((N,\mathrm{g})\) is undirected. That is, ij and ji represent the same link. Two agents \(i,j\in N\) are neighboring each other if and only if \(ij\in \mathrm{g}\). We also use \(N_i\) to denote the set of neighbors of agent i, i.e., \(N_i=\{j\in N: ij\in \mathrm{g}\}\). In simultaneous BRD, time elapses discretely. Agents’ types do not change over time, but their actions may change. We use

$$\begin{aligned} {\mathbf {x}}(0)=(x_1(0),x_2(0),\ldots ,x_n(0))\in \{0,1\}^n \end{aligned}$$

to denote the initial action profile, where \(x_i(0)\) is the initial action of agent i. We define \(x_i(t)\in \{0,1\}\) as the action of agent i at time t, and write \({\mathbf {x}}(t)=(x_1(t),x_2(t),\ldots ,x_n(t))\).

Define \(L_i({\mathbf {x}}(t))\subseteq N_i\) as the subset of neighbors at time t that are liked by agent i. Since conformists like those taking the same action while rebels the opposite, we have

$$\begin{aligned} L_i({\mathbf {x}}(t))=\left\{ \begin{array}{ll}\{j\in N_i: x_j(t)=x_i(t)\} &{}\quad if\ T_i=C,\\ \{j\in N_i: x_j(t)\ne x_i(t)\}&{}\quad if\ T_i=R.\end{array}\right. \end{aligned}$$

Set \(D_i({\mathbf {x}}(t))=N_i{\setminus } L_i({\mathbf {x}}(t))\) as the subset of neighbors that are disliked by agent i at time t. According to the payoff settings in Fig. 1, agent i gains 1 unit from interacting with each neighboring agent in \(L_i({\mathbf {x}}(t))\) but loses 1 unit from interacting with each of those in \(D_i({\mathbf {x}}(t))\). Thus, given an action profile \({\mathbf {x}}(t)\) at time t, the utility of agent i is given by

$$\begin{aligned} u_i({\mathbf {x}}(t))=|L_i({\mathbf {x}}(t))|-|D_i({\mathbf {x}}(t))|. \end{aligned}$$

When \(u_i({\mathbf {x}}(t)\ge 0\), we may also say that agent i is satisfied (at time t). Otherwise, she is called unsatisfied. Under the simultaneous BRD, agent i switches her action at time t if and only if she is unsatisfied:

$$\begin{aligned} x_i(t+1)=\left\{ \begin{array}{ll}1-x_i(t)&{}\quad if\ u_i({\mathbf {x}}(t))<0,\\ x_i(t)&{}\quad otherwise.\end{array}\right. \end{aligned}$$

The focus of our analysis is on the limit cycle, a basic concept from dynamic systems. Intuitively, a limit cycle is an ordered set of action profiles such that starting from each of these profiles the system will evolve into its immediate successor (the successor of the last action profile is the first one). Since there are a total of finitely many action profiles in the NMP and simultaneous BRD is deterministic, some profile will reappear after a certain number of steps. From that time on, everything will be repeated just as that state is reached for the first time. Therefore, limit cycle always exists and it is unique for any NMP and an initial action profile.

In the rest of this paper, we also use \({\mathcal {I}}=(N,\mathrm{g},{\mathbf {T}},{\mathbf {x}}(0))\) to denote an initialized NMP. Below is the formal definition of a limit cycle.

Definition 1

Let \({\mathcal {I}}=(N,\mathrm{g},{\mathbf {T}},{\mathbf {x}}(0))\) be an initialized NMP, and \(t({\mathcal {I}})\ge 1\) be the first time such that

$$\begin{aligned} {\mathbf {x}}(t({\mathcal {I}}))\in \{{\mathbf {x}}(0),{\mathbf {x}}(1),\ldots , {\mathbf {x}}(t({\mathcal {I}})-1)\}. \end{aligned}$$

Let \(r({\mathcal {I}})<t({\mathcal {I}})\) be the time such that \({\mathbf {x}}(t({\mathcal {I}}))={\mathbf {x}}(r({\mathcal {I}}))\). We call the ordered set of states \(({\mathbf {x}}(r({\mathcal {I}})),{\mathbf {x}}(r({\mathcal {I}})+1),\ldots , {\mathbf {x}}(t({\mathcal {I}})-1))\) the limit cycle of \({\mathcal {I}}\).

In the rest of this paper, we use \(\ell ({\mathcal {I}})=t({\mathcal {I}})-r({\mathcal {I}})\) to denote the length of the limit cycle (LLC for short) of \({\mathcal {I}}\). When time goes beyond \(r({\mathcal {I}})\), we say that simultaneous BRD enters the limit cycle. For an initialized NMP \({\mathcal {I}}\), it is clear that a PNE will be eventually reached via simultaneous BRD if and only if \(\ell ({\mathcal {I}})=1\), in which case we also say that the simultaneous BRD converges.

Example 1

Matching pennies, when viewed as an NMP with exactly one conformist and one rebel who are connected, has \(\hbox {LLC}=4\). See Fig. 2 for an illustration.

Fig. 2
figure 2

\(\text {LLC}=4\) in matching pennies. Conformists and rebels are represented by circles and triangles, respectively. Actions are indicated by colors: black for 1 and white for 0

Remark 1

A closely related model to NMP that has been extensively studied is the threshold model (Granovetter 1978). In the threshold model, there is a threshold for each agent and she prefers action 1 over action 0 if and only if the number of her neighbors choosing action 1 is greater than or equal to this threshold. When the threshold is one half of the degree for each agent, the corresponding simultaneous BRD is also known as the majority dynamic. It is worth noting that the preference of a conformist in NMP is not exactly the same as in the majority dynamic. The subtle difference lies in the tie-breaking rules: when the number of action 1 neighbors of an agent is exactly the threshold, action 1 is biased in the majority dynamic, but action 1 and action 0 are symmetric in NMP. It has been pointed out that the tie-breaking rule may be very critical in global behaviors of the dynamic and other tie-breaking rules have also been studied in the literature (Gärtner and Zehmakan 2017).

3 Simultaneous BRD: two benchmark cases

In this section, we consider two benchmark cases. (i) The case where either all agents are conformists (a network coordination game) or all agents are rebels (a network anti-coordination game). Note that in this case all interactions are inner-type ones, i.e., there is no conformist-rebel edge. (ii) The case where all interactions are cross-type ones, i.e., all edges are conformist-rebel ones. In terms of homophily, the first benchmark is the completely homophilic case and the second benchmark is the completely heterophilic case.Footnote 8

3.1 The completely homophilic case

It is well-known that any threshold model has an LLC of either one or two via a Lyapunov function method (Poljak and Sura 1983; Goles 1987; Berninghaus and Schwalbe 1996). It was pointed out by Cannings (2009) that the same Lyapunov function works for anti-threshold models, deriving an LLC of either one or two. However, as we have pointed out in Remark 1, there is a subtle difference between the tie-breaking rules of a threshold model and the network coordination game. As a result, the Lyapunov functions constructed in Poljak and Sura (1983), Goles (1987) and Berninghaus and Schwalbe (1996) do not apply to the network coordination game or the network anti-coordination game. Yet, the \(LLC\le 2\) property still holds.

Theorem 1

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP such that all agents are of the same type. Then LLC equals 1 or 2.

Our proof is based on the Lyapunov function applied in a recent paper by Haslegrave and Cannings (2017), who studied a mixture of a threshold model and an anti-threshold model, which has different tie-breaking rules with our model.

For each agent \(i\in N\) and time \(t\ge 1\), we use \(n_i^1(t)\) to denote the number of i’s neighbors that use action 1 at time t. The following observations are critical.

Observation 1

Suppose i is a conformist. (i) If \(x_i(t+1)=1,\) then \(n_i^1(t) \ge n_i/2\). (ii) If \(x_i(t+1)=0,\) then \(n_i^1(t)\le n_i/2\).

Note that the other way around is incorrect in both cases of the observation, because \(x_i(t+1)=x_i(t)\) when \(n_i^1(t)=n_i/2\). In addition, if \(n_i\) is odd, the two inequalities are both strict. We have an analogous observation for rebels.

Observation 2

Suppose i is a rebel. (i) If \(x_i(t+1)=1,\) then \(n_i^1(t) \le n_i/2\). (ii) If \(x_i(t+1)=0,\) then \(n_i^1(t)\ge n_i/2\).

Using the above two observations, it can be shown that the proof of Haslegrave and Cannings (2017) can be adapted for Theorem 1 (Appendix A).

Remark 2

Theorem 1 may not be valid in the general case, even if there is only one conformist-rebel edge. The example in Fig. 3 provides an illustration.

Fig. 3
figure 3

\(\text {LLC}=6\) in an NMP with one conformist-rebel edge. Conformists and rebels are represented by circles and triangles, respectively. Actions are indicated by colors: black for 1 and white for 0. Unhappy faces indicate that the corresponding agents are unsatisfied

3.2 The completely heterophilic case

As illustrated in Fig. 2, matching pennies always has \(\hbox {LLC}=4\) under simultaneous BRD. Theorem 2 of this subsection is a generalization of this simple fact. Before presenting this result, we provide a formal definition of complete heterophily.

Definition 2

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP. If \(T_i\ne T_j\) for all \(ij\in \mathrm{g}\), i.e., the neighbors of any conformist are all rebels and the neighbors of any rebel are all conformists, we say that G is completely heterophilic.

If G is completely heterophilic, then the underlying network \((N,\mathrm{g})\) is a bipartite network with one side of all conformists and the other side of all rebels.Footnote 9 This structure is a natural generalization of that in the matching pennies game, because only matching pennies will be played in this NMP but the other two base games do not occur.

Theorem 2

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be a completely heterophilic NMP. If no two even-degreed agents (if any) are neighboring each other,  then \(\hbox {LLC}=4\).

Proof

We prove in two steps. In the first step we show via a new Lyapunov function that \(x_i(t-1)\ne x_i(t+1)\) for any agent i with an odd degree and large enough t such that a limit cycle is reached at time \(t-1\). This implies that the period of the action series of agent i in the limit cycle is 4. In the second step, we show that the action series of an even-degreed agent is of either period 1 or period 4, if her neighbours are all odd-degreed. The proof is completed by combining these two steps.

Step 1: Denote \(\alpha _C(t)=\sum _{i\in C: x_i(t)=1}n_i^1(t-1)\) as the number of ij links such that \(i\in C, x_i(t)=1\) and \(j \in R, x_j(t-1)=1\), and \(\beta _C(t)=\sum _{i\in C: x_i(t)=1 \text{ or } x_i(t-1)=0}n_i/2\). Analogously, denote \(\alpha _R(t)=\sum _{i\in R: x_i(t)=1}n_i^1(t-1)\) as the number of ij links such that \(i\in R, x_i(t)=1\) and \(j \in C, x_j(t-1)=1\), and \(\beta _R(t)=\sum _{i\in R: x_i(t)=1 \text{ or } x_i(t-1)=0}n_i/2\). Define a Lyapunov function

$$\begin{aligned} \gamma (t)=[\alpha _C(t)-\beta _C(t)]-[\alpha _R(t)-\beta _R(t)]. \end{aligned}$$

It can be checked that

where

$$\begin{aligned} C_{1\rightarrow 1}(t)=\{i\in C: x_i(t-1)=1, x_i(t+1)=1\} \end{aligned}$$

is the set of conformists who take action 1 at time \(t-1\) and time \(t+1\), and \(C_{0\rightarrow 0}(t), R_{1\rightarrow 1}(t)\) and \(R_{0\rightarrow 0}(t)\) are analogously defined.

By Observations 1 and 2, \(\gamma (t+1)-\gamma (t)\) is always nonnegative. Consequently, \(\gamma (t)\) must be a constant in a limit cycle. In particular, if \(n_i\) is odd, then we must have \(x_i(t-1)\ne x_i(t+1)\) in the limit cycle, i.e, \(C_{1\rightarrow 1}(t), C_{0\rightarrow 0}(t), R_{1\rightarrow 1}(t)\) and \(R_{0\rightarrow 0}(t)\) are all empty. This implies that the action series of agent i has the form of \(00110011\cdots \) in the limit cycle, i.e., with a period of 4.

Step 2: We now consider the case that every neighbour of an even-degreed agent j has an odd degree. Suppose that agent j has k neighbours, among whom \(k_{j0}(t)\) agents use action 0 and \(k_{j1}(t)\) agents use action 1 at time t. Let \(k_{j}(t)=k_{j0}(t)-k_{j1}(t)\) be the value of the number of j’s action 0 neighbors minus the number of j’s action 1 neighbors at time t. From Step 1, the period of the action series of any of agent j’s neighbor i is 4 in a limit cycle, leading to \(x_i(t-1)\ne x_i(t+1)\). This implies that \(k_{j0}(t-1)=k_{j1}(t+1)\) and \(k_{j1}(t-1)=k_{j0}(t+1)\), and hence \(k_{j}(t)=-k_{j}(t+2)=k_{j}(t+4)\). For simplicity, we denote the time series of \(k_{j}(t)\) by \([k_{j}(t), k_{j}(t+1), -k_{j}(t), -k_{j}(t+1)]\).

The action of agent j at time \(t+1\), \(x_j(t+1)\), is determined by the type of j, \(x_j(t)\) and the sign of \(k_{j}(t)\). Since \(k_{j}(t)\) and \(k_{j}(t+1)\) can be either positive, negative or 0, there are 9 cases for the sign series of \([k_{j}(t), k_{j}(t+1), -k_{j}(t), -k_{j}(t+1)]\). For each case, it can be checked that the action series of agent j displays either period 1 or period 4, regardless of j’s type or \(x_j(t)\) (we omit the easy but tedious details). This implies that \(\hbox {LLC}=4\) if there is no link between even-degreed agents. \(\square \)

Remark 3

The condition that even-degreed agents are not neighboring each other is necessary to the correctness of Theorem 2. In fact, \(\hbox {LLC}=12\) is possible even if there are exactly two such agents, as shown in Fig. 4.

Fig. 4
figure 4

\(\text {LLC}=12\) in a completely heterophilic NMP with two even-degreed agents that are neighboring each other. Conformists and rebels are represented by circles and triangles, respectively. Actions are indicated by colors: black for 1 and white for 0. The third conformist and the third rebel are the only even-degreed agents and they are neighboring each other. Unhappy faces indicate that the corresponding agents are unsatisfied

4 Simultaneous BRD: three special network structures

In this section, we consider three special network structures, lines, rings, and stars.Footnote 10 The simultaneous BRD of NMP can be understood more clearly with these networks. Note first that if \((N,\mathrm{g})\) is a line or a ring, then complete heterophily means that conformists and rebels are alternate; and for star networks, complete heterophily means that all the peripheral agents are of the opposite type to the central one. The following basic results will be frequently used in later discussions.

Lemma 1

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP,  \({\mathbf {x}}(0)\) an initial action profile,  and \({\mathcal {I}}=(N,\mathrm{g},{\mathbf {T}},{\mathbf {x}}(0))\) the corresponding initialized NMP. Suppose also \(n\ge 2\).

  1. (a)

    G is an exact potential game if and only if there is no conformist-rebel link; 

  2. (b)

    If no agent is initially satisfied,  i.e.,  \(u_i({\mathbf {x}}(0))<0\) for all \(i\in N,\) then \(\ell ({\mathcal {I}})=2;\)

  3. (c)

    If \((N,\mathrm{g})\) is a line,  then G does not have a PNE if and only if it is completely heterophilic; 

  4. (d)

    If \((N,\mathrm{g})\) is a ring,  then G does not have a PNE if and only if it is completely heterophilic and n / 2 is odd; 

  5. (e)

    If \((N,\mathrm{g})\) is a star,  then G has a PNE if and only if at least one half of the peripheral agents are of the same type as the central agent.

Lemma 1(a) is provided in Zhang et al. (2018). In particular, Lemma 1(a) implies that a PNE exists in NMP when there is only one type of agents. Lemma 1(b) is trivial: because when no agent is initially satisfied, they switch to the other actions simultaneously, leading to a new state where no agent is satisfied either. In the next round, all agents switch back to the initial state. Therefore, the system keeps oscillating between the two states, all agents are in eternal flux and no agent is ever satisfied. Evidently, coordination failure is caused by simultaneous moves. Lemma 1(c) and Lemma 1(d) are proved in Cao and Yang (2014). Lemma 1(e) is simple: since each peripheral agent has the central agent as her unique neighbor, the utility of the central agent in any PNE must be the number of agents that are of the same type minus that of the opposite type; this happens if and only if at least one half of the peripheral agents are of the same type as the central agent.

Certain simple local structures turn out to play crucial roles in the convergence of BRDs. To ease later discussions, we introduce them formally in the definitions below.

Definition 3

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP, and the underlying network \((N,\mathrm{g})\) be a line or a ring. Any component of two adjacent agents that are of the same type is called a weak cluster, and G is called weakly clustered if it has at least one weak cluster. Obviously, G is weakly clustered if and only if it is not completely heterophilic.

Definition 4

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP, and the underlying network \((N,\mathrm{g})\) be a line. A strong cluster of G is defined as any of the following components: (i) the left-most two agents that are of the same type; (ii) the right-most two agents that are of the same type; and (iii) three adjacent agents that are of the same type. G is called strongly clustered if it has at least one strong cluster.

Definition 5

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP, and the underlying network \((N,\mathrm{g})\) be a ring. A strong cluster of G is defined as a component of three adjacent agents that are of the same type. G is called strongly clustered if it has at least one strong cluster.

Since NMP with any network structure has \(\hbox {LLC}\le 2\) (Theorem 1) and is an exact potential game when there is only one type of agents (Lemma 1(a)), it is intuitive that the existence of clusters may be in favor of the convergence of BRDs. It is indeed the case.

4.1 Lines

Below are the main results of this subsection, which give a fairly complete characterization for the simultaneous BRD of NMP with line structures.

Theorem 3

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP with the underlying network \((N,\mathrm{g})\) being a line,  \({\mathbf {x}}(0)\) an initial action profile,  and \({\mathcal {I}}=(N,\mathrm{g},{\mathbf {T}},{\mathbf {x}}(0))\) the corresponding initialized NMP. Suppose also \(n\ge 2\).

  1. (a)

    If G is completely heterophilic,  then \(\ell ({\mathcal {I}})=4\) for all \({\mathbf {x}}(0);\)

  2. (b)

    If G is not completely heterophilic,  i.e.,  it is weakly clustered,  then \(\ell ({\mathcal {I}})\in \{1,2,4\}\) for all \({\mathbf {x}}(0)\). In particular, 

    1. (i)

      \(\ell ({\mathcal {I}})=2\) if and only if all agents are of the same type and they are all initially unsatisfied,  i.e.,  \(u_i({\mathbf {x}}(0))<0\) for all \(i\in N;\)

    2. (ii)

      If there exists a weak cluster such that the corresponding two agents initially like each other,  then \(\ell ({\mathcal {I}})=1;\)

    3. (iii)

      If G is strongly clustered and the condition in (i) is not true,  then \(\ell ({\mathcal {I}})=1;\)

    4. (iv)

      In all the other cases that G is weakly clustered,  i.e.,  G is not strongly clustered and the condition in (ii) is not true,  \(\ell ({\mathcal {I}})\in \{1,4\}\).

Proofs of the above results are relatively easy, except cases (a) and (b)(iv), whose basic ideas are sketched in the following. Detailed rigorous proofs of all results of Theorem 3 are presented in Appendices B.1 and B.2.

The proofs of cases (a) and (b)(iv) rely crucially on two technical lemmas, Lemma 6 and Lemma 7 (Appendix B.2). Suppose now the underlying network is a completely heterophilic line. In the two lemmas, we focus on utility sequences rather than on action profiles. Roughly speaking, Lemma 6 states that certain utility sequences will never occur. Lemma 7 shows further that no agent has utility zero in the limit cycle.

The basic idea to prove Theorem 3(a) goes as follows. Recall first that a PNE does not exist in this case (Lemma 1(d)). By Lemma 7, none of the agents has a utility of zero in the limit cycle. Due to Lemma 6, 2 will never be adjacent with 2 and neither will \(-2\) be adjacent with \(-2\) in any utility sequence. Thus, it can be concluded that, the whole sequence of utilities will consist of alternate 2 and \(-2\) (except for the ending agents, whose utilities are 1 or \(-1\)) once a limit cycle is entered. And hence at any step of the limit cycle, either none conformist is satisfied but every rebel is, or none rebel is satisfied but every conformist is. Based on this argument, it can be checked easily that \(\hbox {LLC}=4\).

To prove Theorem 3(b)(iv), a “cutting technique” is explored. Observe first that, due to (b)(i) and (b)(iii), NMPs discussed in this case cannot be strongly clustered. Since \(\hbox {LLC}=2\) implies that all agents are of the same type and further that G is strongly clustered, we must have \(\ell ({\mathcal {I}})\ne 2\). In addition, there are no three adjacent agents that are of the same type, and neither the left-most two agents nor the right-most two ones are of the same type. So, if we cut the line at all the conformist-conformist links and at all the rebel-rebel ones into sublines, then each subline is completely heterophilic, and no isolated agent will be produced. To prove Theorem 3(b)(iv), it suffices to show that “\(\ell ({\mathcal {I}})\ge 3\) implies \(\ell ({\mathcal {I}})=4\)”. Suppose now \(\ell ({\mathcal {I}})\ge 3\). Since each of the sublines has \(\hbox {LLC}=4\) (Theorem 3(a)), and piecing them together does not affect the dynamic of each other, provided that \(\ell ({\mathcal {I}})\ge 3\) (Lemma 5 in Appendix B.1), we complete the proof of Theorem 3(b)(iv).

Theorem 3(b)(i), though simple, possesses several features that are in general not owned by other network structures (e.g., the star, see Theorem 5(a)): to have \(\hbox {LLC}=2\), it must be true that (i) the limit cycle is entered immediately at the very beginning of the dynamic. There is no possibility to evolve for certain periods of time and then enter a length two limit cycle; (ii) no agent is ever satisfied; and (iii) all agents are of the same type.

Remark 4

Let G be an NMP with the underlying network being a line. Combining Theorem 3 with Lemma 1(c) yields an interesting property: if G does not possess any PNE, then \(\hbox {LLC}=4\) for all initial action profiles. Note that this property is not covered by Theorem 2.

4.2 Rings

The results in Theorem 3(b) for lines are still valid for rings. We summarize them in the following theorem.

Theorem 4

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP with the underlying network \((N,\mathrm{g})\) being a ring,  \({\mathbf {x}}(0)\) an initial action profile,  and \({\mathcal {I}}=(N,\mathrm{g},{\mathbf {T}},{\mathbf {x}}(0))\) the corresponding initialized NMP. Suppose also \(n\ge 3\). If G is weakly clustered,  then \(\ell ({\mathcal {I}})\in \{1,2,4\}\) for all \({\mathbf {x}}(0)\). In particular, 

  1. (i)

    \(\ell ({\mathcal {I}})=2\) if and only if all agents are of the same type and they are all initially unsatisfied,  i.e.,  \(u_i({\mathbf {x}}(0))<0\) for all \(i\in N;\)

  2. (ii)

    If G is weakly clustered and the two agents in some cluster initially like each other,  then \(\ell ({\mathcal {I}})=1;\)

  3. (iii)

    If G is strongly clustered and the condition in (i) is not true,  then \(\ell ({\mathcal {I}})=1;\)

  4. (iv)

    In the other cases that G is weakly clustered,  i.e.,  G is not strongly clustered and the condition in (ii) is not true,  we have \(\ell ({\mathcal {I}})\in \{1,4\}\).

Note that when the number of agents n is odd, G with a ring structure cannot be completely heterophilic. Therefore, Theorem 4 provides a fairly complete characterization for the simultaneous BRD of NMP on ring structures when n is odd. The characterization is not as complete when n is even.

Proofs of Theorem 4(i), (ii) and (iii) are the same as of their counterparts in Theorem 3. However, the result that is parallel to Theorem 3(a) does not hold for rings: if the ring is completely heterophilic, we do not have in general that \(\hbox {LLC}=4\) (see Fig. 5 for a counter example). The key reason is that rings do not have ending nodes, and thus we are unable to get a result that is parallel to Lemma 7 (Appendix B.1), which plays a crucial role in the proof of Theorem 3(a).

Fortunately, a similar cutting technique can still be applied. This technique helps us to cut a ring into completely heterophilic lines without affecting its LLC, when \(\ell ({\mathcal {I}})\ge 3\) (note that while cutting a line for k times leads to \(k+1\) smaller separate lines, cutting a ring for k times leads to k separate lines). Using a similar argument as in the proof of Theorem 3(b)(iv), we can show that Theorem 4(iv) is valid. The detailed proof is presented in Appendix B.1.

Remark 5

When n is even, numerical simulations suggest that simultaneous BRD on completely heterophilic rings behaves very differently from that on completely heterophilic lines. To be specific, (i) if n / 2 is also even, then LLC could be 1, 4 and n, (ii) if n / 2 is odd, then LLC could be 4 and 2n. Figure 5 illustrates two limit cycles with lengths n and 2n, respectively.

Fig. 5
figure 5

Limit cycles for completely heterophilic rings. Conformists and rebels are represented by circles and triangles, respectively. Actions are indicated by colors: black for 1 and white for 0. In the left example, \(n=8\) and \(\hbox {LLC}=n\). In the right example, \(n=6\) and \(\hbox {LLC}=2n\). Unhappy faces indicate that the corresponding agents are unsatisfied

4.3 Stars

When the underlying network is a star, simultaneous BRD can be understood even more clearly than with lines and rings.

Theorem 5

Let \(G=(N,\mathrm{g},{\mathbf {T}})\) be an NMP with the underlying network \((N,\mathrm{g})\) being a star,  \({\mathbf {x}}(0)\) an initial action profile,  and \({\mathcal {I}}=(N,\mathrm{g},{\mathbf {T}},{\mathbf {x}}(0))\) the corresponding initialized NMP. Suppose also \(n\ge 4\).

  1. (a)

    If more than one half of the peripheral agents are of the opposite type to the central agent,  then \(\ell ({\mathcal {I}})=4\) for all \({\mathbf {x}}(0);\)

  2. (b)

    If more than one half of the peripheral agents are of the same type as the central agent,  and initially the central agent is unsatisfied,  then \(\ell ({\mathcal {I}})=2;\)

  3. (c)

    In all the other cases,  \(\ell ({\mathcal {I}})=1\).

The proof of Theorem 5(a) has some similarity to Theorem3(a). It can be shown that either only the central agent is satisfied or only the central agent is unsatisfied in the limit cycle. Using this fact, it can be checked easily that \(\hbox {LLC}=4\). Limit cycles in this case are illustrated in Fig. 6.

Theorem 5(b) is parallel to Theorem 3(b)(i) and Theorem 4(i), all characterizing the \(\hbox {LLC}=2\) case. The condition used in Theorem 5(b) is weaker than the one used for lines and rings, which requires all agents be initially unsatisfied (and hence be of the same type), as for the general case in Lemma 1(c). The reason that this weaker condition is still sufficient for deriving \(\hbox {LLC}=2\) is that, for star networks, after one round of deviations of the initially unsatisfied agents, the peripheral ones, no matter whether initially satisfied or not, are all unsatisfied. Since the peripheral agents that are of the same type as the central agent outnumber those of the opposite type, it follows that the central agent is still unsatisfied at the second step, leading to the general condition in Lemma 1(c) that no agent is satisfied. This shows that Theorem 5(b) is correct.

Fig. 6
figure 6

Illustrations of the \(\text {LLC}=4\) cases for star networks. Conformists and rebels are represented by circles and triangles, respectively. Actions are indicated by colors: black for 1 and white for 0

To prove Theorem 5(c), suppose now the central agent is a conformist. Due to Theorem 5(a)(b), there are two cases to discuss, (i) more than one half of the peripheral agents are conformists, and initially the central conformist is satisfied, and (ii) precisely one half of the peripheral agents are conformists and one half of them are rebels. Case (i) is easy: at the first step only possibly some of the peripheral agents deviate, but not the central conformist. Thus, at the second step, all of the peripheral agents are satisfied, and consequently so is the central conformist, leading to the convergence of the simultaneous BRD. To show case (ii), suppose on the contrary that \(\ell ({\mathcal {I}})\ge 2\). Then, there exists some time t in the limit cycle such that the central conformist is unsatisfied; suppose w.l.o.g. that she takes action 1. It can be observed that, peripheral agents that are of the same type must take the same action in the limit cycle (Lemma 8(a) in Appendix B.3). It follows that all of the peripheral agents take action 0 at time t. At time \(t+1\), all of the peripheral conformists switch to action 1 and the central conformist switches to action 0, making the central conformist satisfied but all the peripheral ones unsatisfied. Thus, at time \(t+2\), all the peripheral agents switch, making all agents happy and \(\ell ({\mathcal {I}})=1\), contradicting the hypothesis that \(\ell ({\mathcal {I}})\ge 2\). The other scenario that the central agent is a rebel can be similarly proved.

The detailed proof of Theorem 5 is presented in Appendix B.3. Due to Theorem 5, we can see that the central agent plays a much more important role than the peripheral ones. This is very reasonable.

Remark 6

Let G be an NMP with the underlying network being a star. Combining Theorem 5 with Lemma 1(e) yields a property similar to the one in Remark 4 for lines: if G does not possess a PNE, then \(\hbox {LLC}=4\) for all initial action profiles.

4.4 A summary

We have seen the crucial roles of weak and strong clusters. Other factors that might affect LLC include network size, population composition, network structure, and initial action profile.

As discussed in Remark 5, network size may sometimes play a very critical role. LLC could be arbitrarily long as the network size n goes to infinity. However, this happens only to ring networks in our study, not to the line or star networks, where LLC is always bounded above by 4. Even for the ring networks, an LLC that is larger than 4 may only possibly occur when the ring is completely heterophilic. Nevertheless, being completely heterophilic with ring structures is very rare if the configuration of agents’ types is randomly determined. To be more specific, let each agent be a conformist with probability \(c\in [0,1]\) and be a rebel with probability \(1-c\). We use \(\mathbb {T}\) to denote this random type configuration. Note that there are \(2^n\) configurations in total. Thus, a random type configuration is a distribution of these \(2^n\) configurations, where the probability of a configuration with \(n_C\) conformists and \(n-n_C\) rebels is \(c^{n_C}(1-c)^{n-n_C}\). Then, the probability that a ring is completely heterophilic (i.e., \(T_i \ne T_{i+1}\) for \(i=1,\ldots ,n-1\) in the type configuration \(\mathbb {T}\)) is \(2c^{n/2}(1-c)^{n/2}\), which tends to 0 as n goes to infinity (assume for simplicity that n is even). Thus, if the network has a line, ring, or star structure, then in almost all cases its size does not affect LLC.

To discuss the effects of other factors, we use a similar probability approach. Note that when \(c=1/2\), the random type configuration \(\mathbb {T}\) obeys the uniform distribution. A general c makes us able to investigate the factor of population composition. For any real number \(x\in [0,1]\), denote

$$\begin{aligned} \overline{x}=\max \{x,1-x\} \quad \text {and}\quad \underline{x}=\min \{x,1-x\}. \end{aligned}$$

Observe that \(\underline{x}\le 0.5 \le \overline{x}\) and \(\overline{x}+\underline{x}=1\) for all \(x\in [0,1]\).

We also use \(\mathbb {X}\) as a random initial action profile such that each agent is independently assigned an initial action 0 with probability \(b\in [0,1]\) and action 1 with probability \(1-b\). There are \(2^n\) initial action profiles in total. Thus, similarly to \(\mathbb {T}\), \(\mathbb {X}\) is a distribution of these \(2^n\) action profiles, where the probability of an action profile with \(n_0\) agents using action 0 and \(n-n_0\) agents using action 1 is \(b^{n_0}(1-b)^{n-n_0}\). In the following, we also use subscripts c and b for the probability function Prob[\(\cdot \)] to indicate the distribution parameters in \(\mathbb {T}\) and \(\mathbb {X}\).

Theorem 6

The following results on the probabilities of LLC hold for all \(c,b\in [0,1]\).

  1. (a)

    If the underlying network \((N,\mathrm{g})\) is a line or a ring,  then

    $$\begin{aligned} \lim _{n\rightarrow \infty } \text {Prob}_{c} \left[ \ell (N,g,\mathbb {T},{\mathbf {x}}(0))=1, \forall {\mathbf {x}}(0)\in \{0,1\}^n\right] =1. \end{aligned}$$
  2. (b)

    If the underlying network \((N,\mathrm{g})\) is a star,  then

    $$\begin{aligned}&\text{ Prob }_{c} \left[ \ell (N,g,\mathbb {T},{\mathbf {x}}(0))=m, \forall {\mathbf {x}}(0)\in \{0,1\}^n\right] =0,\quad \forall m\ne 4,\\&\lim _{n\rightarrow \infty } \text{ Prob }_{c} \left[ \ell (N,g,\mathbb {T},{\mathbf {x}}(0))=4, \forall {\mathbf {x}}(0)\in \{0,1\}^n\right] =\underline{c},\\&\lim _{n\rightarrow \infty } \text{ Prob }_{c,b} \left[ \ell (N,g,\mathbb {T},\mathbb {X})=2\right] =\overline{c}(c\underline{b}+(1-c)\overline{b}),\\&\lim _{n\rightarrow \infty } \text{ Prob }_{c,b} \left[ \ell (N,g,\mathbb {T},\mathbb {X})=1\right] =\overline{c}(c\overline{b}+(1-c)\underline{b}). \end{aligned}$$

Using Theorems 3 and 4, Theorem 6(a) can be easily shown, because the probability that there is at least one strong cluster tends to 1, as n goes to infinity. Using Theorem 5, the correctness of Theorem 6(b) is also easy. The detailed proof is provided in Appendix B.4.

Observe that when type configuration and initial action profile are uniformly selected, i.e., \(c=b=0.5\), the probabilities in the latter three formulas of Theorem 6(b) are respectively 0.5, 0.25 and 0.25. According to Theorem 6, we know that network structure alone can play a very important role. If we are told that the underlying network is a line or a ring, then almost always \(\hbox {LLC}=1\), namely, the simultaneous BRD converges. If it is known that the network is a star and type configuration and initial action profile are uniformly selected, then \(\hbox {LLC}=4\) with probability 0.5.

When the network is a star, population composition plays some role in determining LLC: All the latter three probabilities are significantly affected by it. A not so intuitive fact is that the roles of conformists and rebels are sometimes symmetric, which is not the case in Zhang et al. (2018), where, under the dynamic of sequential BRD, they found that conformists tend to widen the gap between the number of agents choosing action 0 and the number of agents choosing action 1, while rebels tend to narrow this gap. For example, the second formula in Theorem 6(b) indicates that the roles of the proportion of conformists and the proportion of rebels are symmetric in deciding the probability of \(\hbox {LLC}=4\): this probability reaches its highest value when the two proportions are the most balanced (\(c=0.5\)) and reaches its lowest value when the two proportions are the most unbalanced (\(c\in \{0, 1\}\)).

For star networks, the population composition parameter c has to work together with the initial action parameter b in order to decide the probabilities of \(\hbox {LLC}=2\) and \(\hbox {LLC}=1\). For example, when \(c=1\), i.e., all of the agents are conformists, the probability for \(\hbox {LLC}=1\) is approximately \(\overline{b}\), which achieves its highest value when all agents initially take the same action (\(b\in \{0,1\}\)), and it achieves its lowest value when each agent uniformly chooses between action 0 and action 1 (\(b=0.5)\).

5 Conclusion

In this paper, we have studied the matching pennies game on networks. Two benchmark cases and three simple network structures have been investigated. Our results demonstrate that network structures play a very critical role for both the simultaneous and sequential BRD. Indeed, even some very simple local structures, i.e., weak and strong clusters, may play crucial roles. These results may help us understand various phenomena in the real world including fashion.

The NMP is a relatively new model that is concise, interesting and useful. Much work is needed for more completely understanding the model and for exploring applications. First, how often is it to have \(\hbox {LLC}=1\) in NMPs with a single type of agents? Second, it is meaningful to understand completely heterophilic NMPs with at least two linked even-degreed agents. Finally, it is interesting to investigate properties of other familiar dynamics, such as fictitious play and logit response, for NMP.