1: \documentclass[a4paper,11pt]{article}
2: \usepackage{amsfonts,latexsym,amssymb}
3: \parskip 1mm
4:
5:
6: \baselineskip=24pt
7:
8: \input amssymb.sty
9: \usepackage{epsfig}
10:
11: \usepackage[applemac]{inputenc}
12: \usepackage{amsfonts}
13: %\usepackage[dvips]{graphicx}
14:
15: %\advance\voffset by -25mm
16: \advance\hoffset by -25mm
17: \setlength{\textwidth}{160mm}
18: \setlength{\textheight}{240mm}
19: \pagestyle{empty}
20:
21: \def \Z{{\mathbb Z}}
22: \def \N{{\mathbb N}}
23: \def\C{{\mathbb C}}
24: \def\Q{{\mathbb Q}}
25: \def \R{{\mathbb R}}
26: \def\E{{\mathbb E}}
27: \def\B{{\mathbb B}}
28: \def\I{{\mathbb I}}
29: \def\ep{\varepsilon}
30: \renewcommand{\le}{\leqslant}
31: \renewcommand{\ge}{\geqslant}
32: \renewcommand{\hat}{\widehat}
33: \renewcommand{\phi}{\varphi}
34: \renewcommand{\P}{P_{r_N,N}}
35:
36:
37: \newcommand{\eps}{\varepsilon}
38: %\newcommand{\R}{\mathbf{R}}
39: %\newcommand{\N}{\mathbf{N}}
40: %\newcommand{\E}{\mathbf{E}}
41: %\newcommand{\Z}{\mathbf{Z}}
42: %\newcommand{\un}{\mathbf{1}}
43: \newcommand{\dd}{{\mathrm d}}
44: \newcommand{\ee}{{\mathrm e}}
45: \newcommand{\law}{{\cal L}}
46:
47: \parskip 1mm
48:
49:
50: \baselineskip=24pt
51:
52:
53: \newtheorem{lemma}{Lemma}
54: \newtheorem{theorem}{Theorem}
55: \newtheorem{corollary}{Corollary}
56: \newtheorem{assumption}{Assumption}
57: \newtheorem{project}{Project}
58: \newtheorem{definition}{Definition}
59: \newtheorem{remark}{Remark}
60: \newtheorem{example}{Example}
61: \newtheorem{proposition}{Proposition}
62: \def\pih{\pi_{h}}
63:
64: \begin{document}
65: %\baselineskip=24pt
66: \title{\bf Stochastic gene expression\\
67: in switching environments}
68: \author{ Martin Gander, Christian Mazza\thanks{
69: Section de Math\'ematiques, 2-4 Rue du
70: Li\`evre, CP64, CH-1211 Gen\`eve 4, martin.gander@math.unige.ch, christian.mazza@math.unige.ch}
71: and Hansklaus Rummler\thanks{D\'epartement de
72: Math\'ematiques, Chemin du Mus\'ee 23, CH-1700 Fribourg,
73: hansklaus.rummler@unifr.ch}}
74: \date{ February 2, 2005}
75:
76:
77: \maketitle
78:
79:
80:
81:
82: \setcounter{page}{0}
83:
84:
85: \thispagestyle{empty}
86:
87:
88:
89: \begin{abstract}
90:
91: We study a stochastic model proposed recently in the genetic literature
92: to explain the heterogeneity of cell populations or of gene products.
93: Cells are located in two colonies, whose sizes fluctuate as
94: birth with migration processes in switching environment.
95: We prove that there is a range of parameters where
96: heterogeneity induces a larger mean fitness.
97: \end{abstract}
98:
99: \noindent {\it Keywords:} Gene expression, recursive chain, ergodic, stationary
100: measure
101:
102:
103: \vfill\vfill
104:
105:
106:
107: \vfill
108:
109:
110: \pagebreak
111:
112:
113: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
114:
115:
116:
117:
118:
119:
120:
121:
122:
123:
124:
125:
126:
127:
128: \section{Introduction}\label{s.results}
129:
130: In \cite{Thattai}, the authors introduce a model for stochastic gene
131: expression to study the heterogeneity of cell populations. They assume
132: that the cells, or for example the product of some gene, can be in two
133: distinct states or colonies. Let $X(t)$ and $Y(t)$ be the sizes of
134: these colonies, which are here considered as birth with migration
135: processes. We assume that the birth rates are either $\gamma_1$ or
136: $\gamma_0$ with $\triangle\gamma=\gamma_1-\gamma_0>0$, and that the
137: associated migration rates $k_1$ and $k_0$ are such that $k_0\ge k_1$,
138: that is cells located in the colony having the smaller birth rate
139: migrate at a higher rate to the colony with the higher birth rate than
140: the other way round.
141:
142: If the birth and migration rates are assigned once and for all to a
143: corresponding colony (e.g. $\gamma_0$ and $k_0$ to $X$, and $\gamma_1$
144: and $k_1$ to $Y$), then the mean sizes $n_0(t)=\E(X(t))$ and
145: $n_1(t)=\E(Y(t))$ satisfy the pair of differential equations (see
146: \cite{Renshaw73} or \cite{Renshaw91})
147: \begin{equation}\label{eqn0}
148: \begin{array}{rcl}
149: \displaystyle
150: \frac{{\rm d}n_0(t)}{{\rm d}t} & = & (\gamma_0-k_0)n_0(t)+k_1 n_1(t),\\
151: \displaystyle
152: \frac{{\rm d}n_1(t)}{{\rm d}t} & = & (\gamma_1-k_1)n_1(t)+k_0 n_0(t).
153: \end{array}
154: \end{equation}
155: According to \cite{Thattai}, we say that cells of the first colony
156: represented by $X(t)$ are {\it unfit} (they have the lower birth
157: rates), and conversely that cells of the second colony represented by
158: $Y(t)$ are {\it fit}. The proportion of fit cells in the global
159: population, $y(t)=n_1(t)/(n_1(t)+n_0(t))$, $t\ge 0$, satisfies the
160: non-linear differential equation
161: \begin{equation}\label{Fit}
162: \frac{{\rm d}y(t)}{{\rm d}t}=
163: k_0+(\triangle\gamma-k_0-k_1)y(t)-\triangle\gamma y(t)^2.
164: \end{equation}
165:
166: Then, as $t\to\infty$, $y(t)\longrightarrow y_1$, where
167: $\frac{1}{2}\le y_1\le 1$ follows directly from (\ref{Fit}), see
168: Section \ref{s.Stationary}. This describes the equilibrium value of
169: the proportion of fit cells in a non-changing environment. Fixing the
170: values of the parameters $k_0$, $\gamma_0$ and $\gamma_1$, we can ask
171: for the value of $0\le k_1\le k_0$ which maximizes the proportion of
172: fit cells, i.e. the equilibrium value of $y(t)$: the optimal strategy
173: is to keep all the fit cells in the fit state, that is to set their
174: migration rate to zero, $k_1=0$. This leads to $y_1=1$, and
175: thus the optimal solution would be a homogeneous population.
176:
177: Observations reveal however that most cell populations are not
178: homogeneous; to explain this, the authors of \cite{Thattai} propose to
179: introduce a small modification in the model by allowing environmental
180: changes (for related questions in this context, see
181: e.g. \cite{Renshaw91}), and show through Monte-Carlo simulations that
182: the homogeneous solution $k_1=0$ is then not always optimal. The idea
183: in their model is to allow the birth and migration rates to switch at
184: random times from one colony to the other, so that cells in the fit
185: colony become unfit and vice versa. If for example an environmental
186: change occurs at some random time $T_1>0$ ($T_0=0$), then the function
187: $f_1(t)$ representing the proportion of fit cells solves (\ref{Fit})
188: up to time $T_1$, and just after $T_1$, say at time $T_1+0$, the fit
189: cells corresponding to $Y(t)$ become unfit and vice versa. The
190: proportion of fit cells $f_1(t)$ is then switched to
191: $f_1(T_1+0)=1-f_1(T_1)$. After $T_1$, the random process
192: $\{f_1(t)\}_{t\ge 0}$ solves (\ref{Fit}) with initial data
193: $f_1(T_1+0)$ at time $T_1+0$, until a new environmental change occurs,
194: say at time $T_2>T_1$. There is a new switch, and the process is
195: again solution of (\ref{Fit}), until a new event occurs and so on.
196:
197: In \cite{Thattai}, the fluctuations of the environment are modeled
198: using a renewal process; the instants $T_i$, $i\ge 0$, are such that
199: the sequence of random variables $\{t_i\}_{i\ge 1}$ given by
200: $t_i\equiv T_i-T_{i-1}$, $i\ge 1$, is i.i.d. distributed according to
201: some law $\mu$ on $\R^+$. The authors then use Monte-Carlo simulations
202: to estimate the limiting value of the time averages along trajectories
203: of the process $f_1(t)$, of the form
204: $$
205: S_N=\frac{1}{T_N}\int_0^{T_{N}}f_1(s)\,{\rm d}s.
206: $$
207: This limiting average value is denoted by ${\rm Av}(f_1)_{k_1}$ to
208: express its dependency on the migration rate $k_1<k_0$, when all the
209: remaining parameters are fixed. Their simulations indicate that there
210: is a range of parameters ($k_0$ not too large) such that
211: $$
212: {\rm Av}(f_1)_{k_1>0}>{\rm Av}(f_1)_{k_1=0},
213: $$
214: which means that heterogeneous populations are more adapted than
215: homogeneous ones in a switching environment.
216:
217: In this paper, we study mathematically the limiting behavior of the
218: stochastic process $f_1(t)$ and the associated time average $S_N$ by
219: giving its stationary measure, and we provide mathematical formulas and
220: numerical solutions, which might be of interest in practical
221: laboratory experiments (see e.g. \cite{Thattai}).
222:
223: Our technique uses the process $X_k=f_1(T_k-0)$, $X_0=f_1(0)$, which
224: is such that $X_{k+1}=\phi_{t_{k+1}}(1-X_k)$, for some mapping
225: $\phi_t(x)$ (see Definition \ref{MarkovChain}). $(X_k)_{k\ge 0}$ is a
226: stochastic recursive Markov chain, and $S_N$ can be expressed as an
227: additive functional of the trajectory of $(X_k)_{1\le k\le N}$. In
228: Section \ref{s.Stationary}, we recall a Theorem from \cite{Diaconis}
229: on the convergence of stochastic recursive chains, which applies in
230: this setting. We give conditions ensuring the existence and uniqueness
231: of a stationary measure $\pi$, as well as geometric ergodicity. In
232: Section \ref{Stationarity}, we consider the case where $\mu$ is
233: exponential of parameter $\kappa >0$, and show that $\pi$ has a ${\cal
234: C}^{\infty}$ density $P$ with respect to Lebesgue measure. We
235: furthermore prove in Theorem \ref{ergodic} that a multiple $G$ of $P$
236: solves a second order differential equation with weak
237: singularities. Proposition \ref{DiffSolutions} provides series
238: expansions for $P$, which are necessary to derive properties of $P$
239: near the singularities. In Section \ref{s.numerical}, we show
240: numerical solutions, using the series expansions of Proposition
241: \ref{DiffSolutions} to start the numerical integration. We provide
242: an example where ${\rm Av}(f_1)_{k_1>0}>{\rm
243: Av}(f_1)_{k_1=0}$, which shows that it can be better to allow fit
244: cells to migrate to the unfit state than to conserve all the fit cells
245: in the fit state in such a switching environment. This is a regime
246: where it is suitable for the colonies to anticipate bad hypothetical
247: future events.
248:
249: \section{Convergence of recursive chains}\label{s.Stationary}
250:
251: We first give some basic results for the differential equation
252: (\ref{Fit}). The right hand side of (\ref{Fit}) can be factored into
253: $-\triangle\gamma (y-y_0)(y-y_1)$, where
254: $y_0=(\triangle\gamma-k_0-k_1)-\sqrt{d})/(2\triangle\gamma)<0$,
255: $y_1=((\triangle\gamma-k_0-k_1)+\sqrt{d})/(2\triangle\gamma)>0$, and
256: $d=(\triangle\gamma-k_0-k_1)^2+4k_0\triangle\gamma$. Then $k_0>k_1$
257: implies that $ 0<1-y_1<\frac{1}{2}<y_1<1, $ and that the derivative
258: ${\rm d}f_1(t)/{\rm d}t$ is positive when $f_1(t)$ is in the interval
259: $[0,y_1)$, negative in $(y_1,1]$, and it vanishes for $f_1(t)=y_1$.
260: It is not hard to check that any realization of the trajectory
261: $\{f_1(t)\}_{t\ge 0}$, with initial data $f_1(0)\in I=(1-y_1,y_1)$
262: will remain forever in $I$, and that any trajectory starting in the
263: interval $I^c=[0,1]\setminus I$ will enter $I$ after an almost surely
264: finite time. (However, $f_1(0)=y_1$ implies $f_1(t)\equiv y_1$.) We
265: thus restrict our study to the interval $I$.
266:
267: Given $t\in\R^+$, we define the mapping $\phi_t:I\longrightarrow I$
268: such that $\phi_t(x)$ is the value of the solution of (\ref{Fit}) at
269: time $t$ when starting at $x\in I$ at time $t_0=0$. Using separation
270: of variables for (\ref{Fit}), we obtain
271: %$$
272: % \int_x^{\phi_t(x)}\frac{{\rm d}y}{\triangle\gamma
273: % (y_1-y)(y-y_0)}=\int_0^t\,{\rm d}s=t,
274: %$$
275: %and thus one has
276: the relation
277: \begin{equation}\label{BasicRelation}
278: \frac{\phi_t(x)-y_0}{y_1-\phi_t(x)}=\frac{x-y_0}{y_1-x}\exp(\beta t),
279: \end{equation}
280: where we set $\beta=\triangle\gamma (y_1-y_0)$. Given $u\in I$, let
281: $\delta t(u,y)$ denote the time interval the orbit of the dynamical
282: system (\ref{Fit}) needs to join $u$ and $y$, $y\ge u$, when starting
283: at time $t=0$ at $u$. Then
284: \begin{equation}\label{TimeInterval}
285: \beta\delta t(u,y)=\ln\Big(\frac{(y-y_0)(y_1-u)}{(y_1-y)(u-y_0)}\Big).
286: \end{equation}
287:
288: \begin{definition}\label{MarkovChain}
289: Given $X_0=f_1(0)\in I$, consider the Markov chain with values in $I$
290: defined by
291: $$
292: X_{k+1}=\phi_{t_{k+1}}(1-X_k),
293: $$
294: where the sequence of random variables $\{t_k\}_{k\in\N^+}$ is
295: i.i.d. distributed according to some law $\mu$ on $\R^+$. This
296: Markov chain describes the evolution of $f_1(T_k-0)$, at the
297: instants just before the switches, with $T_{k+1}-T_k=t_{k+1}$.
298: \end{definition}
299: We first recall and adapt results of \cite{Diaconis} on the
300: convergence of such Markov chains, also called {\it stochastic
301: recursive chains}, see e.g. \cite{Borovkov}. The general setting is
302: described by a complete separable metric space $(S,\rho)$, the set of
303: values taken by the Markov chain, a family of mappings $f_{\theta}:\
304: S\longrightarrow S$, indexed by parameters $\theta$ living in some
305: parameter space ${\Theta}$, and a probability measure $\mu$ on
306: ${\Theta}$. Given an i.i.d.\ sequence of random elements $\theta_n$,
307: $n\ge 1$, of law $\mu$, we can consider the Markov chain
308: $(X_n)_{n\in\N}$ given by $X_{n+1}=f_{\theta_{n+1}}(X_n)$. The
309: following Theorem gives conditions for the existence and uniqueness of
310: a stationary measure (Theorem 1.1 of \cite{Diaconis}). In what
311: follows, $P^{(n)}(x,{\rm d}y)$ denotes the law of the Markov chain
312: $X_n$ and $\rho[P^{(n)}(x,\cdot),\pi]$ is the Prokhorov metric, see
313: below.
314: \begin{theorem}\label{DiacConv}
315: Assume that the family of functions $f_{\theta}$, $\theta\in \Theta$
316: is Lipschitz with
317: $$
318: \rho(f_{\theta}(x),f_{\theta}(y))\le K_{\theta}\rho(x,y),\ \ x,y\in
319: S,
320: $$
321: $\forall \theta\in \Theta$. Assume furthermore that
322: \begin{equation}\label{TrivCondi}
323: \int K_{\theta} \mu({\rm d}\theta)<\infty,\ \ \int
324: \rho(f_{\theta}(x_0),x_0)\mu({{\rm d}\theta})<\infty,
325: \end{equation}
326: for some $x_0\in S$, and that
327: \begin{equation}\label{Contraction}
328: \int \ln(K_{\theta})\mu({{\rm d}\theta})<0.
329: \end{equation}
330: Then
331: \begin{itemize}
332: \item The Markov chain has a unique stationary distribution $\pi$,
333: \item $\rho[P^{(n)}(x,\cdot),\pi]\le A_x r^n$, for constants $A_x$
334: and $r$ with $0<A_x<\infty$ and $0<r<1$; this bound holds for all
335: times $n$ and all starting positions $x$,
336: \item the constant $r$ does not depend on $n$ or $x$; the constant
337: $A_x$ does not depend on $n$, and $A_x<a+b\rho(x,x_0)$, where
338: $0<a,b<\infty$.
339: \end{itemize}
340: \end{theorem}
341: In our setting, $S$ is given by $I$ and the parameter set
342: $\Theta$ is just $\R^+$. The Prokhorov distance
343: $d_n:=\rho[P^{(n)}(X_0,\cdot),\pi]$ is the infimum of the $\delta >0$
344: such that
345: \begin{equation}\label{Prok}
346: P^{(n)}(X_0,C)<\pi(C_{\delta})+\delta\quad
347: \mbox{and}\quad \pi(C)<P^{(n)}(X_0,C_{\delta})+\delta,
348: \end{equation}
349: where $C$ runs over the Borel sets of $I$ and, for given
350: $C\in\B(I)$, $C_{\delta}$ denotes the set of points of $I$
351: whose distance from $C$ is less than $\delta$ (see Section 5.1 of
352: \cite{Diaconis}). Condition (\ref{Contraction}) means that the
353: functions $f_{\theta}$ are contractions in the average. We first
354: express this condition in our setting: for $t\in \Theta=\R^+$ and
355: $u\in I=S$, the mapping $\phi_t(u)$ is given explicitly by
356: \begin{equation}\label{ExplicitPhi}
357: \phi_t(u)=\frac{y_0(y_1-u)+y_1(u-y_0)\exp(\beta t)}{y_1-u
358: +(u-y_0)\exp(\beta t)}.
359: \end{equation}
360: Setting $f_t(x)=\phi_t(1-x)$, we obtain
361: \begin{lemma}\label{Lip}
362: For all $t\in\R^+$,
363: $$
364: \frac{{\rm d}}{{\rm d}x}f_t(x)=-\frac{(y_1-y_0)^2\exp(\beta t)}
365: {(y_1-1+x+(1-x-y_0)\exp(\beta t))^2},
366: $$
367: $$
368: K_t := \sup_{x\in I}\vert \frac{{\rm d}}{{\rm d}x}f_t(x)\vert
369: =\frac{(y_1-y_0)^2\exp(\beta t)}{(2y_1-1+(1-y_1-y_0)\exp(\beta
370: t))^2}.
371: $$
372: If $\mu$ is exponential of parameter $\kappa >0$, and
373: $\alpha=\kappa/\beta$, then the conditions given in (\ref{TrivCondi})
374: are satisfied, and
375: $$
376: \int_{\R^+}\kappa \exp(-\kappa t)\ln(K_t) {\rm d}t=
377: -\alpha-2z \int_0^{\infty}\frac{\exp(-(1+\alpha)t)}
378: {1-z \exp(-t)}{\rm d}t,
379: $$
380: where we set $z=-(2y_1-1)/(1-y_1-y_0)<0$. Condition
381: (\ref{Contraction}) is thus satisfied if
382: \begin{equation}\label{Condi}
383: -\alpha-2z \int_0^{\infty}\frac{\exp(-(1+\alpha)t)}{1-z \exp(-t)}{\rm d}t<0.
384: \end{equation}
385: \end{lemma}
386: \begin{remark}
387: When $\vert z\vert\le 1$, the integral
388: $\int_0^{\infty}(\exp(-(1+\alpha)t))/(1-z \exp(-t)){\rm d}t$ is the
389: Lerch Phi function $\Phi(z,s,v)=\sum_{n\ge 0}(v+n)^{-s}z^n$, with
390: $s=1$ and $v=1+\alpha$, and is also equal to Gauss's Hypergeometric
391: function $_2{\rm F}_1(1,1+\alpha;2+\alpha;z)/(1+\alpha)$ (see e.g.
392: \cite{Erdelyi}, chap. 1.11).
393: \end{remark}
394: {\bf Proof}:
395: Taking the derivative of (\ref{ExplicitPhi}) with
396: respect to $u$, we obtain
397: $$
398: \frac{{\rm d}}{{\rm d}u}\phi_t(u)=
399: \frac{(y_1-y_0)^2\exp(\beta t)}{\bigl(y_1-u+(u-y_0)\exp(\beta t)\bigr)^2}
400: $$
401: and thus
402: $$
403: f'_t(x)=-\frac{\rm d}{{\rm
404: d}u}\phi_t(1-x)=-\frac{(y_1-y_0)^2\exp(\beta t)}
405: {(y_1-1+x+(1-x-y_0)\exp(\beta t))^2}<0,
406: $$
407: as required. The expression for $K_t$ follows from direct
408: computation.
409:
410:
411:
412:
413: \section{Convergence to stationarity in Poissonian environments}
414: \label{Stationarity}
415:
416: Assume that $\mu$ is exponential of parameter $\kappa >0$. We will
417: see in the sequel that the stationary measure $\pi$ has, under some
418: conditions, a density $P(y)$ such that with
419: $Q(y)=((y-y_0)/(y_1-y))^{\alpha}$, where $\alpha=\kappa/\beta$, the
420: function $G(y)=P(y)Q(y)(y-y_0)(y_1-y)$ satisfies the differential
421: equation
422: \begin{equation}\label{Diff}
423: G''(y)+ U(y)G'(y)+ V(y)G(y)=0,
424: \end{equation}
425: where
426: $\tilde y_0=1-y_0$, $\tilde y_1 =1-y_1$,
427: \begin{equation}\label{Conda}
428: U(y)=\frac{\alpha+1}{y-\tilde
429: y_0}-\frac{\alpha-1}{y-\tilde
430: y_1}+\frac{\alpha}{y-y_1}-\frac{\alpha}{y-y_0},
431: \end{equation}
432: and
433: \begin{equation}\label{Condb}
434: V(y)=\frac{\alpha^2
435: (y_1-y_0)^2}{(y-y_0)(y-y_1)(y-\tilde y_0)(y-\tilde y_1)}.
436: \end{equation}
437: The following proposition will therefore be useful:
438: \begin{proposition}\label{DiffSolutions}
439: The solutions of the second order homogeneous linear differential
440: equation (\ref{Diff}) are analytic on the interval
441: $I=(\tilde y_1,y_1)$. Two fundamental solutions $\tilde
442: G_1(y)$, $\tilde G_2(y)$ are
443: \begin{itemize}
444: \item $\tilde G_1(y)=(y-\tilde y_1)^\alpha \tilde W_1(y)$, where
445: $\tilde W_1(y)$ is analytic on $(\tilde y_1-\delta,y_1)$ for some
446: $\delta>0$ and with $\tilde W_1(\tilde y_1)=1$.
447:
448: \item $\displaystyle \tilde G_2(y)=\left\{
449: \begin{array}{ll}
450: \tilde W_2(y),& {\rm if}\ \alpha \not\in\Z,\\
451: \tilde W_2(y)
452: +\tilde C\tilde G_1(y)\ln(y-\tilde y_1),& {\rm if}\ \alpha\in\Z,
453: \end{array}
454: \right.$\\
455: with $\tilde W_2(y)$ analytic on $(\tilde
456: y_1-\delta,y_1)$ for some $\delta>0$, $\tilde W_2(\tilde y_1)=1$
457: and $\tilde C\in\R$.
458: \end{itemize}
459: Another set of two fundamental solutions $G_1(y)$, $G_2(y)$ is
460: \begin{itemize}
461: \item $G_1(y)=(y_1-y)^{1-\alpha}W_1(y)$, where $ W_1(y)$ is analytic
462: on $(\tilde y_1,y_1+\delta)$ for some $\delta>0$ and with
463: $W_1(y_1)=1$.
464:
465: \item
466: $\displaystyle G_2(y)=\left\{
467: \begin{array}{ll}
468: W_2(y),& {\rm if}\ \alpha \not\in\Z,\\
469: W_2(y)
470: +CG_1(y)\ln(y_1-y),& {\rm if}\ \alpha\in\Z,
471: \end{array}
472: \right.$\\
473: with $W_2(y)$ analytic on $(\tilde y_1,y_1+\delta)$
474: for some $\delta>0$, $W_2(y_1)=1$ and $C\in\R$.
475: \end{itemize}
476:
477: \end{proposition}
478: In the appendix, we show this result for completeness, and also how
479: these fundamental solutions can be computed by series expansion about
480: $\tilde y_1$ and $y_1$ respectively.
481:
482: \begin{theorem}\label{ergodic}
483: Assume that
484: $$
485: -\alpha-2z \int_0^{\infty}\frac{\exp(-(1+\alpha)t)}
486: {1-z \exp(-t)}{\rm d}t <0,
487: $$
488: where $z=-(2y_1-1)/(1-y_1-y_0)<0$. Then the Markov chain $X_k$ from
489: Definition \ref{MarkovChain}, with initial data $X_0\in I=(1-y_1,y_1)$
490: has a unique stationary distribution $\pi$ of ${\cal C}^{\infty}$
491: density
492: $$
493: P(y)=\frac{Q(y)^{-1}(y-\tilde y_1)^{\alpha}
494: \tilde W_1(y)/(y_1-y)/(y-y_0)}{\int_I
495: Q(z)^{-1}(z-\tilde y_1)^{\alpha}
496: \tilde W_1(z)/(y_1-y)/(y-y_0)\,{\rm d}z}.
497: $$
498: Here, $Q(y)=\Big(\frac{y-y_0}{y_1-y}\Big)^{\alpha}$, where
499: $\alpha=\kappa/\beta$, $\tilde W_1(y)$ is the analytic function on
500: $(\tilde y_1-\delta,y_1)$ with $\tilde W(\tilde y_1)=1$, such
501: that $\tilde G_1(y)=(y-\tilde y_1)^\alpha \tilde W_1(y)$ is a
502: solution of the differential equation (\ref{Diff}). In the
503: neighborhood of $y=y_1$, this solution is such that $0<\lim_{y\to
504: y_1}\tilde W_1(y)<+\infty$. Finally, the behavior of the density $P$ near
505: $y_1$ is given by $(y_1-y)^{\alpha -1}$, and thus converges when
506: $\alpha\ge 1$ and diverges toward $+\infty$ when $\alpha <1$. Let
507: $f(x)=x$ and $g(x)=\ln((x-1+y_1)/(y_1-x))$ be defined on $I$. Then $
508: g \in {\rm L}^1(I,\B(I),\pi)$ with
509: \begin{equation}\label{Moment}
510: \E_{\pi}(f)=y_0+\frac{\kappa}{\triangle\gamma}\E_{\pi}(g).
511: \end{equation}
512: \end{theorem}
513: \begin{remark}
514: Relation (\ref{Moment}) will be useful
515: when considering time averages for Monte-Carlo simulations, see
516: Section \ref{Average}.
517: \end{remark}
518: {\bf Proof}:
519: The existence and uniqueness of the stationary measure follows
520: from Theorem \ref{DiacConv} and Lemma \ref{Lip}.
521: Let $Y$ be a random variable of law $\pi$, and
522: let $T$ be exponential of parameter $\kappa >0$, independent
523: of $Y$. In the stationary regime,
524: $Y=_{{\cal L}} \phi_T(1-Y)$.
525: Let $F(y)=P(Y<y)$. Then
526: $$
527: F(y)=\int_{I\times\R^+}\pi({\rm d}v)\kappa \exp(-\kappa t)
528: \I(\phi_t(1-v)<y) {\rm d}t,
529: $$
530: where $\I(\cdot)$ denotes the indicator function. For given $y\in I$,
531: the time variable $t$ is restricted to the interval $0\le t< \delta
532: t(\tilde y_1,y)$ , see (\ref{TimeInterval}). Thus
533: $$
534: F(y)=\int_{0}^{\delta t(1-y_1,y)}\kappa \exp(-\kappa t)\int_I\pi({\rm d}v)
535: \I(\phi_t(1-v)<y) {\rm d}t .
536: $$
537: For given $t$ in this interval, the set of $v\in I$ with $\phi_t(1-v)<
538: y$ is given by
539: $$
540: \{v\in I;\ 1-v<\frac{y_1(y-y_0)+\exp(\beta t)(y_1-y)y_0}
541: {y-y_0+\exp(\beta t)(y_1-y)}\}.
542: $$
543: It follows that $\int_I\pi({\rm d}v)\I(\phi_t(1-v)<y)=1-F(1-u)$, where
544: we set $u=\bigl(y_1(y-y_0)+\exp(\beta t)(y_1-y)y_0\bigr)/
545: \bigl(y-y_0+\exp(\beta t)(y_1-y)\bigr)$, with $t=\delta t(u,y)$. We
546: make the change of variable $t=\delta t (u,y)$ with
547: $$
548: \frac{{\rm d}t}{{\rm d}u}=-\frac{y_1-y_0}{\beta(y_1-u)(u-y_0)}.
549: $$
550: Then
551: $$
552: F(y)= \alpha\Big(\frac{y_1-y}{y-y_0}\Big)^{\alpha}\int_{1-y_1}^y
553: \frac{y_1-y_0}{(y_1-u)(u-y_0)}
554: \Big(\frac{u-y_0}{y_1-u}\Big)^{\alpha}(1-F(1-u)){\rm d}u.
555: $$
556: This is a fixed point equation for the distribution function $F$. We
557: use it for proving that the probability measure $\pi$ has a ${\cal
558: C}^{\infty}$ density on the interval $I$. First notice that $F$ is
559: monotonically increasing and integrable on $I$. The above relation
560: then shows that $F$ is continuous on $I$. Using again this argument
561: recursively, one sees that $F$ is the integral of a continuous function
562: and is therefore differentiable, with a continuous derivative. The
563: ${\cal C}^{\infty}$ differentiability is obtained by iterating this
564: argument. Let $P$ be the ${\cal C}^{\infty}$ density of $\pi$ with
565: respect to Lebesgue measure. Our strategy runs as follows: We use the
566: fixed point relation to show that a multiple $G$ of $P$ satisfies a
567: second order differential equation, which has only weak singularities,
568: and then deduce properties of $P$ with the help of Proposition
569: \ref{DiffSolutions}.
570:
571: For given $v\in I$, the time variable $t$ is restricted to the
572: interval
573: $$
574: 0\le t\le \delta t(u,y)=
575: \ln\bigl((y-y_0)(y_1-y)/(y_1-y)(u-y_0)\bigr)/\beta,
576: $$
577: where $u=1-v$ (see \ref{TimeInterval}). It follows that
578: $$
579: F(y)=\int_{1-y}^{y_1}P(v){\rm d}v\int_{0}^{\delta t(u,y)}\kappa
580: \exp(-\kappa t){\rm d}t,
581: $$
582: with
583: \begin{eqnarray*}
584: P(y)=\frac{{\rm d}F(y)}{{\rm d}y}&=&\int_{1-y}^{y_1}P(v) {\rm d}v
585: \kappa \exp(-\kappa \delta t(u,y))\frac{{\rm d}\delta t(u,y)}{{\rm
586: d}y}\\
587: &=&\alpha \int_{1-y}^{y_1}{\rm d}v P(v)
588: \frac{Q(u)}{Q(y)}\frac{(y_1-y_0)}{(y-y_0)(y_1-y)},
589: \end{eqnarray*}
590: where we set $Q(y)=((y-y_0)/(y_1-y))^{\alpha}$. Using $u=1-v$ and
591: setting $G(y)=P(y)Q(y)(y-y_0)(y_1-y)$, one gets
592: \begin{equation}\label{relc}
593: G(y)=\int_{1-y_1}^y G(1-u)R(u)H(u)\,{\rm d}u,
594: \end{equation}
595: where $R(u)=\alpha Q(u)Q(1-u)^{-1}$ is such that $R(1-u)=\alpha^2/
596: R(u)$, and $H(u)=(y_1-y_0)/(y_1-1+u)/(1-u-y_0)$. Taking the
597: derivative gives
598: \begin{equation}\label{Deriv}
599: G'(y)=G(1-y)R(y)H(y),
600: \end{equation}
601: or
602: $$
603: G(1-y)=G'(y)R(y)^{-1}H(y)^{-1}=\alpha^{-2}G'(y) R(1-y)/H(y).
604: $$
605: Taking a second derivative then gives
606: $$
607: G''(y)+\frac{{\rm d}}{{\rm d}y}\ln(\frac{R(1-y)}
608: {H(y)})G'(y)+\alpha^2H(y)H(1-y) G(y)=0.
609: $$
610: and simplifying the terms leads to (\ref{Diff}). We see that
611: $R(u)H(u)\sim (u-1+y_1)^{\alpha -1}$, as $u\to 1-y_1$. The exponents associated with the
612: fundamental solutions are $\rho = 0$ or $\alpha$ in the neighborhood
613: of $y=1-y_1$ and $\rho'=0$ or $1-\alpha$ near $y=y_1$.
614:
615: Assume first that $\alpha\not\in \N^+$.
616: We first check the behavior of $G$ in a neighborhood of
617: $y=\tilde y_1$. Set $y=\tilde y_1+\ep$, $\ep >0$,
618: with $1-y=y_1-\ep$.
619: Proposition \ref{DiffSolutions} shows that $G$
620: is a linear combination $G(y)=\tilde A \ep^{\alpha}\tilde W_1(y)
621: +\tilde B \tilde W_2(y)$, for constants $\tilde A$, $\tilde B\in\R$.
622: Similarly, $G(1-y)=A \ep^{1-\alpha} W_1(1-y)+ B W_2(1-y)$,
623: for real constants $A$ and $B$. As $\ep\to 0$, the right hand side
624: of (\ref{relc}) behaves like $\ep^{\alpha}G(y_1-\ep)\to 0$.
625: Suppose that $\tilde B\ne 0$. Then $G(y)\sim \tilde B \tilde W_2(y)\ne 0$,
626: and (\ref{relc}) can't be satisfied. One must thus have $\tilde B =0$,
627: so that $G(y)=\tilde A \ep^{\alpha}\tilde W_1(y)$. When $\alpha >1$,
628: (\ref{relc}) implies that $A=0$. It follows that, for arbitrary $\alpha >0$,
629: $\lim_{y\to y_1}G(y)=B W_2(y_1)\ne 0$, and that
630: $G(\tilde y_1+\ep)\sim \tilde A \ep^{\alpha}\tilde W_1(\tilde y_1)$, $\ep\to 0$,
631: as required. The corresponding result for $P$ follows.
632:
633: Suppose that $\alpha\in\N^+$. The right hand side of (\ref{relc}) behaves like
634: $$F(\ep):=\ep^{\alpha}(A \ep^{1-\alpha}W_1(y_1)+B(W_2(y_1)+C \ep^{\alpha-1}W_1(y_1)\ln(\ep))),$$
635: with $F(\ep)\to 0$ as $\ep\to 0$, and $G(\tilde y_1+\ep)$ behaves like
636: $$\tilde F(\ep):=\tilde A \ep^{\alpha}\tilde W_1(\tilde y_1)+\tilde B(\tilde W_2(\tilde y_1)+\tilde C \ep^{\alpha}\tilde W_1(\tilde y_1)
637: \ln(\ep)).$$
638: One has $\tilde F(\ep)\sim \tilde B \tilde W_2(\tilde y_1)$, $\ep\to 0$, when $\tilde B\ne 0$ and
639: $\tilde F(\ep)\sim \tilde A \ep^{\alpha}\tilde W_1(\tilde y_1)$, when $\tilde B =0$.
640: (\ref{relc}) shows that necessarily $\tilde B =0$. Suppose that $\alpha =1$. Then one must
641: have $BC=0$, implying the existence of the limit $\lim_{y\to y_1}G(y)\ne 0$. When
642: $\alpha >1$, $A=0$, $B\ne 0$, and $\lim_{y\to y_1}G(y)=B W_2(y_1)$, as required.
643:
644:
645:
646:
647:
648:
649:
650:
651:
652:
653:
654:
655:
656:
657:
658:
659: Finally, we check the
660: identity (\ref{Moment}). First $g\in L^1(I,\B(I),\pi)$ follows from
661: the behavior of the density $P$ at the boundaries of $I$, as described
662: above. Next,
663: $$
664: \E_{\pi}(g)=\int_I \ln(\frac{y-1+y_1}{y_1-y})P(y){\rm d}y,
665: $$
666: where $J:=\int_I\ln(y-1+y_1)P(y){\rm d}y$ is such that
667: \begin{eqnarray*}
668: J&=&\int_I\ln(y-1+y_1)G(y)Q(y)^{-1}\frac{H(1-y)}{y_1-y_0}{\rm d}y\\
669: &=&\frac{1}{y_1-y_0}\int_I\ln(y_1-u)G(1-u)Q(1-u)^{-1}H(u){\rm d}u\\
670: &=&\frac{1}{\alpha(y_1-y_0)}\int_I
671: \frac{\ln(y_1-u)}{Q(u)}G(1-u)R(u)H(u){\rm d}u\\
672: &=&\frac{1}{\alpha(y_1-y_0)}\int_I
673: \frac{\ln(y_1-u)}{Q(u)}G'(u){\rm d}u\\
674: &=&\frac{1}{\alpha(y_1-y_0)}\Big(G(u)
675: \frac{\ln(y_1-u)}{Q(u)}\Big\vert_{1-y_1}^{y_1}-\int_IG(u)
676: \Big(\frac{\ln(y_1-u)}{Q(u)}\Big)'{\rm d}u\Big)\\
677: &=&\frac{1}{\alpha(y_1-y_0)}\int_I\frac{G(u)}{Q(u)}
678: \frac{(u-y_0)}{(y_1-u)(u-y_0)}{\rm d}u
679: +\int_I\ln(y_1-u)P(u){\rm d}u,
680: \end{eqnarray*}
681: where we use (\ref{Deriv}). It follows that
682: $$
683: \E_{\pi}(g)=\frac{1}{\alpha(y_1-y_0)}
684: \E_{\pi}(f)-\frac{y_0}{ \alpha(y_1-y_0)},
685: $$
686: proving (\ref{Moment}) since $\alpha=\kappa/(\triangle\gamma(y_1-y_0))$.
687:
688:
689: \begin{corollary}\label{LimitingLaw}
690: Assume that condition (\ref{Condi}) holds. Then, as $t\to +\infty$,
691: the law of the stochastic process $f_1(t)$, $t\ge 0$, $f_1(0)\in I$,
692: converges toward the stationary measure $\pi$ of density $P$ of the
693: Markov Chain $X_k$.
694: \end{corollary}
695:
696: \noindent {\bf Proof}:
697: Given $t\in\R^+$, let $t_*$ be the last renewal time before $t$, and
698: set $S_*=t-t_*$. When the length of the overlapping random interval
699: is exponential, $S_*$ is also exponential. In the stationary regime,
700: or equivalently for large $t$, one has the identity in law
701: $f_1(t)=_{\cal L} \phi_{S_*}(1-X)$, where $X$ is distributed according
702: to $\pi$, and the result follows.
703:
704:
705: \section{Time averages\label{Average}}
706:
707: When the conclusions of Theorem \ref{DiacConv} hold, the chain $X_k$
708: has a unique stationary probability measure $\pi$, and $\sum_{k=1}^n
709: g(X_k)/n$ converges a.s. toward the expectation of $g$ under $\pi$,
710: for any function $g$ in $L^1(I,\B(I),\pi)$, (see e.g. \cite{Breiman}).
711: In \cite{Thattai}, the authors use Monte-Carlo methods based on the
712: process $f_1(t)$, $t\ge 0$, to estimate the mean fitness by
713: considering the time average
714: \begin{equation}\label{TimeAverage}
715: S_N=\frac{1}{T_{N}}\int_0^{T_{N}}f_1(s)\,{\rm d}s,
716: \end{equation}
717: where $N$ is a fixed number of renewal periods.
718:
719: \begin{lemma}\label{Formula}
720: Let $N\in\N^+$. Given a realization $0=T_0<T_1<\cdots<T_N$ of the
721: renewal process, we have
722: \begin{equation}\label{TimeAverageB}
723: \frac{1}{T_N}\int_0^{T_N}f_1(s)\,{\rm d}s =y_0+\frac{(y_1-y_0)}{\beta T_N}
724: \ln\Big(\prod_{i=1}^ N\frac{X_{i-1}-(1-y_1)}{y_1-X_i}\Big).
725: \end{equation}
726: \end{lemma}
727:
728: \noindent {\bf Proof}:
729: Consider the integrals
730: $$
731: \int_{T_{i-1}}^{T_i}f_1(s)\,{\rm d}s,
732: $$
733: where $f_1(T_{i-1}+0)=1-X_{i-1}$ and $f_1(T_i-0)=X_i$. The value of
734: $y(s)=f_1(T_{i-1}+s)$, $s\in(0,T_i-T_{i-1})$ is given implicitly by
735: (\ref{BasicRelation}); Therefore
736: $$
737: y(s)=\frac{y_0(y_1-u)+y_1(u-y_0)\exp(\beta s)}
738: {y_1-u+(u-y_0)\exp(\beta s)},
739: $$
740: where we set $u=1-X_{i-1}$, and thus, after a longer but not difficult
741: computation, one obtains
742: $$
743: \int_{T_{i-1}}^{T_i}f_1(s)\,{\rm d}s=
744: y_0(T_i-T_{i-1})+\frac{y_1-y_0}\beta
745: \ln\left(\frac{y_1-u+(u-y_0)\exp\bigl(\beta(T_i-T_{i-1})\bigr)}
746: {y_1-y_0}\right),
747: $$
748: and the result follows, since
749: \begin{eqnarray*}
750: y_1-u+(u-y_0)\exp(\beta(T_{i}-T_{i-1}))&=&(y_1-u)\bigl(1+\frac{u-y_0}{
751: y_1-u}\exp(\beta t_i)\bigr)\\
752: &=&\frac{(y_1-(1-X_{i-1}))(y_1-y_0)}{y_1-X_i}.
753: \end{eqnarray*}
754:
755: \begin{theorem}\label{ConvergenceTimeAverage}
756: Suppose that $\mu$ is exponential of parameter $\kappa >0$, and assume
757: (\ref{Condi}). Let $f(x)=x$ and $g(x)=\ln((x-1+y_1)/(y_1-x))$ be
758: defined on $I$. Then
759: $$
760: \lim_{N\to\infty}\frac{1}{T_N}\int_0^{T_N}f_1(s)\,{\rm
761: d}s=y_0+\frac{\kappa}{\triangle\gamma}
762: \E_{\pi}(g)=\E_{\pi}(f),\ \ a.s.
763: $$
764: \end{theorem}
765:
766: \noindent {\bf Proof}:
767: From equation (\ref{TimeAverageB}), we obtain
768: $$
769: \frac{1}{T_N}\int_0^{T_N}f_1(s)\,{\rm d}s=y_0+\frac{(y_1-y_0)}{\beta
770: T_N}\ln(\frac{X_0-1+y_1}{y_1-X_N})
771: +\frac{(y_1-y_0)}{\beta T_N}\sum_{i=1}^{N-1}g(X_i).
772: $$
773: As $T_N$ is a renewal process with exponential inter arrival times of
774: parameter $\kappa$, it follows that $T_N/N$ converges a.s. toward
775: $1/\kappa$. Next, $ g\in L^1(I,\B(I),\pi)$ follows from the behavior
776: of the density $P$ at the boundaries of $I$, as described in Theorem
777: \ref{ergodic}. From Proposition \ref{DiffSolutions} and Theorem
778: \ref{ergodic}, the behavior of $P$ in the neighborhood of $y=1-y_1$ is
779: given by $(y-1+y_1)^{\rho_1}$ where $\rho_1=\alpha$ and by
780: $(y_1-y)^{\rho_2+\alpha-1}$ in the neighborhood of $y=y_1$, where
781: $\rho_2= 0$. The Markov chain $X_k $ is geometrically ergodic, and
782: thus the last term converges a.s. toward
783: $(\kappa/(\triangle\gamma))\E_{\pi}(g)$. We finally check that
784: $\ln(y_1-X_N)/N$ converges a.s. toward 0. Given $\epsilon >0$,
785: consider the probability
786: \begin{eqnarray*}
787: P(\vert \ln(y_1-X_N)\vert > N\epsilon)&=&P(\ln(y_1-X_N)<-N\epsilon)\\
788: &=&P(X_N>y_1-\exp(-N\epsilon))=P^{(N)}(X_0,A_N),
789: \end{eqnarray*}
790: where $A_N=\{x>y_1-\exp(-N\epsilon)\}$. Using the behavior of $P$ in
791: the neighborhood of $y=y_1$, one gets that $\pi(A_N)\le M
792: (\exp(-\epsilon N))^{\rho_2+\alpha}$, for some positive constant
793: $M$. Let $\gamma_N:=\vert P^{(N)}(X_0,A_N)-\pi(A_N)\vert$, and let
794: $d_N$ be the Prokhorov distance defined in (\ref{Prok}). If
795: $\pi(A_N)\ge P^{(N)}(X_0,A_N)$, then $\gamma_N\le \pi(A_N)$. If
796: $\pi(A_N)\le P^{(N)}(X_0,A_N)$, one has $P^{(N)}(X_0,A_N)\le
797: \pi((A_N)_{d_N})+d_N$, and it follows that
798: \begin{eqnarray*}
799: \gamma_N= P^{(N)}(X_0,A_N)-\pi(A_N)&\le& \pi((A_N)_{d_N})-\pi(A_N)+d_N\\
800: &=& \int_{y_1-\exp(-\epsilon N)-d_N}^{y_1-\exp(-\epsilon
801: N)}P(y) {\rm d}y +d_N\\
802: &\le&d_N +D(\exp(-\epsilon
803: N)^{\rho_2+\alpha}-(d_N+\exp(-\epsilon N))^{\rho_2+\alpha}),
804: \end{eqnarray*}
805: for some positive constant $D>0$. Theorem \ref{DiacConv} gives that
806: $$
807: P(\vert \ln(y_1-X_N)\vert >\epsilon N)\le\vert
808: P^{(N)}(X_0,A_N)-\pi(A_N)\vert +\pi(A_N)
809: \le h(X_0)\lambda^N,
810: $$
811: for some bounded function $h$ and a positive number $0<\lambda<1$.
812: The result then follows from the Borel-Cantelli Lemma. The last
813: identity is (\ref{Moment}) of Theorem \ref{ergodic}.
814:
815:
816: \section{Numerical Examples}\label{s.numerical}
817:
818: We now compute the density $P$ given in Theorem \ref{ergodic}
819: numerically. To do so, we solve the differential equation (\ref{Diff})
820: numerically, starting in the neighborhood of the singular point
821: $y=\tilde y_1=1-y_1$. Proposition \ref{DiffSolutions} and Theorem
822: \ref{ergodic} show that $\lim_{y\to \tilde y_1}P(y)=0$, and that the
823: first derivative of $P$ behaves like $(y-\tilde y_1)^{\alpha -1}$,
824: which goes to $+\infty$ when $\alpha <1$. We start the numerical
825: solution at the point $y=\tilde y_1+\varepsilon$, where $\varepsilon
826: >0$ is small, and use the initial conditions $G(\tilde
827: y_1+\varepsilon)$ and $G'(\tilde y_1+\varepsilon)$ from the series
828: expansions given in Proposition \ref{DiffSolutions}. In addition, we
829: use the numerical integration procedure to compute the integral to
830: scale the density $P$, by adding an additional ordinary differential
831: equation to (\ref{Diff}).
832:
833: We show in Figures \ref{fig12} to \ref{fig0} the results obtained for
834: five different sets of parameters.
835: \begin{figure}
836: \centering
837: \includegraphics[width=0.49\textwidth]{fig1.eps}
838: \includegraphics[width=0.49\textwidth]{fig2.eps}
839: \caption{Density $P$ on the left when $\kappa=10$,
840: $\triangle\gamma=1$, $k_0=0.4$, $k_1=0.05$, and on the right when
841: $\kappa =1.5$, $\triangle\gamma=1$, $k_0=0.1$, $k_1=0.05$.}
842: \label{fig12}
843: \end{figure}
844: \begin{figure}
845: \centering
846: \includegraphics[width=0.49\textwidth]{fig3.eps}
847: \includegraphics[width=0.49\textwidth]{fig4.eps}
848: \caption{Density $P$ on the left when $\kappa =10$,
849: $\triangle\gamma=9$, $k_0=0.1$,
850: $k_1=0$. $Av(f_1)_{k_1=0}=0.553274111$, and on the right when
851: $\kappa =10$, $\triangle\gamma=9$, $k_0=0.1$,
852: $k_1=0.05$. $Av(f_1)_{k_1=0.05}=0.55672212$. Clearly the average
853: fitness is larger when $k_1=0.05$ than when $k_1=0$.}
854: \label{fig34}
855: \end{figure}
856: \begin{figure}
857: \centering
858: \includegraphics[width=0.49\textwidth]{fig0.eps}
859: \caption{Density $P$ when $\kappa=5$,
860: $\triangle\gamma=3$, $k_0=3$, $k_1=1$, a case where
861: $\alpha<1$.}
862: \label{fig0}
863: \end{figure}
864: In all the figures, we show the computed solution $G$ of the
865: differential equation (\ref{Diff}) in dashed, the computed density $P$
866: as a solid line, and the results of a Monte-Carlo simulation with
867: 100'000 samples as circles. The density from the theory and the
868: Monte-Carlo simulations agree very well. It is interesting to see in
869: Figures \ref{fig12} and \ref{fig34} the variety of densities that can
870: be generated by this simple model. Figure \ref{fig34} contains a case
871: where increasing $k_1$ increases the overall fitness of the
872: population. Figure \ref{fig0} finally shows a case where $\alpha<1$. We
873: note that the numerical integration out of the singularity can be
874: challenging. In particular, for the first case in Figure \ref{fig12},
875: the standard ode45 from Matlab needed very small absolute tolerances
876: to succeed with the integration for $\varepsilon<1e-2$. A more robust
877: method turned out to be DOPRI853, see \cite{Hairer}.
878:
879: \pagebreak
880:
881: \section{Appendix}
882:
883: In this appendix we show for completeness the proof of Proposition
884: \ref{DiffSolutions} and describe a method how to solve the
885: differential equation (\ref{Diff}) (see also \cite{Jaenich},
886: pp.~317-321). This equation is of the form
887: $$G''(y)+U(y)G'(y)+V(y)G(y)=0,$$
888: where the functions $U(y)$ and $V(y)$ are meromorphic in the complex plane
889: with four poles of order one at
890: $y_0<\tilde y_1:=1-y_1<y_1<\tilde y_0:=1-y_0$.
891: The solutions are therefore analytic in the open disc of radius
892: $(y_1-\tilde y_1)/2$ centered at $1/2$.
893: We look for real solutions in
894: the interval
895: $I=(\tilde y_1,y_1)$. In order to simplify calculations, we use
896: the variable transformation
897: \begin{equation} \label{transform}
898: y=\tilde y_1+(y_1-\tilde y_1)z,\qquad z=\frac{y-\tilde
899: y_1}{y_1-\tilde y_1}
900: \end{equation}
901: and set $g(z):=G(y)$.
902: With this transformation, the differential equation (\ref{Diff}) becomes
903: \begin{equation}\label{newDiff}
904: g''(z)+u(z)g'(z)+v(z)g(z)=0,
905: \end{equation}
906: where
907: $u$ and $v$ have four poles of order one at the points $-b<0<1<1+b$ with
908: $b=(\tilde y_1-y_0)/(y_1-\tilde y_1)$:
909: $$u(z)=\frac{1-\alpha}z+\frac\alpha{z-1}-\frac\alpha{z+b}+\frac{\alpha
910: +1}{z-(1+b)},\quad
911: v(z)=
912: \frac{\alpha(1+b)^2}{z(1-z)(z+b)(1+b-z)}.$$
913: We can therefore rewrite this equation as
914: \begin{equation} \label{genDiff}
915: g''(z)+\frac{h(z)}zg'(z)+\frac{k(z)}{z^2}g(z)=0,
916: \end{equation}
917: where $h(z)$ and $k(z)$ are analytic in the disc of radius $\min\{1,b\}$
918: centered at 0:
919: $$h(z)=\sum_{n=0}^\infty \alpha_nz^n,\qquad k(z)=\sum_{n=0}^\infty
920: \beta_nz^n.$$
921: Multiplying the equation (\ref{genDiff}) by $z^2$ we get an
922: equivalent equation
923: which can be written as
924: \begin{equation} \label{opDiff}
925: {\rm L}(g):=({\mu_z}^2{\rm D}^2+\mu_h\mu_z{\rm D}+\mu_k)(g)=0,
926: \end{equation}
927: where ${\rm D}$ denotes differentiation and $\mu_f$ multiplication
928: by a function $f(z)$.
929: Looking for solutions of the form
930: $$g(z)=z^\rho w(z),\qquad w(z)=1+\sum_{n=1}^\infty w_nz^n,$$
931: we may identify the function $g(z)$ with the infinite row
932: $[w]=[1,w_1,w_2,w_3,\ldots]$ and write (\ref{opDiff}) in matrix form:
933: \begin{equation} \label{linDiff}
934: [{\rm L^\rho}][w]^{\rm T}=0.
935: \end{equation}
936: If we write ${\rm L}$
937: as
938: ${\rm L}=(\mu_z{\rm D}+\mu_{h-1})\mu_z{\rm D}+\mu_k$, we get the lower
939: triangular matrix
940: $$[{\rm L^\rho}]=\left[\begin{array}{ccccc}
941: \rho(\rho+\alpha_0-1)+\beta_0&0&\ldots\\
942: \rho \alpha_1+\beta_1&(\rho+1)(\rho+\alpha_0)+\beta_0&0&\ldots\\
943: \rho \alpha_2+\beta_2&(\rho+1)\alpha_1+
944: \beta_1&(\rho+2)(\rho+1+\alpha_0)+\beta_0&0&\ldots\\
945: \vdots&\vdots&\vdots&\vdots&\ddots
946: \end{array}\right].
947: $$
948: A solution $[w]=[1,w_1,w_2,\ldots]$ of the linear system (\ref{linDiff})
949: exists if and only if $L_{00}^\rho=0$. This is the so-called indicial
950: equation for $\rho$. From now on we shall no longer treat the general case
951: but only the case corresponding to our differential equation
952: (\ref{newDiff}). In this case $\alpha_0=1-\alpha$ and $\beta_0=0$. So the
953: indicial equation is
954: $\rho(\rho-\alpha)=0$
955: and yields the two characteristic exponents
956: $\rho_1=\alpha$ and $\rho_2=0$.
957: We shall write $L_{ij}^\nu$ instead of
958: $L_{ij}^{\rho_\nu}$.
959:
960: For $\rho=\rho_1$, the solution
961: $[w^{(1)}]=[1,w_1^{(1)},w_2^{(1)},\ldots]$ may be calculated by the
962: recursion scheme
963: $$w_0^{(1)}=1,
964: \quad w_n^{(1)}=\frac{-1}{
965: L_{nn}^1}\left(\sum_{j=0}^{n-1}L_{nj}^1w_j^{(1)}
966: \right)\ {\rm for}\ n\ge1.$$
967: With these coefficients $w_n^{(1)}$, the function
968: $$g_1(z)=z^{\rho_1}\left(1+\sum_{n=1}^\infty w_n^{(1)}z^n\right)$$
969: is a solution of (\ref{newDiff}). From the general theory of linear
970: differential equations in the
971: complex plane it follows that $g_1$ is analytic in the disc of radius
972: $1/2$ centered
973: at $1/2$, but the power series for $w_1(z)$ might have a convergence
974: radius $0<\delta<1$.
975:
976: If $\alpha$ is not an integer, another solution $g_2(z)$, linearly
977: independent of
978: $g_1(z)$, can be obtained in the same way from $\rho=\rho_2=0$. If, however,
979: $\alpha$ is an integer, the corresponding matrix has the entry ${\rm
980: L}_{nn}^2=0$
981: for $n=\alpha$, and we look in this case for a solution $g_2(z)$ of the form
982: $g_2(z)=1+\sum_{n\ge1} w_n^{(2)}z^n+Cg_1(z)\ln z$.
983: As $g_1$ is a solution, the terms in ${\rm L}(g_2)$
984: containing $\ln z$ cancel and the function
985: $w^{(2)}(z)=1+\sum_{n\ge1}w_n^{(2)}z^n$ must satisfy the equation
986: $${\rm L}(w^{(2)})=-C(2\mu_z{\rm D}+\mu_{h-1})(g_1).$$
987: Identifying $w^{(2)}(z)$ with the infinite row
988: $[w^{(2)}]=[1,w_1^{(2)},w_2^{(2)},\ldots]$,
989: we can write this in matrix form
990: \begin{equation} \label{linDiff2}
991: [{\rm L^2}][w^{(2)}]^{\rm T}=-C[v_1,v_2,\ldots]^{\rm T}.
992: \end{equation}
993: For the right-hand side one checks easily that $v_j=0$ for
994: $j=0,\ldots,\alpha-1$ and
995: $v_\alpha=\alpha$. Therefore we can resolve the inhomogeneous linear
996: system (\ref{linDiff2}) in the following way:
997: \begin{enumerate}
998:
999: \item
1000: We determine $w_j^{(2)}$ for $j\le\alpha$ in the same way as
1001: $w_j^{(1)}$.
1002:
1003: \item
1004: We set $w_\alpha^{(2)}:=0$ and determine the constant $C$ by the equation
1005: $\sum_{j=0}^{\alpha-1} L_{\alpha,j}^{(2)}w_j{(2)}=-Cv_{\alpha}$.
1006:
1007: \item
1008: We determine the coefficients $w_n^{(2)}$ for $n>\alpha$ by the recursion
1009: formula
1010: $$w_n^{(2)}=\frac{-1}{
1011: L_{nn}^2}\left(Cv_n+\sum_{j=0}^{n-1}L_{nj}^2w_j^{(2)}
1012: \right)\ {\rm for}\ n\ge\alpha+1.$$
1013:
1014: \end{enumerate}
1015:
1016: We shall not go into further details, for example present concrete formulas
1017: expressing the $v_n$ by the $w_n^{(1)}$, because we don't really need
1018: the solution
1019: $g_2$ of (\ref{genDiff}) in our case, as we have shown in the proof of Theorem
1020: \ref{ergodic}.
1021:
1022: Using the variable transformation (\ref{transform}) we get the
1023: solutions $\tilde G_j(y)$ of
1024: the original differential equation (\ref{Diff}), in particular
1025: $$\tilde G_1(y)=(y_1-\tilde y_1)^\alpha g_1\left(\frac{y-\tilde
1026: y_1}{y_1-\tilde y_1}\right)
1027: =(y-\tilde y_1)^\alpha \tilde W_1(y)=(y-\tilde y_1)^\alpha\left(
1028: 1+\sum_{n=1}^\infty
1029: \frac{w_n^{(1)}}{(y_1-\tilde y_1)^n}\right).$$
1030:
1031: In order to find fundamental solutions near the singularity $y_1$, we can
1032: apply the same method once more, but using the variable transformation
1033: $$y=y_1-(y_1-\tilde y_1)z,\qquad z=\frac{y_1-y}{y_1-\tilde y_1}.$$ One easily
1034: checks that in this case the indicial equation is
1035: $\rho(\rho+\alpha-1)=0$
1036: and that therefore the two characteristic exponents at $y_1$ are
1037: $\rho'_1=1-\alpha\ {\rm and}\ \rho'_2=0.$ We obtain thus the second fundamental
1038: system of solutions $G_1(y)$ and $G_2(y)$.
1039:
1040:
1041:
1042: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1043:
1044: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1045:
1046:
1047:
1048: \noindent{\bf Acknowledgment}: The authors thank their collaborator and friend Gerhard Wanner for
1049: many fruitful discussions concerning numerical solutions of differential
1050: equations.
1051:
1052:
1053:
1054:
1055: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1056:
1057: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1058: \begin{thebibliography}{3}
1059: \bibitem{Borovkov}
1060: {\sc Borovkov, A.}(1998).
1061: \emph{ Ergodicity and Stability of Stochastic Processes},
1062: Wiley Series in Probability and Statistics, New-York.
1063: \bibitem{Breiman}
1064: {\sc Breiman, L.} (1960).
1065: The strong law of large numbers for a class of Markov chains.
1066: \emph{Ann. Math. Stat.}, {\bf 31,} 801--803.
1067: \bibitem{Diaconis}
1068: {\sc Diaconis, P. and Freedman, D.}
1069: Iterated Random Functions.
1070: \emph{Siam Review}, {\bf 41,} 45--76.
1071: %\bibitem{Doob}
1072: %J.~L.~Doob
1073: %\newblock {\em Stochastic Processes}
1074: %\newblock Wiley, New York. 1953.
1075: \bibitem{Erdelyi}
1076: {\sc Erdelyi, A.}
1077: \emph{ Higher Transcendental Functions.} (1953).
1078: Bateman Manuscript Project. Vol.1. McGraw-Hill
1079: \bibitem{Hairer}
1080: {\sc Hairer, H., Norsett, S. and Wanner, G.} (2000)
1081: \emph{ Solving Ordinary Differential Equations I, Nonstiff Problems},
1082: Springer Series in Computational Mathematics. Springer.
1083: Second Edition.
1084: \bibitem{Jaenich}
1085: {\sc J\"anich, K.} (2001).
1086: \emph{ Analysis f\"ur Physiker und Ingenieure},
1087: Springer Verlag.
1088: \bibitem{Renshaw73}
1089: {\sc Renshaw, E.} (1973)
1090: Birth, death and migration processes.
1091: \emph{Biometrika}, {\bf 59,} 49--69.
1092: \bibitem{Renshaw91}
1093: {\sc Renshaw, E} (1991)
1094: \emph{ Modelling Biological Populations in Space and Time},
1095: Cambridge Studies in Mathematical Biology. Cambridge
1096: University Press.
1097: \bibitem{Thattai}
1098: {\sc Thattai, M. and van Oudenaarden, A.}(2004)
1099: Stochastic Gene Expression in Fluctuating Environments
1100: \emph{Genetics}, {\bf 167,} 523--530.
1101: \end{thebibliography}
1102:
1103:
1104: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1105:
1106: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1107:
1108:
1109: \end{document}
1110:
1111: