nlin0505032/rep.tex
1: \documentclass{article}
2: 
3: \usepackage{amsmath}
4: \usepackage{graphicx}
5: 
6: \begin{document}
7: 
8: \title{Replicator dynamics with mutations for games with a continuous
9: strategy space}
10: \author{M.Ruijgrok$\thanks{%
11: Mathematics Institute, Utrecht University, Utrecht, The Netherlands. e-mail
12: address: ruijgrok@math.uu.nl}$ and Th.W. Ruijgrok$\thanks{%
13: Institute of Theoretical Physics, Utrecht University, Utrecht, The
14: Netherlands.}$}
15: \date{May 2005}
16: \maketitle
17: 
18: \begin{abstract}
19: A partial differential equation is derived, describing the replicator
20: dynamics with mutations of games with a continuous strategy space. This
21: equation is then applied to continuous versions of symmetric 2x2 games, such
22: as the Prisoners Dilemma, Hawk-Dove and Coordination games, and to the
23: Ultimatum Game. In the latter case, we find that adding even a small mutation
24: term to the replicator equation leads to a solution where the average offer is
25: significantly larger than zero.
26: \end{abstract}
27: 
28: \section{Introduction}
29: 
30: Evolutionary game dynamics is a fast developing field, with applications in
31: biology, economics, sociology and anthropology. Background material and
32: countless references can be found in the monographs by Weibull \cite{WB},
33: Fudenburg \& Levine \cite{FL}, Samuelson \cite{SA}, Hofbauer \& Sigmund \cite%
34: {HS1}, Gintis \cite {GIN}, Cressman \cite{CRS} and Vincent \& Brown \cite{VB} or in the survey
35: paper by Hofbauer \& Sigmund \cite{HS2}. The standard ingredients of
36: evolutionary game dynamics are a population of players, an $n$-person game, a
37: set of strategies and a rule to update the distribution of strategies from
38: one generation to the next. Within this general setting several variations
39: are possible: time can be discrete or continuous, populations can be finite
40: or infinite, the game can have a finite or infinite number of strategies.
41: Also, there are several choices for the updating rule, the most popular of
42: which are Adaptive Dynamics (See e.g. Diekmann \cite{DI}) and Replicator
43: Dynamics, introduced by Taylor \& Jonker \cite{TJ}. In this paper we
44: consider a model with continuous time, an infinite population and a 2-person
45: game, where each participant has a continuum of strategies to choose from.
46: The update rule we use is the replicator dynamics, with deterministic
47: mutations. Although this model has been alluded to by Dieckmann \cite{DIC},
48: where a hierarchy of evolutionary models is presented, its details have not
49: been worked out before.
50: 
51: The strategy space that we consider is a subset of $R^{n}$, which in all
52: applications is compact. In section 2 we first consider replicator dynamics
53: with a deterministic mutation term on a finite set of strategies. Using a
54: procedure that is familiar in Statistical Mechanics, we make a transition to
55: an infinite, continuous, strategy-space. The state-space of the model is
56: then no longer discrete, but consists of distributions over the strategies.
57: We derive a partial differential equation, together with appropriate
58: boundary conditions, that describes the time-evolution of this distribution,
59: given an initial distribution. Without the mutation term, the equation we
60: derive was previously studied by Cressman \cite{CR} and Oechsler \& Riedel 
61: \cite{OR}. Our addition of a mutation term is new and leads to completely
62: different dynamics.
63: 
64: In section 3, the equation is applied to symmetric 2x2 games, where we need
65: to extend the original two-strategy set to a continuous one. This can be
66: done by allowing for strategies that play one of the pure strategies with a
67: certain probability. We compare our results with those of \ Vaughan \cite{VA}%
68: , who analyses the replicator dynamics for the pure strategies, to which a
69: stochastic perturbation term is added, leading to a Fokker-Planck equation.
70: There are similarities between the two models, but also striking
71: differences. In the stochastic mutation version, there is always only one
72: attracting, stationary, distribution, which for small mutations converges to
73: a point-distribution (Dirac-delta function). In other words, all players
74: eventually use the same strategy. In the deterministic mutation version,
75: however, we find the possibility of two attracting stationary distributions
76: existing simultaneously. Also, in certain cases the limiting distribution
77: for small mutations is not concentrated on a point, but has the whole
78: strategy space as support.
79: 
80: A more complicated example, namely the Ultimatum Game, is treated in section
81: 4. In this game, a sum of money is to be split by two players. The first
82: player proposes the split and the second player then has the choice to
83: accept the split or refuse, in which case both players get nothing. The
84: solution offered by standard game theory is for the second player to accept
85: any amount (something is always better than nothing) and for the first
86: player, therefore, to offer the lowest possible amount. In our model we find
87: that the offers converge to a Gauss-like distribution around a mean that is
88: not equal to zero. The position of the mean and the width of the
89: distribution seem to converge to zero when the rate of mutation vanishes.
90: However, this convergenge is slow, so that even for very small values of the
91: mutation rate, the average offer is well above zero. Also, the dynamics
92: shows two time-scales. A random starting distribution is initially attracted
93: to a member of a certain set of distributions, which can have an average
94: offer much larger than zero. Then, on a time-scale inversely proportional to
95: the rate of mutation, the solution converges to the, unique, stationary
96: solution.
97: 
98: In section 5 we discuss the results and suggest some topics for further
99: research.
100: 
101: \section{From the discrete to the continuous equation}
102: 
103: The equation for the replicator dynamics with continuous strategy space will
104: be derived by a limiting process, starting from the equations for the
105: discrete case.
106: 
107: Following Hofbauer and Sigmund \cite{HS1}, we consider an infinite
108: population of players and a set of strategies $S_{1},\cdots ,S_{N}$. When a
109: player $A$ opts for strategy $S_{i}$ against $B,$ who uses $S_{j}$, the
110: payoff to $A$ is taken to be $M_{ij},$ and the payoff to $B$ is $M_{ji}.$
111: Consequently, if $p_{j}(t)$ is the fraction of the whole population that at
112: time $t$ plays the strategy $S_{j}$, the average payoff to each player using 
113: $S_{i}$ is equal to 
114: \begin{equation}
115: \Pi _{i}(\mathbf{p})\equiv \sum_{j=1}^{N}M_{ij}p_{j}.  \label{indpay}
116: \end{equation}
117: The average payoff for the whole population therefore is 
118: \begin{equation}
119: \overline{\Pi }(\mathbf{p})=\sum_{i=1}^{N}\Pi _{i}(\mathbf{p}%
120: )p_{i}=\sum_{i,j=1}^{N}p_{i}M_{ij}p_{j}.  \label{avfit}
121: \end{equation}
122: Now in the course of time, the fraction $p_{i}$ changes at a rate which is
123: proportional to the difference between the payoff to $p_{i}$ and the average
124: payoff for the whole population. In addition we assume that for each player
125: of $\ S_{j}$ there is a transition probability per unit time to make a
126: spontaneous transfer to strategy $S_{i}$ at a rate given by $W_{ij}$. In
127: this way the \textit{discrete replicator equation with mutation} is derived: 
128: \begin{equation}
129: \dot{p}_{i}=(\Pi _{i}(\mathbf{p})-\overline{\Pi }(\mathbf{p}%
130: ))p_{i}+\sum_{j=1}^{N}(W_{ij}p_{j}-W_{ji}p_{i}).  \label{drep}
131: \end{equation}
132: It is easy to show that $S(t)\equiv \sum\limits_{i=1}^{N}p_{i}(t)=1$ for all 
133: $t$ if $S(0)=1.$
134: 
135: In transforming to a continous strategy space, we replace the discrete index 
136: $i$ by a continuous variable $s\in D\subset R^{n}$ and the variables $%
137: p_{i}(t)$ by a probability distribution $P(s,t)$. The payoff matrix $M_{ij}$
138: must now be replaced by a payoff function $M(s,s^{\prime })$, which gives
139: the payoff to strategy $s$ when playing strategy $s^{\prime }$.
140: 
141: Eq.(\ref{drep}) now takes the form 
142: \begin{equation}
143: \frac{\partial P(s,t)}{\partial t}=(\Pi (s,P)-\overline{\Pi }(P))P(s,t)+%
144: \mathcal{M}(P,s)  \label{crep}
145: \end{equation}%
146: in which the mutation term is equal to 
147: \begin{equation}
148: \mathcal{M}(P,s)=\int [W(s|s^{\prime })P(s^{\prime },t)-W(s^{\prime
149: }|s)P(s,t)]\,\mathrm{d}s^{\prime }.  \label{mut}
150: \end{equation}%
151: The average payoff for strategy $s$ and the total average payoff are given
152: by 
153: \begin{equation}
154: \Pi (s,P)=\int M(s,s^{\prime })P(s^{\prime },t)\,\mathrm{d}s^{\prime }
155: \label{Pisp}
156: \end{equation}%
157: and 
158: \begin{equation}
159: \overline{\Pi }(P)=\int \Pi (s,P)P(s,t)\,\mathrm{d}s.  \label{Pip}
160: \end{equation}%
161: Let us now restrict ourselves to mutations in which only small changes in
162: the strategies occur and apply the method which is used to derive the
163: Fokker-Planck equation from a master equation \cite{VK}. For simplicity of
164: presentation, we restrict ourselves to the case of a one-dimensional
165: strategy space. Define $\xi $ by $s^{\prime }=s-\xi $ and write the mutation
166: rate as a function of $s$ and $\xi $%
167: \begin{equation}
168: W(s^{\prime }|s)=\widetilde{W}(s,s-s^{\prime })=\widetilde{W}(s,\xi )
169: \label{mut1}
170: \end{equation}%
171: and so 
172: \begin{equation}
173: W(s|s^{\prime })=\widetilde{W}(s^{\prime },s^{\prime }-s)=\widetilde{W}%
174: (s-\xi ,-\xi ).  \label{mut2}
175: \end{equation}%
176: Assume now that $\widetilde{W}(s,\xi )$ varies slowly in the first variable $%
177: s$ and that due to the mutations only small variations in the strategies
178: will occur. Then $\widetilde{W}(s,\xi )$ is only nonvanishing when $\xi $ is
179: small. In the mutation term (\ref{mut}), which can be written as 
180: \begin{equation}
181: \mathcal{M}(P,s)=\int [\widetilde{W}(s-\xi ,-\xi )P(s-\xi ,t)-\widetilde{W}%
182: (s,\xi )P(s,t)]\,\mathrm{{d}\xi ,}  \label{mut3}
183: \end{equation}%
184: we can now expand the dependence on the first variable in powers of $\xi $
185: and obtain 
186: \begin{eqnarray}
187: \mathcal{M}(P,s) &=&\int [\widetilde{W}(s,-\xi )P(s,t)-\xi \frac{\partial }{%
188: \partial s}\{\widetilde{W}(s,-\xi )P(s,t)\}+  \label{mut4} \\
189: &&+\frac{1}{2}\xi ^{2}\frac{\partial ^{2}}{\partial s^{2}}\{\widetilde{W}%
190: (s,-\xi )P(s,t)\}+\cdots -\widetilde{W}(s,\xi )P(s,t)]\,\mathrm{{d}\xi .} 
191: \notag
192: \end{eqnarray}%
193: Because $\int [\widetilde{W}(s,-\xi )-\widetilde{W}(s,\xi )]P(s,t)\mathrm{{d}%
194: \xi =0}$, the first and last term cancel, so that to second order we are
195: left with 
196: \begin{equation}
197: \mathcal{M}(P,s)=\frac{\partial }{\partial s}\{\alpha _{1}(s)P(s,t)\}+\frac{1%
198: }{2}\frac{\partial ^{2}}{\partial s^{2}}\{\alpha _{2}(s)P(s,t)\},
199: \label{mut5}
200: \end{equation}%
201: in which 
202: \begin{equation}
203: \alpha _{1}(s)=-\int \xi \widetilde{W}(s,\xi )\,\mathrm{d}\xi {\qquad }\text{%
204: and}{\qquad }\alpha _{2}(s)=\int \xi ^{2}\widetilde{W}(s,\xi )\,\mathrm{d}%
205: \xi .  \label{alf}
206: \end{equation}%
207: We will further simplify the equations by assuming that the average change
208: in strategy due to mutations is equal to zero, so $\alpha _{1}(s)=0,$ and
209: that the average of the square of this change is constant, so $\alpha
210: _{2}(s)=2\sigma .$
211: 
212: The final form of the \textit{continuous replicator equation }(\ref{crep})%
213: \textit{\ }then becomes 
214: \begin{equation}
215: \frac{\partial P(s,t)}{\partial t}=(\Pi (s,P(t))-\overline{\Pi }%
216: (P(t)))P(s,t)+\sigma \Delta P(s,t),  \label{rep}
217: \end{equation}%
218: where we have restored the correct dimensionality of the strategy space by
219: replacing $\frac{\partial ^{2}}{\partial s^{2}}$ by the \textit{n}%
220: -dimensional Laplace operator. $P(s,t)$ should satisfy $P(s,t)\geq 0$ and $%
221: S(t)\equiv \int_{D}P(s,t)\,$d$s,$ should be equal to unity at all times.
222: This last condition is fulfilled when we choose Neumann, or reflecting, boundary
223: conditions:
224: 
225: \begin{equation}
226: \mathbf{n}.\nabla P(s,t)|_{\partial D}=0,  \label{bound}
227: \end{equation}%
228: where $\mathbf{n}$ is the normal to the boundary $\partial D$ of the domain $%
229: D$.
230: 
231: Indeed, integrating (\ref{rep}) over $D$ and using (\ref{bound}) we find 
232: \begin{equation}
233: \frac{dS(t)}{dt}=\overline{\Pi }(P)(1-S(t)),  \label{dst}
234: \end{equation}%
235: showing that if $S(0)=1$ then $S(t)=1$ for all times. In the case that $%
236: \sigma =0$, condition (\ref{bound}) is not required to ensure that $S(t)$
237: remain constant in time.
238: 
239: 
240: 
241: Equation (\ref{rep}) is a nonlinear reaction-diffusion
242: equation, where the reaction term $\Pi (s,P(t))-\overline{\Pi }(P(t))$ is
243: nonlocal. On the function space of twice continuous space-differentiable and
244: once time-differentiable functions, we can show that the solution of (\ref%
245: {rep}) exists for all times. This follows from the assumption that $%
246: M(s,s^{\prime })$ is bounded on $D$, so that $\left\vert \Pi (s,P(t))-%
247: \overline{\Pi }(P(t))\right\vert \leq \int_{D\times D}\left\vert
248: M(s,s^{\prime })\right\vert P(s,t)P(s^{\prime },t)\,\mathrm{d}s\,\mathrm{d}%
249: s^{\prime }+\int_{D}\left\vert M(s,s^{\prime })\right\vert P(s^{\prime },t)\,%
250: \mathrm{d}s^{\prime }\leq $
251: 
252: $\max \left\vert M(s,s^{\prime })\right\vert \left\{ \int_{D\times
253: D}P(s,t)P(s^{\prime },t)\,\mathrm{d}s\,\mathrm{d}s^{\prime
254: }+\int_{D}P(s^{\prime },t)\,\mathrm{d}s^{\prime }\right\} $
255: 
256: $=2\max \left\vert M(s,s^{\prime })\right\vert $. 
257: 
258: Standard comparison theorems for parabolic equations (Pao \cite{PAO}) complete the
259: proof. Also, by standard positivity results for parabolic equations, it can
260: be shown that the when the initial distribution $P(s,0)\geq 0,$ then $P(s,t)\geq 0$
261: for all times $t.$
262: 
263: Numerical simulations suggest that even stronger results hold. In
264: particular, we suspect that the solution of Eqs.(\ref{rep}), (\ref{bound}) is
265: uniformly (in space and time) bounded in terms of the sup-norm of the initial distribution. For one-dimensional strategy spaces, we speculate that the solution will always
266: converge to a stationary solution. 
267: 
268: For $\sigma =0$, Eq.(\ref{rep}) has been studied by Cressman \cite{CR} and
269: Oechsler \& Riedel \cite{OR}. They show that Eq.(\ref{rep}) has a unique
270: solution, for all times, on a large space of distributions, containing
271: amongst others the Dirac-delta distributions.
272: \newpage
273: \section{Symmetric 2x2 games}
274: 
275: In symmetric $2\times 2$ games there are two possible strategies, denoted by 
276: $I$ and $II$. The payoff to player $A$ is given by the payoff matrix
277: 
278: \begin{center}
279: $M$= 
280: \begin{tabular}{|l||l|l|}
281: \hline
282: A{$\backslash$}B & I & II \\ \hline\hline
283: I & a & b \\ \hline
284: II & c & d \\ \hline
285: \end{tabular}
286: .
287: \end{center}
288: 
289: \subsection{Discrete replicator dynamics}
290: 
291: The discrete replicator dynamics associated with this game consists of an
292: infinite population where a fraction $x_{1}(t)$ plays the pure strategy $I$
293: and a a fraction $x_{2}(t)$ plays the pure strategy $II$ . The payoffs to
294: strategies $I$ and $II$ are given by: 
295: \begin{equation}
296: \Pi _{I}(x_{1},x_{2})=ax_{1}+bx_{2}\qquad \text{and}\qquad \Pi
297: _{II}(x_{1},x_{2})=cx_{1}+dx_{2}.
298: \end{equation}%
299: The average payoff to the total population is then 
300: \begin{equation}
301: \overline{\Pi }(x_{1},x_{2})=\Pi _{I}(x_{1},x_{2})x_{1}+\Pi
302: _{II}(x_{1},x_{2})x_{2}=ax_{1}^{2}+(b+c)x_{1}x_{2}+dx_{2}^{2}.
303: \end{equation}%
304: With a mutation rate matrix of the form $W=\sigma \left( 
305: \begin{array}{ll}
306: 0 & 1 \\ 
307: 1 & 0%
308: \end{array}%
309: \right) $, $\sigma >0$ and using $x_{1}(t)+x_{2}(t)=1$, this leads to the
310: following equation for $x_{1}(t):$%
311: \begin{equation}
312: \dot{x}_{1}=x_{1}(1-x_{1})(B+(A-B)x_{1})+\sigma (1-2x_{1}),  \label{discrrep}
313: \end{equation}%
314: where $A=a-c$ and $B=b-d$. The solutions of Eq.(\ref{discrrep}) for $\sigma
315: =0$ are summarised in figure 1.
316: 
317: \begin{figure}[hbt]
318: \centering
319: \includegraphics[width=8cm]{f1.eps}
320: \caption{The four quadrants of the parameter space}
321: \end{figure}
322: 
323: A typical example of a game in the quadrant $A>0,B>0$ is the classic
324: Prisoners Dilemma, where $a=1,b=5,c=0,d=3$. The strategy $I$ corresponds to
325: 'defect' and strategy $II$ to 'cooperate'. Figure 1 shows that for $\sigma
326: =0 $ the discrete replicator dynamics Eq.(\ref{discrrep}) predicts a final
327: outcome of $x_{1}=1$, $x_{2}=0$, or 'All Defect'. The effect of the mutation
328: is to shift the stable solution to a slightly lower value of $x_{1}$.
329: 
330: The Hawk-Dove game (also known as Chicken) has $a=(G-C)/2,b=G,c=0,d=G/2$,
331: where now strategy $I$ corresponds to 'hawk' (or 'never back down') and $II$
332: to 'dove' ('allways back down'). When the cost $C$ to the loser of a
333: hawk-hawk fight is larger than the gain $G$ a hawk makes when confronting a
334: dove, we have $A<0,B>0$. It follows from Eq.(\ref{discrrep}) that for $%
335: \sigma =0$ the solution tends to the stable equilibrium $x_{1}=B/(B-A)=G/C$,
336: describing a population where the strategies co-exist. Also in this case,
337: the effect of the mutation term is restricted to a small shift in the
338: location of the stable equilibrium.
339: 
340: The quadrant with $A>0,B<0$ is the domain of the Coordination Games,
341: exemplified by the situation $a=2,b=0,c=0,d=1.$ In a Coordination Game it is
342: advantageous for both players to play the same strategy. In a $2\times 2$
343: game, this leads to two equilibria. Note, however, that in the example
344: mentioned here the situation where both play strategy $I$ is superior to the
345: one where both play $II.$ For $\sigma =0$ there are two stable solutions and
346: one unstable one, and the final outcome depends on the initial situation.
347: When $x_{1}(0)<\frac{1}{3}$ the solution will ultimately tend to $x_{1}=0$,
348: otherwise to $x_{1}=1$. In other words, when the initial fraction of
349: strategy $I$ players is too small, the final population will consist
350: exclusively of strategy $II$ players, even though this is the less
351: attractive of the two equilibria.
352: 
353: In the case of Coordination Games, the effect of mutation can qualitatively
354: change this picture. With the above given values of $a,b,c$ and $d$, we have
355: plotted the right-hand side of Eq.(\ref{discrrep}) for several values of $%
356: \sigma $. (figure 2).
357: 
358: \begin{figure}[hbt]
359: \centering
360: \includegraphics[width=7cm]{f2.eps}
361: \caption{The
362: bifurcation property of Eq.(\protect\ref{discrrep})}
363: \end{figure}
364: 
365: From this
366: picture it follows that for values of $\sigma $ larger than a critical value 
367: $\sigma _{c}$ (in this example $\sigma _{c}\approx 0.11$) the only
368: equilibrium left is the optimal one near $x_{1}=1$, which is also the
369: globally attracting solution.
370: 
371: \subsection{Discrete replicator dynamics with a stochastic term}
372: 
373: In \cite{VA}, Eq.(\ref{discrrep}) is studied, where instead of a
374: deterministic mutation term, a small noise term is added. This leads to the
375: stochastic equation:
376: 
377: \begin{equation}
378: dx=G(x)dt+2\sigma \,dW,  \label{discrstoch}
379: \end{equation}
380: where $G(x)=x(1-x)(B+(A-B)x)$ and $W(t)$ denotes a Wiener process with zero
381: mean and unit variance.
382: 
383: In this model, the mutation from one strategy to another does not happen at
384: a fixed rate, as in the model described by Eq.(\ref{discrrep}), but rather
385: the fraction $x_{1}(t)$ is changed by a small random amount per time-step.
386: To describe the outcome of Eq.(\ref{discrstoch}), we consider the evolution
387: of the probability density $f(x,t).$ The probability that at time $t$ the
388: fraction of strategy $I$ players lies in the interval $\left[ x,x+\Delta x%
389: \right] $ is given by $f(x,t)\Delta x$. The equation for $f(x,t)$ is the
390: Fokker-Planck equation \cite{VK}: 
391: \begin{equation}
392: \frac{\partial f(x,t)}{\partial t}=-\frac{\partial }{\partial x}%
393: (G(x)f(x,t))+\sigma \frac{\partial ^{2}f(x,t)}{\partial x^{2}}.  \label{FP1}
394: \end{equation}
395: We assume reflecting boundaries, which yield as boundary conditions: 
396: \begin{equation}
397: \frac{\partial f(x=0,t)}{\partial x}=\frac{\partial f(x=1,t)}{\partial x}=0.
398: \label{bcFP}
399: \end{equation}
400: Eq.(\ref{FP1}) with conditions (\ref{bcFP}) has a unique, attracting,
401: stationary distribution $f^{\ast }(x)$ \cite{VA}.
402: 
403: It is easy to see that this equilibrium distribution is given by 
404: \begin{equation}
405: f^{\ast }(x)=C\exp (\frac{1}{\sigma }\int_{0}^{x}G(x^{\prime })\, \mathrm{d}%
406: x^{\prime }),  \label{fster}
407: \end{equation}
408: where the constant $C$ is determined by the condition $\int_{0}^{1}f^{\ast
409: }(x)dx=1$.
410: 
411: Differentiating $f^{\ast }(x)$ once yields 
412: \begin{equation}
413: \frac{df^{\ast }(x)}{dx}=\frac{C}{\sigma }G(x)\exp (\frac{1}{\sigma }%
414: \int_{0}^{x}G(x^{\prime })\,\mathrm{d}x^{\prime }),
415: \end{equation}%
416: from which it follows that the extrema of $f^{\ast }(x)$ are the zeroes of $%
417: G(x)$, which are the equilibria of the discrete replicator equation (\ref%
418: {discrrep}) with $\sigma =0$. Differentiating once more gives 
419: \begin{equation}
420: \frac{d^{2}f^{\ast }(x)}{dx^{2}}=\frac{C}{\sigma }(G^{\prime }(x)+\frac{%
421: G^{2}(x)}{\sigma })\exp (\frac{1}{\sigma }\int_{0}^{x}G(x^{\prime })\,%
422: \mathrm{d}x^{\prime }),
423: \end{equation}%
424: so at an equilibrium point $x_{e}$ we have that $\frac{d^{2}f^{\ast
425: }(x=x_{e})}{dx^{2}}=K\,G^{\prime }(x_{e})$ with $K>0$. Therefore, a stable
426: equilibrium of Eq.(\ref{discrrep}) with $\sigma =0$ corresponds to a maximum
427: of $f^{\ast }(x)$, and an unstable equilibrium to a minimum.
428: 
429: In the limit $\sigma \rightarrow 0$, the stationary distribution $f^{\ast
430: }(x)$ will tend to a point-distribution, where the total probability is
431: concentrated on one point. In the case of the Hawk-Dove Game and the
432: Prisoners Dilemma it is clear that this point-distribution is concentrated
433: on the unique stable equilibrium of the corresponding discrete replicator
434: equation. For Coordination Games, there are two stable equilibria in the
435: discrete case. Correspondingly, the stationary distribution has two local
436: maximuma, namely at $x=0$ and at $x=1$. In \cite{VA} it is proved that for $%
437: \sigma \rightarrow 0$ one of these two maxima will eventually dominate. The
438: equilibrium that finally emerges is the one which has the largest basin of
439: attraction in the discrete case, and is known in the game-theory literature
440: as the risk-dominant equilibrium.
441: 
442: \subsection{Continuous replicator dynamics}
443: 
444: The two strategy set of $2\times 2$ games can be extended to a continuum of
445: strategies, each of which indicated by a real number $x\,\epsilon \,[0,1].$
446: For the payoff function $M(x,x^{\prime })$ we choose a simple interpolation
447: between the four payoff values of the symmetric $2\times 2$ game: 
448: \begin{equation}
449: M(x,x^{\prime })=xx^{\prime }a+x(1-x^{\prime })b+(1-x)x^{\prime
450: }c+(1-x)(1-x^{\prime })d.  \label{mpay}
451: \end{equation}
452: 
453: This game can be considered as the underlying discrete $2\times 2$ game
454: where now mixed strategies are allowed, in the following sense. A strategy $%
455: x\epsilon \lbrack 0,1]$ means that the player will use pure strategy $I$
456: with probability $x$. We now assume that two players, using strategies $x$
457: and $x^{\prime }$ respectively, at one encounter play each other a large
458: number of times. Eq.(\ref{mpay}) then gives the expectation value of the
459: payoff to the first player. Because our players will soon become aware of
460: the Law of Large Numbers, they won't bother with playing against each other,
461: but at an encounter simply settle for the payoff given by Eq.(\ref{mpay}),
462: making this a deterministic game.
463: 
464: The expressions for $\Pi (x,P)$ and $\overline{\Pi }(P)$ are easy to
465: calculate: 
466: \begin{eqnarray*}
467: \Pi (x,P) &=&\int_{0}^{1}M(x,x^{\prime })P(x^{\prime },t)\,\mathrm{d}x \\
468: &=&\int_{0}^{1}(xx^{\prime }a+x(1-x^{\prime })b+(1-x)x^{\prime
469: }c+(1-x)(1-x^{\prime })d)P(x^{\prime },t)\,\mathrm{d}x^{\prime } \\
470: &=&d+(b-d)x+(c-d)\overline{x}(t)+(a-b-c+d)x\overline{x}(t)
471: \end{eqnarray*}%
472: in which 
473: \begin{equation}
474: \overline{x}(t)=\int_{0}^{1}xP(x,t)\,\mathrm{d}x  \label{xav}
475: \end{equation}%
476: is the average strategy. Then 
477: \begin{eqnarray*}
478: \overline{\Pi }(P) &=&\int_{0}^{1}\Pi (x,P)P(x,t)\,\mathrm{{d}x=} \\
479: &=&d+(b+c-2d)\overline{x}(t)+(a-b-c+d)\overline{x}^{2}(t),
480: \end{eqnarray*}%
481: and 
482: \begin{eqnarray}
483: \Pi (x,P)-\overline{\Pi }(P) &=&(b-d)x-(b-d)\overline{x}(t)+(a-b-c+d)(x%
484: \overline{x}(t)-\overline{x}^{2}(t))  \notag \\
485: &=&(B+(A-B)\overline{x}(t))(x-\overline{x}(t)),  \label{reac}
486: \end{eqnarray}%
487: where $A$ and $B$ are the same parameters as defined in section 3.1.
488: 
489: Eq.(\ref{rep}) for $P(x,t)$ now becomes: 
490: \begin{equation}
491: \frac{\partial P(x,t)}{\partial t}=(B+(A-B)\overline{x}(t))(x-\overline{x}%
492: (t))P(x,t)+\sigma \frac{\partial ^{2}P(x,t)}{\partial x^{2}},  \label{full}
493: \end{equation}
494: with boundary conditions: 
495: \begin{equation}
496: \left( \frac{\partial P(x,t)}{\partial x}\right) _{x=0}=\left( \frac{%
497: \partial P(x,t)}{\partial x}\right) _{x=1}=0,  \label{bcpd}
498: \end{equation}
499: and an intial distribution $P(x,0)$.
500: 
501: \subsubsection{The equation without mutation}
502: 
503: As was noted before, when considering Eq.(\ref{full}) without mutation, i.e.
504: with $\sigma =0$, it is not necessary to impose the boundary conditions (\ref%
505: {bcpd}) in order to ensure that $\int_{0}^{1}P(x,t)\,$d$x$ remain constant.
506: In \cite{CR} it is shown that 
507: \begin{equation}
508: \frac{\partial P(x,t)}{\partial t}=(B+(A-B)\overline{x}(t))(x-\overline{x}%
509: (t))P(x,t),\;\;\;P(x,0)=P_{0}(x)  \label{geenmut1d}
510: \end{equation}%
511: has a unique solution for all $t>0$. In this subsection we will analyse the
512: asymptotic behaviour of the solution of Eq.(\ref{geenmut1d}) as $%
513: t\rightarrow \infty $.
514: 
515: Firstly, the equation for the average $\overline{x}(t)$ is given by: 
516: \begin{equation}
517: \frac{d\overline{x}}{dt}=(B+(A-B)\overline{x}(t))(\overline{x^{2}}(t)-(%
518: \overline{x}(t))^{2}),\;\;\;x(0)=\int_{0}^{1}xP_{0}(x)\mathrm{d}x,
519: \label{xstreep}
520: \end{equation}
521: where $\overline{x^{2}}(t)=\int_{0}^{1}x^{2}P(x,t)$d$x$. The factor $%
522: \overline{x^{2}}(t)-(\overline{x}(t))^{2}$ is always positive. For the four
523: different regions of the $(A,B)$ parameter plane, as illustrated in figure
524: 1, Eq.(\ref{xstreep}) implies the following.
525: 
526: \begin{description}
527: \item $(A,B)\in I$: in this case $(B+(A-B)\overline{x}(t))>0$ for all $t$,
528: so $\overline{x}(t)$ is an increasing function. Because $P(x,t)=0$ for $%
529: x\notin \lbrack 0,1]$, the distribution will accumulate at $x=1$ and $%
530: \lim_{t\rightarrow \infty }\overline{x}(t)=1$.
531: 
532: \item $(A,B)\in II$: Eq.(\ref{xstreep}) now has an attractive fixed point at 
533: $\overline{x}=B/(B-A)$. Therefore, $\lim_{t\rightarrow \infty }\overline{x}%
534: (t)=B/(B-A)$
535: 
536: \item $(A,B)\in III$: similar to $I$, but now $\lim_{t\rightarrow \infty }%
537: \overline{x}(t)=0$.
538: 
539: \item $(A,B)\in IV:$Eq.(\ref{xstreep}) has a repelling fixed point at $%
540: \overline{x}=B/(B-A)$. It follows that $\lim_{t\rightarrow \infty }\overline{%
541: x}(t)=0$ if $\overline{x}(0)<B/(B-A)$ and $\lim_{t\rightarrow \infty }%
542: \overline{x}(t)=1$ if $\overline{x}(0)>B/(B-A).$
543: \end{description}
544: 
545: \bigskip
546: 
547: In the cases $I$, $III$ and $IV$ the limiting distributions $P_{\infty
548: }(x)=\lim_{t\rightarrow \infty }P(x,t)$ are point-distributions concentrated
549: on one of the endpoints of the domain $[0,1]$. This is comparable with the
550: results in the model described in section 3.2 when $\sigma \rightarrow 0$:
551: eventually all players will play either one or the other of the pure
552: strategies. The situation is quite different for case $II$, i.e. for
553: Hawk-Dove type games. We will show that in this case, $P_{\infty }(x)$
554: depends on the initial condition $P_{0}(x)$, but is in general non-zero on
555: all of $[0,1]$. This is a dramatic difference with the above mentioned
556: model. There, the population is divided into a fraction $B/(B-A)$ of the
557: population who play pure strategy $I$ and the rest who play strategy $II$.
558: In our model, where the players have access to a continuum of mixed
559: strategies, we do not find that everybody plays the mixed strategy $%
560: x=B/(B-A) $ (play strategy $I$ with probability $B/(B-A)$), as might be
561: expected. Rather, the final outcome is a population who's members play a
562: broad range of strategies, although the average value of the strategies
563: played is $\overline{x}=B/(B-A)$. The details are as follows.
564: 
565: \bigskip Every distribution $P(x)$ with $\int_{0}^{1}xP(x)\, \mathrm{d}x=%
566: \overline{x}=B/(B-A)$ is a solution of Eq.(\ref{geenmut1d}). We will now
567: show that this set of distributions is attractive.\newline
568: Let $\lambda (t)=(B+(A-B)\overline{x}(t))$. The solution of Eq.(\ref%
569: {geenmut1d}) then is 
570: \begin{equation}
571: P(x,t)=P_{0}(x)\exp (\int_{0}^{t}\lambda (t^{\prime })(x-\overline{x}%
572: (t^{\prime }))\,\mathrm{{d}t.}  \label{solgeenmut}
573: \end{equation}
574: Using $\int_{0}^{1}P(x,t)$d$x=1$ and writing $g(t)=\int_{0}^{t}\lambda
575: (t^{\prime })$d$t$, it follows that 
576: \begin{equation}
577: \exp (\int_{0}^{t}\lambda (t^{\prime })\overline{x}(t^{\prime
578: })dt)=\int_{0}^{1}P_{0}(x)\,e^{xg(t)\,\,}\mathrm{{d}x.}  \label{lambda}
579: \end{equation}
580: Therefore 
581: \begin{equation}
582: P(x,t)=\frac{P_{0}(x)\,e^{xg(t)}}{\int_{0}^{1}P_{0}(x)\,e^{xg(t)}\,\mathrm{{d%
583: }x}}.  \label{solgeenmut2}
584: \end{equation}
585: Differentiating relation (\ref{lambda}) with respect to $t$ yields 
586: \begin{equation*}
587: \lambda (t)\,\overline{x}(t)\exp (\int_{0}^{t}\lambda (t^{\prime })\overline{%
588: x}(t^{\prime })\,\mathrm{d}t)=g^{\prime
589: }(t)\int_{0}^{1}xP_{0}(x)\,e^{xg(t)}\,\mathrm{d}x.
590: \end{equation*}
591: Using $g^{\prime }(t)=\lambda (t)$, and Eq.(\ref{lambda}), this reduces to 
592: \begin{equation}
593: \,\overline{x}(t)=\frac{\int_{0}^{1}xP_{0}(x)\,e^{xg(t)}\,\mathrm{d}x}{%
594: \int_{0}^{1}P_{0}(x)\,e^{xg(t)}\,\mathrm{d}x}.  \label{solxstreep}
595: \end{equation}
596: Since $\lim_{t\rightarrow \infty }\overline{x}(t)=B/(B-A)$, then $%
597: g(t)=\int_{0}^{t}(B+(A-B)\overline{x}(t^{\prime }))\,$d$t^{\prime }$ either
598: tends to a finite limit or diverges. Using the change of variable $u=x\,g(t)$%
599: , we can write:
600: 
601: \begin{equation*}
602: \,\overline{x}(t)=\frac{1}{g(t)}\frac{\int_{0}^{g(t)}uP_{0}(\frac{u}{g(t)}%
603: )\,e^{u}\,\mathrm{{d}u}}{\int_{0}^{g(t)}P_{0}(\frac{u}{g(t)})\,e^{u}\,%
604: \mathrm{{d}u}}.
605: \end{equation*}
606: Therefore, when $g(t)\rightarrow \infty $, then 
607: \begin{equation*}
608: \,\overline{x}(t)=\lim_{\varepsilon \downarrow 0}\frac{\varepsilon
609: \int_{0}^{1/\varepsilon }uP_{0}(\varepsilon u)\,e^{u}\,\mathrm{d}u}{%
610: \int_{0}^{1/\varepsilon }P_{0}(\varepsilon u)\,e^{u}\,\mathrm{d}u}%
611: =\lim_{\varepsilon \downarrow 0}\frac{\varepsilon \int_{0}^{1/\varepsilon
612: }u\,e^{u}\,\mathrm{d}u}{\int_{0}^{1/\varepsilon }\,e^{u}\,\mathrm{d}u}=1.
613: \end{equation*}
614: By a similar argument, when $g(t)\rightarrow -\infty $ then $\overline{x}%
615: (t)\rightarrow 0$. Since $\lim_{t\rightarrow \infty }\overline{x}(t)=B/(B-A)$
616: is neither $0$ nor $1$, it follows that $\lim_{t\rightarrow \infty
617: }g(t)=g_{0}$ is finite. Therefore, as $t\rightarrow \infty $ the solution
618: Eq.(\ref{solgeenmut2}) tends to 
619: \begin{equation*}
620: P_{\infty }(x)=\frac{P_{0}(x)\,e^{xg_{0}}}{\int_{0}^{1}P_{0}(x)\,e^{xg_{0}}\,%
621: \mathrm{d}x},
622: \end{equation*}
623: which is clearly not a point-distribution.
624: 
625: The value of $g_{0}$ can be found from Eq.(\ref{solxstreep}), which in the
626: limit $t\rightarrow \infty $ reads: 
627: \begin{equation*}
628: \frac{B}{B-A}=\frac{\int_{0}^{1}xP_{0}(x)\,e^{xg_{0}}\,\mathrm{d}x}{%
629: \int_{0}^{1}P_{0}(x)\,e^{xg_{0}}\,\mathrm{d}x}.
630: \end{equation*}%
631: This equation for $g_{0}$ can be solved numerically for given values of $A$
632: and $B$ and a given initial distribution $P_{0}(x)$. In figure 3, two examples,
633: both with $A=-3$ and $B=7,$ are given for different initial distributions,
634: shown in the left column. In the right column the final distributions are
635: plotted.
636: 
637: \begin{figure}[hbt]
638: \centering
639: \includegraphics[width=8cm]{f3.eps}
640: \caption{Initial and final distributions
641: without mutations}
642: \end{figure}
643: 
644: 
645: 
646: \subsubsection{Stationary solutions of the full equation}
647: 
648: Numerical experiments show that all solutions of the full equation (\ref%
649: {full}) converge to a time-independent solution. The equation for these
650: stationary solutions is given by: 
651: \begin{equation}
652: \sigma \frac{d^{2}P(x)}{dx^{2}}+(B+(A-B)\overline{x})(x-\overline{x})P(x)=0,
653: \label{fullstat}
654: \end{equation}
655: \begin{equation}
656: P^{\prime }(0)=P^{\prime }(1)=0,\qquad P(x)\geq 0,\qquad \int_{0}^{1}P(x)\,%
657: \mathrm{{d}}x=1. \label{statbound}
658: \end{equation}
659: Rather than giving a definition of $\overline{x}$ in terms of $P(x)$, we
660: take $\overline{x}$ to be a free parameter and impose the conditions (\ref%
661: {statbound}). Integrating Eq. (\ref{fullstat}) over $x$ from $0$ to $1$ then
662: yields: 
663: \begin{equation}
664: (B+(A-B)\overline{x})(\int_{0}^{1}xP(x)\,\mathrm{d}x-\overline{x})=0.
665: \end{equation}
666: Assuming that $B+(A-B)\overline{x}\neq 0$, the equality 
667: \begin{equation}
668: \overline{x}=\int_{0}^{1}xP(x)\,\mathrm{d}x.
669: \end{equation}
670: then follows automatically.
671: 
672: Let $s(\overline{x})=sign(B+(A-B)\overline{x})$ and $\kappa (\overline{x})=|%
673: \frac{B+(A-B)\overline{x}}{\sigma }|^{1/3}$. Then the solution of Eq. (\ref%
674: {fullstat}) is given by: 
675: \begin{equation}
676: P(x)=a\,Ai[-s(\overline{x})\kappa (\overline{x})(x-\overline{x})]+b\,Bi[-s(%
677: \overline{x})\kappa (\overline{x})(x-\overline{x})],  \label{solgen}
678: \end{equation}%
679: where the Airy functions $Ai(z)$ and $Bi(z)$ are the standard linearly
680: independent solutions of $y^{\prime \prime }(z)-zy(z)=0$. These functions
681: are plotted in figure 4.
682: 
683: \begin{figure}[hbt]
684: \centering
685: \includegraphics[width=7cm]{f4.eps}
686: \caption{Airy
687: functions $Ai(z)$ and $Bi(z)$}
688: \end{figure}
689: 
690: The position $%
691: -\eta _{0}$ of the first maximum of $Ai(z)$ is indicated by a vertical line
692: segment. The curve for $Bi(z)$ is dashed. Imposing the boundary conditions $%
693: P^{\prime }(0)=P^{\prime }(1)=0$ gives 
694: \begin{equation*}
695: a\,Ai^{\prime }[s(\overline{x})\kappa (\overline{x})\overline{x}%
696: ]+b\,Bi^{\prime }[s(\overline{x})\kappa (\overline{x})\overline{x}]=0
697: \end{equation*}%
698: \begin{equation*}
699: a\,Ai^{\prime }[s(\overline{x})\kappa (\overline{x})(\overline{x}%
700: -1)]+b\,Bi^{\prime }[s(\overline{x})\kappa (\overline{x})(\overline{x}-1)]=0,
701: \end{equation*}%
702: so that a non-trivial solution only exists if $\overline{x}$ is a solution
703: to the \textquotedblright eigenvalue equation\textquotedblright : 
704: \begin{equation}
705: Ai^{\prime }[s(\overline{x})\kappa (\overline{x})\overline{x}]Bi^{\prime }[s(%
706: \overline{x})\kappa (\overline{x})(\overline{x}-1)]-Bi^{\prime }[s(\overline{%
707: x})\kappa (\overline{x})\overline{x}]Ai^{\prime }[s(\overline{x})\kappa (%
708: \overline{x})(\overline{x}-1)]=0.  \label{eigenval}
709: \end{equation}%
710: Corresponding to a solution $\overline{x}$ of Eq. (\ref{eigenval}), we find
711: that the solution (\ref{solgen}) can be written as: 
712: \begin{gather}
713: P(x)=  \label{solspec} \\
714: =c(Bi^{\prime }[s(\overline{x})\kappa (\overline{x})\overline{x}]Ai[s(%
715: \overline{x})\kappa (\overline{x})(\overline{x}-x)]-Ai^{\prime }[s(\overline{%
716: x})\kappa (\overline{x})\overline{x}]Bi[s(\overline{x})\kappa (\overline{x})(%
717: \overline{x}-x)]),  \notag
718: \end{gather}%
719: in which $c$ is determined by the normalisation condition. We found that,
720: although the eigenvalue equation (\ref{eigenval}) can have many solutions,
721: there will be at most three that correspond to a distribution (\ref{solspec}%
722: ) with the property that $P(x)\geq 0$ for all $x\in \lbrack 0,1]$. For
723: values of A and B in the regions I and II of figure 1, typified by PD- and
724: HD-games, there is only one solution. The distributions calculated for $%
725: \sigma =0.001,$ are shown in figure 5. 
726: 
727: \begin{figure}[hbt]
728: \centering
729: \includegraphics[width=12cm]{f5.eps}
730: \caption{Stationary distributions for PD- and HD-games}
731: \end{figure}
732: 
733: 
734: For the PD-case
735: the parameters are $A=1,$ $B=2$ and $\overline{x}=0.901.$ For the HD-case
736: they are $A=-2,$ $B=1$ and $\overline{x}=0.342.$ For (A,B)$\,\epsilon \,IV$, 
737: \textit{i.e.,} for Coordination Games, there exist three solutions for this
738: value of $\sigma $. The corresponding distributions are plotted in one
739: picture (figure 6).
740: 
741: \begin{figure}[hbt]
742: \centering
743: \includegraphics[width=7cm]{f6.eps}
744: \caption{Stationary
745: distributions for CG-game}
746: \end{figure}
747: 
748: With $A=2$ and $%
749: B=-1$ we find for the average strategies $\overline{x}=0.118,$ $\overline{x}%
750: =0.323$ and $\overline{x}=0.915.$
751: 
752: We now study the behaviour of the solution of (\ref{fullstat}) and (\ref%
753: {statbound}) for $\sigma \rightarrow 0$. First consider the case that $%
754: |B+(A-B)\overline{x}|>const$, independent of $\sigma $. Then $\kappa (%
755: \overline{x})=|\frac{B+(A-B)\overline{x}}{\sigma }|^{1/3}\rightarrow \infty $
756: as $\sigma \rightarrow 0$. From figure 4 it is clear that if $s(\overline{x}%
757: )=sign(B+(A-B)\overline{x})=1$, then $Bi^{\prime }[s(\overline{x})\kappa (%
758: \overline{x})\overline{x}]\rightarrow \infty $, $Bi^{\prime }[s(\overline{x}%
759: )\kappa (\overline{x})(\overline{x}-1)]$ remains bounded and $Ai^{\prime }[s(%
760: \overline{x})\kappa (\overline{x})\overline{x}]\rightarrow 0$, as $\sigma
761: \rightarrow 0$. This implies that $\overline{x}$ can only be a solution of
762: Eq.(\ref{eigenval}) if $Ai^{\prime }[s(\overline{x})\kappa (\overline{x})(%
763: \overline{x}-1)]\rightarrow 0$. From figure 4, it then follows that $\kappa (%
764: \overline{x})(\overline{x}-1)\rightarrow -\eta _{0}$, where $\eta
765: _{0}=1.01879\cdots $ is the smallest positive solution of $Ai^{\prime
766: }(-\eta _{0})=0,$ so that $\overline{x}\rightarrow 1$ and $\kappa (\overline{%
767: x})\rightarrow |\frac{A}{\sigma }|^{1/3}$. More precisely, we have: 
768: \begin{equation*}
769: \overline{x}=1-|\frac{\sigma }{A}|^{1/3}\,\eta _{0}\qquad as\quad \sigma
770: \rightarrow 0
771: \end{equation*}%
772: This solution is consistent with the assumption $s(\overline{x})=1$ if and
773: only if $A>0$, as in the Prisoners Dilemma and the Cooperation Game. The
774: corresponding asymptotic expression for $P(x)$ becomes: 
775: \begin{equation*}
776: P(x)=c\,Ai[-\kappa (\overline{x})(x-\overline{x})]=c\,Ai[(\frac{A}{\sigma }%
777: )^{1/3}(1-x)-\eta _{0}].
778: \end{equation*}%
779: Note that for $\sigma \rightarrow 0$ this distribution becomes sharply
780: peaked at $x=1$, with the width of the peak proportional to $\sigma ^{1/3}$.
781: 
782: By a similar reasoning, it is found that for $B<0$ (as in the Cooperation
783: Game) there exists a solution such that 
784: \begin{equation*}
785: \overline{x}=\,|\frac{\sigma }{B}|^{1/3}\,\eta _{0}\qquad \text{and}\quad
786: P(x)=c\,Ai[|\frac{B}{\sigma }|^{1/3}x-\eta _{0}]
787: \end{equation*}%
788: as $\sigma \rightarrow 0$.
789: 
790: In the previous section we found that for $\sigma =0$ the Hawk-Dove Game and
791: the Cooperation Game have solutions with $\overline{x}=\frac{B}{B-A}$, which
792: we call a central solution. This motivates us to look for solutions for
793: which $\kappa (\overline{x})=|\frac{B+(A-B)\overline{x}}{\sigma }|^{1/3}$
794: remains bounded as as $\sigma \rightarrow 0$. We therefore assume that 
795: \begin{equation}
796: \overline{x}=\frac{B}{B-A}+\frac{\alpha }{B-A}\sigma +...\qquad as\quad
797: \sigma \rightarrow 0,  \label{asxtreep}
798: \end{equation}%
799: where $\alpha \in R$ is as yet unknown. Substituting Eq.(\ref{asxtreep})
800: into Eq.(\ref{eigenval}) and taking the limit $\sigma \rightarrow 0$ yields: 
801: \begin{equation}
802: Ai^{\prime }[\frac{B}{B-A}\alpha ^{1/3}]Bi^{\prime }[\frac{A}{B-A}\alpha
803: ^{1/3}]-Bi^{\prime }[\frac{B}{B-A}\alpha ^{1/3}]Ai^{\prime }[\frac{A}{B-A}%
804: \alpha ^{1/3}]=0,  \label{alpha}
805: \end{equation}%
806: where $\alpha ^{1/3}$ is understood to mean $sign(\alpha )|\alpha |^{1/3}$.
807: This is the equation from which $\alpha $ must be solved. Although Eq.(\ref%
808: {alpha}) has many zeroes, only one corresponds to a positive distribution
809: given by: 
810: \begin{equation}
811: P(x)=c(Bi^{\prime }[\beta \alpha ^{1/3}]Ai[\alpha ^{1/3}(x-\beta
812: )]-Ai^{\prime }[\beta \alpha ^{1/3}]Bi[\alpha ^{1/3}(x-\beta )]),
813: \label{palpha}
814: \end{equation}%
815: with $\beta =\frac{B}{B-A}.$ For $A=-2$ and $B=1,$ (a Hawk-Dove game), it is
816: plotted in figure 7.
817: 
818: \begin{figure}[hbt]
819: \centering
820: \includegraphics[width=6cm]{f7.eps}
821: \caption{Asymptotic
822: HD-distribution for $\protect\sigma \rightarrow 0$}
823: \end{figure}
824: 
825: 
826: We note that Eqs. (\ref{alpha}) and (\ref{palpha}) are invariant under $%
827: A\rightarrow -A$, $B\rightarrow -B$. Therefore, as $\sigma \rightarrow 0$,
828: we find the same central solution for the Coordination Game with $A=2$ and $%
829: B=-1$. From numerical simulations we find, however, that the central
830: solution is stable in the Hawk-Dove Game, but unstable in the Coordination
831: Game.
832: 
833: \bigskip
834: 
835: Away from the limit $\sigma \rightarrow 0$, we can track the fate of the
836: central solution as $\sigma $ grows. Since the full eigenvalue-equation (\ref%
837: {eigenval}) is not invariant under $A\rightarrow -A$, $B\rightarrow -B$, the
838: situation is different for Hawk-Dove Games as opposed to Coordination Games.
839: For Hawk-Dove Games, we find that the central solution persists for all
840: values of the diffusion coefficient $\sigma $. However, for Coordination
841: Games we find that above a critical value $\sigma _{c}$ of $\sigma $, the
842: unstable central solution disappears, together with one of the two stable
843: solutions, leaving only one attracting, stationary solution. When $A<|B|$,
844: the solution at $x=1$ remains, and when $A>|B|$, the solution near $x=0$
845: survives. This can be summarised by saying that for large enough values of $%
846: \sigma $, the only attracting solution is a stationary solution near the
847: risk-dominated solution of the discrete equation. The bifurcation process is
848: illustrated in figure 8,where for two values of $\sigma $ the left-hand-side of Eq.(\ref{eigenval}) is plotted as a
849: function of $\overline{x}(\sigma ).$ For $\sigma =0.0035$ this function has
850: three zeroes, whereas for $\sigma =0.0039$ there is only one$.$ This closely
851: resembles the situation in the discrete case as described in section 2.
852: However, the critical values for $\sigma $ in the discrete case and in the
853: continuous case are not comparable ($0.11$ vs. $0.0037$).
854: 
855: 
856: \begin{figure}[hbt]
857: \centering
858: \includegraphics[width=8cm]{f8.eps}
859: \caption{Bifurcation of CG-games}
860: \end{figure}
861: 
862: 
863: \subsubsection{Summary}
864: 
865: Here we summarise and compare the results of the three types of games, using
866: the discrete model, the discrete model with a stochastic term and the
867: continuous model with deterministic mutation.
868: 
869: The Prisoners Dilemma is straightforward. In the discrete case the
870: population will eventually play All Defect. This does not change when the
871: strategy-space is made continuous and adding deterministic mutation simply
872: changes the limiting delta-distribution to a finite peak with width
873: proportional to $\sigma ^{1/3}$. The contrast with the corresponding
874: stochastic equation is that there the width of the peak is narrower and
875: proportional to $\sigma ^{1/2}$.
876: 
877: In the discrete version of the Coordination Game, the population will
878: eventually play either $x_{1}=0$ or $x_{1}=1$, where the final outcome
879: depends on the intial condition. If the mutation rate is larger than a
880: certain threshhold, only the stable solution around the risk-dominant
881: solution remains. This result remains the same in the continuous case.
882: Depending on the initial distribution of strategies, the final distribution
883: will be sharply peaked (width of peak proportional to $\sigma ^{1/3}$)
884: around either $x=0$ or $x=1$. When the mutation rate becomes larger than a
885: critical value, only the distribution around the risk-dominant solution
886: remains. We note that this critical value is much smaller in the continuous
887: case than in the discrete case. The stochastic equation has only one,
888: attracting, stationary solution. When the mutation rate $\sigma $ is small,
889: this distribution is almost completely concentrated around the risk dominant
890: solution, where again the width of the distribution is proportional to $%
891: \sigma ^{1/2}$.
892: 
893: The Hawk-Dove Game shows the following behaviour. In the discrete case,
894: there is an asymptotically stable solution for all values of $\sigma >0$,
895: which for $\sigma =0$ has the value $x_{1}=B/(B-A)$. The stochastic equation
896: has a unique attracting stationary solution, peaked around $x=B/(B-A)$ and
897: with a width proportional to $\sigma ^{1/2}$. Without mutation, $\sigma =0$,
898: the continuous equation has an asymptotically stable invariant set of
899: solutions, consisting of all distributions $P(x)$ with average $\overline{x}%
900: =B/(B-A)$. Depending on the initial distribution, the solution converges to
901: a member of this set. When the mutation term is added, $\sigma >0$, only
902: one, attracting, solution remains. In the limit $\sigma \rightarrow 0$, this
903: solution converges to (\ref{palpha}). This solution is not concentrated on a
904: single point, but has the whole interval $[0,1]$ as support. Such a solution
905: is sometimes referred to as polymorphic. For small $\sigma $, numerical
906: experiments show that a starting distribution initially converges to a
907: distribution close to the solution it would reach if $\sigma =0$, but then
908: slowly (on a time-scale of $1/\sigma $) evolves to the unique limiting
909: solution. This behaviour resembles that of a singulary perturbed ordinary
910: differential equation, which in the unperturbed case has an attracting set
911: of fixed points, and where after adding the perturbation, the attracting
912: invariant set survives. In this invariant set we would have, in this
913: analogy, one attracting fixed point left.
914: 
915: \section{The Ultimatum Game}
916: 
917: This game has attracted a great deal of interest, mainly as a model to
918: explain the occurrence of strong reciprocity in populations of supposedly
919: selfish individuals.(Binmore \cite{BI}, Fehr \& Gachter \cite{FG}). Here
920: "strong reciprocity" means the willingness to share, but also to punish
921: egotistical behaviour in others, even at a cost to oneself. See Bowles \&
922: Gintis \cite{BG}.
923: 
924: The game is played by two players. The first player is given a certain
925: amount of money and proposes a split of this money with the second player.
926: This second player has the choice between accepting the offer, or rejecting
927: it, in which case neither of the two players will receive anything. An
928: obvious strategy for the second player is to accept every offer, since
929: something is better than nothing. Realising this, the first player will
930: maximise his share by offering the lowest possible amount to the second
931: player. This (combined) strategy is sometimes referred to as
932: \textquotedblright the rational solution\textquotedblright\ (Page \& Nowak 
933: \cite{PN}), \textquotedblright the subgame-perfect
934: equilibrium\textquotedblright\ (Seymour \cite{SE}) or the strategy of \
935: \textquotedblright Homo Economicus\textquotedblright\ (Bowles \& Gintis \cite%
936: {BO}).
937: 
938: An evolutionary version of this game, taking into account mutations, was
939: studied in \cite{PN} and by Nowak et al. \cite{NPS}, using adaptive
940: dynamics. In adaptive dynamics models, the population is assumed to always
941: be monomorphic, \textit{i.e.}, everybody plays the same strategy. Every now
942: and then, a mutant is introduced. If the strategy of the mutant is more
943: succesful than the resident strategy, it will quickly spread in the
944: population, thus becoming the new resident strategy. For the Ultimatum Game
945: it was found that in the absence of any restrictions, the solution of the
946: adaptive dynamics model indeed converged to the \textquotedblright rational
947: solution\textquotedblright .
948: 
949: The Ultimatum Game was also studied by Seymour \cite{SE}, using replicator
950: dynamics. He included mutations as a given, exogeneous term and found that
951: other solutions can emerge, far from the ''subgame-perfect solution'',
952: depending on the form and intensity of the mutation function.
953: 
954: \bigskip
955: 
956: We model the Ultimatum Game as follows. The strategy space is $S=[0,1]\times
957: \lbrack 0,1]$, where a stategy $s=(x,y)\in S$ means that the player, in the
958: role of nr. 1, will offer a fraction $x$, while in the role of nr. 2 he will
959: reject any offer lower than $y$.
960: 
961: In one round, the players will play the role of nr. 1 and nr. 2 alternately.
962: This leads to a payoff function giving the payoff to strategy $s$ when
963: playing strategy $s^{\prime }$ (the factor $1/2$ has been omitted): 
964: \begin{equation}
965: M(s,s^{\prime })=M(x,y|x^{\prime },y^{\prime })=(1-x)\Theta (x-y^{\prime
966: })+x^{\prime }\Theta (x^{\prime }-y),  \label{mug}
967: \end{equation}%
968: in which the Heaviside function is defined by 
969: \begin{equation}
970: \Theta (z)=\left\{ 
971: \begin{array}{c}
972: 1\quad \text{if}\mathrm{\quad }z \geq 0 \\ 
973: 0\quad \text{if}\mathrm{\quad }z<0%
974: \end{array}%
975: \right.  \label{th}
976: \end{equation}
977: 
978: Before writing down the full replicator equation (\ref{rep}) for this case
979: we first introduce a number of abbreviations: 
980: \begin{equation}
981: H(x,t)=\int_{0}^{1}P(x,y,t)\,\mathrm{d}y\quad \text{and}\quad
982: V(y,t)=\int_{0}^{1}P(x,y,t)\,\mathrm{d}x,  \label{HV}
983: \end{equation}%
984: which are normalised as 
985: \begin{equation}
986: \int_{0}^{1}H(x,t)\,\mathrm{d}x=\int_{0}^{1}V(y,t)\,\mathrm{d}y=1.
987: \label{HVN}
988: \end{equation}%
989: $Q(x,t)$ and $R(y,t)$ are defined by 
990: \begin{equation}
991: Q(x,t)=\int_{0}^{x}V(y,t)\,\mathrm{d}y\quad \text{and}\quad
992: R(y,t)=\int_{y}^{1}xH(x,t)\,\mathrm{d}x,  \label{QR}
993: \end{equation}%
994: so that 
995: \begin{equation}
996: Q(0,t)=0\quad \text{and}\quad Q(1,t)=1\quad \text{and}\quad R(1,t)=0.
997: \label{QR1}
998: \end{equation}%
999: In terms of these functions the local and global averages take the form 
1000: \begin{equation}
1001: \Pi (x,y,P)=(1-x)Q(x,t)+R(y,t)  \label{UGp}
1002: \end{equation}%
1003: and 
1004: \begin{equation}
1005: \overline{\Pi }(P)=c_{1}(t)+c_{2}(t),  \label{pcc}
1006: \end{equation}%
1007: with 
1008: \begin{equation}
1009: c_{1}(t)=\int_{0}^{1}(1-x)Q(x,t)H(x,t)\,\mathrm{d}x\quad \quad
1010: c_{2}(t)=\int_{0}^{1}R(y,t)V(y,t)\,\mathrm{d}y.  \label{cc}
1011: \end{equation}%
1012: At last the replicator equation becomes 
1013: \begin{eqnarray}
1014: \frac{\partial P(x,y,t)}{\partial t}
1015: &=&[(1-x)Q(x,t)+R(y,t)-c_{1}(t)-c_{2}(t)]P(x,y,t)+  \label{UGrep} \\
1016: &&+\sigma \Delta P(x,y,t),  \notag
1017: \end{eqnarray}%
1018: in which $\Delta $ is the two-dimensional Laplace operator. The boundary
1019: condition is 
1020: \begin{equation}
1021: \nabla P\cdot \mathbf{n}=0\quad \text{on\ the\ boundary.}  \label{bnd}
1022: \end{equation}%
1023: In what follows we will restrict ourselves to solutions which can be written
1024: as the product of two normalised functions of $(x,t)$ and of $(y,t)$
1025: respectively. It \ then necessarily follows that 
1026: \begin{equation}
1027: P(x,y,t)=H(x,t)V(y,t).  \label{phv}
1028: \end{equation}%
1029: With this restriction we easily show, by integrating Eq.(\ref{UGrep}) over $%
1030: y $, that 
1031: \begin{equation}
1032: \frac{\partial H(x,t)}{\partial t}=[(1-x)Q(x,t)-c_{1}(t)]H(x,t)+\sigma \frac{%
1033: \partial ^{2}H(x,t)}{\partial x^{2}}.  \label{Hrep}
1034: \end{equation}%
1035: The boundary conditions are 
1036: \begin{equation}
1037: \frac{\partial H(x=0,t)}{\partial x}=\frac{\partial H(x=1,t)}{\partial x}%
1038: =0\quad \text{for\ all}\mathrm{\;}t.  \label{Hbnd}
1039: \end{equation}%
1040: Integration over $x$ of Eq.(\ref{UGrep}) leads to 
1041: \begin{equation}
1042: \frac{\partial V(y,t)}{\partial t}=[R(y,t)-c_{2}(t)]V(y,t)+\sigma \frac{%
1043: \partial ^{2}V(y,t)}{\partial y^{2}},  \label{Vrep}
1044: \end{equation}%
1045: with boundary conditions 
1046: \begin{equation}
1047: \frac{\partial V(y=0,t)}{\partial y}=\frac{\partial V(y=1,t)}{\partial y}%
1048: =0\quad \text{for\ all}\mathrm{\;}t.  \label{Vbnd}
1049: \end{equation}%
1050: We note that for $\sigma =0$ these equations are the same as those studied
1051: in \cite{SE}, where it is assumed that there are two separate populations of
1052: players, one where all members always play the role of nr. 1 and the other
1053: with nr. 2 players.
1054: 
1055: \subsection{The equation without mutation}
1056: 
1057: For $\sigma =0$ the equations become: 
1058: \begin{eqnarray}
1059: \frac{\partial H(x,t)}{\partial t} &=&[(1-x)Q(x,t)-c_{1}(t)]H(x,t)  \notag \\
1060: \frac{\partial V(y,t)}{\partial t} &=&[R(y,t)-c_{2}(t)]V(y,t).  \label{nomut}
1061: \end{eqnarray}%
1062: Similar to the Hawk-Dove Game, we identify a set of stationary solutions: 
1063: \begin{eqnarray}
1064: H_{0}(x) &=&\delta (x-\overline{x})  \notag \\
1065: V_{0}(y) &=&\left\{ 
1066: \begin{array}{c}
1067: v(y)\quad \text{if}\quad y<\overline{x} \\ 
1068: 0\quad \quad \text{if}\quad y>\overline{x}%
1069: \end{array}%
1070: \right.  \notag \\
1071: c_{1} &=&1-\overline{x}\quad ,\quad c_{2}=\overline{x}  \label{stat}
1072: \end{eqnarray}%
1073: where $v(y)$ is an arbitrary function with $\int_{0}^{\overline{x}}v(y)\,$d$%
1074: y=1$ and $\delta (z)$ is the Dirac-$\delta $ distribution. The solution (\ref{stat}) is easily checked, by
1075: noting that $R(y)=\overline{x}\Theta (\overline{x}-y)$ and that $Q(\overline{%
1076: x})=1$. We have also used $%
1077: z\,\delta (z)\equiv 0$. The interpretation of this solution is clear: player nr. 1 always
1078: offers $\overline{x}$, so player nr. 2 will always receive this amount, as
1079: long as his acceptence threshhold is below $\overline{x}$. The average payoff
1080: is therefore $\overline{\Pi }=c_{1}+c_{2}=1$ and any distribution of the $y$%
1081: -values below $\overline{x}$ is stationary, given this distribution of $x$.
1082: The limit $\overline{x}\rightarrow 0$ corresponds to the subgame-perfect
1083: solution.
1084: 
1085: To show that (\ref{stat}) represents an attracting set of solutions, we use
1086: the same reasoning as in section 3.3.1, and find 
1087: \begin{equation}
1088: \exp [\int_{0}^{t}c_{1}(t^{\prime })\,\mathrm{d}t^{\prime
1089: }]=\int_{0}^{1}H_{0}(x)\,\exp [(1-x)\int_{0}^{t}Q(x,t^{\prime })\,\mathrm{d}%
1090: t^{\prime }]\,\mathrm{d}x\equiv h(t)  \label{expc1}
1091: \end{equation}%
1092: \begin{equation}
1093: H(x,t)=\frac{H_{0}(x)}{h\left( t\right) }\exp
1094: [(1-x)\int_{0}^{t}Q(x,t^{\prime })\,\mathrm{d}t^{\prime }]  \label{Huitdr}
1095: \end{equation}%
1096: \begin{equation}
1097: \mathrm{\exp }[\int_{0}^{t}c_{2}(t^{\prime })\,\mathrm{d}t^{\prime
1098: }]=\int_{0}^{1}V_{0}(y)\,\exp [\int_{0}^{t}R(x,t^{\prime })\,\mathrm{d}%
1099: t^{\prime }]\,\mathrm{d}y\equiv v(t)  \label{expc2}
1100: \end{equation}%
1101: \begin{equation}
1102: V(y,t)=\frac{V_{0}(y)}{v(t)}\exp [\int_{0}^{t}R(y,t^{\prime })\,\mathrm{d}%
1103: t^{\prime }].  \label{Vuitdr}
1104: \end{equation}%
1105: By differentiating Eq. (\ref{expc1}) and (\ref{expc2}), respectively, we
1106: obtain: 
1107: \begin{equation}
1108: c_{1}(t)=\frac{1}{h(t)}\int_{0}^{1}(1-x)Q(x,t)H_{0}(x)\,\exp
1109: [(1-x)\int_{0}^{t}Q(x,t^{\prime })\,\mathrm{d}t^{\prime }]\,\mathrm{d}x
1110: \label{c1}
1111: \end{equation}%
1112: \begin{equation}
1113: c_{2}(t)=\frac{1}{v(t)}\int_{0}^{1}R(y,t)V_{0}(y)\,\exp
1114: [\int_{0}^{t}R(y,t^{\prime })\,\mathrm{d}t^{\prime }]\,\mathrm{d}y.
1115: \label{c2}
1116: \end{equation}%
1117: Assuming that $V(y,t)$ converges to a stationary distribution (as all
1118: numerical results show), then $(1-x)\int_{0}^{t}Q(x,t^{\prime })\,\mathrm{d}%
1119: t^{\prime }$ converges to a function with a finite number of isolated local
1120: maxima. One of these, say $x=\overline{x}$, is the absolute maximum, and it
1121: follows from Eq.(\ref{Huitdr}) that $H(x,t)$ converges to $\delta (x-%
1122: \overline{x})$. From Eq.(\ref{QR}) it follows that $R(y,t)$ converges to $%
1123: R(y)=\overline{x}\,\Theta (\overline{x}-y)$ and $V(y,t)$ converges to $%
1124: c\,V_{0}(y)\Theta (\overline{x}-y)$, with $c$ a normalization constant. From
1125: Eq.(\ref{cc}) it follows that $c_{1}(t)$ converges to $1-\overline{x}$ and $%
1126: c_{2}(t)$ to $\overline{x}$.\newline
1127: The above considerations show that if the solution of Eq.(\ref{nomut})
1128: converges to a stationary solution, it must be a member of the invariant set
1129: (\ref{stat}). However, the value of $\overline{x}$ cannot be predicted from
1130: the above formula's. Numerical solution of Eq.(\ref{nomut}) shows that for
1131: random initial distributions of $H(x,0)$ and $V(y,0)$ on the whole interval $%
1132: [0,1],$ the functions $H(x,t)$ and $V(y,t)$ indeed approach the form of Eq.(%
1133: \ref{stat}) for $t\rightarrow \infty $. The average strategy $\overline{x},$
1134: based on $100$ simulations, takes values between $0.12$ and $0.30,$ with a
1135: mean value equal to $0.22$ and a standard deviation of $0.04.$ We note that
1136: a uniform distribution of both $H(x,0)$ and $V(y,0)$ also leads to a value
1137: of $\overline{x}=0.22.$
1138: 
1139: \subsection{The equation with mutation}
1140: 
1141: We have found numerically that when $\sigma >0$, all solutions of the full equations (\ref{Hrep}) and (\ref{Vrep}) tend
1142: to a unique solution of the stationary equations: 
1143: \begin{equation}
1144: \sigma \frac{{d}^{2}H(x)}{{\ d}x^{2}}+[(1-x)Q(x)-c_{1}]H(x)=0  \label{Hh}
1145: \end{equation}%
1146: and 
1147: \begin{equation}
1148: \sigma \frac{{\ d}^{2}V(y)}{{\ d}y^{2}}+[R(y)-c_{2}]V(y)=0.  \label{Vh}
1149: \end{equation}%
1150: The boundary values are those of Eqs.(\ref{Hbnd}) and(\ref{Vbnd}).\newline
1151: Unfortunately, we have not been able to find closed form expressions for the
1152: solutions of these equations. There are two ways to approximate the stationary solution, which lead to the same
1153: result. First, we numerically solved the full equations(\ref{Hrep}) and (\ref{Vrep}), by
1154: discretising space, solving the resulting coupled set of ordinary
1155: differential equations and considering the solution as $t \rightarrow \infty$. In the second method, we define the following seven functions 
1156: \begin{eqnarray}
1157: F_{1}(x) &=&H(x),\quad F_{2}(x)=\frac{dH(x)}{dx},\quad F_{3}(x)=V(x),\quad
1158: F_{4}(x)=\frac{dV(x)}{dx},  \notag \\
1159: F_{5}(x) &=&Q(x),\quad F_{6}(x)=R(x),\quad F_{7}(x)=\int_{x}^{1}H(x^{\prime
1160: })\,\mathrm{{d}x^{\prime }.}  \label{f7}
1161: \end{eqnarray}%
1162: In terms of these functions and with $k=1/\sigma $ ,the stationary equations
1163: can now be written as 
1164: \begin{eqnarray}
1165: \frac{dF_{1}(x)}{dx} &=&F_{2}(x)  \notag \\
1166: \frac{dF_{2}(x)}{dx} &=&-k[(1-x)F_{5}(x)-c_{1}]F_{1}(x)  \notag \\
1167: \frac{dF_{3}(x)}{dx} &=&F_{4}(x)  \notag \\
1168: \frac{dF_{4}(x)}{dx} &=&-k[F_{6}(x)-c_{2}]F_{3}(x)  \label{feq7} \\
1169: \frac{dF_{5}(x)}{dx} &=&F_{3}(x)  \notag \\
1170: \frac{dF_{6}(x)}{dx} &=&-xF_{1}(x)  \notag \\
1171: \frac{dF_{7}(x)}{dx} &=&-F_{1}(x)  \notag
1172: \end{eqnarray}%
1173: These equation can be solved numerically by starting the integration from
1174: the following values at $x=1$%
1175: \begin{eqnarray}
1176: F_{1}(1) &=&a,\quad F_{2}(1)=0,\quad F_{3}(1)=b,\quad F_{4}(1)=0,  \notag \\
1177: F_{5}(1) &=&1,\quad F_{6}(1)=0,\quad F_{7}(1)=0,  \label{fbnd}
1178: \end{eqnarray}%
1179: and using a standard routine to arrive at the values of these functions in $%
1180: x=0.$ The numbers ($a$,$b$,$c_{1}$,$c_{2})$ are as yet unknown. They should
1181: be chosen in such a way that the boundary conditions at $x=0$ be satisfied, 
1182: \textit{i.e.,} 
1183: \begin{equation}
1184: (F_{2}(0),F_{4}(0),F_{5}(0),F_{7}(0))=(0,0,0,1).  \label{fcon}
1185: \end{equation}%
1186: This matching of four numbers by varying four other numbers should be
1187: possible in many ways. It turns out, however, that the requirement of
1188: positivity of $F_{1}(x)$ and $F_{3}(x)$ in the whole interval $[0,1]$ makes
1189: the solution unique. A root finding routine of Mathematica does the job. For 
1190: $\sigma =0.001$ the stationary solutions $H(x)$ and $V(y)$ are shown in
1191: figure 9. 
1192: 
1193: \begin{figure}[hbt]
1194: \centering
1195: \includegraphics[width=12cm]{f9.eps}
1196: \caption{Stationary solution
1197: of Eqs.(\protect\ref{Hh}) and (\protect\ref{Vh}) for $\protect\sigma =0.001$}
1198: \end{figure}
1199: 
1200: We note that $%
1201: H(x)$ has a Gauss-like distribution around a mean value $\overline{x}=0.3172$%
1202: , while $V(y)$ is approximated by the right half of a Gaussian, with its maximum
1203: at $y=0$. For smaller values of $\sigma $, the value of $%
1204: \overline{x}(\sigma )$ and the width of the peak of the $H(x)$-distribution
1205: decrease, but the shape of the distributions is otherwise unchanged.
1206: 
1207: For values of \ $\sigma $ down to $10^{-9}$ we have calculated the $\sigma $%
1208: -dependence of $\overline{x}.$ In figure 10
1209: 
1210: 
1211: \begin{figure}[hbt]
1212: \centering
1213: \includegraphics[width=6cm]{f10.eps}
1214: \caption{Log-log plot of $\overline{x}$ as function of $\protect\sigma $}
1215: \end{figure}
1216: 
1217: a log-log plot
1218: of this dependence is shown. A good fit of the data points is given by $%
1219: \overline{x}(\sigma )\approx 1.64\times \,\,\sigma ^{0.23}$.
1220: 
1221: The numerical solution of the full time dependent equations reveals a
1222: dynamical pattern similar to the Hawk-Dove Game of section 2. Initially a
1223: distribution approaches the attracting set (\ref{stat}), after which it
1224: slowly converges to the unique stationary solution, on a time-scale of $%
1225: 1/\sigma $ .
1226: 
1227: \section{Conclusions}
1228: 
1229: In this paper we have generalised the replicator dynamics of games with
1230: deterministic mutations, as described in \cite{HS1}, to the situation where
1231: players have access to a continuous strategy space. The resulting equation (%
1232: \ref{rep}) has a well defined solution, which, however, is not easy to
1233: analyse in general.
1234: 
1235: Our first example, the continuous version of $2\times 2$ symmetric games,
1236: already illustrates a number of interesting and perhaps unexpected features
1237: of this equation. Although there is no a priori reason to believe that the
1238: continuous and the discrete strategy version of the same $2\times 2$ game
1239: have anything to do with each other, the similarities between some results
1240: warrant our surprise at the differences in others. The continuous Prisoners
1241: Dilemma and the Coordination Game behave similar to their discrete case
1242: counterparts: the final state is a monomorphic distribution, where every
1243: player in the population plays one of the two pure strategies. The extension
1244: to a continuous strategy space and the inclusion of mutation (which has the
1245: form of a diffusion term) only leads to the existence of small variations
1246: around the single peak of the final distribution. Also in the continuous
1247: Coordination Game we encounter, as in the discrete case, a threshold value
1248: for the mutation term separating a regime with two attracting solutions from
1249: one where only a single attractor exists.
1250: 
1251: The difference occurs in the Hawk-Dove game. In the discrete version there
1252: is a stable equilibrium with Hawks and Doves coexisting. In the continuous
1253: version this does not translate into a monomorphic distribution around the
1254: mixed strategy corresponding to this equilibrium. Rather, in the unique
1255: limiting distribution the whole range of mixed strategies from pure Hawk to
1256: pure Dove is represented. The attraction to this stationary solution occurs
1257: on two time scales. On a fast time scale, the solution is attracted to the
1258: set of distributions with average corresponding to the equilibrium mixed
1259: strategy of the discrete case. Then, on a slow time scale proportional to
1260: the inverse of the mutation rate, the solution converges to the unique
1261: attractor. This two timescales phenomenon was observed in numerical
1262: simulations, and is currently awaiting a more thorough analysis. Also, there
1263: are many interesting games with three or more strategies (for instance Rock,
1264: Scissors, Paper) of which the continuous version can hold more surprises.%
1265: \newline
1266: \newline
1267: The results of the second example, the Ultimatum Game, are of great interest
1268: to the debate around strong reciprocity and how it could have evolved. Our
1269: model shows that replicator dynamics and a small mutation term can lead to a
1270: final outcome far from the subgame-perfect solution. Take, for instance, a
1271: mutation rate of $10^{-3}.$  So in every time-interval, all players vary their strategies
1272: according to the cold rules of self-interest, after which a small fraction of $%
1273: 0.1\%$ of the population, change their strategy just a little bit.
1274: Then we find that an initial population consisting almost entirely of
1275: cynical misers (accept everything and offer nothing), eventually turns into
1276: a world where the average offer is more than $30\%$!
1277: 
1278: This surprising result can be explained in the following way. Consider a
1279: situation where all proposers offer only a small share to their opponent and these
1280: opponents all have an acceptance threshhold lower than this offer. Now, due to
1281: mutation, some acceptors will demand a share that is slightly larger than what
1282: is being offered. Normally, this would be a suicidal strategy. However, also due
1283: to mutation, there will be amongst the proposers a small set who are willing to
1284: offer slightly more than their colleagues. 
1285: On the one hand, these fairer-minded proposers earn slightly less from the bulk
1286: of the acceptors, but on the other hand they are the only ones to profit from the small group of
1287: high-minded mutants on the acceptor-side. The net result can be that the second
1288: effect dominates and that there will be a tendency towards higher offers.
1289: 
1290: 
1291: For sufficiently small mutation rates the dynamics of the Ultimatum Game show the same structure
1292: with two timescales as the continuous Hawk-Dove game. An initial
1293: distribution is quickly attracted to a distribution where the offers are
1294: sharply peaked, and then slowly converges to the unique stationary solution.
1295: 
1296: In this case too, a more rigorous mathematical analysis is required for a
1297: better understanding of the model. In particular it would be nice to be able
1298: to calculate the value of the exponent in the formula relating the mutation
1299: rate and the average value of the offers, which in this paper we derived
1300: from numerical simulations. For this purpose singular perturbation theory
1301: seems to be an appropriate tool. Furthermore, we have only considered
1302: mutation rates that are the same for the proposers as for the acceptors.
1303: Differentiating between these may also lead to a fuller understanding.
1304: 
1305: \begin{thebibliography}{99}
1306: \bibitem{WB} Weibull, J., \textit{Evolutionary Game Theory,} MIT Press, 1995.
1307: 
1308: \bibitem{FL} Fudenburg, D. and Levine, D., \textit{The Theory of Learning in
1309: Games,} MIT Press, 1998.
1310: 
1311: \bibitem{SA} Samuelson, L., \textit{Evolutionary Games and Equilibrium
1312: Selection,} MIT Press, 1998.
1313: 
1314: \bibitem{HS1} Hofbauer, J. and Sigmund, K., \textit{Evolutionary Games and
1315: Population Dynamics,} Cambridge University Press, 1998.
1316: 
1317: \bibitem{GIN} Gintis, H., \textit{Game Theory Evolving,} Princeton University
1318: Press, 2000.
1319: \bibitem{CRS} Cressman, R., \textit{Evolutionary Dynamics and Extensive Form
1320: Games,} MIT Press, 2003.
1321: 
1322: \bibitem{VB} Vincent, Th. and Brown, J., \textit{Evolutionary Game Theory,
1323: Natural Selection and Darwinian Dynamics,} Cambridge University Press, 2005.
1324: 
1325: \bibitem{HS2} Hofbauer, J. and Sigmund, K., \textit{Evolutionary Game
1326: Dynamics,} Bulletin AMS 40, 479-519, 2003.
1327: 
1328: \bibitem{DI} Diekmann, O., \textit{A beginners guide to adaptive dynamics,}
1329: in \textit{Mathematical Modelling of Population Dynamics, }R.Rudnicki, ed.,
1330: 47-86. Banach Center Publications , Vol.63, Institute of Mathematics, Polish
1331: Academy of Sciences.
1332: 
1333: \bibitem{TJ} Taylor, P.D. and Jonker, L., \textit{Evolutionary stable
1334: strategies and game dynamics,}Mathematical Bioscience, 40, 145-156, 1978.
1335: 
1336: \bibitem{DIC} Dieckmann, U., \textit{Can adaptive dynamics invade?,} Trends
1337: in Ecology and Evolution, 12, 367-370, 1998.
1338: 
1339: \bibitem{VK} Van Kampen, N.G., \textit{Stochastic Processes in Physics and
1340: Chemistry,} North-Holland, 1992.
1341: 
1342: \bibitem{PAO} Pao, C.V., \textit{Nonlinear Parabolic and Elliptic Equations,}
1343: Plenum Press, New York, 1992.
1344: 
1345: 
1346: \bibitem{CR} Cressman, R., \textit{Dynamic stability of the replicator
1347: equation with continuous strategy space,} IIASA Report IR-04-017, 2004.
1348: 
1349: \bibitem{OR} Oechsler, J. and Riedel, F., \textit{Evolutionary dynamics on
1350: infinite strategy spaces,} Economic Theory, 17(1), 141-162, 2001.
1351: 
1352: \bibitem{VA} Vaughan, R., \textit{Evolutive equilibrium selection.
1353: 1:Symmetric two-player binary choice games, }ELSE working paper, 2004.
1354: 
1355: \bibitem{BI} Binmore, K.G., \textit{Playing Fair: Game Theory and the Social
1356: Contract I and II. }Cambridge, MA: MIT Press, 1998.
1357: 
1358: \bibitem{FG} Fehr, E. and Gachter, S., \textit{Fairness and retaliation: the
1359: economics of reciprocity,} Journal of economic perspectives, 14(3), 159-182,
1360: 2000.
1361: 
1362: \bibitem{BG} Bowles, S. and Gintis, H., \textit{The evolution of strong
1363: reciprocity: cooperation in heterogeneous populations.} Theoretical
1364: Population Biology, 65, 17-28, 2004.
1365: 
1366: \bibitem{PN} Page, K.M. and Nowak, M., \textit{A generalized adaptive
1367: dynamics framework can describe the evolutionary Ultimatum Game,} Journal of
1368: Theoretical Biology, 209, 173-179, 2000.
1369: 
1370: \bibitem{SE} Seymour, R.M., \textit{Stationary distributions of noisy
1371: replicator dynamics in the Ultimatum Game}, Journal of Mathematical
1372: Sociology, 24(3), 193-243, 2000.
1373: 
1374: \bibitem{BO} Bowles, S. and Gintis, H., \textit{Homo reciprocans,} Nature,
1375: 415,\textbf{\ } 125-128, 2002.
1376: 
1377: \bibitem{NPS} Nowak, M., Page, K.M. and Sigmund, K., \textit{Fairness versus
1378: reason in the Ultimatum Game,} Science, 289, 1773-1775, 2000.
1379: \end{thebibliography}
1380: 
1381: \end{document}
1382: