This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

Spatial effect on stochastic dynamics of bistable evolutionary games

, and

Published 15 October 2014 © 2014 IOP Publishing Ltd and SISSA Medialab srl
, , Citation Kohaku H Z So et al J. Stat. Mech. (2014) P10020 DOI 10.1088/1742-5468/2014/10/P10020

1742-5468/2014/10/P10020

Abstract

We consider the lifetimes of metastable states in bistable evolutionary games (coordination games), and examine how they are affected by spatial structure. A semiclassical approximation based on a path integral method is applied to stochastic evolutionary game dynamics with and without spatial structure, and the lifetimes of the metastable states are evaluated. It is shown that the population dependence of the lifetimes is qualitatively different in these two models. Our result indicates that spatial structure can accelerate the transitions between metastable states.

Export citation and abstract BibTeX RIS

1. Introduction

Evolutionary game theory provides a mathematical framework for analysing conflicts of interests between individuals. The theory is based on the principle of evolution, i.e. strategies that perform well in the population increase in their abundance but those that perform poorly are wiped out. It has been applied in many disciplines such as social sciences [1, 2] and evolutionary biology [35]. Recently, methods developed in statistical physics have been utilised in research in this interdisciplinary field [610]

Evolutionary game theory has often been formulated by a deterministic process by assuming an infinitely large population [5, 11]. For example, replicator dynamics describe the deterministic change of frequencies of strategies [12]. The resulting dynamics are, in most cases, frequency-dependent, which means that a winning strategy depends on the current distribution of strategies in the population.

More recently, stochastic models have been applied to analyse evolutionary game dynamics in a finite population [4, 13]. Examples include evolutionary game-theoretic versions of the Wright–Fisher and Moran processes [14, 15], which are historically known as models in population genetics [16]. Analyses of stochastic models reveal the effect of various random factors, such as demographic stochasticity or recursive mutations, on game dynamics.

One of the most important applications of stochastic game dynamics is the study of transitions among multiple 'metastable states' that would correspond to fixed points (hence equilibria) in a deterministic description. A metastable state of a stochastic process is a state where the system remains for a sufficiently long time. By applying stochastic game dynamics and examining which metastable state has the longest lifetime, one can reveal which metastable state is most likely to be realised [4], [1721].

Another dimension of studies in evolutionary game theory is the incorporation of spatial structure into the model to explore the interplay between the game and space [6], [9], [2224]. There are numerous studies in this direction, which suggest that the speed of evolution as well as its outcome can be significantly influenced by the presence of spatial structure.

In this paper, we study how spatial structure affects the transition between metastable states in a class of evolutionary games called coordination games [17]. In coordination games, there are two strategies, A and B. The payoff of an A-individual is higher when matched with another A-individual than when matched with a B-individual. Similarly, the payoff of a B-individual is higher when matched with another B-individual than when matched with an A-individual. Cooperative hunting is one biological example of such a game, where two hunters hunt either prey A or B; hunting the same prey together greatly improves the success, so coordination is the better strategy. This game has bistability, because both the all-A state (= everyone adopts strategy A) and the all-B state (= everyone adopts strategy B) are stable fixed points of the deterministic evolutionary game dynamics. When a small 'mutation', i.e. a small chance that the individuals will randomly change their strategy, is introduced, the two states (A-dominant and B-dominant) are realised as metastable states and transitions between them may occur, i.e. they have finite lifetimes. When the system is large, such transitions are extremely rare.

To evaluate the lifetimes of metastable states, we adopt a semiclassical (WKB) approximation [2530], which is an analog of the WKB approximation used in quantum mechanics. This method is known to be suitable for evaluating the probability of rare events caused by large fluctuations [30], because it can appropriately take into account an exponentially small tail of probability distribution beyond the Gaussian fluctuation considered in the Fokker–Planck approximation or the van Kampen expansion [31]. The semiclassical approximation has been used for evaluating the probabilities of rare events in the context of extinction phenomena in ecology [3236]. For evolutionary games, although it has been applied to the analysis of fixation [10], transition between metastable states, especially how it is affected by spatial structure has not been studied so far.

Here we consider the spatial effect on the transition between metastable states in coordination games. To investigate the spatial effect, we analyse two types of evolutionary game models: (i) a model without any spatial structure (the 'well-mixed' model) and (ii) a model with a 1D spatial structure (the 'spatial' model). For each model, we calculate the lifetime of each metastable state (which will be defined in section 2.3) based on the semiclassical approximation and discuss the effect of spatial structure by comparing them. We show that the spatial structure qualitatively changes the population size dependence of the lifetimes and that this change is caused by the presence of nucleation processes, i.e. by transitions that occur via a 'critical nucleus' allowed only in the spatial model. In addition, we clarify that nucleation processes can occur only when the system's length exceeds a characteristic length. Although we demonstrate the spatial effect using a specific model, qualitatively the same result is expected to hold for general bistable evolutionary games.

This paper is organized as follows: In sections 2 and 3, we evaluate the lifetimes of the metastable states for the well-mixed and the spatial models, respectively. In section 4, by comparing the lifetimes in the two models, we discuss the spatial effect as well as its intuitive description. Section 5 concludes the paper. Readers who are less interested in the details of the calculations may directly proceed to sections 2.1, 3.1 and 4.

2. The well-mixed model

In this section, we consider the transitions between metastable states in a model without spatial structure ('well-mixed model'). The analysis and discussion presented in this section provide an important basis for studying spatial effect in the following sections.

We define the model in section 2.1 and in section 2.2, we consider the dynamics of the expectation value of the number (or proportion) of the strategy-A individuals and confirm the bistability of the model. We evaluate the lifetimes of the metastable states by the path integral method and the semiclassical approximation in section 2.3 and discuss their model parameter dependence in section 2.4.

2.1. Model

Let us assume that we have a population composed of N (≫ 1) individuals, n of which follow strategy A and N − n of which follow strategy B. We assume that the population is well-mixed, i.e. all the individuals play games with all the other individuals.

We describe the games performed between the individuals using a payoff matrix

Equation (1)

which specifies payoffs of the games in the following manner: The game between two A individuals gives both of them a payoff a. The game between an A individual and a B individual gives them payoffs b and c respectively. The game between two B individuals gives both of them a payoff d. Then, when the number of A individuals is n, the respective average payoffs of A and B individuals are ΠA(n/N) and ΠB(n/N), where

Equation (2)

Here, q := n/N is the proportion of the A individuals in the population (note that we allowed self-interaction for simplicity). In this paper, we consider only a bistable 'coordination game', i.e. we impose the following conditions:

Equation (3)

The condition means that, to obtain a higher payoff, it is always best to play the same strategy as the opponent.

As an update rule, we adopt a 'pairwise comparison process' [37, 38]. In this process, two individuals, called a focal individual and a role individual, are selected randomly at a rate λ. The focal individual adopts the strategy of the role individual with probability p(Δ Π) := (1 + w Δ Π)/2, where Δ Π := Πr − Πf and Πf and Πr are the payoffs of the focal and role individuals respectively. The parameter w controls the effect of the payoffs on the spread of strategies and is hence called 'the intensity of selection'. Thus the strategy yielding a higher payoff is more likely to be imitated by others. We assume w ∈ [0, 1] and 0 < ac < 1, 0 < db < 1 to guarantee p(Δ Π) ∈ [0, 1].

To avoid fixation, we introduce mutations (i.e. the possibility that imitation fails): If the role individual is A (resp. B), the focal individual adopts strategy B (resp. A) with a small probability μA (resp. μB). We assume that the mutation rates are so small that the dynamics of the expectation value of n remains bistable (see section 2.2).

The parameters used in the present paper are summarized in table 1.

Table 1. Parameters used in the well-mixed model.

Characters Meaning
λ the rate at which the strategy update occurs
w the intensity of selection (w ∈ [0, 1])
μA mutation probability from A to B
μB mutation probability from B to A
a, b, c, d elements of the payoff matrix

We define the transition rates for the processes, in which the number of A individuals increase or decrease by one, as λ W+(n/N) and λ W(n/N), respectively, where W± are the dimensionless transition rates. By adopting the update rule described above, W± are given as follows:

Equation (4)

Equation (5)

The system is then described by a continuous time Markov process with these transition rates. Note that these transition rates depend on payoff matrix elements only through a − c and d − b because

Equation (6)

The master equation for P(n, t), the probability that the number of A is n at time t, can be written as

Equation (7)

2.2. Deterministic dynamics of the expectation value

Let us first consider the dynamics of the expectation value of n, which we denote by 〈n(t)〉. In the limit N → , the stochastic fluctuations can be neglected and 〈n(t)〉 obeys the deterministic equation

Equation (8)

It is convenient to introduce the new variable q := n/N, which represents the proportion of A in the population and rewrite the equation as, with dimensionless time τ := λ t/N,

Equation (9)

For the model considered here, W+(q) − W(q) is a cubic function of q:

Equation (10)

Figure 1 shows a graph of W+(q) − W(q) and the flow of q determined by (9).

Figure 1.

Figure 1. The phase portrait of (9), the deterministic dynamics of the well-mixed model. Here, q is the proportion of A individuals in the system. The full circles denote fixed points, the dashed lines denote W+(q) − W(q) and arrows on the axis indicate the flow of q. It can be seen that q1 and q3 are stable and that q2 is unstable. Thus, the system exhibits bistability. Parameters: a − c = 0.4, d − b = 1.0, w = 0.5, μA = μB = 0.01.

Standard image High-resolution image

The fixed points of the deterministic equation (9) are determined by the condition W+(q) − W(q) = 0, i.e. the transition rates of both processes balance. For coordination games (a > c, b < d) without mutation (μA = μB = 0), it can be easily shown that the system has three fixed points: two stable fixed points q1 = 0 (= everyone adopts strategy B) and q3 = 1 (= everyone adopts strategy A) and one unstable fixed point q2 = (d − b)/(a − c+d − b) (= mixture of strategy A and B). The positions of the fixed points shift slightly when small mutations are introduced. However, the system remains bistable, i.e. it has two stable fixed points (q1 and q3) and one unstable fixed point (q2) in the range [0, 1] (see figure 1). Hereafter, we assume that the mutation rates μA and μB are sufficiently small so that the system remains bistable.

2.3. Transitions between metastable states

As discussed above, the expectation value of n exhibits bistability. In addition, as can be seen from the definition of the model, fixation to one strategy is impossible because of mutations. Thus, although the system remains at states q1 or q3 for an extremely long time, large fluctuations can occasionally carry the system from one state to the other. Therefore, states q1 and q3 are metastable and have a long but finite lifetime.

To demonstrate this feature, we performed a Monte Carlo simulation of the stochastic process described by the master equation (7) based on the Gillespie algorithm [39]. We show an example of stochastic time evolution of the number of A individuals for N = 50 in figure 2, which shows that while the system stays around the two stable fixed points (q1 and q3) for a long time, there are rare transitions between these two metastable states due to stochastic fluctuations. These transitions are expected to be very rare when N is large, because fluctuations around metastable states are suppressed.

Figure 2.

Figure 2. Time evolution of q:= n/N (the proportion of A individuals in the population) obtained by a Monte Carlo simulation of (7), the stochastic dynamics of the well-mixed model. Dashed lines denote q1, q2 and q3. The system clearly exhibits metastability: although the system stays around the metastable states (B-dominant state q1 or A-dominant state q3) for a long time, it occasionally undergoes transition between these states. Parameters: N = 50, a − c = d − b = 0.5, w = 0.4, μA = μB = 0.01. For these parameters, the positions of the metastable states are q1 ≃ 0.0259, q3 ≃ 0.974.

Standard image High-resolution image

In this paper, we evaluate the lifetimes of these metastable states, which are defined to be the mean waiting times until the system escapes from the given metastable state and undergoes transition to the other metastable state. In the following section, we mainly focus on the lifetime of the metastable state q3, because the lifetime of the metastable state q1 can be calculated in the same manner.

2.3.1. Path integral expression.

The lifetime of the metastable state q3 is calculated from Pq3, wm(T), the probability that the system stays around q3 from τ = 0 to τ = T (≫ 1) without ever visiting the other metastable state q1. Pq3, wm(T) decays exponentially with T and the inverse of the decay rate gives a lifetime of q3.

Pq3, wm(T) can be approximated by the probability that q(T) = q3 and q(τ)≠ q1 (∀τ ∈ [0, T]) given that q (0) = q3. We express the latter probability using the path integral formulation of stochastic processes [30, 40], in which the probability of given paths can be expressed as a summation, with some weight, over the paths. With this technique, we obtain (see appendix A for the derivation)

Equation (11)

Equation (12)

Equation (13)

where W± are given by (4) and (5), p is a new variable conjugate to q and $\int^{\prime}_{q(0) = q_3, q(T) = q_3} \mathcal{D}q \mathcal{D}p$ represents the (restricted) summation over all the paths (q, p) satisfying q (0) = q(T) = q3 and q(τ)≠ q1 (∀τ ∈ [0, T]).

The meaning of this expression is that the probability density of each path q (·), p (·) is exp(−NS[q (·), p (·)]) and that the desired probability is a weighted sum over all paths satisfying the conditions. Note that the present formalism is an analog of the path integral in quantum mechanics. From this analogy, S and H0 are called an action and a Hamiltonian, respectively. The path integral expression is useful for the analysis of N ≫ 1 cases as discussed in the next section.

2.3.2. Semiclassical approximation.

We will now evaluate the path integral expression (11) under the assumption N ≫ 1. We adopt a semiclassical approximation, in which the path integral is approximated by contribution from the stationary paths (q, p) of action S (i.e. δS[q, p] = 0) and the fluctuations around them because, when N is large, only the stationary paths contribute dominantly to the path integral. This approximation is an extension of the steepest descent method in evaluating integrals and an analog of the WKB (or semiclassical) approximation in quantum mechanics. It is known that the semiclassical approximation is suitable for treating rare events [30].

The stationary condition δS = 0 yields differential equations:

Equation (14)

Equation (15)

(where $W^{\prime}_{\pm}(q) := {\rm d}W_{\pm}(q)/{\rm d}q)$ subjected to the boundary conditions q (0) = q(T) = q3 and the condition q(τ)≠ q1 (∀τ ∈ [0, T]). Because the equations have the same form as Hamilton's equations of motion in analytical mechanics, the solutions (q, p) satisfying these differential equations are called 'classical' trajectories in the following discussion. Figure 3 shows phase portraits of the flow of these equations. Note that H0(q, p) is conserved along each trajectory.

Figure 3.

Figure 3. Phase portraits of the equations of motion (14) and (15). q is the fraction of A individuals in the system and p is a variable conjugate to q. The arrows indicate the flows of (14) and (15), three green circles denote fixed points and coloured curves show the contours of the Hamiltonian. Left: bold lines denote activation trajectories pa (see (16)) Right: bold lines denote closed trajectories that determine the lifetimes of the metastable states. The red trajectory (right) determines the lifetime of q3, while the blue trajectory (left) determines the lifetime of q1. The area shaded in red (right) is equal to Sq3, wm (see (17)). Parameters: a − c = 0.4, d − b = 1.0, w = 0.5, μA = μB = 0.01.

Standard image High-resolution image

There are two important types of trajectories. The first type are the trajectories on the horizontal line of p = 0. These trajectories represent the dynamics of the expectation value of q, because the equation of motion (14) for p = 0 coincides with the equation of 〈q〉 (9). Note that H0(q, p) = 0 for these trajectories.

The second type are the trajectories shown by bold lines in the left panel of figure 3, connecting stable fixed points ((q1, 0) and (q3, 0)) and an unstable fixed point ((q2, 0)). These two trajectories are called 'activation trajectories' [29, 34]. Because the Hamiltonian is always constant on the connected trajectories, H0(q, p) = 0 holds on the activation trajectories. Hence, the shape of the activation trajectories can be collectively expressed as p = pa(q), where

Equation (16)

which can be obtained from the condition H(q, pa(q)) = 0 (see (13)).

In the semiclassical approximation of Pq3, wm, classical trajectories should be properly chosen so that the conditions q (0) = q(T) = q3 and q(τ)≠ q1 (∀τ ∈ [0, T]) are satisfied. Therefore, the classical trajectories relevant for the evaluation of Pq3, wm are restricted to (i) a trivial solution (q(τ), p(τ)) ≡(q3, 0) and (ii) nontrivial solutions circulating on the closed trajectory through (q2, 0) and (q3, 0) (the red closed trajectory shown in the right panel of figure 3), i.e. the closed trajectory composed of an activation trajectory and a p = 0 trajectory. Note that only these trajectories spend a significant amount of time going around and can satisfy the boundary conditions q (0) = q(T) = q3 for T ≫ 1. These solutions are called 'bounce solutions' [30].

The action of the trivial solution is zero, whereas the action of nontrivial solutions are given by n Sq3, wm, where n (= 1, 2, 3, ...) is the number of the rotation and Sq3, wm is the action per one cyclic motion on the closed trajectory described above. Using the zero-energy condition H0(q, p) = 0 and the expression for the activation trajectories pa(q), Sq3, wm can be expressed as

Equation (17)

which is equal to the area shaded in red in the right panel of figure 3.

Summing up all the contributions from the bounce solutions, we obtain

Equation (18)

Here, the prefactor K is a positive value determined by Gaussian integrals around bounce solutions.

Thus, the lifetime of the metastable state q3, τq3, wm, can be expressed as

Equation (19)

The lifetime of the other metastable state q1 can be evaluated in the same manner, except that the other closed trajectory (the blue trajectory in the right panel of figure 3) should be used:

Equation (20)

Equation (21)

where the prefactor K' is a positive quantity (determined by Gaussian integrals around bounce solutions around q1).

It can be seen from (19) and (20) that the lifetimes are extremely long (τqi, wm≫ 1) under the condition N ≫ 1 assumed in this paper. The lifetimes strongly depend on the action Sqi, wm because of the presence of a large factor N in the exponent. In this paper, we focus on the exponent of the lifetimes and ignore the weak parameter dependence of the prefactors K and K'.

2.4. Result

In this section, we briefly summarize the parameter dependence of the lifetime τqi, wm(i ∈ {1, 3}).

2.4.1. N dependence.

Because action Sqi, wm is positive and independent of N, the lifetimes τqi, wm∝exp(N Sqi, wm) (i ∈ {1, 3}) increase exponentially with N. This result is consistent with the previous one, which showed that the fixation probability in the coordination game decays exponentially with N [41]. This kind of exponential dependence of some 'lifetimes' on population size is known for various models in which the population is well-mixed. Examples include ecological models (the mean time to the extinction of a population) [34, 35, 42], evolutionary games (the mean time to fixation in anti-coordination games) [10] and general reaction models [29].

2.4.2. w dependence.

Figure 4 shows Sq1, wm and Sq3, wm as the functions of w (the intensity of selection). As can be seen from the figure, the action grows with w. This result indicates that, when N ≫ 1, the lifetimes τqi, wm∝exp(N Sqi, wm) (i ∈ {1, 3}) increase rapidly as natural selection becomes strong.

Figure 4.

Figure 4. w dependence of Sq1, wm and Sq3, wm. Sqi, wm increases with w. Parameters: a − c = 0.4, d − b = 1.0, μA = μB = 0.005.

Standard image High-resolution image

2.4.3. Payoff matrix dependence.

As can be seen from (4)–(6), the stochastic dynamics of the evolutionary game considered here depends on payoff matrix elements only through a − c and d − b. The dependence of Sq1, wm and Sq3, wm on these two parameters is shown in figure 5 (only the region 0.2 < ac < 0.8 and 0.2 < db < 0.8 is plotted because outside this region there are parameter sets for which the system is not bistable).

Figure 5.

Figure 5. The payoff matrix dependence of Sq1, wm(left) and Sq3, wm(right). Parameters: w = 0.8, μA = μB = 0.005.

Standard image High-resolution image

This dependence can be intuitively explained in the following manner: The larger a − c is, the more advantageous A individuals are in a population dominated by A, compared with B individuals in the same population, and the longer the lifetime of q3. Almost the same discussion applies to d − b and the lifetime of q1. Note that when μA = μB, a symmetry relation Sq1, wm(a − c, d − b) = Sq3, wm(d − b, a − c) holds.

3. Spatial model

On the basis of the analysis presented in the previous section, in this section we examine the transition between metastable states in a model with spatial structure ('spatial model'). The discussion proceeds almost parallel to that of the previous section. The model is defined in section 3.1. In section 3.2, we analyse the behaviour of the expectation value and determine steady states and their stability. In section 3.3, we evaluate the lifetimes of metastable states using the path integral expression and the semiclassical approximation. In section 3.4, we show how these lifetimes depend on the system size.

3.1. Model

We consider a 1D array of M patches. Let l be the separation between neighbouring patches and $\tilde{L} := Ml$ be the length of the system. We impose a periodic boundary condition. In each patch, there are Np (≫ 1) individuals, who take either strategy A or strategy B and change their strategies according to the same evolutionary game as described in section 2. We further assume that the individuals can move between neighbouring patches. This migration process is modelled as the 'swapping' of individuals so that the number of individuals per patch is conserved (see figure 6).

Figure 6.

Figure 6. A schematic diagram of the spatial model. The small circles denote individuals (red: A individuals, blue: B individuals). Individuals play games and reproduce inside each patch and can move between neighbouring patches.

Standard image High-resolution image

Let ni be the number of A individuals in the ith patch (i ∈ {1, 2, ···, M}). The parameters of the evolutionary rule in each patch are the same as those described in section 2 (see table 1). The transition rates of ni due to the strategy update can be written as λ W+(ni/Np) and λ W(ni/Np), where W± are given in (4) and (5), respectively.

The migration (swapping) processes are defined as follows: A pair of neighbouring patches, i and j, is randomly chosen at a rate σ. One individual is randomly picked up from each patch and they are swapped. The rate of the process (ni, nj) → (ni −1, nj +1) due to swapping is given by σ Wm(nj/Np, nj/Np), where

Equation (22)

With these transition rates, the master equation for the probability distribution P (n, t) on the population configuration n = (n1, n2, ···, nM) is written as

Equation (23)

where 〈i, j〉 indicates neighbouring patches and

Equation (24)

3.2. Dynamics of the expectation value

In this section, we derive a differential equation for the expectation values of ni and examine its steady solutions and their stability, which play an important role in understanding the transitions between metastable states in the spatial model.

3.2.1. Deterministic equations.

Let 〈ni〉 be the expectation value of ni. Then, 〈ni〉(i ∈ {1, 2, ···, M}) obey

Equation (25)

By introducing new variables qi := ni/Np (the proportion of A individuals in the ith patch) and a rescaled time τ := λ t/Np, we obtain

Equation (26)

In this paper, we assume σλ, i.e. migration processes occur sufficiently faster than strategy update processes. Under this assumption, 〈qi〉 changes smoothly as a function of i and can be expressed by a function of a continuum spatial degree of freedom x:

Equation (27)

The second term on the right hand side of (26) is approximated by D2 q/∂x2, where D := σ l2/λ is a diffusion constant. By introducing the dimensionless space variable $\xi := x/ \sqrt{D}$ , we arrive at the following reaction diffusion equation:

Equation (28)

where $L := \tilde{L}/\sqrt{D}$ is the rescaled system size.

3.2.2. Steady solutions.

During the analysis of the well-mixed model discussed in section 2, fixed points and their stability played important roles. In the spatial model, steady solutions will play similar roles.

Let qs(·) be a steady solution of (28). qs obeys an ordinary differential equation

Equation (29)

Integrating this equation yields

Equation (30)

Equation (31)

By regarding ξ as the 'time', we can describe qs as a coordinate of a particle moving in one dimension, subjected to potential V. Hence, the trajectories of qs on the (q, dq/dξ) plane can be expressed by the contours of 'energy' E defined in (30) (dotted lines in the left panel of figure 7). Because of the periodic boundary condition qs(ξ+L) = qs(ξ), qs must be a closed orbit or a point on the (q, dq/dξ) plane. Therefore, the steady solutions must be

  • (a)  
    uniform solutions: qs(ξ) ≡ qi(i ∈ {1, 2, 3}) (recall that qi are the solutions of W+(q) − W(q) = 0 defined in section 2.2) or
  • (b)  
    non-uniform periodic solutions: closed orbits surrounding q2.
Figure 7.

Figure 7. Steady solutions of (28). Left: a 'phase portrait' of (29). The solid lines represent critical nuclei. The coloured bar denotes the value of E defined in (30). Right: a critical nucleus (L = 40). Lc is the typical scale of a critical nucleus. Parameters: a − c = 0.4, d − b = 1.0, w = 0.8, μA = μB = 0.005 (Lc ≃ 13.5 and V(q1) > V(q3) in this case). The critical nuclei were numerically calculated by applying Newton's method to a discretized version of (29).

Standard image High-resolution image

First, we consider the uniform solutions, which always exist regardless of system size L. Because $W^{\prime}_{+}(q) - W^{\prime}_{-}(q)$ is negative for q1, q3 and positive for q2 (see figure 1), two solutions, qs(ξ) ≡ q1 and qs(ξ) ≡ q3, are linearly stable and qs(ξ) ≡ q2 is linearly unstable; the former solution corresponds to the two metastable states, whereas the latter solution corresponds to the 'marginal' state located at the boundary q2, below which the system goes to q1 and above which the system goes to q3.

Next, we examine the properties of the non-uniform solutions. The forms of these solutions depend on the system size L, because the period of the solution must coincide with L:

Equation (32)

Here, qmin and qmax are the minimum and the maximum of q in the trajectory, which is determined by E. The left panel of figure 7 shows the phase portrait for the case of V(q1) > V(q3), where the non-uniform solution in the limit of L →  corresponds to the homoclinic orbit starting from q3 and ending at q3 (the blue line in the figure). The right panel of figure 7 shows the spatial profile of a non-uniform solution for the same parameter set as the left panel. As can be seen from the figure, the solution represents a 'nucleus' of B individuals surrounded by a region dominated by A individuals. Note that for the opposite case (i.e. V(q1) < V(q3)), the form of the non-uniform solution becomes 'upside down' compared to the right panel of figure 7, i.e. it represents a nucleus of A individuals surrounded by a region dominated by B individuals.

It can be shown that the obtained non-uniform solutions are unstable, i.e. even an infinitesimally small perturbation drives the system away from the non-uniform solutions and toward the stable states q1 or q3 (see appendix B). Because of such 'critical' behaviour, these non-uniform steady solutions are called 'critical nuclei' [36] and we denote them by qcn.

As L decreases, both qmin and qmax approach q2 and at a critical length Lc, they coalesce with q2. This indicates that critical nuclei do not exist when L is smaller than Lc, which is calculated, by linearizing (29) around q2, as

Equation (33)

Lc gives the characteristic length scale of the critical nuclei and plays an important role in considering the spatial effect on the transitions between metastable states, as discussed later in section 4.

3.3. Transitions between metastable states

As shown in the previous section, the spatial model considered deterministically has two stable steady solutions, i.e. the uniform solutions qs(ξ) ≡ q1 and qs(ξ) ≡ q3, which we henceforth simply call q1 and q3, respectively. If stochasticity is taken into account, these solutions correspond to metastable states: although the system stays at q1 or q3 for an extremely long time, it occasionally undergoes transitions from one state to the other as a result of large stochastic fluctuations. In this section, we evaluate the lifetimes of these metastable states. We show the calculation of the lifetime of q3 (the lifetime of q1 can be calculated in the same manner).

3.3.1. Path integral expression.

To evaluate the lifetime of a metastable state q3, we calculate Pq3, sp(T), the probability that the system stays around q3 from τ = 0 to τ = T (≫ 1) without ever visiting the other metastable state q1. Pq3, sp(T) decays exponentially with T and the inverse of the decay rate gives the lifetime of q3.

The path integral formalism, which was applied to the well-mixed model in section 2, is also applicable to the spatial model (see appendix A for the detail). Pq3, sp can be approximated by the probability that q (·, T) = q3 and q (·, τ)≠ q1 (∀τ ∈ [0, T]) given that q (·, 0) = q3. By adopting a continuum description assuming σλ and using a dimensionless spatial coordinate $\xi := jl/\sqrt{D} \ \ ( j \in \left\{1,2, \cdots, M \right\} )$ and the rescaled system size $L := \tilde{L}/\sqrt{D}$ , we obtain

Equation (34)

Equation (35)

Equation (36)

Equation (37)

where the prime in the path integral indicates the restriction to paths that satisfy q (·, τ)≠ q1 (∀τ ∈ [0, T]).

3.3.2. Semiclassical approximation.

The semiclasslcal approximation can be applied in almost the same manner as was performed in section 2.3.2 for the well-mixed model. We consider stationary solutions of the action represented by (35). The stationary condition δS = 0 leads to partial differential equations

Equation (38)

Equation (39)

subjected to a boundary condition q (·, 0) = q (·, T) = q3. Note that when p = 0, the equation for q coincides with (28), i.e. the equation for the expectation value.

We now calculate Pq3, sp(T) (T ≫ 1) by considering solutions to the equations of motion (38) and (39), under the boundary condition q(ξ, 0) = q (·, T) = q3 and the restriction that q never visits q1. Unlike the well-mixed model, it is not easy to obtain such solutions in the spatial model. However, we can infer qualitative features of these solutions from the analysis of the well-mixed model, where the solution is composed of two trajectories: a trajectory moving from q3 to q2 in the p < 0 region and a trajectory moving from q2 to q3 in the p = 0 region (see section 2.3.2, figure 3 right panel). Note that the unstable fixed point q2 plays the role of a 'watershed', i.e. a dividing point between the two stable states q1 and q3. In the spatial model, there are two kinds of unstable solution: (a) the uniform solution q(ξ) ≡ q2 and (b) the critical nucleus. Therefore, it is inferred that there are two kinds of bounce solution:

  • (a)  
    a solution which is initially q(ξ) ≡ q3, then changes into q(ξ) ≡ q2 and finally returns to q(ξ) ≡ q3 (see figure 8 left panel), and
  • (b)  
    a solution which is initially q(ξ) ≡ q3, then changes into q = qcn and finally returns to q(ξ) ≡ q3 (see figure 8 right panel).
Figure 8.

Figure 8. Schematic representation of bounce solutions. Left: a uniform bounce solution α. The system changes from q3 to q2 and returns back to q3 uniformly in space. Right: a non-uniform bounce solution β. The system changes from q3 to qcn and returns back to q3.

Standard image High-resolution image

We call the former a 'uniform bounce solution' (or solution α) and the latter a 'non-uniform bounce solution' (or solution β). It is clear that the non-uniform bounce solution exists only when L > Lc.

Let $S_{\alpha, q_3}$ and $S_{\beta, q_3}$ be the action of the bounce solutions α and β calculated from (35)–(37). If L < Lc, α is the only possible bounce solution. Hence, Pq3, sp(T) is calculated to be

Equation (40)

where $K_{\alpha,q_3}$ is a prefactor determined by the Gaussian integrals around the stationary solution α. Then, τq3, sp, the lifetime of the metastable state q3 can be expressed as $\tau_{q_3, {\rm sp}} \simeq K_{\alpha, q_3}^{-1} \exp\left( N_p \sqrt{\sigma/\lambda} S_{\alpha, q_3} \right)$ . If L > Lc, there are bounce solutions α and β. By summing up all the contributions from these two bounce solutions, we obtain

Equation (41)

Equation (42)

where $K_{\alpha, q_3}$ and $K_{\beta, q_3}$ are prefactors determined by the Gaussian integrals around α and β respectively. Then,

Equation (43)

Under the condition $N_p\sqrt{\sigma/\lambda} \gg 1$ assumed in this paper, the lifetime is well-approximated by one of the two terms, which has the smaller value of the action. Let γ be either α or β, which corresponds to the smaller value of action (e.g. if $S_{\alpha, q_3} > S_{\beta, q_3}$ , then γ = β. In fact, for the case treated in the next section, we obtain γ = β). Then, we obtain $\tau_{q_3, {\rm sp}} \simeq K_{\gamma, q_3}^{-1} \exp \left( N_p \sqrt{\sigma/\lambda} S_{\gamma, q_3} \right)$ . Thus, the expression for the lifetime is summarized as

Equation (44)

3.4. Results

To discuss the L dependence of the lifetimes, we numerically calculated the action of the bounce solutions (see appendix C for the detail). If two kinds of bounce solution exist (i.e. if L > Lc), we choose the one with the smaller action. The action obtained is denoted by Sqi, sp(i ∈ {1, 3}).

Figure 9 shows the L dependence of the action calculated in this manner for a specific parameter set, for which V(q1) > V(q3) holds. If L < Lc, both Sq1, sp(L) and Sq3, sp(L) increase linearly with the system size L. A different feature appears when the system size exceeds the threshold length Lc, where the non-uniform bounce solutions become important; Sq3, sp is now independent of L, whereas Sq1, sp continues increasing with L. A detailed discussion of these results and their intuitive meaning is provided in the next section.

Figure 9.

Figure 9. L dependence of the action. Left: Sq1, sp, action of the bounce solution that determines the lifetime of q1. Right: Sq3, sp, action of the bounce solution that determines the lifetime of q3. The dashed line denotes Lc. Sq1, sp keeps increasing with L, whereas Sq3, sp takes constant value when L > Lc. Parameters: a − c = 0.4, d − b = 1.0, w = 0.8, μA = μB = 0.005. For this parameter set, Lc ≃ 13.5 and V(q1) > V(q3) hold.

Standard image High-resolution image

4. Comparison and discussion

In this section, we discuss the spatial effect on the transitions between metastable states by comparing the result of the well-mixed model (section 2) and that of the spatial model (section 3).

In section 4.1, we discuss qualitative difference in the population size dependence of the lifetimes between these two studied models. In section 4.2, we present an intuition which explains the difference based on a 'nucleation' process. In these two sections, we restrict ourselves to the case in which V(q1) > V(q3) unless otherwise specified (for the opposite case, the same discussion holds by exchanging q1 and q3). In section 4.3, the parameter dependence of the characteristic length scale Lc is discussed in detail. Section 4.4 is devoted to the interpretation of the spatial effect in terms of migration rate. We close this section with a remark on which metastable state is long–lived.

4.1. Spatial effect on strategy selection

First, we summarize the results for the well-mixed model (section 2). For the evolutionary game dynamics considered here (coordination games with mutations), there are two metastable states denoted by q1 and q3, where q ∈ [0, 1] is the proportion of A individuals in the system; the state q1 corresponds to the population dominated by strategy B and q3 corresponds to the one dominated by strategy A. Although the system spends an extremely long time at q1 or q3, it can undergo a transition from one state to the other, via an unstable state q2 (satisfying q1 < q2 < q3). The lifetime of the metastable states q1 and q3 are calculated to be

Equation (45)

where Sq1, wm and Sq3, wm are the actions of the bounce solution defined by (21) and (17), respectively. Because Sq1, wm and Sq3, wm are positive and independent of N, the lifetimes τqi, wm increase exponentially with the population size N.

Next, we summarize the results for the spatial model (section 3). The lifetimes of the metastable states q1 and q3 in the spatial model are evaluated to be

Equation (46)

where Sq1, sp and Sq3, sp are the positive functions of the system size $L = M\sqrt{\lambda/\sigma}$ and Np is the population number per patch. Note that L is proportional to the total population number NpM, where M is the number of patches. As discussed in section 3.2.2, the spatial model has a characteristic length scale Lc. For small systems (L < Lc), Sq3, sp grows linearly with L, whereas for large systems (L > Lc), it becomes independent of L (see the right panel of figure 9). Thus, from (46) one can conclude that the lifetime τq3, sp grows exponentially with L for L < Lc, while it is independent of L for L > Lc. In other words, while the model is effectively well-mixed when L < Lc, it behaves differently from the well-mixed model when L > Lc. We stress that this difference is quite drastic due to the large factor $N_p \sqrt{\sigma / \lambda} \ (\gg 1)$ appearing in (46). On the other hand, Sq1, sp has a linear dependence on L for both L < Lc and L > Lc (see the left panel of figure 9). This implies that there is no qualitative difference in the system size dependence of τq1, sp between the two regions L < Lc and L > Lc, which is in clear contrast to what is observed for τq3, sp.

Note that the difference between the lifetimes of q1 and q3 originates from that of V(q1) and V(q3), where V is defined in (31). If V(q1) < V(q3) (the opposite to the above situation), the roles of q1 and q3 are reversed so that the lifetime of q1 takes a constant value for L > Lc while the lifetime of q3 keeps growing (the intuitive reason for this behaviour is provided in the next subsection). The condition V(q1) > V(q3) (resp. V(q1) < V(q3)) can be rewritten as A1 > A3 (resp. A1 < A3), where

Equation (47)

Equation (48)

Note that A1 and A3 correspond to the two areas enclosed by the curve W+(q) − W(q) and the horizontal axis in figure 1. Without mutation (μA = μB = 0), the condition A1 > A3 (resp. A1 < A3) is clearly equivalent to the condition q2 > 1/2 (resp. q2 < 1/2), where q2 = (d − b)/(a − c+d − b), because W+(q) − W(q) is a cubic function of q. Thus, the condition V(q1) > V(q3) (resp. V(q1) < V(q3)) is equivalent to d − b > a − c (resp. d − b < a − c). With a small mutation, although the precise condition is slightly modified, the above discussion still holds approximately.

The observation above shows that in the limit of small mutation, the lifetimes of the metastable states are simply characterized by the relative size of a − c and d − b in the original game (1). The value a − c (> 0) measures the coordination advantage of playing A over playing B when matched with an A individual. Similarly, the value d − b (> 0) measures the coordination advantage of playing B over playing A when matched with a B individual. The strategy with the larger (resp. smaller) coordination advantage is called a 'risk-dominant' (resp. 'risk-dominated') strategy in game theory [4] (e.g. if d − b > a − c, strategy B is risk-dominant and strategy A is risk-dominated). Therefore, our results are summarized as follows: the lifetime of the metastable state of the risk-dominant strategy keeps growing exponentially with the system size, while that of the risk-dominated one saturates after the system size reaches a certain threshold. Hence (when the system is sufficiently large) spatial structure drastically accelerates the selection of the strategy with a larger coordination advantage (= risk-dominant strategy), facilitates its invasion against a risk-dominated strategy and gives a prediction of how natural selection resolves coordination problems in biology.

A similar spatial effect was already found by Ellison for an evolutionary game model (Theorem 3 in [22]). However, a direct comparison with his result is not appropriate because his model assumes one individual per each patch, while our model assumes Np (≫ 1) individuals per patch.

4.2. The origin of the spatial effect: an intuition

Here we intuitively discuss the origin of the qualitative differences in the system size dependence of the lifetimes between the well-mixed model and the spatial model. We first point out that during the transition from one metastable state to the other, the system must cross a 'dividing point', which is a marginal unstable steady solution of the system located between two metastable states. Once the system has crossed this dividing point, it evolves with a very high probability along a 'deterministic' path (the trajectory of the expectation value satisfying (9) or (28)), which quickly carries the system to the other metastable state. This view implies that the lifetime of a metastable state is mostly determined by the difficulty of crossing a dividing point from a given metastable state. Then, the system size dependence of the lifetimes can be intuitively explained as follows (again, we restrict our consideration to the case V(q1) > V(q3)):

  • In the well-mixed model or the spatial model for L < Lc, the unstable state q2 plays the role of a dividing point. To make a transition from one metastable state to the other, the entire system has to cross the dividing point q2 (figure 10(a)). As the system size increases, the probability that the system changes from a given metastable state to the dividing point decays exponentially with N (or L), resulting in the exponential dependence of the lifetimes on N (or L).
  • In the spatial model for L > Lc, the critical nucleus also plays the role of a dividing point. This indicates that the transition from one metastable state to the other can be caused by 'nucleation', i.e. transition via the critical nucleus. The details of the nucleation process depend on the direction of transition.
    • For a transition from q3 to q1 to occur, it is sufficient for the system to change from q3 to the critical nucleus, which means that the change of only a part of the system, whose length scale is approximately Lc, is sufficient (see figure 10(b) and the spatial form of the critical nucleus in the right panel of figure 7). The difficulty of this change does not depend on L and hence neither does the difficulty of transition from q3 to q1. Therefore, the lifetime of q3 does not grow with L.
    • For a transition from q1 to q3 to occur, it is necessary for the system to change from q1 to the critical nucleus, which means that almost the entire system has to change (figure 10(c)). This change becomes more difficult as L grows and so does the transition from q1 to q3. Hence, the lifetime of q1 grows exponentially with L.
Figure 10.

Figure 10. Schematic figures illustrating the transition between metastable states in the spatial model when V(q1) > V(q3). Red and blue circles represent A and B individuals, respectively. (a) In the well-mixed model or the spatial model for L < Lc, transition between metastable states requires the whole system to change (almost uniformly) to cross the dividing point q2. In the spatial model for L > Lc, the transition can be triggered by crossing a critical nucleus. (b) For the transition from q3 to q1 to occur, only a part of the system, whose length scale is approximately Lc, has to change. (c) For the transition from q1 to q3 to occur, almost the whole system has to change.

Standard image High-resolution image

It should be noted that the existence of two kinds of transition paths (shown in figure 10) is directly related to two kinds of bounce solution, i.e. the uniform bounce solution and the non-uniform bounce solution, discussed in section 3.3.2. We stress that the essential change of the transition process is described by the change in the bounce solutions.

Note that when V(q1) < V(q3), the behaviour of the lifetime of q1 and q3 is reversed. This is because the form of the critical nucleus becomes 'upside down' compared with the right panel of figure 7, so that the roles of strategy A (red) and B (blue) in figure 10 should be replaced by each other.

To confirm the intuition described above, we performed a Monte Carlo simulation of the spatial model, specified by the master equation (23) employing the Gillespie algorithm [39]. Figure 11 shows results of the simulation. For a small system (the upper panel, L < Lc), the transition from one metastable state (the red region) to the other state (the blue region) occurs almost uniformly. For a large system (the lower panel, L > Lc), however, the transition proceeds via nucleation; transition occurs first in a small region and then it spreads over the whole system.

Figure 11.

Figure 11. Results of a Monte Carlo simulation of master equation (23). Top: M = 35, Bottom: M = 140. The colour shows the number of A individuals in each patch (Np = 40). Parameters: Np = 40, a − c = 0.4, d − b = 1.0, w = 0.1, μA = μB = 0.005, λ = Np, σ = 2Np (for a detailed description of the parameters, see section 2.1 and 3.1). For this parameter set, Npq1 ≃ 1.01, Npq3 ≃ 36.8 and Lc ≃ 50.1 (from (33)). In terms of the patch number M, Lc ≃ 50.1 is equivalent to $M_c := \sqrt{\sigma/\lambda}L_c \simeq 70.8$ . We can see that for a small system (the upper figure, M < Mc), the transition process proceeds almost uniformly in space, while in a large system (the lower figure, M > Mc), the transition starts from a small region and then spreads across the whole system (nucleation), as explained in figure 10.

Standard image High-resolution image

We point out that the intuition presented above suggests the generality of our results; although we have demonstrated a spatial effect on bistable evolutionary games with a specific update rule, the qualitative behaviour of the lifetimes is expected to be the same for models with other update rules as long as the underlying evolutionary game is bistable. This is because our results rely on the presence of a critical nucleus (which has a characteristic length scale Lc), which in turn originates from the structure of the phase portrait (see figures 1 and 7). It can easily be shown that models with other update rules (e.g. a pairwise comparison process with other forms of p(Δ Π) or the Moran process) also generate phase portraits with the same structure as figures 1 and 7, provided that the payoff matrix elements satisfy (3) and mutation rates are sufficiently small.

4.3. Characteristic length scale Lc

The characteristic length Lc in the spatial model is an important quantity that determines the population size dependence of the lifetimes. In this section, we discuss its parameter dependence in detail. The characteristic length Lc is calculated in section 3.2.2 and is given by

Equation (49)

To observe the parameter dependence more clearly, we derive an approximate expression for Lc: because q1, q2 and q3 are the solutions of the cubic equation W+(q) − W+(q) = 0, we obtain an alternative expression for W+(q) − W(q) (see (10)):

Equation (50)

By using (49), we obtain

Equation (51)

Assuming that μA and μB are sufficiently small, we arrive at a concise approximate form

Equation (52)

from which we can conclude that Lc decreases with w, a − c and d − b.

Figure 12 shows the w dependence of Lc, where the circles and solid line denote the results obtained from a numerical evaluation of (49) and the approximate expression (52), respectively. One can see that the present approximation works well as w increases.

Figure 12.

Figure 12. w dependence of Lc(a − c = 0.4, d − b = 1.0, μA = μB = 0.005). circle: equation (49), solid line: approximate expression (52).

Standard image High-resolution image

Figure 13 shows the payoff matrix dependence of Lc; the left and right panels are obtained from the numerical evaluation of (49) and the approximate expression (52), respectively. One can see that Lc decreases as a − c and d − b increase and that the approximate expression (52) captures the qualitative feature well.

Figure 13.

Figure 13. Payoff matrix dependence of Lc(w = 0.8, μA = μB = 0.005) Left: equation (49), Right: approximate expression (52).

Standard image High-resolution image

4.4. Migration rate dependence

For a comparison with real systems, it is important to judge whether the system under consideration can be regarded as well-mixed or spatial. In this section, we show the criterion in terms of migration rate σ and examine the effect of σ on the lifetimes of the metastable states.

As was pointed out in section 4.1, the system is effectively well-mixed if L < Lc. Because L is defined by $L := \tilde{L}/\sqrt{D} = M \sqrt{\lambda/\sigma}$ (M: the number of patches), the condition L < Lc is equivalent to

Equation (53)

Note that Lc depends only on the game theoretical parameters given in table 1 but not on σ or λ (see (49)). Therefore, in terms of the migration rate σ, the system is effectively well-mixed if

Equation (54)

i.e. if the ratio between migration rate and update rate is larger than the value determined by M and the game theoretical parameters. On the other hand, when σ is small so that (54) is not satisfied, the system should be described as the spatial model. Then, the transition between the metastable states is affected by the nucleation process described in section 4.2.

Finally, we discuss the σ dependence of the lifetimes of the metastable states. We focus on the exponents of the lifetimes

Equation (55)

(see (46)). The lifetimes take constant values when σ is large so that (54) is satisfied, i.e. when the system is effectively well-mixed. This is because Sqi, sp(L) grows linearly with L when L < Lc and therefore factors depending on σ are cancelled out (see (55)). When σ becomes smaller than the critical value in (54), the lifetimes decrease as σ decreases. Thus, it can be concluded that spatial structure accelerates the transition between metastable states, leading to a drastic reduction of the lifetimes compared to those of the well-mixed model.

4.5. Which is the long-lived metastable state?

When we study coordination games, it is important to determine which metastable state has a longer lifetime, because it gives a criterion for which one of A and B is better off [17]. We briefly discuss the spatial effect on this problem.

For the well-mixed model, the metastable state with the larger value of action Sqi, wm has a longer lifetime as can be seen from (45). For the spatial model with a sufficient size (L > Lc), on the other hand, the metastable state with the larger value of potential V(qi) has a longer lifetime (see section 3.4 and 4.1).

In the limit of small mutation rates (μA, μB → 0), it can easily be shown that these two conditions coincide: V(q1) > V(q3) and Sq1, wm > Sq3, wm are, respectively, equivalent to the condition d − b > a − c, suggesting that the risk-dominant strategy has a longer lifetime, which is consistent with previous studies [17, 22]. However, in the presence of asymmetric mutation (μA, μB > 0, μAμB), the conditions are not necessarily equivalent. This observation indicates that the long–lived metastable state may switch from one state to the other as the system size increases. A detailed discussion of this problem is left for a future study.

5. Summary

In this paper, we have evaluated the lifetimes of metastable states in bistable evolutionary games by utilising the path integral method and the semiclassical approximation. It has been shown that spatial structure qualitatively changes the system size dependence of the lifetimes of the metastable states: For the model without spatial structure (the well-mixed model), we have shown that the lifetimes of the metastable states grow exponentially with the total population size N. On the other hand, for the model with spatial structure (the spatial model), we have shown that there is a threshold length Lc across which the system size L (∝ the total population size) dependence of the lifetimes change. For L < Lc, the lifetimes of the two metastable states grow exponentially with the system size, whereas for L > Lc, the lifetime of one metastable state remains constant, while that of the other continues growing exponentially with L. This significant change in the system size dependence can be intuitively explained by the presence of critical nuclei; for large systems (L > Lc) the transition is induced by 'nucleation' via critical nuclei.

We stress that the present method allows a semi-quantitative calculation of the lifetimes taking into account large fluctuations. Although we considered specific models, the present method can be easily applied to other models of the evolutionary game theory. Extension toward evolutionary games on higher-dimensional lattices or complex networks is an important future problem.

Appendix A.: Stochastic processes and path integral

In this appendix, we will derive a path integral expression for the transition probability used in section 2.3.1 and 3.3.1. The derivation is based on [40].

We consider a general continuous time Markov process on the discrete states $n\in\mathbb{N}$ . The rates at which processes $n \rightarrow n+ r \ (r \in R \subset \mathbb{Z} - \left\{0 \right\})$ occur are given by Wr(n). By setting W_±1(n) = W±(n/N) and Wr(n) = 0 (for |r|⩾2), we obtain the well-mixed model discussed in section 2. Let t ∈ [0, T] be time. Our aim is to calculate p(nfin, T | nini, 0), the probability that n = nfin at t = T given that n = nini at t = 0 (other conditions can be added, such as restriction of the path).

First, we discretize the time interval [0, T] into K (≫ 1) small fractions:

Equation (A.1)

where n0 := nini and nK := nfin. Let P(Δ n | n) be the probability that the system changes from n to n +Δ n in one discrete time step. Then, p(nfin, T | nini, 0) can be approximated by

Equation (A.2)

By substituting

Equation (A.3)

we obtain

Equation (A.4)

Since

Equation (A.5)

holds for a sufficiently small time interval Δ t, we arrive at

Equation (A.6)

Taking the limit K →  yields

Equation (A.7)

Equation (A.8)

Equation (A.9)

The derivation for models with spatial degrees of freedom proceeds almost in parallel with the derivation shown above.

We consider a 1D array of M patches with a periodic boundary condition. The state of the system is specified by n := (n1, n2, ···, nM). We assume that local processes occur as specified above: the process ${\bit n} \rightarrow {\bit n}+ r{\bit e}_i \ (r \in R \subset \mathbb{Z} - \left\{0 \right\})$ occurs at rate Wr(ni), where

Equation (A.10)

In addition to the local processes, we assume there are migration processes between neighbouring patches: for |i − j| = 1, the process n → nei + ej occurs at a rate Wm(ni, nj). By replicating the discussion above, we obtain

Equation (A.11)

Since

Equation (A.12)

holds for sufficiently small Δ t, we arrive at

Equation (A.13)

where

Equation (A.14)

Appendix B.: Stability of non-uniform solutions

In this appendix, we outline the proof that a non-uniform steady solution for (28) is unstable. Let qcn(·) be a non-uniform solution for (28). Linearization of (28) around qcn by substituting q(ξ, τ) = qcn(ξ) + δq(ξ, τ) yields

Equation (B.1)

If all the eigenvalues of $\mathcal{L}$ are negative, qcn is linearly stable and otherwise qcn is linearly unstable. It can be easily shown that δq0 := dqcn/dξ is an eigenfunction with a zero eigenvalue (called a zero mode) of $\mathcal{L}$ . Because δq0 has two nodes, one can show (with the help of the theory of periodic Sturm–Liouville problems [43]) that there must be one nodeless eigenmode with a positive eigenvalue. Hence, qcn is linearly unstable.

The above result indicates that a non-uniform steady solution represents a marginal state located at the boundary between the metastable states q1 and q3. If a positive perturbation δq(ξ) (>0for allξ) is added to qcn, the system is driven toward q3 following the deterministic equation (28). On the other hand, if a negative perturbation δq(ξ) (<0for allξ) is added, the system is driven toward q1.

Appendix C.: Numerical calculation

In this appendix, we briefly describe the numerical method used to obtain the action of the bounce solution in the spatial model (section 3.3.2). As in section 3.3.2, we calculate bounce solutions that determines the lifetime of q3 (almost the same discussion applies to q1).

First, note that for both the solutions α and β, it is not necessary to consider the second half of the transition (transition from q2 to q3 for the bounce solution α and transition from qcn to q3 for the bounce solution β) since p ≡ 0 for these trajectories and hence they do not contribute to the action (see (35)-(37)). This is because they correspond to trajectories obeying the 'deterministic' partial differential equation (28), which can be obtained from (38) and (39) by setting p ≡ 0. Hence, it is sufficient to consider the first half of the transition of the the bounce solutions α and β, which we call solutions α' and β' respectively, for the calculation of $S_{\alpha} (= S_{\alpha^{\prime}})$ and $S_{\beta} (= S_{\beta^{\prime}})$ . These solutions are obtained by solving the partial differential equations (38) and (39)

Equation (C.1)

Equation (C.2)

where F(q, p) := epW+(q) − epW(q) and $G(q,p) := -({\rm e}^{p}-1)W^{\prime}_{+}(q) - ({\rm e}^{-p}-1)W^{\prime}_{-}(q)$ , subjected to boundary conditions

where T is taken to be sufficiently large (T ≫ 1). This problem was solved numerically as follows.

We discretize time and space as

Equation (C.3)

Equation (C.4)

Equation (C.5)

where NT and NL are the division numbers of the time and spatial coordinates, respectively. Note that q0,j, qNT−1, j(j ∈ {0, 1, ···, NL−1}) are given as the boundary conditions. We then discretize (C.1) and (C.2) into the following difference equations: For i ∈ {1, 2, ···, NT −2}, j ∈ {0, 1, ···, NL−1}

Equation (C.6)

Equation (C.7)

For i = 0, j ∈ {0, 1, ···, NL −1},

Equation (C.8)

For i = NT−1, j ∈ {0, 1, ···, NL−1},

Equation (C.9)

The original problem of solving partial differential equations is now reduced to finding zeros of a function $f: \mathbb{R}^{2N_L (N_T -1)} \rightarrow \mathbb{R}^{2N_L (N_T -1)}$ . This problem can be solved numerically by Newton's method. Solving boundary value problems by applying Newton's method is called the relaxation method [44].

Figure C1 shows the calculated bounce solutions which determine the lifetime of q3. It can easily be seen from (38) and (39) that the uniform bounce solution is the same as the activation trajectory of the well-mixed model. This is because, if q and p do not depend on ξ, (38) and (39) coincide with the equations of motion for the well-mixed model (14) and (15), respectively. Bounce solutions which determine the lifetime of q1 can be calculated in almost the same way by changing the boundary conditions and are shown in figure C2.

Figure C1.

Figure C1. Bounce solutions which determine the lifetime of q3. Left: uniform bounce solution α'(L = 10.5). Right: nonuniform bounce solution β'(L = 20.5). Parameters: a − c = 0.4, d − b = 1.0, w = 0.8, μA = μB = 0.005. For this parameter set, Lc ≃ 13.5 and the fixed points are q1 = 0.0031, q2 = 0.72, q3 = 0.99.

Standard image High-resolution image
Figure C2.

Figure C2. Bounce solutions which determine the lifetime of q1. Left: uniform bounce solution (L = 10.5). Right: nonuniform bounce solution (L = 20.5). Parameters: a − c = 0.4, d − b = 1.0, w = 0.8, μA = μB = 0.005. For this parameter set, Lc ≃ 13.5 and the fixed points are q1 = 0.0031, q2 = 0.72, q3 = 0.99.

Standard image High-resolution image
Please wait… references are loading.
10.1088/1742-5468/2014/10/P10020