1: \documentclass[matbio]{svjour}
2: \usepackage{amsmath,amssymb,times}
3:
4: \usepackage{bbm,epic}
5:
6: % mathematical symbols
7: \newcommand\B{B}
8: \newcommand\C{\mathrm{C}}
9: \newcommand\ccint[1]{\mathopen[ #1 \mathclose]}
10: \newcommand\coint[1]{\mathopen[ #1 \mathclose[}
11: \newcommand\ooint[1]{\mathopen] #1 \mathclose[}
12: \newcommand\dist{\operatorname{dist}}
13: \newcommand\dd[1]{\,\mathrm{d}#1}
14: \newcommand\eps\varepsilon
15: \newcommand\Id{\mathbbm{1}}
16: \newcommand\Ll[1]{\ell^{#1}}
17: \newcommand\M{\mathcal{M}}
18: \newcommand\NN{\mathbb{N}}
19: \newcommand\NNzer{\NN_0}
20: \newcommand\ord{\operatorname{\scriptstyle\mathcal{O}}}
21: \newcommand\R{\mathcal{R}}
22: \newcommand\Ra{\tilde{\R}}
23: \newcommand\RR{\mathbb{R}}
24: \newcommand\RRnn{\RR_{\ge0}}
25: \newcommand\X{X}
26: \renewcommand\le\leqslant
27: \renewcommand\ge\geqslant
28:
29:
30: % mathematical typesetting
31: \allowdisplaybreaks
32: \newcommand{\es}{\hspace*{0.5pt}} % space before punctuation marks in numbered equations
33: \newcommand\is\, % space between indices, e.g. M_{i \is i+1}
34: \newcommand\vc[1]{\boldsymbol{#1}} % vector
35: \newcommand\vp{\vphantom{()}}
36:
37: % typesetting
38: \newcommand\pr[1]{\textup{(}#1\textup{)}}
39: \newcommand\textdef[1]{\emph{#1}}
40:
41: % references to equations, figures, etc.
42: \newcommand\dref[1]{Definition~\textup{\ref{def:#1}}}
43: \newcommand\Dref[1]{Definition~\textup{\ref{def:#1}}}
44: \newcommand\Eref[1]{Equation~\textup{(\ref{eq:#1})}}
45: \newcommand\eref[1]{\textup{(\ref{eq:#1})}}
46: \newcommand\fref[1]{Figure~\textup{\ref{fig:#1}}}
47: \newcommand\Fref[1]{Figure~\textup{\ref{fig:#1}}}
48: \newcommand\lref[1]{Lemma~\textup{\ref{lem:#1}}}
49: \newcommand\Lref[1]{Lemma~\textup{\ref{lem:#1}}}
50: \newcommand\pref[1]{Proposition~\textup{\ref{prop:#1}}}
51: \newcommand\Pref[1]{Proposition~\textup{\ref{prop:#1}}}
52: \newcommand\sref[1]{Section~\textup{\ref{sec:#1}}}
53: \newcommand\Sref[1]{Section~\textup{\ref{sec:#1}}}
54: \newcommand\thref[1]{Theorem~\textup{\ref{thm:#1}}}
55: \newcommand\Thref[1]{Theorem~\textup{\ref{thm:#1}}}
56:
57: % abbreviations
58: \makeatletter
59: \newcommand\cf{cf.\@ifnextchar,{}{\ }}
60: \newcommand\eg{e.g.\@ifnextchar,{}{, }}
61: \newcommand\ie{i.e.\@ifnextchar,{}{, }}
62: \makeatother
63: \newcommand\chaps{Chs.~}
64: \newcommand\cor{Cor.~}
65: \newcommand\Eqs{Equations }
66: \newcommand\lem{Lemma~}
67: \newcommand\propositions{Propositions }
68: \newcommand\prop{Prop.~}
69: \newcommand\remk{Rem.~}
70: \newcommand\sect{Sec.~}
71: \newcommand\sections{Sections }
72: \newcommand\thm{Thm.~}
73: \newcommand\thms{Thms.~}
74: \newcommand\weaks{\mbox{weak-$*$} }
75: \newcommand\Wrt{With respect to }
76: \newcommand\wrt{with respect to }
77:
78: \let\oldcite\cite
79: \def\cite{\upshape\oldcite}
80: \smartqed
81: \makeatletter
82: \newif\if@showqed
83: \@showqedtrue
84: \global\@namedef{endproof}{\if@showqed\qed\fi\global\@showqedtrue\@endtheorem}
85: \makeatother
86: % Modified from amsthm.sty:
87: \makeatletter
88: \newlength\qedraise
89: \newcommand\qedhere{\@ifnextchar[{\qed@here}{\qed@here[0pt]}%] bracket matching
90: }
91: \def\qed@here[#1]{%
92: \global\setlength{\qedraise}{#1}%
93: {\@xp\aftergroup\csname\@currenvir @qed\endcsname}\global\@showqedfalse}%
94: \def\displaymath@qed{%
95: \relax
96: \ifmmode
97: \ifinner \aftergroup\linebox@qed
98: \else
99: \eqno
100: \let\eqno\relax \let\leqno\relax \let\veqno\relax
101: \raisebox{\qedraise}{\qed}%
102: \fi
103: \else
104: \aftergroup\linebox@qed
105: \fi
106: }
107: \@xp\let\csname equation*@qed\endcsname\displaymath@qed
108: \@xp\let\csname multline*@qed\endcsname\displaymath@qed
109: \def\align@qed{\tag*{\raisebox{\qedraise}{\qed}}}
110: \@xp\let\csname align*@qed\endcsname\align@qed
111: \def\linebox@qed{\hfil\hbox{\qed}\hfilneg}
112: \makeatother
113:
114: \begin{document}
115:
116: \title{Unequal Crossover Dynamics \\ in Discrete and Continuous Time}
117: \author{Oliver Redner \and Michael Baake}
118: \institute{Institut f\"ur Mathematik, Univ.\ Greifswald,
119: Jahnstr.\ 15a, 17487 Greifswald, Germany
120: \and
121: Fakult\"at f\"ur Mathematik. Univ.\ Bielefeld, Pf.\ 100131, 33501 Bielefeld, Germany
122: \and
123: \email{redner@uni-greifswald.de}
124: \and
125: \email{mbaake@mathematik.uni-bielefeld.de}}
126: \keywords{Unequal crossover -- Recombination -- Quadratic operators --
127: Probability generating functions -- Lyapunov functions}
128: \makeatletter
129: \def\@date{}
130: \makeatother
131: \def\copyleft{}
132: \def\makeheadbox{}
133: \maketitle
134:
135: \begin{abstract}
136: We analyze a class of models for unequal crossover (UC) of sequences
137: containing sections with repeated units that may differ in length.
138: In these, the probability of an `imperfect' alignment, in which the
139: shorter sequence has $d$ units without a partner in the longer one,
140: scales like $q^d$ as compared to `perfect' alignments where all
141: these copies are paired. The class is parameterized by this penalty
142: factor $q$. An effectively infinite population size and thus
143: deterministic dynamics is assumed. For the extreme cases $q=0$ and
144: $q=1$, and any initial distribution whose moments satisfy certain
145: conditions, we prove the convergence to one of the known fixed
146: points, uniquely determined by the mean copy number, in both
147: discrete and continuous time. For the intermediate parameter
148: values, the existence of fixed points is shown.
149: \end{abstract}
150:
151:
152: \section{Introduction}
153: \label{sec:intro}
154:
155: Recombination is a by-product of (sexual) reproduction that leads
156: to the mixing of parental genes by exchanging genes (or sequence
157: parts) between homologous chromosomes (or DNA strands). This is
158: achieved through an alignment of the corresponding sequences,
159: along with crossover events which lead to reciprocal exchange
160: of the induced segments. In this process, imperfect alignment
161: may result in sequences that differ in length form the parental
162: ones; this is known as {\em unequal crossover\/} (UC). Imperfect
163: alignment is facilitated by the presence of repeated elements
164: (as is observed in some rDNA sequences, compare \cite{GrLi}), and
165: is believed to be an important driving mechanism for their evolution.
166: The repeated elements may follow an evolutionary course independent of
167: each other and thus give rise to evolutionary innovation. For a detailed
168: discussion of these topics, see \cite{ShAt,ShWa} and references therein.
169:
170: This article is concerned with a class of models for UC, originally
171: investigated by Shpak and Atteson \cite{ShAt} for discrete time, which
172: is built on preceding work by Ohta \cite{Oht} and Walsh \cite{Wsh}
173: (see \cite{ShAt} for further references). Starting from their partly heuristic
174: results, we prove various existence and uniqueness theorems and
175: analyze the convergence properties, both in discrete and in continuous
176: time. This will require a rather careful mathematical development
177: because the dynamical systems are infinite dimensional.
178:
179: In this model class, one considers individuals whose genetic
180: sequences contain a section with repeated units. These may vary in
181: number, $i\in\NNzer = \{0,1,2,3,\ldots\}$, where $i=0$ is explicitly
182: allowed, corresponding to no unit being present (yet). The
183: composition of these sections (\wrt mutations that might have
184: occurred) and the rest of the sequence are ignored here.
185:
186: In the course of time, recombination events happen, each of which
187: basically consists of three steps. First, independent pairs are formed at random
188: (in equidistant time steps, or at a fixed rate). Then, their respective
189: sections are randomly aligned, possibly with imperfections in form of `overhangs',
190: according to some probability distribution for the various possibilities.
191: Finally, both sequences are cut at an arbitrary common position between two adjacent
192: building blocks, with uniform distribution for the cut positions, and
193: their right (or left) fragments are interchanged. This so-called
194: unequal crossover is schematically depicted in \fref{ureco}.
195: Obviously, the total number of relevant units is conserved in each
196: event.
197:
198: \begin{figure}[t]
199: \begin{center}
200: \unitlength=1mm
201: \begin{picture}(85,11)(0,0)
202: \matrixput(0.625,9.5)(2.5,0){4}(0,0){1}{\line(1,0){1.25}}
203: \matrixput(10,8)(5,0){9}(0,3){2}{\line(1,0){5}}
204: \matrixput(10,8)(5,0){10}(0,3){1}{\line(0,1){3}}
205: \matrixput(60,8)(5,0){3}(0,3){2}{\line(1,0){5}}
206: \matrixput(60,8)(5,0){4}(0,3){1}{\line(0,1){3}}
207: \matrixput(75.625,9.5)(2.5,0){4}(0,0){1}{\line(1,0){1.25}}
208: \matrixput(0.625,1.5)(2.5,0){10}(0,0){1}{\line(1,0){1.25}}
209: \matrixput(25,0)(5,0){6}(0,3){2}{\line(1,0){5}}
210: \matrixput(25,0)(5,0){7}(0,3){1}{\line(0,1){3}}
211: \matrixput(60,0)(5,0){2}(0,3){2}{\line(1,0){5}}
212: \matrixput(60,0)(5,0){3}(0,3){1}{\line(0,1){3}}
213: \matrixput(70.625,1.5)(2.5,0){6}(0,0){1}{\line(1,0){1.25}}
214: \qbezier(55,9.5)(57.5,5.5)(59.9,1.5)
215: \qbezier(55,1.5)(57.5,5.5)(59.9,9.5)
216: \end{picture}
217: \caption{Snapshot of an unequal crossover event as described in the text.
218: Rectangles denote the relevant blocks, while the dashed lines
219: indicate possible extensions with other elements that are
220: disregarded here.}
221: \label{fig:ureco}
222: \end{center}
223: \end{figure}
224:
225: We assume the population size to be (effectively) infinite. (Concerning
226: finite populations, see the remarks in \sref{remarks}.) Then, almost
227: surely in the probabilistic sense, compare \cite[Sec.\ 11.2]{EtKu86},
228: the population is described by the deterministic time evolution of
229: a probability measure $\vc{p} \in \M_1^+(\NNzer)$, which we
230: identify with an element $\vc{p} = (p_k^{})_{k\in\NNzer}^{}$ in the
231: appropriate subset of $\Ll{1}(\NNzer)$. Since we will not consider any
232: genotype space other than $\NNzer$ in this article, reference to it will be
233: omitted in what follows. These spaces are complete in the metric derived from
234: the usual $\Ll{1}$ norm, which is the same as the total variation norm here.
235: The metric is denoted by
236: \begin{equation}
237: \label{eq:urgpmetric}
238: d(\vc{p},\vc{q}) =
239: \|\vc{p} - \vc{q}\|^{}_1 = \sum_{k\ge0} |p_k^{} - q_k^{}| \es.
240: \end{equation}
241:
242: With this notation, the above process is described by the recombinator
243: \begin{equation}
244: \label{eq:reco}
245: \R(\vc{p})_i^{} =
246: \frac{1}{\|\vc{p}\|^{}_1}
247: \sum_{j,k,\ell\ge0} \, T_{ij,k\ell}^{} \, p_k^{} \, p_\ell^{} \es.
248: \end{equation}
249: Here, $T_{ij,k\ell} \ge 0$ denotes the probability that, given a pair
250: $(k,\ell)$, this pair turns into $(i,j)$. Consequently, for normalization, we require
251: \begin{equation}
252: \label{eq:sumt}
253: \sum_{i,j\ge0} T_{ij,k\ell} = 1
254: \qquad\text{for all $k, \ell \in \NNzer$.}
255: \end{equation}
256: The factor $p_k^{} \, p_\ell^{}$ in \eref{reco} describes the
257: probability that a pair $(k,\ell)$ is formed, \ie we assume that two
258: individuals are chosen independently from the population. We assume
259: further that $T_{ij,k\ell} = T_{ji,k\ell} = T_{ij,\ell k}$, \ie that
260: $T_{ij,k\ell}$ is symmetric \wrt both index pairs, which is
261: reasonable. Then, the summation over $j$ in \eref{reco}
262: represents the breaking-up
263: of the pairs after the recombination event. These two ingredients of
264: the dynamics constitute what is known as (\textdef{instant})
265: \textdef{mixing} and are responsible for the quadratic nature of the
266: iteration process.
267:
268: As mentioned above, we will only consider processes that conserve the
269: total copy number in each event, \ie $T^{(q)}_{ij,k\ell} > 0$ for $i+j
270: = k+\ell$ only. To\-geth\-er with the normalization \eref{sumt} and
271: the symmetry condition from above, this yields the (otherwise weaker)
272: condition
273: \begin{equation}
274: \label{eq:urcons}
275: \sum_{i\ge0} i \, T_{ij,k\ell}
276: = \sum_{i,j\ge0} \frac{i+j}{2} \, T_{ij,k\ell}
277: = \frac{k+\ell}{2} \es,
278: \end{equation}
279: which implies conservation of the \emph{mean copy number} in the population,
280: \begin{displaymath}
281: \sum_{i\ge0} i \, \R(\vc{p})_i^{}
282: = \sum_{i,j,k,\ell\ge0} i \, T^{(q)}_{ij,k\ell} \, p_k^{} \, p_\ell^{}
283: = \sum_{k,\ell\ge0} \frac{k+\ell}{2} \, p_k^{} \, p_\ell^{}
284: = \sum_{k\ge0} k \, p_k^{} \es.
285: \end{displaymath}
286:
287: Condition \eref{sumt} and the presence of the prefactor
288: $1/\|\vc{p}\|^{}_1$ in \eref{reco} make $\R$ norm non-increasing, \ie
289: $\|\R(\vc{x})\|^{}_1 \le \|\vc{x}\|^{}_1$, and positive homogeneous of
290: degree 1, \ie $\R(a\vc{x}) = |a| \R(\vc{x})$, for all $\vc{x} \in \Ll{1}$
291: and $a \in \RR$. Furthermore, $\R$ is a positive operator with
292: $\|\R(\vc{x})\|^{}_1 = \|\vc{x}\|^{}_1$ for all positive elements
293: $\vc{x} \in \Ll{1}$. Thus, it is guaranteed that $\R$ maps $\M_r^+$,
294: the space of positive measures of total mass $r$, into itself.
295: This space is complete in the
296: topology induced by the norm $\|.\|_1^{}$, \ie by the metric $d$ from
297: \eref{urgpmetric}. (For $r=1$, of course, the prefactor
298: is redundant but ensures numerical stability of an iteration with
299: $\R$.)
300:
301: Given an initial configuration $\vc{p}_0^{} = \vc{p}(t\mathbin=0)$,
302: the dynamics may be taken in discrete time steps, with subsequent
303: generations,
304: \begin{equation}
305: \label{eq:urdisctime}
306: \vc{p}(t+1) = \R(\vc{p}(t)) \es,
307: \qquad t\in\NNzer \es.
308: \end{equation}
309: Our treatment of this case will be set up in a way that also allows
310: for a generalization of the results to the analogous process in
311: continuous time, where generations are overlapping,
312: \begin{equation}
313: \label{eq:urconttime}
314: \tfrac{\dd{}}{\dd{t}} \vc{p}(t) = \varrho \, (\R-\Id)(\vc{p}(t)) \es,
315: \qquad t\in\RRnn \es.
316: \end{equation}
317: Obviously, the (positive) parameter $\varrho$ in \eref{urconttime} only leads
318: to a rescaling of the time $t$. We therefore choose $\varrho=1$ without loss
319: of generality. Furthermore, the formulation of discrete versus continuous
320: time dynamics in \eref{urdisctime} and \eref{urconttime} is chosen
321: so that the fixed points of \eref{urdisctime} are identical to the equilibria
322: of \eref{urconttime}, regardless of $\varrho$. This can easily be verified by
323: a direct calculation. In what follows, we will thus use the term \emph{fixed
324: point} for both discrete and continuous dynamics.
325:
326: In the UC model, one distinguishes `perfect' alignments, in which each
327: unit in the shorter sequence has a partner in the longer sequence, and
328: `imperfect' alignments, with `overhangs' of the shorter sequence relative
329: to the longer one. To come to a reasonable probability distribution for
330: the various possiblities, the first are taken to be equally probable among
331: each other, while the latter are penalized by a
332: factor $q^d$ relative to the first, where $q \in \ccint{0,1}$ is a
333: model parameter and $d$ is the length of the overhang (at most the entire
334: length of the shorter sequence; in the example of \fref{ureco}, interpreted
335: as a snapshot right after the crossover event took place, we
336: have $d=1$). In the extreme case $q=0$, only perfect alignments may
337: occur, whereas for $q=1$ overhangs are not penalized at all and one
338: obtains the uniform distribution on the possibilities. For
339: obvious reasons, the first case is dubbed \textdef{internal UC}, the
340: second \textdef{random UC} \cite{ShAt}.
341:
342: It is now straightforward, though a bit tedious, to derive the transition
343: probabilities $T^{(q)}_{ij,k\ell}$. To this end, one has to trace what
344: happens in steps two and three of the recombination event only, while
345: the random formation of pairs does not enter here.
346: This has been done in \cite{ShAt} and need not be repeated.
347: However, in view of our above remarks, it is desirable to rewrite
348: the findings in a way that reflects the natural symmetry properties
349: of the $T^{(q)}_{ij,k\ell}$. In compact notation, this leads to the
350: transition probabilities
351: \begin{equation}
352: \label{eq:urtq}
353: T^{(q)}_{ij,k\ell} =
354: C^{(q)}_{k\ell} \, \delta_{i+j,k+\ell} \, (1+\min\{k,\ell,i,j\}) \,
355: q^{0 \vee (k\wedge\ell - i \wedge j)} \es,
356: \end{equation}
357: where $k \vee \ell := \max\{k,\ell\}$, $k \wedge \ell :=
358: \min\{k,\ell\}$, and $0^0 = 1$. The $C^{(q)}_{k\ell}$ are chosen
359: such that \eref{sumt} holds, \ie $\sum_{i,j\ge0} T^{(q)}_{ij,k\ell} =
360: 1$, and are hence symmetric in $k$ and $\ell$. Explicitly, they read
361: (see also \cite[\sect 2.1]{ShAt})
362: \begin{displaymath}
363: C^{(q)}_{k\ell} =
364: \frac{(1-q)^2}{
365: (k\wedge\ell+1)(|k-\ell|+1)(1-q)^2 +
366: 2q(k\wedge\ell - (k\wedge\ell+1)q + q^{k\wedge\ell+1})} \es.
367: \end{displaymath}
368: Note further that the total number of units is indeed conserved in
369: each event and that the process is symmetric within both pairs. Hence
370: \eref{urcons} is satisfied.
371:
372: Let us briefly come back to the question of `discrete' versus `continuous'
373: time, which are considered simultaneously for good reasons.
374: Common to both is the nonlinearity that stems from the probability that a
375: certain (random) pair is formed in the first place. Then,
376: for the discrete time dynamics \eref{urdisctime}, the $T^{(q)}_{ij,k\ell}$
377: have the direct meaning of the transition probability that, given a pair
378: $(k,\ell)$, this turns into a pair $(i,j)$. In contrast, for the
379: continuous time dynamics \eqref{eq:urconttime}, the number $T^{(q)}_{ij,k\ell}$
380: is to be understood as the probability to obtain a pair $(i,j)$
381: \emph{conditioned} on a recombination event with a pair of type
382: $(k,\ell)$, of which each recombines at the same rate. In probabilistic
383: terminology, the $\R$ of \eref{urdisctime} is the discrete time
384: skeleton of the process in continuous time, also called the
385: embedded discrete time process.
386:
387:
388: \bigskip\noindent
389: The aim of this article is to find answers to the following questions:
390: \begin{enumerate}
391: \renewcommand\labelenumi{\theenumi.}
392: \item Are there fixed points of the dynamics?
393: \item Given the mean copy number $m$, is there a unique fixed point?
394: \item If so, under which conditions and in which sense does an initial
395: distribution converge to this fixed point under time evolution?
396: \end{enumerate}
397: Of course, the trivial fixed point with $p_0^{} = 1$ and $p_k^{} = 0$
398: for $k>0$ always exists, which we generally exclude from our
399: considerations. But even then, the answer to the first question is
400: positive for general operators of the form \eref{reco} that satisfy
401: \eref{sumt} and some rather natural further condition. This is
402: discussed in \sref{urgeneral}. For the extreme cases $q=0$ (internal
403: UC) and $q=1$ (random UC), fixed points are known explicitly for every
404: $m$ and it has been conjectured \cite{ShAt} that, under mild
405: conditions, also questions 2 and 3 can be answered positively for all
406: values of $q \in \ccint{0,1}$. Indeed, for both extreme cases, norm
407: convergence of the population distribution to the fixed points can be
408: shown, which is done in \sections \ref{sec:urq0} and \ref{sec:urq1},
409: respectively. Since the dynamical systems involved are infinite
410: dimensional, a careful analysis of compactness properties is needed
411: for rigorous answers. The proofs for $q=1$ are based on alternative
412: representations of probability measures via generating functions,
413: presented in \sref{urrepr}. For the intermediate parameter regime, we
414: can only show that there exists a fixed point for every $m$, but
415: neither its uniqueness nor convergence to it, see \sref{urqgen}. Some
416: remarks in \sref{remarks} conclude this article.
417:
418:
419: \section{Existence of fixed points}
420: \label{sec:urgeneral}
421:
422: Let us begin by stating the following general fact.
423: \begin{proposition}
424: \label{prop:recolipschitz}
425: If the recombinator\/ $\R$ of \eref{reco} satisfies \eref{sumt},
426: then the global Lipschitz condition
427: \begin{displaymath}
428: \|\R(\vc{x}) - \R(\vc{y})\|^{}_1 \le C \|\vc{x} - \vc{y}\|^{}_1
429: \end{displaymath}
430: is satisfied, with constant\/ $C=3$ on\/ $\Ll{1}$, respectively\/
431: $C=2$ if\/ $\vc{x}$, $\vc{y} \in \M_r$.
432: \end{proposition}
433: \begin{proof}
434: Let $\vc{x}$, $\vc{y} \in \Ll{1}$ be non-zero (otherwise the
435: statement is trivial). Then,
436: \begin{align*}
437: \|\R(\vc{x}) - \R(\vc{y})\|^{}_1 &=
438: \sum_{i\ge0} \left|
439: \sum_{j,k,\ell\ge0} T_{ij,k\ell}^{} \left(
440: \frac{x_k^{}\,x_\ell^{}}{\|\vc{x}\|^{}_1} -
441: \frac{y_k^{}\,y_\ell^{}}{\|\vc{y}\|^{}_1} \right) \right| \\
442: &\le \sum_{k,\ell\ge0} \left|
443: \frac{x_k^{}\,x_\ell^{}}{\|\vc{x}\|^{}_1} -
444: \frac{y_k^{}\,y_\ell^{}}{\|\vc{y}\|^{}_1} \right|
445: \sum_{i,j\ge0} T_{ij,k\ell} \\
446: &= \sum_{k,\ell\ge0} \left|
447: \frac{x_k^{}\,x_\ell^{}}{\|\vc{x}\|^{}_1} -
448: \frac{x_k^{}\,y_\ell^{}}{\|\vc{x}\|^{}_1} +
449: \frac{x_k^{}\,y_\ell^{}}{\|\vc{x}\|^{}_1} -
450: \frac{y_k^{}\,y_\ell^{}}{\|\vc{y}\|^{}_1} \right| \\
451: &\le \sum_{k,\ell\ge0} \left(
452: \frac{|x_k^{}|}{\|\vc{x}\|^{}_1} |x_\ell^{} - y_\ell^{}| +
453: |y_\ell^{}| \left| \frac{x_k^{}}{\|\vc{x}\|^{}_1} -
454: \frac{y_k^{}}{\|\vc{y}\|^{}_1} \right| \right) \\
455: &= \|\vc{x}-\vc{y}\|^{}_1 +
456: \frac{1}{\|\vc{x}\|^{}_1} \bigl\|
457: \|\vc{y}\|^{}_1 \vc{x} - \|\vc{x}\|^{}_1 \vc{y} \bigr\|^{}_1 \es.
458: \end{align*}
459: The last term becomes
460: \begin{multline*}
461: \frac{1}{\|\vc{x}\|^{}_1} \bigl\|
462: \|\vc{y}\|^{}_1 \vc{x} - \|\vc{x}\|^{}_1 \vc{y} \bigr\|^{}_1 \\
463: = \frac{1}{\|\vc{x}\|^{}_1} \bigl\|
464: (\|\vc{y}\|^{}_1 - \|\vc{x}\|^{}_1) \vc{x} +
465: \|\vc{x}\|^{}_1 (\vc{x} - \vc{y}) \bigr\|^{}_1
466: \le 2 \|\vc{x}-\vc{y}\|^{}_1 \es,
467: \end{multline*}
468: from which $\|\R(\vc{x}) - \R(\vc{y})\|^{}_1 \le 3\|\vc{x} -
469: \vc{y}\|^{}_1$ follows for $\vc{x}$, $\vc{y} \in \Ll{1}$. If
470: $\vc{x}$, $\vc{y} \in \M_r$, the above calculation simplifies to
471: $\|\R(\vc{x}) - \R(\vc{y})\|^{}_1 \le 2\|\vc{x} - \vc{y}\|^{}_1$.
472: \end{proof}
473: In continuous time, this is a sufficient condition for the existence
474: and uniqueness of a solution of the initial value problem
475: \eref{urconttime}, \cf \cite[\thms 7.6 and 10.3]{Ama}. Another useful
476: notion in this respect is the following.
477: \begin{definition}{\cite[\sect 18]{Ama}}
478: \label{def:lyapunov}
479: Let\/ $Y$ be an open subset of a Banach space\/ $E$ and let\/ $f
480: \colon Y \to E$ satisfy a \pr{local} Lipschitz condition. A
481: continuous function\/ $L$ from\/ $X \subset Y$ to\/ $\RR$ is called
482: a \textdef{Lyapunov function} for the initial value problem
483: \begin{displaymath}
484: \tfrac{\dd{}}{\dd{t}} \vc{x}(t) = f(\vc{x}(t)) \es,
485: \qquad \vc{x}(0) = \vc{x}_0 \in X \es,
486: \end{displaymath}
487: if the \textdef{orbital derivative}\/ $\dot{L}(\vc{x}_0) :=
488: \liminf_{t\to 0^+} \frac1t \bigl( L(\vc{x}(t)) - L(\vc{x}_0) \bigr)$
489: satisfies
490: \begin{equation}
491: \label{eq:dotl}
492: \dot{L}(\vc{x}_0) \le 0
493: \end{equation}
494: for all initial conditions\/ $\vc{x}_0 \in X$.
495: \end{definition}
496: If further $\dot{L}(\vc{x}_{\mathrm{F}}) = 0$ for a single fixed point
497: $\vc{x}_{\mathrm{F}}$ only, then $L$ is called a \textdef{strict}
498: Lyapunov function. If a Lyapunov function exists, we have
499: \begin{theorem}{\cite[\thm 17.2 and \cor 18.4]{Ama}}
500: \label{thm:lyapunov}
501: With the notation of \dref{lyapunov}, assume that there is a
502: Lyapunov function\/ $L$, that the set\/ $X$ is closed, and that, for
503: an initial condition\/ $\vc{x}_0 \in X$, the set\/ $\{ \vc{x}(t) :
504: t\in\RRnn, \text{$\vc{x}(t)$ exists} \}$ is relatively compact in\/
505: $X$. Then,\/ $\vc{x}(t)$ exists for all\/ $t\ge0$ and
506: \begin{displaymath}
507: \lim_{t\to\infty} \dist(\vc{x}(t), X_L) = 0 \es,
508: \end{displaymath}
509: where\/ $\dist(\vc{x},X_L) = \inf_{\vc{y} \in X_L} \|\vc{x} -
510: \vc{y}\|$ and\/ $X_L$ denotes the largest invariant subset of\/ $\{
511: \vc{x} \in X : \dot{L}(\vc{x}) = 0 \}$ \pr{in forward and backward
512: time}. \hspace*{\fill}\qed
513: \end{theorem}
514: Obviously, if $L$ is a strict Lyapunov function, we have $X_L =
515: \{\vc{x}_{\mathrm{F}}\}$ and this theorem implies $d(\vc{x}(t),
516: \vc{x}_{\mathrm{F}}) \to 0$ as $t \to \infty$.
517:
518: Returning to the original question of the existence of fixed points,
519: we now recall the following facts, compare \cite{BilC,Shir} for
520: details.
521: \begin{proposition}{\cite[\cor to \thm V.1.5]{Yos}}
522: \label{prop:vaguenorm}
523: Assume the sequence\/ $\bigl(\vc{p}^{(n)}\bigr)$ in\/ $\M_1^+$ to
524: converge in the \weaks topology \pr{\ie pointwise, or vaguely} to
525: some\/ $\vc{p} \in \M_1^+$, \ie
526: \begin{displaymath}
527: \lim_{n\to\infty} p^{(n)}_k = p_k^{}
528: \quad\text{for all\, $k\in\NNzer$} \es,
529: \quad\text{with\, $p_k^{}\ge0$ \; and\;
530: $\textstyle\sum_{k\ge0} p_k^{} = 1$} \es.
531: \end{displaymath}
532: Then, it also converges weakly \pr{in the probabilistic sense} and in
533: total variation, \ie $\lim_{n\to\infty} \|\vc{p}^{(n)} -
534: \vc{p}\|^{}_1 = 0$.\hspace*{\fill}\qed
535: \end{proposition}
536: \begin{proposition}
537: \label{prop:urgfixed}
538: Assume that the recombinator\/ $\R$ from \eref{reco} satisfies
539: \eref{sumt} and has a convex, \weaks closed invariant set\/ $M
540: \subset \M_1^+$, \ie $\R(M) \subset M$, that is\/ \emph{tight}, \ie for
541: every\/ $\eps>0$ there is an\/ $m\in\NNzer$ such that\/ $\sum_{k \ge
542: m} p_k^{} < \eps$ for every\/ $\vc{p} \in M$. Then, $\R$ has a
543: fixed point in\/ $M$.
544: \end{proposition}
545: \begin{proof}
546: Prohorov's theorem \cite[\thm III.2.1]{Shir} states that tightness
547: and relative compactness in the \weaks topology are equivalent (see
548: also \cite[\chaps 1.1 and 1.5]{BilC}). In our case, $M$ is tight
549: and \weaks closed, therefore, due to \pref{vaguenorm}, norm compact.
550: Furthermore, $M$ is convex and $\R$ is (norm) continuous by
551: \pref{recolipschitz}. Thus, the claim follows from the
552: Leray--Schauder--Tychonov fixed point theorem \cite[\thm
553: V.19]{ReSi}.
554: \end{proof}
555: \Wrt the UC model, we will see that such compact invariant subsets
556: indeed exist.
557:
558:
559: \section{Internal unequal crossover}
560: \label{sec:urq0}
561:
562: After these preliminaries, let us begin with the case of internal UC
563: with perfect alignment only, \ie $q=0$ in \eref{urtq}. This case is
564: the simplest because, in each recombination event, no sequences exceeding
565: the longer of the participating sequences can be formed. Here, on $\M_1^+$,
566: the recombinator \eref{reco} simplifies to
567: \begin{equation}
568: \label{eq:ur0reco}
569: \R_0(\vc{p})_i^{} =
570: \sum_{\substack{k,\ell\ge0 \\[0.4\baselineskip]
571: k\wedge\ell \le i \le k\vee\ell}}
572: \frac{p_k^{} \, p_\ell^{}}{1+|k-\ell|} \es.
573: \end{equation}
574: From now on, we write $\R_q$ rather than $\R$ whenever we look at a
575: recombinator with (fixed) parameter $q$. It is instructive to
576: generalize the notion of reversibility (or detailed balance, compare
577: \cite[(4.1)]{ShAt}).
578: \begin{definition}
579: We call a probability measure\/ $\vc{p} \in \M_1^+$
580: \textdef{reversible} for a recombinator\/ $\R$ of the form
581: \eref{reco} if, for all\/ $i, j, k, \ell \ge 0$,
582: \begin{equation}
583: \label{eq:urrev}
584: T_{ij,k\ell}^{} \, p_k^{} \, p_\ell^{} =
585: T_{k\ell,ij}^{} \, p_i^{} \, p_j^{} \es.
586: \end{equation}
587: \end{definition}
588: The relevance of this concept is evident from the following property.
589: \begin{lemma}
590: \label{lem:urrev}
591: If\/ $\vc{p} \in \M_1^+$ is reversible for\/ $\R$, it is also a
592: fixed point of\/ $\R$.
593: \end{lemma}
594: \begin{proof}
595: Assume $\vc{p}$ to be reversible. Then, by \eref{sumt},
596: \begin{displaymath}
597: \R(\vc{p})_i^{} =
598: \sum_{j,k,\ell\ge0} \, T_{ij,k\ell}^{} \, p_k^{} \, p_\ell^{} =
599: \sum_{j,k,\ell\ge0} \, T_{k\ell,ij}^{} \, p_i^{} \, p_j^{} =
600: p_i^{} \sum_{j\ge0} p_j^{} = p_i^{} \es. \qedhere[-13.9pt]
601: \end{displaymath}
602: \end{proof}
603: So, in our search for fixed points, we start by looking for solutions
604: of \eref{urrev}. Since, for $q=0$, forward and backward transition
605: probabilities are simultaneously non-zero only if $\{i,j\} =
606: \{k,\ell\} \subset \{n,n+1\}$ for some $n$, the components $p_k^{}$
607: may only be positive on this small set as well. By the following
608: proposition, this indeed characterizes all fixed points for $q=0$.
609: \begin{proposition}
610: \label{prop:ur0unique}
611: A probability measure\/ $\vc{p} \in \M_1^+$ is a fixed point of\/
612: $\R_0$ if and only if its mean copy number\/ $m = \sum_{k\ge0}
613: k\,p_k^{}$ is finite,\/ $p_{\lfloor m \rfloor}^{} = \lfloor m
614: \rfloor + 1 - m$, \linebreak $p_{\lceil m \rceil}^{} = m + 1 -
615: \lceil m \rceil$, and\/ $p_k^{} = 0$ for all other\/ $k$. This
616: includes the case that\/ $m$ is integer and\/ $p_{\lfloor m
617: \rfloor}^{} = p_{\lceil m \rceil}^{} = p_m^{} = 1$.
618: \end{proposition}
619: \begin{proof}
620: The `if' part was stated in \cite[\sect 4.1]{ShAt} and follows
621: easily by insertion into \eref{ur0reco} or \eref{urrev}. For the
622: `only if' part, let $i$ denote the smallest integer such that
623: $p_i^{} > 0$. Then,
624: \begin{displaymath}
625: \R(\vc{p})_i^{} =
626: p_i^2 + 2 p_i^{} \sum_{\ell\ge1} \frac{p_{i+\ell}^{}}{1+\ell} =
627: p_i^{} \left( p_i^{} + p_{i+1}^{} +
628: \sum_{\ell\ge2} \frac{2}{\ell+1} p_{i+\ell}^{} \right) \le
629: p_i^{} \es,
630: \end{displaymath}
631: where the last step follows since $\frac{2}{\ell+1} < 1$ in the last
632: sum, with equality if and only if $p_k^{} = 0$ for all $k \ge i+2$.
633: This implies $m < \infty$ and the uniqueness of $\vc{p}$ (given $m$)
634: with the non-zero frequencies as claimed.
635: \end{proof}
636:
637: It it possible to analyze the case of internal UC on the basis of the
638: compact sets to be introduced below in \sref{urrepr}. However, as J.
639: Hofbauer pointed out to us \cite{HofPriv}, it is more natural to start
640: with a larger compact set to be introduced in \eref{m1mc}. Our main
641: result in this section is thus
642: \begin{theorem}
643: \label{thm:ur0conv}
644: Assume that, for the initial condition\/ $\vc{p}(0)$ and fixed\/
645: $r>1$, the\/ $r$-th moment exists,\/ $\sum_{k\ge0} k^r p_k^{}(0) <
646: \infty$. Then, $m = \sum_{k\ge0} k \, p_k^{}(0)$ is finite and,
647: both in discrete and in continuous time,\/ $\lim_{t\to\infty}
648: \|\vc{p}(t) - \vc{p}\|^{}_1 = 0$ with the appropriate fixed point\/
649: $\vc{p}$ from \pref{ur0unique}.
650: \end{theorem}
651: The proof relies on the following lemma, which slightly modifies and
652: completes the convergence arguments of \cite[\sect 4.1]{ShAt}, puts
653: them on rigorous grounds, and extends them to continuous time.
654: \begin{lemma}
655: \label{lem:ur0dp}
656: Let\/ $r>1$ be arbitrary, but fixed. Consider the set of
657: probability measures with fixed mean\/ $m<\infty$ and a centered\/
658: $r$-th moment bounded by\/ $C<\infty\es$,
659: \begin{equation}
660: \label{eq:m1mc}
661: \M_{1,m,C}^+ = \{ \vc{p}\in\M_1^+ : \sum_{k\ge0} k\,p_k^{} = m,\,
662: M_r(\vc{p}) \le C \} \es,
663: \end{equation}
664: equipped with \pr{the metric induced by} the total variation norm,
665: where
666: \begin{equation}
667: \label{eq:ur0dp}
668: M_s(\vc{p}) = \sum_{k\ge0} |k-m|^s \, p_k^{}
669: \end{equation}
670: for\/ $s \in \{1,r\}$. This is a compact and convex space. Both\/
671: $M_1$ and\/ $M_r$ satisfy\/ $M_s(\R_0(\vc{p})) \le M_s(\vc{p})$,
672: with equality if and only if\/ $\vc{p}$ is a fixed point of\/
673: $\R_0$. Furthermore,\/ $M_1$ is a continuous mapping from\/
674: $\M_{1,m,C}^+$ to\/ $\RRnn$ and a Lyapunov function for the dynamics
675: in continuous time.
676: \end{lemma}
677: \begin{proof}
678: Let a sequence $(\vc{p}^{(n)}) \subset \M_{1,m,C}^+$ be given and
679: consider the random variables $\vc{f}^{(n)} = (k)_{k\in\NNzer}^{}$
680: on the probability spaces $(\NNzer,\vc{p}^{(n)})$. Their
681: expectation values are equal to $m$, which, by Markov's inequality
682: \cite[p.~599]{Shir}, implies the tightness of the sequence
683: $(\vc{p}^{(n)})$. Hence, by Prohorov's theorem \cite[\thm
684: III.2.1]{Shir} (see also \cite[\chaps 1.1 and 1.5]{BilC}), it
685: contains a convergent subsequence $(\vc{p}^{(n_i)})$ (recall that,
686: by \pref{vaguenorm}, norm and pointwise convergence are equivalent
687: in this case). Let $\widetilde{\vc{p}} \in \M_1^+$ denote its limit and
688: $\widetilde{\vc{f}} = (k)_{k\in\NNzer}^{}$ a random variable on
689: $(\NNzer,\widetilde{\vc{p}})$, to which the $\vc{f}^{(n_i)}$ converge in
690: distribution. Since $r>1$, the $\vc{f}^{(n_i)}$ are uniformly
691: integrable by Markov's and H\"older's inequalities. Hence, due to
692: \cite[\lem 3.11]{Kal}, their expectation values, which all equal
693: $m$, converge to the one of $\widetilde{\vc{f}}$, which is thus $m$ as
694: well. Consider now the random variables $\vc{g}^{(n_i)} =
695: \widetilde{\vc{g}} = (|k-m|^r)_{k\in\NNzer}^{}$ on
696: $(\NNzer,\vc{p}^{(n)})$ and $(\NNzer,\widetilde{\vc{p}})$, respectively.
697: The expectation values of the $\vc{g}^{(n_i)}$ are bounded by $C$,
698: which, again by \cite[\lem 3.11]{Kal}, is then also an upper bound
699: for the expectation value of $\widetilde{\vc{g}}$ (to which the
700: $\vc{g}^{(n_i)}$ converge in distribution). This proves the
701: compactness of $\M_{1,m,C}^+$. The convexity is obvious.
702:
703: \Wrt the second statement, consider
704: \begin{equation}
705: \label{eq:ur0dr0}
706: \begin{aligned}
707: M_s(\R_0(\vc{p})) &=
708: \sum_{i\ge0} \!
709: \sum_{\substack{k,\ell\ge0 \\[0.4\baselineskip]
710: k\wedge\ell \le i \le k\vee\ell}} \!
711: \frac{|i-m|^s}{1+|k-\ell|} \, p_k^{} \, p_\ell^{} \\
712: &= \sum_{k,\ell\ge0} \frac{p_k^{} \, p_\ell^{}}{1+|k-\ell|}
713: \, \frac12 \sum_{i=k\wedge\ell}^{k\vee\ell}
714: (|i-m|^s + |k+\ell-i-m|^s) \es.
715: \end{aligned}
716: \end{equation}
717: For notational convenience, let $j = k+\ell-i$. We now show
718: \begin{equation}
719: \label{eq:ur0ijklm}
720: |i-m|^s + |k+\ell-i-m|^s \le |k-m|^s + |\ell-m|^s \es.
721: \end{equation}
722: If $\{k,\ell\} = \{i,j\}$, then \eref{ur0ijklm} holds with equality.
723: Otherwise, assume, without loss of generality, that $k < i \le j <
724: \ell$. If $m \le k$ or $m \ge \ell$, we have equality for $s=1$ but
725: a strict inequality for $s=r$ due to the convexity of $x \mapsto
726: x^r$. (For $s=1$, this describes the fact that a recombination
727: event between two sequences that are both longer or both shorter
728: than the mean does not change their mean distance to the mean copy
729: number.) In the remaining cases, the inequality is strict as well.
730: Hence, $M_s(\R_0(\vc{p})) \le M_s(\vc{p})$ with equality if and only
731: if $\vc{p}$ is a fixed point of $\R_0$, since otherwise the sum in
732: \eref{ur0dr0} contains at least one term for which \eref{ur0ijklm}
733: holds as a strict inequality.
734:
735: To see that $M_1$ is continuous, select a converging sequence
736: $(\vc{p}^{(n)})$ in $\M_{1,m,C}^+$ and consider the random variables
737: $\vc{h}^{(n)} = (|k-m|)_{k\in\NNzer}^{}$ on $(\NNzer,\vc{p}^{(n)})$.
738: As above, the latter are uniformly integrable, from which the
739: continuity of $M_1$ follows. Since $M_1(\vc{p})$ is linear in
740: $\vc{p}$ and thus infinitely differentiable, so is the solution
741: $\vc{p}(t)$ for every initial condition $\vc{p}_0 \in \M_{1,m,C}^+$,
742: compare \cite[\thm 9.5 and \remk 9.6(b)]{Ama}. Therefore, we have
743: \begin{displaymath}
744: \dot{M_1}(\vc{p}_0) =
745: \liminf_{t\to0^+} \frac{M_1(\vc{p}(t)) - M_1(\vc{p}_0)}{t} =
746: M_1(\R_0(\vc{p}_0)) - M_1(\vc{p}_0) \le 0 \es,
747: \end{displaymath}
748: again with equality if and only if $\vc{p}_0$ is a fixed point.
749: Thus, $M_1$ is a Lyapunov function.
750: \end{proof}
751: \begin{theopargself}
752: \begin{proof}[of \thref{ur0conv}]
753: By assumption, the $r$-th moment of $\vc{p}(0)$ exists, which is
754: equivalent to the existence of the centered $r$-th moment by
755: Minkowski's inequality \cite[\sect II.6.6]{Shir}. This obviously
756: implies the existence of the mean $m$. By \lref{ur0dp},
757: $\vc{p}(t) \in \M_{1,m,C}^+$ follows for all $t\ge0$,
758: directly for discrete time and via a satisfied subtangent
759: condition \cite[\thm VI.2.1]{Mar} (see also \cite[\thm 16.5]{Ama})
760: for continuous time. In the discrete case, due to the compactness
761: of $\M_{1,m,C}^+$, there is a convergent subsequence
762: $(\vc{p}(t_i))$ with some limit $\vc{p}$. Consider now the mean
763: distance $M_1$ to the mean copy number from \eref{ur0dp}. If
764: $\lim_{t\to\infty} \vc{p}(t) = \vc{p}$, we have, due to the
765: continuity of $M_1$ and $\R_0$,
766: \begin{displaymath}
767: M_1(\R_0(\vc{p})) = \lim_{t\to\infty} M_1(\R_0(\vc{p}(t))) =
768: \lim_{t\to\infty} M_1(\vc{p}(t+1)) = M_1(\vc{p}) \es,
769: \end{displaymath}
770: thus $\vc{p}$ is a fixed point by \lref{ur0dp}. Otherwise, there
771: are two convergent subsequences $(\vc{p}(t_i))$, with limit
772: $\vc{p}$, and $(\vc{p}(s_i))$, with limit $\vc{q}$, whose indices
773: alternate, $t_i < s_i < t_{i+1}$. Then, we also have
774: $M_1(\R_0(\vc{p}(t_i))) \ge M_1(\vc{p}(s_i))$ and
775: $M_1(\R_0(\vc{p}(s_i))) \ge M_1(\vc{p}(t_{i+1}))$, and therefore
776: \begin{multline*}
777: M_1(\vc{p}) \ge M_1(\R_0(\vc{p})) =
778: \lim_{i\to\infty} M_1(\R_0(\vc{p}(t_i))) \ge
779: \lim_{i\to\infty} M_1(\vc{p}(s_i))
780: = M_1(\vc{q}) \\
781: \ge M_1(\R_0(\vc{q})) =
782: \lim_{i\to\infty} M_1(\R_0(\vc{p}(s_i))) \ge
783: \lim_{i\to\infty} M_1(\vc{p}(t_{i+1})) = M_1(\vc{p}) \es.
784: \end{multline*}
785: Thus, both $\vc{p}$ and $\vc{q}$ are fixed points by \lref{ur0dp}
786: and hence equal by \pref{ur0unique}. In continuous time, the
787: claim follows from \thref{lyapunov} since $M_1$ is a Lyapunov
788: function by \lref{ur0dp}.
789: \end{proof}
790: \end{theopargself}
791:
792: Note that, for $q=0$, the recombinator can be expressed in terms of
793: explicit frequencies $\pi_{k,\ell}$ of fragment pairs before
794: concatenation (with copy numbers $k$ and $\ell$) as $\R_0(\vc{p})_i^{}
795: = \sum_{j=0}^i \pi_{j,i-j}$. However, we have, so far, not been able
796: to use this for a simplification of the above treatment.
797:
798:
799: \section{Alternative probability representations}
800: \label{sec:urrepr}
801:
802: Our next goal is to find the analogue of \thref{ur0conv} for the case of
803: $q=1$ (random UC).
804: Whereas the convergence arguments for the case $q=0$ relied on a compact set
805: of probability measure defined via the $r$-th moment, we are not (yet) able to
806: extend this approach to $q>0$. Instead, we will consider, as an alternative
807: representation for a probability measure $\vc{p} \in \M_1^+$, the generating
808: function
809: \begin{equation}
810: \label{eq:urgpsip}
811: \psi(z) = \sum_{k\ge0} p_k^{} z^k \es,
812: \end{equation}
813: for which $\psi(1) = \|\vc{p}\|^{}_1 = 1$ and the radius of
814: convergence is at least 1. We will restrict our discussion to such
815: $\vc{p}$ for which $\limsup_{k\to\infty} \sqrt[k]{p_k^{}} < 1$. Then,
816: the radius of convergence, $\rho(\psi) = 1/\limsup_{k\to\infty}
817: \sqrt[k]{p_k^{}}$ by Hadamard's formula \cite[10.5]{RudRCA}, is larger
818: than $1$. This is, biologically, no restriction since for any
819: `realistic' system there are only finitely many non-zero $p_k^{}$ (and
820: therefore $\rho(\psi) = \infty$). Mathematically, this condition
821: ensures the existence of all moments and enables us to go back and
822: forth between the probability measure and its generating function,
823: even when looked at $\psi(z)$ only in the vicinity of $z=1$ (see
824: \pref{urgainxad} below and \cite[\sect II.12]{Shir}). By abuse of
825: notation, we define the induced recombinator for these generating
826: functions as
827: \begin{equation}
828: \label{eq:urgrecopsi}
829: \R(\psi)(z) = \sum_{k\ge0} \R(\vc{p})_k^{} \, z^k \es.
830: \end{equation}
831: In general, with the exception of the case $q=1$, we do not know any
832: simple expression for $\R(\psi)$ in terms of $\psi$. Nevertheless,
833: \eref{urgrecopsi} will be central to our further analysis.
834:
835: It is advantageous to use the local expansion around $z=1$, written in
836: the form
837: \begin{equation}
838: \label{eq:urgpsi1}
839: \psi(z) = \sum_{k\ge0} (k+1) a_k (z-1)^k \es,
840: \end{equation}
841: whose coefficients are given by
842: \begin{equation}
843: \label{eq:urga}
844: a_k = \frac{1}{(k+1)!} \psi^{(k)}(1) =
845: \frac{1}{k+1} \sum_{\ell \ge k} \binom{\ell}{k} \, p_\ell^{} =:
846: \vc{a}(\vc{p})_k^{} \ge 0 \es.
847: \end{equation}
848: In particular, $a_0 = 1$ and $a_1 = \frac12 \sum_{\ell\ge0} \ell \,
849: p_\ell^{}$. This definition of $a_k$ is size biased, and will become
850: clear from the simplified dynamics for $q=1$ that results from it.
851: For the sake of compact notation, we use $\vc{a} =
852: (a_k)^{}_{k\in\NNzer}$ both for the coefficients and for the mapping.
853: The coefficients $\vc{a}$ are elements of the following compact,
854: convex metric space.
855: \begin{definition}
856: \label{def:xad}
857: For fixed\/ $\alpha$ and\/ $\delta$ with\/ $0 < \alpha \le \delta <
858: \infty$, let
859: \begin{displaymath}
860: \X_{\alpha,\delta} =
861: \{ \vc{a} = (a_k)_{k\in\NNzer} : a_0 = 1,\, a_1 = \alpha,\,
862: \text{$0 \le a_k \le \delta^k$ for $k\ge2$} \} \es.
863: \end{displaymath}
864: On this space, define the metric
865: \begin{equation}
866: \label{eq:urgmetric}
867: d(\vc{a},\vc{b}) = \sum_{k\ge0} d_k |a_k-b_k|
868: \end{equation}
869: with\/ $d_k = (\gamma/\delta)^k$ for some\/ $0<\gamma<\frac13$.
870: \end{definition}
871: It is obvious that $d$ is indeed a metric and that
872: $\X_{\alpha,\delta}$ is a convex set, \ie we have $\eta\,\vc{a} +
873: (1-\eta)\vc{b} \in \X_{\alpha,\delta}$ for all $\vc{a}$, $\vc{b} \in
874: \X_{\alpha,\delta}$ and $\eta \in \ccint{0,1}$. Note that we use the
875: same symbol $d$ as in \eref{urgpmetric} since it will always be clear
876: which metric is meant. The space $\X_{\alpha,\delta}$ is naturally
877: embedded in the Banach space (\cf \cite[\sect 24.I]{WalEn})
878: \begin{equation}
879: \label{eq:hdelta}
880: H_{\gamma/\delta} = \{ \vc{x}\in\RR^{\NNzer} : \|\vc{x}\| < \infty \}
881: \end{equation}
882: with the norm $\|\vc{x}\| = \sum_{k\ge0} (\gamma/\delta)^k |x_k|$, for
883: $\gamma$ and $\delta$ as in \dref{xad}. In particular,
884: $d(\vc{a},\vc{b}) = \|\vc{a}-\vc{b}\|$. Furthermore, we have the
885: following two propositions.
886: \begin{proposition}
887: \label{prop:urgxadcomp}
888: The space\/ $\X_{\alpha,\delta}$ is compact in the metric\/ $d$ of
889: \eref{urgmetric}.
890: \end{proposition}
891: \begin{proof}
892: In metric spaces, compactness and sequential compactness are
893: equivalent, compare \cite[\thm II.3.8]{Lan}. Hence, let
894: $(\vc{a}^{(n)})$ be any sequence in $\X_{\alpha,\delta}$. By
895: assumption, $a^{(n)}_0 \equiv 1$ and $a^{(n)}_1 \equiv \alpha$.
896: Furthermore, each element sequence $(a_k^{(n)}) \subset
897: \ccint{0,\delta^k}$ has a convergent subsequence. We now
898: inductively define, for every $k$, a convergent subsequence
899: $(a_k^{(n_{k,i})})$, with limit $a_k$, such that the indices
900: $\{n_{k,i} : i\in\NN\}$ are a subset of the preceding indices
901: $\{n_{k-1,i} : i\in\NN\}$. This way, we can proceed to a `diagonal'
902: sequence $(\vc{a}^{(n_{i,i})})$. The latter is now shown to
903: converge to $\vc{a} = (a_k)$, which is obviously an element of
904: $\X_{\alpha,\delta}$. To this end, let $\eps>0$ be given. Choose
905: $m$ large enough such that $\sum_{k>m} (2\gamma)^k < \eps/2$, and
906: then $i$ such that $\sum_{k=0}^m d_k^{} |a^{(n_{i,i})}_k - a_k^{}| <
907: \eps/2$. Then
908: \begin{equation}
909: \label{eq:urgxadcomp}
910: \begin{aligned}
911: d(\vc{a}^{(n_{i,i})},\vc{a}) &=
912: \sum_{k=2}^m d_k |a_k^{(n_{i,i})}-a_k^{}| +
913: \sum_{k>m} d_k |a_k^{(n_{i,i})}-a_k^{}| \\
914: &< \frac\eps2 + \sum_{k>m} (2\gamma)^k <
915: \eps \es,
916: \end{aligned}
917: \end{equation}
918: which proves the claim.
919: \end{proof}
920: \begin{proposition}
921: \label{prop:urgainxad}
922: If\/ $\limsup_{k\to\infty} \sqrt[k]{p_k^{}} < 1$, the coefficients\/
923: $a_k$ from \eref{urga} exist and\/ $\vc{a}(\vc{p}) \in
924: \X_{\alpha,\delta}$ with\/ $\alpha = \vc{a}(\vc{p})^{}_1 = \frac12 m
925: = \frac12 \sum_{k\ge0} k\,p_k^{}$ and some\/ $\delta$. Conversely,
926: if\/ $\vc{p}(\vc{a}) \in \X_{\alpha,\delta}$ for some\/ $\alpha$,
927: $\delta$, one has\/ $\limsup_{k\to\infty} \sqrt[k]{p_k^{}} < 1$.
928: \end{proposition}
929: For a proof, we need the following
930: \begin{lemma}
931: \label{lem:urgradii}
932: Let\/ $f_0(z) = \sum_{k\ge0} c_k z^k$ be a power series with
933: non-negative coefficients\/ $c_k$ and\/ $f_x(z) = \sum_{k\ge0}
934: \frac{1}{k!} f_0^{(k)}(x) (z-x)^k$ the expansion of\/ $f_0$ around
935: some\/ $x \in \coint{0,\rho(f_0)}$. Then, $\rho(f_0) = x +
936: \rho(f_x)$, including the case that both radii of convergence are
937: infinite.
938: \end{lemma}
939: \begin{proof}
940: Since the open disc $\B_x(\rho(f_0)-x)$ is entirely included in
941: $\B_0(\rho(f_0))$, the inequality $\rho(f_x) \ge \rho(f_0)-x$
942: immediately follows from the theorem of representability by power
943: series \cite[\thm 10.16]{RudRCA}. Consider now the power series
944: $f_{xe^{i\varphi}}(z) = \sum_{k\ge0} \frac{1}{k!}
945: f_0^{(k)}(xe^{i\varphi}) (z-xe^{i\varphi})^k$ with arbitrary
946: $\varphi\in\coint{0,2\pi}$. Its coefficients satisfy
947: $|f_0^{(k)}(xe^{i\varphi})| \le \sum_{n \ge k} \frac{n!}{(n-k)!}
948: c_k x^{n-k} = f_0^{(k)}(x)$ due to the non-negativity of the $c_k$.
949: This implies $\rho(f_{xe^{i\varphi}}) \ge \rho(f_x)$ by Hadamard's
950: formula. Therefore, $f_0$ admits an analytic continuation on
951: $\B_0(x+\rho(f_x))$, the uniqueness of which follows from the
952: monodromy theorem \cite[\thm 16.16]{RudRCA}. The theorem of
953: representability by power series then implies the inequality
954: $\rho(f_0) \ge x+\rho(f_x)$, which, together with the opposite
955: inequality above, proves the claim.
956: \end{proof}
957: \begin{theopargself}
958: \begin{proof}[of \pref{urgainxad}]
959: The assumption implies $\rho(\psi) > 1$ for $\psi$ from
960: \eref{urgpsip}. Then, from \lref{urgradii}, we know that
961: $\limsup_{k\to\infty} \sqrt[k]{(k+1)a_k} < \infty$. Since
962: furthermore $a_k \le (k+1) a_k$, also $\limsup_{k\to\infty}
963: \sqrt[k]{a_k} < \infty$, so there is an upper bound $\delta$ for
964: $\sqrt[k]{a_k}$ and thus $\vc{a}(\vc{p}) \in \X_{\alpha,\delta}$.
965: The converse statement follows from \eref{urga} and
966: \lref{urgradii}.
967: \end{proof}
968: \end{theopargself}
969: Therefore, any mapping from $\X_{\alpha,\delta}$ into itself that is
970: continuous \wrt the metric $d$ from \eref{urgmetric} has a fixed point
971: by the Leray--Schauder--Tychonov theorem \cite[\thm V.19]{ReSi}.
972:
973: Note further that each $\X_{\alpha,\delta}$ contains a maximal element
974: \wrt the partial order $\vc{a}\le\vc{b}$ defined by $a_k \le b_k$ for
975: all $k\in\NNzer$, which is given by
976: $(1,\alpha,\delta^2,\delta^3,\ldots)$. This property finally leads to
977: \begin{proposition}
978: \label{prop:urgpadcomp}
979: The space\/ $P_{\alpha,\delta} := \{ \vc{p}\in\M_1^+ :
980: \vc{a}(\vc{p})\in\X_{\alpha,\delta} \}$, equipped with \pr{the
981: metric induced by} the total variation norm, is compact and
982: convex.
983: \end{proposition}
984: The proof is based on the following two lemmas.
985: \begin{lemma}
986: \label{lem:urglocbound}
987: For any subset of\/ $P_{\alpha,\delta}$, the corresponding
988: generating functions from \eref{urgpsip} are locally bounded on\/
989: $\B_{1+1/\delta}(0)$.
990: \end{lemma}
991: \begin{proof}
992: It is sufficient to show boundedness on every compact $K \subset
993: \B_{1+1/\delta}(0)$, see \cite[\sect 7.1]{Rem}. Thus, let such a
994: $K$ be given and fix $r \in \coint{0,\frac1\delta}$ so that $K$ is
995: contained in $\overline{\B_{1+r}(0)}$. Then, for every $\vc{p} \in
996: P_{\alpha,\delta}$ and every $z \in K$,
997: \begin{align*}
998: |\psi(z)| &= \Bigl|\sum_{k\ge0} p_k^{} z^k\Bigr| \le
999: \sum_{k\ge0} p_k^{} (1+r)^k = \psi(1+r) \\
1000: &= \sum_{k\ge0} (k+1) \vc{a}(\vc{p})_k^{} \, r^k \le
1001: 1 + 2\,\alpha\,r + \sum_{k\ge2} (k+1) (r\delta)^k <
1002: \infty \es,
1003: \end{align*}
1004: where $r\delta < 1$ was used. This needed to be shown.
1005: \end{proof}
1006: \begin{lemma}
1007: \label{lem:urgpnormconv}
1008: If, for a sequence\/ $(\vc{p}^{(n)}) \subset P_{\alpha,\delta}$, the
1009: coefficients\/ $\vc{a}^{(n)} = \vc{a}(\vc{p}^{(n)})$ from
1010: \eref{urga} converge to some\/ $\vc{a}$ \wrt the metric\/ $d$ from
1011: \eref{urgmetric}, then the generating functions\/ $\psi_n$ from
1012: \eref{urgpsip} converge compactly to some\/ $\psi$ with\/ $\psi(z) =
1013: \sum_{k\ge0} p_k^{} z^k$ and the\/ $\vc{p}^{(n)}$ thus converge in
1014: norm to\/ $\vc{p} \in P_{\alpha,\delta}$.
1015: \end{lemma}
1016: \begin{proof}
1017: By to \lref{urglocbound}, the sequence $(\psi_n)$ is locally
1018: bounded in $\B_{1+1/\delta}(0)$. Due to the pointwise convergence
1019: $|a^{(n)}_k - a^{\vp}_k| \le d_k^{-1} d(\vc{a}^{(n)},
1020: \vc{a}) \to 0$, we have
1021: \begin{displaymath}
1022: \psi^{(k)}_n(1) = (k+1)! \, a^{(n)}_k
1023: \;\xrightarrow{n\to\infty}\;
1024: (k+1)! \, a^{\vp}_k = \psi^{(k)}(1)
1025: \end{displaymath}
1026: for every $k\in\NNzer$. Then, the compact convergence $\psi_n \to
1027: \psi$ follows from Vitali's theorem \cite[\thm 7.3.2]{Rem}. In
1028: particular, this implies that $p^{(n)}_k \to p_k^{\vp} \ge 0$ and $1
1029: = \sum_{k\ge0} p^{(n)}_k = \psi_n(1) \to \psi(1) = \sum_{k\ge0}
1030: p_k^{\vp}$, thus $\vc{p} \in \M_1^+$.
1031:
1032: Now, choose $r \in \ooint{1,1+\frac1\delta}$. Then there is, for
1033: every $\eps>0$, an $n_\eps$ such that $\sup_{|z| \le r}
1034: |\psi(z)-\psi_n(z)| < \eps$ for all $n \ge n_\eps$. This implies
1035: \begin{displaymath}
1036: |p^{(n)}_k - p_k^{\vp}| =
1037: \left|\frac{1}{2\pi i} \oint_{|z|=r}
1038: \frac{\psi_n(z) - \psi(z)}{z^{k+1}} \dd{z} \right| < \frac{\eps}{r^k}
1039: \end{displaymath}
1040: for all $n \ge n_\eps$ by Cauchy's integral formula \cite[\thm
1041: 7.3]{LanCA}. Now, let $\eps>0$ be given. Then
1042: \begin{displaymath}
1043: \|\vc{p}^{(n)} - \vc{p}\|^{}_1 =
1044: \sum_{k\ge0} |p^{(n)}_k - p_k^{\vp}| <
1045: \eps \frac{1}{1-\tfrac1r}
1046: \end{displaymath}
1047: for all $n \ge n_\eps$, which proves the claim.
1048: \end{proof}
1049: \begin{theopargself}
1050: \begin{proof}[of \pref{urgpadcomp}]
1051: Let $(\vc{p}^{(n)})$ denote an arbitrary sequence in
1052: $P_{\alpha,\delta}$ and \linebreak $(\vc{a}^{(n)}) =
1053: (\vc{a}(\vc{p}^{(n)}))$ the corresponding sequence in
1054: $\X_{\alpha,\delta}$. Due to \pref{urgxadcomp}, there is a
1055: convergent subsequence $(\vc{a}^{(n_i)})^{}_i$. Then, by
1056: \lref{urgpnormconv}, $(\vc{p}^{(n_i)})^{}_i$ converges in norm to
1057: some $\vc{p} \in P_{\alpha,\delta}$. This proves the compactness
1058: property. The convexity of $P_{\alpha,\delta}$ is a simple
1059: consequence of the convexity of $\M_1^+$, the linearity of the
1060: mapping $\vc{a}$, and the convexity of $\X_{\alpha,\delta}$.
1061: \end{proof}
1062: \end{theopargself}
1063:
1064: Another property of the mapping $\vc{a} \colon P_{\alpha,\delta} \to
1065: \X_{\alpha,\delta}$ is stated in
1066: \begin{lemma}
1067: \label{lem:urgacont}
1068: For every\/ $\alpha$ and\/ $\delta$, the mapping\/ $\vc{a} \colon
1069: P_{\alpha,\delta} \to \X_{\alpha,\delta}$ from \eref{urga} is
1070: continuous \pr{\wrt the total variation norm and the metric\/ $d$}
1071: and injective. Its inverse\/ $\vc{p} \colon
1072: \vc{a}(P_{\alpha,\delta}) \to P_{\alpha,\delta}$ is continuous as
1073: well.
1074: \end{lemma}
1075: \begin{proof}
1076: Let $\vc{p}, \vc{q} \in P_{\alpha,\delta}$ and assume
1077: $\vc{a}(\vc{p}) = \vc{a}(\vc{q})$. Then, as in the proof of
1078: \lref{urgradii}, the uniqueness of the generating function in
1079: $\B_{1+1/\delta}(0)$ follows, and thus $\vc{p} = \vc{q}$, which
1080: proves the injectivity of $\vc{a}$. The other statements follow
1081: from Vitali's theorem \cite[\thm 7.3.2]{Rem}: Norm convergence of a
1082: sequence $(\vc{p}^{(n)})$ in $P_{\alpha,\delta}$ to some $\vc{p}$
1083: implies convergence of its element sequences and thus compact
1084: convergence of the corresponding generating functions $\psi_n$ to
1085: $\psi$, which is given by $\psi(z) = \sum_{k\ge0} p_k^{}\,z^k$.
1086: This, in turn, implies convergence of each sequence
1087: $(\vc{a}(\vc{p}^{(n)})_k)$ to $\vc{a}(\vc{p})_k^{}$, from which, as
1088: in \eref{urgxadcomp}, the convergence $(\vc{a}(\vc{p}^{(n)})) \to
1089: \vc{a}(\vc{p})$ (\wrt $d$) follows. The converse is the statement
1090: of \lref{urgpnormconv} (see also \cite[\prop 1.6.8]{Ped}).
1091: \end{proof}
1092: Note that, if $\rho(\psi)>2$, the inverse of the mapping $\vc{a}$ is
1093: given by
1094: \begin{displaymath}
1095: \vc{p}(\vc{a})_k^{}
1096: = \sum_{\ell \ge k} (-1)^{\ell-k} \binom{\ell}{k} (\ell+1) \, a_\ell^{} \es.
1097: \end{displaymath}
1098:
1099:
1100: \section{Random unequal crossover}
1101: \label{sec:urq1}
1102:
1103: Let us now turn to the random UC model, described by $q=1$ in
1104: \eref{urtq}. Here, the recombinator \eref{reco} simplifies to
1105: \cite[(3.1)]{ShAt}
1106: \begin{equation}
1107: \label{eq:ur1reco}
1108: \R_1(\vc{p})_i^{} =
1109: \sum_{\substack{k,\ell\ge0 \\[0.4\baselineskip] k+\ell \ge i}}
1110: \frac{1+\min\{k,\ell,i,k+\ell-i\}}{(k+1)(\ell+1)}
1111: \, p_k^{} \, p_\ell^{} \es.
1112: \end{equation}
1113: As for internal UC, by \lref{urrev}, the reversibility condition
1114: \eref{urrev} directly leads to an expression for fixed points,
1115: \begin{displaymath}
1116: \frac{p_{k\vphantom{j}}^{}}{k+1} \frac{p_{\ell\vphantom{j}}^{}}{\ell+1} =
1117: \frac{p_{i\vphantom{j}}^{}}{i+1} \frac{p_j^{}}{j+1}
1118: \quad\text{for all } k+\ell = i+j \es.
1119: \end{displaymath}
1120: This has $p_k^{} = C (k+1) x^k$ as a solution, with appropriate
1121: parameter $x$ and normalization constant $C$. Again, it turns out
1122: that all fixed points are given this way.
1123: \begin{proposition}{\cite[\thm A.2]{ShAt}}
1124: \label{prop:ur1fixed}
1125: Every fixed point\/ $\vc{p} \in \M_1^+$ of\/ $\R_1$ is of the form
1126: \begin{equation}
1127: \label{eq:ur1fixed}
1128: p_k^{} =
1129: \left(\frac{2}{m+2}\right)^2 (k+1) \left(\frac{m}{m+2}\right)^k \es,
1130: \end{equation}
1131: where\/ $m = \sum_{k\ge0} k \, p_k^{} \ge 0$.
1132: \hspace*{\fill}\qed
1133: \end{proposition}
1134: One can verify this in several ways, one being a direct inductive
1135: calculation.
1136:
1137: The main result of this section is
1138: \begin{theorem}
1139: \label{thm:ur1conv}
1140: Assume that\/ $\limsup_{k\to\infty} \sqrt[k]{p_k^{}(0)} < 1$. Then,
1141: both in discrete and in continuous time,\/ $\lim_{t\to\infty}
1142: \|\vc{p}(t) - \vc{p}\|^{}_1 = 0$ with the appropriate fixed point\/
1143: $\vc{p}$ from \pref{ur1fixed}.
1144: \end{theorem}
1145: For a proof, we consider the following alternative process, verbally
1146: described in \cite[p.\ 720f]{ShAt}. It is a two-step stick breaking
1147: and glueing procedure which ultimately induces the same (deterministic)
1148: dynamics as random UC, even though the underlying process is rather different.
1149: This will lead to a simple expression for the
1150: induced recombinator of the coefficients $\vc{a}$ from \eref{urga},
1151: which allows for an explicit solution.
1152: \begin{proposition}
1153: Let\/ $\vc{p} \in \M_1^+$. Then,
1154: \begin{equation}
1155: \label{eq:ur1iterqp}
1156: \pi_k^{} = \sum_{\ell \ge k} \frac{1}{\ell+1} p_\ell^{}
1157: \end{equation}
1158: gives a probability measure\/ $\vc{\pi} \in \M_1^+$, and the
1159: recombinator can be written as
1160: \begin{equation}
1161: \label{eq:ur1iterpq}
1162: \R_1(\vc{p})_i^{} =
1163: \sum_{j=0}^i \pi_j^{} \, \pi_{i-j}^{} =
1164: (\vc{\pi}*\vc{\pi})_i^{} \es,
1165: \end{equation}
1166: where\/ $*$ denotes the convolution in\/ $\Ll{1}(\NNzer)$.
1167: \end{proposition}
1168: Here, \eref{ur1iterqp} describes a breaking process in which, without any pairing,
1169: each sequence is cut equally likely between any two of its building blocks. In a
1170: second step, described by \eref{ur1iterpq}, these fragments are paired
1171: randomly and joined (or `glued').
1172: \begin{proof}
1173: It is easily verified that $\vc{\pi}$ is normalized to $1$. \Wrt
1174: \eref{ur1iterpq}, note the following identity for $k+\ell \ge i$,
1175: \begin{displaymath}
1176: |\{j : (i-\ell) \vee 0 \le j \le i \wedge k\}| =
1177: 1+\min\{k,\ell,i,k+\ell-i\} \es,
1178: \end{displaymath}
1179: which can be shown by treating the four cases on the LHS separately.
1180: With this, inserting \eref{ur1iterqp} into the RHS of
1181: \eref{ur1iterpq} yields
1182: \begin{align*}
1183: \sum_{j=0}^i \pi_j^{} \, \pi_{i-j}^{} &=
1184: \sum_{j=0}^i \sum_{k \ge j} \sum_{\ell \ge i-j}
1185: \frac{1}{(k+1)(\ell+1)} \, p_k^{} \, p_\ell^{} \\
1186: &= \sum_{\substack{k,\ell \ge 0 \\[0.4\baselineskip] k+\ell \ge i}}
1187: \frac{1}{(k+1)(\ell+1)} \, p_k^{} \, p_\ell^{}
1188: \sum_{j=(i-\ell)\vee0}^{i \wedge k} 1 \\
1189: &= \sum_{\substack{k,\ell\ge0 \\[0.4\baselineskip] k+\ell \ge i}}
1190: \frac{1+\min\{k,\ell,i,k+\ell-i\}}{(k+1)(\ell+1)}
1191: \, p_k^{} \, p_\ell^{} \,=\,
1192: \R_1(\vc{p})_i^{} \es. \qedhere[-23.4pt]
1193: \end{align*}
1194: \end{proof}
1195:
1196: This nice structure has an analogue on the level of the generating
1197: functions.
1198: \begin{proposition}
1199: Under the assumptions of \thref{ur1conv},
1200: let\/ $\phi(z) = \sum_{k\ge0} \pi_k z^k$ denote the generating
1201: function for\/ $\vc{\pi}$ from \eref{ur1iterqp}. Then,
1202: \begin{displaymath}
1203: \phi(z) = \frac{1}{1-z} \int_z^1 \psi(\zeta) \dd{\zeta}
1204: \qquad\text{and}\qquad
1205: \R_1(\psi)(z) = \phi(z)^2 \es.
1206: \end{displaymath}
1207: \end{proposition}
1208: \begin{proof}
1209: Recall that $\psi(z) = \sum_{k\ge 0} p_k z^k$.
1210: \Eqs \eref{ur1iterqp} and \eref{ur1iterpq} lead to
1211: \begin{align*}
1212: \phi(z) &=
1213: \sum_{k\ge0} \sum_{\ell \ge k} \frac{1}{\ell+1} p_\ell^{} z^k =
1214: \sum_{\ell\ge0} \frac{1}{\ell+1} p_\ell^{} \sum_{k\le\ell} z^k =
1215: \sum_{\ell\ge0} \frac{1}{\ell+1} p_\ell^{} \frac{1-z^{\ell+1}}{1-z}\\
1216: &= \frac{1}{1-z} \sum_{\ell\ge0} p_\ell^{} \frac{1-z^{\ell+1}}{\ell+1} =
1217: \frac{1}{1-z} \sum_{\ell\ge0} p_\ell^{} \int_z^1 \zeta^\ell \dd{\zeta} =
1218: \frac{1}{1-z} \int_z^1 \psi(\zeta) \dd{\zeta}
1219: \end{align*}
1220: and, due to absolute convergence of the series involved,
1221: \begin{align*}
1222: \R_1(\psi)(z) &= \sum_{k\ge0} \R_1(\vc{p})_k^{} z^k =
1223: \sum_{k\ge0} z^k \sum_{\ell=0}^k \pi_\ell \pi_{k-\ell} \\ &=
1224: \sum_{\ell\ge0} \pi_\ell z^\ell \sum_{k\ge\ell} \pi_{k-\ell} z^{k-\ell} =
1225: \phi(z)^2 \es. \qedhere[-13.9pt]
1226: \end{align*}
1227: \end{proof}
1228: The following lemma states that the radius of convergence of $\psi$
1229: does not decrease under the random UC dynamics. Thus, it is ensured
1230: that, if $\rho(\psi) > 1$, also $\R_1(\psi)$ may be described by an
1231: expansion at $z=1$, \ie by coefficients $\vc{a}$.
1232: \begin{lemma}
1233: \label{lem:ur1rhotildepsi}
1234: The radius of convergence of\/ $\R_1(\psi)$ is\/ $\rho(\R_1(\psi))
1235: \ge \rho(\psi)$.
1236: \end{lemma}
1237: \begin{proof}
1238: As $1/\rho(\psi) = \limsup_{k\to\infty} \sqrt[k]{p_k^{}} =: x \le 1$
1239: and $\lim_{k\to\infty} \sqrt[k]{k+1} = 1$, there is a constant $C>0$
1240: with $p_k^{} \le C (k+1) x^k$ for all $k$. Note the identity
1241: \begin{displaymath}
1242: \sum_{j=0}^n (1+\min\{i,j,n-i,n-j\}) =(i+1)(n-i+1)
1243: \end{displaymath}
1244: for $i \le n$, which follows from an elementary calculation. Then,
1245: \eref{ur1reco} implies
1246: \begin{align*}
1247: \R_1(\vc{p})_i^{} &\le
1248: C^2 \sum_{\substack{k,\ell\ge0 \\ k+\ell \ge i}}
1249: x^{k+\ell} (1+\min\{k,\ell,i,k+\ell-i\}) \\
1250: &= C^2 \sum_{n \ge i} x^n \sum_{j=0}^n (1+\min\{i,j,n-i,n-j\}) \\
1251: &= C^2 (i+1) x^i \sum_{\ell\ge0} (\ell+1) \, x^\ell =
1252: \left(\frac{C}{1-x}\right)^2 (i+1) \, x^i \es.
1253: \end{align*}
1254: Accordingly, $\limsup_{k\to\infty} \sqrt[k]{\R_1(\vc{p})_k^{}} \le x
1255: \le 1$, which proves the claim.
1256: \end{proof}
1257:
1258: These results enable us to derive the following expression for the
1259: coefficients $\vc{a}$, using the expansion of \eref{urgpsi1}:
1260: \begin{equation}
1261: \label{eq:ur1recopsi}
1262: \begin{aligned}
1263: \R_1(\psi)(z) &=
1264: \left[ \frac{1}{z-1} \int_1^z \psi(\zeta) \dd{\zeta} \right]^2 \\
1265: &= \left[ \sum_{k\ge0} a_k (z-1)^k \right]^2 =
1266: \sum_{k\ge0} \left( \sum_{n=0}^k a_n a_{k-n} \right) (z-1)^k \es.
1267: \end{aligned}
1268: \end{equation}
1269: So it is natural to define the induced
1270: recombinator
1271: \begin{equation}
1272: \label{eq:ur1recoa}
1273: \Ra_1(\vc{a})_k^{} =
1274: \frac{1}{k+1} \sum_{n=0}^k a_n a_{k-n} \ge 0 \es,
1275: \end{equation}
1276: for which we have
1277: \begin{lemma}
1278: \label{lem:ur1cont}
1279: The recombinator\/ $\Ra_1$ given by \eref{ur1recoa} maps each
1280: space\/ $\X_{\alpha,\delta}$ into itself and is continuous \wrt the
1281: metric\/ $d$ from \eref{urgmetric}.
1282: \end{lemma}
1283: \begin{proof}
1284: Let $\alpha, \delta > 0$ be given and $\vc{a}, \vc{b} \in
1285: \X_{\alpha,\delta}$. Trivially, $\Ra_1(\vc{a})_0^{} = 1$ and
1286: $\Ra_1(\vc{a})_1^{} = \alpha$. For $k\ge2$, $\Ra_1(\vc{a})_k^{} =
1287: \frac{1}{k+1} \sum_{\ell \le k} a_\ell \, a_{k-\ell} \le \delta^k$.
1288: This proves the first statement. For the continuity, note first
1289: that every $\Ra_1(\vc{a})_k^{}$ with $k\ge2$ is continuous as a
1290: mapping from $\X_{\alpha,\delta}$ to $\ccint{0,\delta^k}$. Now, let
1291: $\eps>0$ be given. Choose $n$ large enough so that $\sum_{k>n}
1292: (2\gamma)^k < \eps/2$, where $\gamma$ is the parameter introduced in
1293: \dref{xad}. Then, there is an $\eta>0$ such that $\sum_{k=2}^n
1294: (\gamma/\delta)^k |\Ra_1(\vc{a})_k - \Ra_1(\vc{b})_k| < \eps/2$ for
1295: $\vc{a}, \vc{b} \in \X_{\alpha,\delta}$ with $d(\vc{a},\vc{b}) <
1296: \eta$. Thus, for such $\vc{a}$ and $\vc{b}$,
1297: \begin{displaymath}
1298: d(\Ra_1(\vc{a}), \Ra_1(\vc{b})) \le
1299: \sum_{k=0}^n \left(\frac\gamma\delta\right)^k
1300: |\Ra_1(\vc{a})_k - \Ra_1(\vc{b})_k| +
1301: \sum_{k>n} (2\gamma)^k < \eps \es,
1302: \end{displaymath}
1303: which proves the claim.
1304: \end{proof}
1305: Note that the fixed point equation on the level of the coefficients
1306: $\vc{a}$ is always satisfied for $a_0$ and $a_1$. If $k>1$, one
1307: obtains the recursion
1308: \begin{displaymath}
1309: a_k = \frac{1}{k-1} \, \sum_{n=1}^{k-1} a_n a_{k-n} \es,
1310: \end{displaymath}
1311: which shows that at most one fixed point with given mean can exist.
1312:
1313: Let us now consider the case of discrete time first. Analogously to
1314: \eref{urdisctime}, define $\vc{a}(t) = \vc{a}(\vc{p}(t))$ as the
1315: coefficients belonging to $\vc{p}(t)$, which are assumed to exist. It
1316: is clear from \eref{urgrecopsi}, \eref{ur1recopsi} and \eref{ur1recoa}
1317: that $\vc{a}(t+1) = \Ra_1(\vc{a}(t))$. We then have the following two
1318: propositions.
1319: \begin{proposition}
1320: \label{prop:ur1convdisc}
1321: Assume\/ $\vc{a}(0)$ to exist. Then, in discrete time,
1322: \begin{displaymath}
1323: \lim_{t\to\infty} a_k(t) = \alpha^k
1324: \qquad\text{for all $k\ge0$.}
1325: \end{displaymath}
1326: \end{proposition}
1327: This result indicates that a weaker condition than the one of
1328: \thref{ur1conv} may be sufficient for convergence of $\vc{p}(t)$.
1329: \begin{proof}
1330: Clearly, $a_0(t) \equiv 1$, $a_1(t) \equiv \alpha$. Furthermore, by
1331: the assumption and \eref{ur1recopsi}, the coefficients $a_k(t)$
1332: exist for all $k, t \in \NNzer$. Now, assume that the claim holds
1333: for all $k \le n$ with some $n$ and let $k=n+1$. According to the
1334: general properties of $\limsup$ and $\liminf$, we then have
1335: \begin{align*}
1336: \limsup_{t\to\infty} \, a_k(t+1)
1337: &\le \frac{1}{k+1} \sum_{\ell=0}^k \limsup_{t\to\infty}
1338: \bigl( a_\ell(t) a_{k-\ell}(t) \bigr) \\
1339: &= \frac{k-1}{k+1} \alpha^k + \frac{2}{k+1} \limsup_{t\to\infty}\,a_k(t)
1340: \end{align*}
1341: and analogously with $\ge$ for $\liminf$. This leads to
1342: \begin{displaymath}
1343: \frac{k-1}{k+1} \limsup_{t\to\infty}\,a_k(t) \le \frac{k-1}{k+1} \alpha^k
1344: \le \frac{k-1}{k+1} \liminf_{t\to\infty} a_k(t) \es,
1345: \end{displaymath}
1346: from which the claim follows for all $k \le n+1$ and, by induction
1347: over $n$, for all $k\ge0$.
1348: \end{proof}
1349: \begin{proposition}
1350: \label{prop:ur1contract}
1351: The recombinator\/ $\Ra_1$, acting on\/ $\X_{\alpha,\delta}$, is a
1352: strict contraction \wrt the metric\/ $d$ from \eref{urgmetric}, \ie
1353: there is a $C<1$ such that, for all elements $\vc{a}, \vc{b} \in
1354: \X_{\alpha,\delta}$,
1355: \begin{displaymath}
1356: d(\Ra_1(\vc{a}),\Ra_1(\vc{b})) \le C \, d(\vc{a},\vc{b}) \es.
1357: \end{displaymath}
1358: \end{proposition}
1359: \begin{proof}
1360: First consider, for $k\ge2$, without using the special choice of the
1361: $d_k$,
1362: \begin{equation}
1363: \label{eq:ur1contractineq}
1364: \begin{aligned}
1365: d(\Ra_1(\vc{a}),\Ra_1(\vc{b})) &=
1366: \sum_{k\ge2} d_k \frac{1}{k+1} \Bigl|
1367: \sum_{\ell=0}^k (a_\ell\,a_{k-\ell} - b_\ell\,b_{k-\ell}) \Bigr| \\
1368: &= \sum_{k\ge2} d_k \frac{1}{k+1} \Bigl|
1369: \sum_{\ell=0}^k (a_\ell-b_\ell) (a_{k-\ell}+b_{k-\ell}) \Bigr| \\
1370: &\le \sum_{k\ge2} d_k \frac{2}{k+1}
1371: \sum_{\ell=2}^k \delta^{k-\ell} |a_\ell-b_\ell| \\
1372: &= \sum_{\ell\ge2} d_\ell |a_\ell-b_\ell|
1373: \sum_{k\ge\ell} \frac{2}{k+1} \delta^{k-\ell} \frac{d_k}{d_\ell} \es.
1374: \end{aligned}
1375: \end{equation}
1376: With the choice $d_k = (\gamma/\delta)^k$, where we had $\gamma <
1377: \frac13$, we can find, for $\ell\ge2$, an upper bound for the inner
1378: sum,
1379: \begin{displaymath}
1380: \sum_{k\ge\ell} \frac{2}{k+1} \delta^{k-\ell} \frac{d_k}{d_\ell} \le
1381: \frac23 \sum_{k\ge\ell} \gamma^{k-\ell} = \frac{2}{3-3\gamma} =:
1382: C < 1 \es,
1383: \end{displaymath}
1384: which, together with \eref{ur1contractineq}, proves the claim.
1385: \end{proof}
1386: Together with Banach's fixed point theorem (compare \cite[\thm
1387: V.18]{ReSi}), the two propositions imply that $\vc{a}(t)$ converges to
1388: $(1,\alpha,\alpha^2,\ldots)$ \wrt the metric $d$, and that convergence
1389: is exponentially fast.
1390:
1391: In continuous time, we consider the time derivative of $\vc{a}(t) :=
1392: \vc{a}(\vc{p}(t))$, which is, by \eref{urga},
1393: \begin{equation}
1394: \label{eq:ur1ivpa}
1395: \tfrac{\dd{}}{\dd{t}} \vc{a}(t) =
1396: \tfrac{\dd{}}{\dd{t}} \vc{a}\bigl(\vc{p}(t)\bigr) =
1397: \vc{a}\bigl(\R_1(\vc{p}(t)) - \vc{p}(t)\bigr) =
1398: \Ra_1(\vc{a}(t)) - \vc{a}(t) \es.
1399: \end{equation}
1400: The following lemma ensures, together with \cite[\thm 7.6 and \remk
1401: 7.10(b)]{Ama}, that this initial value problem has a unique solution
1402: for all $\vc{a}(0) = \vc{a}_0 \in \X_{\alpha,\delta}$.
1403: \begin{lemma}
1404: \label{lem:ur1lipschitz}
1405: Consider the Banach space\/ $H_{\gamma/\delta}$ from \eref{hdelta},
1406: with some $0<\gamma<\tfrac13$, and its open subset\/ $Y = \{ \vc{x} \in
1407: H_{\gamma/\delta} : |x_k| < (2\delta)^k \}$. Then, the
1408: recombinator\/ $\Ra_1$ from \eref{ur1recoa} maps\/ $Y$ into itself,
1409: satisfies a global Lipschitz condition, and is bounded on\/ $Y.$
1410: Furthermore, it is infinitely differentiable,\/ $\Ra_1 \in
1411: \C^\infty(Y,Y)$.
1412: \end{lemma}
1413: \begin{proof}
1414: For $\vc{x} \in Y,$ one has $|x_k| < (2\delta)^k$, hence
1415: $|\Ra_1(\vc{x})_k^{}| < (2\delta)^k$, with a similar argument as in
1416: the proof of \lref{ur1cont}. Consequently, $\Ra_1(Y) \subset Y.$
1417: So, let $\vc{x}, \vc{y} \in Y.$ Then, similarly to the proof of
1418: \pref{ur1contract}, one shows the Lipschitz condition
1419: \begin{equation*}
1420: \|\Ra_1(\vc{x}) - \Ra_1(\vc{y})\| \le
1421: \sum_{\ell\ge0} \left(\frac\gamma\delta\right)^\ell |x_\ell - y_\ell|
1422: \sum_{k\ge\ell} \frac{2}{k+1} (2\gamma)^{k-\ell} \le
1423: \frac{2}{1-2\gamma} \|\vc{x}-\vc{y}\|
1424: \end{equation*}
1425: and, since $\|\vc{x}\| < 1/(1-2\gamma)$ in $Y$, the boundedness,
1426: \begin{displaymath}
1427: \|\Ra_1(\vc{x})\| \le \frac{1}{1-2\gamma} \|\vc{x}\| <
1428: \frac{1}{(1-2\gamma)^2} \es.
1429: \end{displaymath}
1430: \Wrt differentiability, consider, for sufficiently small $\vc{h} \in
1431: Y,$
1432: \begin{displaymath}
1433: \Ra_1(\vc{x}+\vc{h})_k^{} =
1434: \Ra_1(\vc{x})_k^{} + \frac{2}{k+1} \sum_{\ell=0}^k x_{k-\ell}\,h_\ell +
1435: \Ra_1(\vc{h})_k^{} \es.
1436: \end{displaymath}
1437: Since
1438: \begin{align*}
1439: \|\Ra_1(\vc{h})\| &\le
1440: \sum_{k\ge0} \left(\frac\gamma\delta\right)^k \frac{1}{k+1}
1441: \sum_{\ell=0}^k |h_{k-\ell}| \, |h_\ell| \\
1442: &= \sum_{\ell\ge0} \left(\frac\gamma\delta\right)^\ell |h_\ell|
1443: \sum_{k\ge\ell} \left(\frac\gamma\delta\right)^{k-\ell}
1444: \frac{|h_{k-\ell}|}{k+1} \le \|\vc{h}\|^2 \es,
1445: \end{align*}
1446: it is clear that $\Ra_1$ is differentiable with linear (and thus
1447: continuous) derivative, whose Jacobi matrix is explicitly
1448: $\Ra_1'(\vc{x})_{k\ell}^{} = \frac{\partial}{\partial x_\ell}
1449: \Ra_1(\vc{x})_k^{} = \frac{2}{k+1} x_{k-\ell}$ if $k\ge\ell$ and
1450: zero otherwise, hence one has $\Ra_1 \in \C^1(Y,Y)$. It is now
1451: trivial to show that $\Ra_1 \in \C^2(Y,Y)$ with constant second
1452: derivative and thus $\Ra_1 \in \C^\infty(Y,Y)$.
1453: \end{proof}
1454: \begin{proposition}
1455: \label{prop:ur1convcont}
1456: If\/ $\vc{a}_0 \in \X_{\alpha,\delta}$ for some\/ $\alpha, \delta$,
1457: then\/ $\vc{a}(t) \in \X_{\alpha,\delta}$ for all\/ $t\ge0$ and\/
1458: $\lim_{t\to\infty} d(\vc{a}(t), \vc{\alpha}) = 0$ with\/
1459: $\vc{\alpha} = (1,\alpha,\alpha^2,\alpha^3,\ldots)$.
1460: \end{proposition}
1461: \begin{proof}
1462: The first statement follows from \cite[\thm VI.2.1]{Mar} (see also
1463: \cite[\thm 16.5]{Ama}) since, due to the convexity of
1464: $\X_{\alpha,\delta}$, we have $\vc{a} + t (\Ra(\vc{a})-\vc{a}) \in
1465: \X_{\alpha,\delta}$ for every $\vc{a} \in \X_{\alpha,\delta}$ and
1466: $t\in\ccint{0,1}$, hence a subtangent condition is satisfied. For
1467: the second, observe that $\Ra_1(\vc{\alpha}) = \vc{\alpha}$. We now
1468: show that
1469: \begin{equation}
1470: \label{eq:ur1lyapunov}
1471: L(\vc{a}_0) = d(\vc{a}_0, \vc{\alpha})
1472: \end{equation}
1473: is a Lyapunov function, \cf \dref{lyapunov}. With the notation of
1474: \lref{ur1lipschitz}, note that the compact metric space
1475: $\X_{\alpha,\delta}$ is contained in the open subset $Y$ of the
1476: Banach space $H_{\gamma/\delta}$. The continuity of $L$ is obvious.
1477: Now, let $\vc{a}_0 \in \X_{\alpha,\delta}$ be given. By
1478: \lref{ur1lipschitz} and \cite[\thm 9.5 and \remk 9.6(b)]{Ama}, the
1479: solution $\vc{a}(t)$ of \eref{ur1ivpa} is infinitely differentiable.
1480: Thus, for $t\in\ccint{0,1}$,
1481: \begin{equation}
1482: \label{eq:ur1lyapuineq}
1483: \begin{aligned}
1484: L(\vc{a}(t)) - L(\vc{a}_0) &=
1485: \|\vc{a}_0 + t (\Ra_1(\vc{a}_0)-\vc{a}_0) + \ord(t) -
1486: \vc{\alpha}\| -
1487: \|\vc{a}_0 - \vc{\alpha}\| \\
1488: &\le t \bigl( \|\Ra_1(\vc{a}_0) - \Ra_1(\vc{\alpha})\| -
1489: \|\vc{a}_0 - \vc{\alpha}\| \bigr) + \ord(t) \es,
1490: \end{aligned}
1491: \end{equation}
1492: where $\ord(t)$ is the usual Landau symbol and represents some
1493: function that vanishes faster than $t$ as $t\to0$. From this, by
1494: the strict contraction property of $\Ra_1$ (\pref{ur1contract}), the
1495: Lyapunov property \eref{dotl} follows, with equality if and only if
1496: $\vc{a}_0 = \vc{\alpha}$. Since $\X_{\alpha,\delta}$ is compact,
1497: \thref{lyapunov} implies the claim.
1498: \end{proof}
1499:
1500: \pagebreak[2]
1501: We are now able to give the previously postponed
1502: \begin{theopargself}
1503: \begin{proof}[of \thref{ur1conv}]
1504: By \pref{urgainxad}, we have $\vc{a}(0) = \vc{a}(\vc{p}(0)) \in
1505: \X_{\alpha,\delta}$ with $\alpha = \frac12 m$ and some $\delta$.
1506: In discrete time, according to \propositions
1507: \ref{prop:ur1convdisc} and \ref{prop:ur1contract} and Banach's
1508: fixed point theorem (compare \cite[\thm V.18]{ReSi}), it then
1509: follows that $\vc{a}(t) \to \vc{\alpha} =
1510: (1,\alpha,\alpha^2,\ldots)$ \wrt the metric $d$. Inserting
1511: \eref{ur1fixed} into \eref{urga} and letting $x = m/(m+2)$ yields
1512: \begin{align*}
1513: a_k^{} &=
1514: \sum_{\ell \ge k} \frac{\ell!}{(\ell-k)!(k+1)!}
1515: (1-x)^2 (\ell+1) x^\ell \\
1516: &= (1-x)^2 \sum_{\ell \ge k} \binom{\ell+1}{k+1} x^\ell =
1517: \left(\frac{x}{1-x}\right)^k = \alpha^k .
1518: \end{align*}
1519: The claim now follows from \lref{urgpnormconv}. Similarly, in
1520: continuous time, the claim follows from \pref{ur1convcont}.
1521: \end{proof}
1522: \end{theopargself}
1523:
1524: Let us finally note
1525: \begin{proposition}
1526: For the dynamics described by \eref{ur1ivpa}, the fixed point\/
1527: $\vc{\alpha}$ from \pref{ur1convcont} is exponentially stable.
1528: \end{proposition}
1529: \begin{proof}
1530: Let $\vc{a}_0 \in \X_{\alpha,\delta}$ be arbitrary. The Lyapunov
1531: function from the proof of \pref{ur1convcont} satisfies, as a
1532: consequence of \eref{ur1lyapuineq} and \pref{ur1contract},
1533: \begin{displaymath}
1534: \dot{L}(\vc{a}_0) \le
1535: d(\Ra_1(\vc{a}_0), \Ra_1(\vc{\alpha})) - d(\vc{a}_0, \vc{\alpha}) \le
1536: - (1-C) \, d(\vc{a}_0, \vc{\alpha}) \es,
1537: \end{displaymath}
1538: with $0 < C < 1$. From this, together with \eref{ur1lyapunov} and
1539: \cite[\thm 18.7]{Ama}, the claim follows.
1540: \end{proof}
1541:
1542: \noindent
1543: \textbf{Remark.}
1544: In a related UC model introduced by Takahata \cite{Tak}, for which
1545: \begin{displaymath}
1546: T_{ij,k\ell} = \delta_{i+j,k+\ell} \frac{1}{k+\ell+1} \es,
1547: \end{displaymath}
1548: the recombinator $\Ra_1$ appears for the coefficients
1549: $\vc{b}(\vc{p})_k^{} = (k+1) \, \vc{a}(\vc{p})_k^{}$, where
1550: $\vc{b}(\vc{p})_1^{}$ is the mean copy number $m$. The above results
1551: then imply, under the appropriate condition on $\vc{p}(0)$, that
1552: $\vc{b}(t) \to (1,m,m^2,\ldots)$ as $t\to\infty$ both in discrete and
1553: in continuous time. This corresponds to convergence of $\vc{p}(t)$ to
1554: the fixed point $\vc{p}$ with $p_k^{} = \frac{1}{m+1}
1555: (\frac{m}{m+1})^k$.
1556:
1557:
1558: \section{The intermediate parameter regime}
1559: \label{sec:urqgen}
1560:
1561: In this section, $q$ may take any value in $\ccint{0,1}$. \Wrt
1562: reversibility of fixed points, one finds
1563: \begin{proposition}
1564: For parameter values\/ $q\in\ooint{0,1}$, any fixed point\/ $\vc{p}
1565: \in \M_1^+$ of the recombinator\/ $\R_q$, given by \eref{reco} and
1566: \eref{urtq}, satisfies\/ $p_k^{}>0$ for all\/ $k\ge0$ \pr{unless it
1567: is the trivial fixed point\/ $\vc{p} = (1,0,0,\ldots)$ we
1568: excluded}. None of these extra fixed points is reversible.
1569: \end{proposition}
1570: \begin{proof}
1571: Let a non-trivial fixed point $\vc{p}$ be given and choose any $n>0$
1572: with $p_n^{}>0$. Observe that $T^{(q)}_{n+1 \is n-1, nn} > 0$ for
1573: $0 < q < 1$ and hence
1574: \begin{displaymath}
1575: p_{n\pm1}^{\vp} = \R_q(\vc{p})_{n\pm1}^{\vp}
1576: = \sum_{j,k,\ell\ge0} T^{(q)}_{n\pm1 \is j,k\ell} \,
1577: p_k^{\vp} \, p_\ell^{\vp}
1578: \ge T^{(q)}_{n+1 \is n-1, nn} \, p_n^{\vp} \, p_n^{\vp} > 0 \es.
1579: \end{displaymath}
1580: The first statement follows now by induction.
1581:
1582: For the second statement,
1583: evaluate the reversibility condition \eref{urrev} for all
1584: combinations of $i$, $j$, $k$, $\ell$ with $i+j = k+\ell \le 4$.
1585: This leads to four independent equations. Three of them can be
1586: transformed to the recursion
1587: \begin{displaymath}
1588: p_k^{}
1589: = \frac{(k+1)q}{2(k-1)+2q} \frac{p_1^{}}{p_0^{}} \, p_{k-1}^{} \es,
1590: \qquad k \in \{2,3,4\} \es,
1591: \end{displaymath}
1592: from which one derives explicit equations for all $p_k^{}$ with $k
1593: \in \{2,3,4\}$ in terms of $p_0^{}$ and $p_1^{}$. Inserting the one
1594: for $p_2^{}$ into the remaining equation yields another equation for
1595: $p_4^{}$ in terms of $p_0^{}$ and $p_1^{}$, which contradicts the
1596: first equation for all $q \in \ooint{0,1}$, as is easily verified.
1597: \end{proof}
1598: So, non-trivial fixed points for $0 < q < 1$ are not reversible, and
1599: thus much more difficult to determine. Our most general result so far
1600: is
1601: \begin{theorem}
1602: \label{thm:urqfixed}
1603: If\/ $\vc{p}(0) \in P_{\alpha,\delta}$ for some\/ $\alpha, \delta$,
1604: then\/ $\vc{p}(t) \in P_{\alpha,\delta}$ for all times\/
1605: $t\in\NNzer$, respectively\/ $t\in\RRnn$, and\/ $\R_q$ has a fixed
1606: point in\/ $P_{\alpha,\delta}$.
1607: \end{theorem}
1608: The proof is based on the fact that $\R_q$ is, in a certain sense,
1609: monotonic in the parameter $q$. This is stated in
1610: \begin{proposition}
1611: \label{prop:urqainxad}
1612: Assume\/ $\vc{a}(\vc{p}) \in \X_{\alpha,\delta}$ for some\/
1613: $\alpha$, $\delta$. Then, \wrt the partial order introduced before
1614: \pref{urgpadcomp}, $\vc{a}(\R_q(\vc{p})) \le
1615: \vc{a}(\R_{q'}(\vc{p}))$ for all\/ $0 \le q \le q' \le 1$. In
1616: particular,\/ $\vc{a}(\R_q(\vc{p})) \in \X_{\alpha,\delta}$ for
1617: all\/ $0 \le q \le 1$.
1618: \end{proposition}
1619: To show this, we need three rather technical lemmas. The first one
1620: collects formal conditions on the difference of two discrete probability
1621: distributions
1622: $T^{(q)}_{ij,k\ell}$ with different parameter values (but $j=k+\ell-i$
1623: and the same fixed $k$, $\ell$). These are then verified in our case.
1624: \begin{lemma}
1625: \label{lem:urqineq}
1626: Let the numbers\/ $x_i\in\RR$ \pr{$0 \le i \le r$ with some\/
1627: $r\in\NNzer$} satisfy the following three conditions:
1628: \begin{gather}
1629: \label{eq:urqineqcond1}
1630: \sum_{i=0}^r x_i = 0 \es.\\
1631: \label{eq:urqineqcond2}
1632: x_{r-i} = x_i \quad\text{for all\/ $0 \le i \le r$.}\\
1633: \label{eq:urqineqcond3}
1634: \text{There is an integer\/ $n$ such that }
1635: \begin{cases} x_i \ge0 \; : \; 0 \le i \le n \\
1636: x_i<0 \; : \; n < i \le \lfloor \frac{r}{2} \rfloor \end{cases}
1637: \!\!\!\!\!.
1638: \end{gather}
1639: Further, let\/ $f_i\in\RR$ \pr{$0 \le i \le r$} be given with
1640: \begin{equation}
1641: \label{eq:urqineqcond4}
1642: 0 \le f_1 - f_0 \le f_2 - f_1 \le \ldots \le f_r - f_{r-1} \es.
1643: \end{equation}
1644: Then, we have
1645: \begin{displaymath}
1646: \sum_{i=0}^r f_i x_i \ge 0 \es.
1647: \end{displaymath}
1648: \end{lemma}
1649: \begin{proof}
1650: Let us first consider the trivial cases. If $x_i\equiv0$,
1651: everything is clear, so let $x_i\not\equiv0$. If $r\le1$ then
1652: $x_i\equiv0$, so let $r\ge2$, and thus $n\le\frac{r}{2}-1$. Define
1653: $x_{\frac{r}{2}} = f_{\frac{r}{2}} = 0$ for odd $r$. Then, we can
1654: write
1655: \begin{equation}
1656: \label{eq:urqineq1}
1657: \sum_{i=0}^r f_i x_i = \sum_{i=0}^n \left( f_i + f_{r-i} \right) x_i +
1658: \sum_{i=n+1}^{\lceil\frac{r}{2}\rceil-1} \left( f_i + f_{r-i}
1659: \right) x_i + f_{\frac{r}{2}} x_{\frac{r}{2}} \es.
1660: \end{equation}
1661: Furthermore, for $r-i \ge i$, due to \eref{urqineqcond4},
1662: \begin{displaymath}
1663: f_i + f_{r-i} = f_{i-1} + f_{r-i+1} + (f_i - f_{i-1}) -
1664: (f_{r-i+1} - f_{r-i}) \le f_{i-1} + f_{r-i+1} \es.
1665: \end{displaymath}
1666: Now, define $C := \sum_{i=0}^n x_i = -
1667: \sum_{i=n+1}^{\lceil\frac{r}{2}\rceil-1} x_i - \tfrac12
1668: x_{\frac{r}{2}} > 0$, and the claim follows with \eref{urqineq1},
1669: since $r-n \ge n+1$ by assumption:
1670: \begin{multline*}
1671: \sum_{i=0}^r f_i x_i \ge
1672: C \left[ f_n + f_{r-n} - f_{n+1} - f_{r-n-1} \right] \\
1673: = C \left[ (f_{r-n} - f_{r-n-1}) - (f_{n+1} - f_n) \right] \ge 0 \es.
1674: \qedhere
1675: \end{multline*}
1676: \end{proof}
1677: \begin{lemma}
1678: \label{lem:urqcoeff}
1679: Let\/ $j\in\NNzer$ be fixed and\/ $f_i = (i)_j^{}$,
1680: $i\in\NNzer$, where\/ $(i)_j$ is the falling factorial, which
1681: equals\/ $1$ for\/ $j=0$ and\/ $i(i-1)\cdots(i-j+1)$ for\/
1682: $j>0$, hence\/ $\frac{i!}{(i-j)!}$ for\/ $i \ge j$. Then
1683: condition \eref{urqineqcond4} is satisfied.
1684: \end{lemma}
1685: \begin{proof}
1686: For $j=0$, condition \eref{urqineqcond4} is trivially true.
1687: Otherwise, each $f_i$ is a polynomial of degree $j$ in $i$ with
1688: zeros $\{0,1,\ldots,j-1\}$, hence we have the equality $0 = f_1 -
1689: f_0 = \ldots = f_{j-1} - f_{j-2}$. Then, for $i \ge j-1$, the
1690: polynomial and all its derivatives are increasing functions since
1691: $\lim_{i\to\infty} f_i = \infty$. Therefore, for $i \ge j-1$, we
1692: have $0 \le f_{i+1} - f_i \le f_{i+2} - f_{i+1}$. Hence
1693: \eref{urqineqcond4} holds.
1694: \end{proof}
1695: \begin{lemma}
1696: \label{lem:urqineqcond123}
1697: For\/ $0 \le q \le q' \le 1$ and all\/ $k, \ell$, Equations
1698: \eref{urqineqcond1}--\eref{urqineqcond3} are true for\/ $r=k+\ell$
1699: and $x_i = T^{(q')}_{ik\ell} - T^{(q)}_{ik\ell}$, where\/
1700: $T^{(q)}_{ik\ell} = T^{(q)}_{ij,k\ell}$ with\/ $j = k+\ell-i$.
1701: \end{lemma}
1702: \begin{proof}
1703: The validity of \eref{urqineqcond1} and \eref{urqineqcond2} is clear
1704: from the normalization \eref{sumt} and the symmetry of the
1705: $T^{(q)}_{ik\ell}$. For \eref{urqineqcond3}, let $k \le \ell$
1706: without loss of generality. In the trivial cases $q=q'$ or $k=0$,
1707: choose $n=\lfloor\frac{r}{2}\rfloor$. Otherwise, $x_i =
1708: T^{(q')}_{ik\ell} - T^{(q)}_{ik\ell} < 0$ for $k \le i \le
1709: \lfloor\frac{r}{2}\rfloor$, since $C^{(q')}_{k\ell} <
1710: C^{(q)}_{k\ell}$, and $x_0 > 0$. For $0 \le i \le k$, consider
1711: \begin{displaymath}
1712: y_i = \frac{x_i}{T^{(q)}_{ik\ell}} + 1
1713: = \frac{C^{(q')}_{k\ell}}{C^{(q)}_{k\ell}}
1714: \left(\frac{q'}{q}\right)^{k-i} \es.
1715: \end{displaymath}
1716: Here, the first factor is less than 1, the second is equal to 1 for
1717: $k=i$, greater than 1 for $0 \le k < i$, and strictly decreasing
1718: with $i$. Since $x_i\ge0$ if and only if $y_i\ge1$, there is an
1719: index $n$ with the properties needed.
1720: \end{proof}
1721: \begin{theopargself}
1722: \begin{proof}[of \pref{urqainxad}]
1723: We assume $0\le q\le q' \le 1$.
1724: Lemmas \ref{lem:urqineq}--\ref{lem:urqineqcond123} imply, for all
1725: $k, \ell, j\in\NNzer$ with $k+\ell \ge j$,
1726: \begin{displaymath}
1727: \sum_{i=j}^{k+\ell} \tfrac{i!}{(i-j)!} T^{(q)}_{ik\ell} \le
1728: \sum_{i=j}^{k+\ell} \tfrac{i!}{(i-j)!} T^{(q')}_{ik\ell} \es.
1729: \end{displaymath}
1730: Then, since $T^{(q)}_{ik\ell} = 0$ for $i >k +\ell$,
1731: \begin{align*}
1732: \vc{a}(\R_q(\vc{p}))_j^{} &= \tfrac{1}{(j+1)!} \sum_{i \ge j}
1733: \tfrac{i!}{(i-j)!} \R_q(\vc{p})_i^{} =
1734: \tfrac{1}{(j+1)!} \sum_{i \ge j} \tfrac{i!}{(i-j)!}
1735: \sum_{k,\ell\ge0} T^{(q)}_{ik\ell} p_k^{\vp} p_\ell^{\vp} \\
1736: &=\tfrac{1}{(j+1)!} \mskip-4mu\sum_{k,\ell\ge0}\mskip-4mu
1737: p_k^{\vp} p_\ell^{\vp} \sum_{i \ge j}
1738: \tfrac{i!}{(i-j)!} T^{(q)}_{ik\ell} \\
1739: &\le \tfrac{1}{(j+1)!} \mskip-4mu\sum_{k,\ell\ge0}\mskip-4mu
1740: p_k^{\vp} p_\ell^{\vp} \sum_{i \ge j}
1741: \tfrac{i!}{(i-j)!} T^{(q')}_{ik\ell}
1742: = \vc{a}(\R_{q'}(\vc{p}))_j^{} \es.
1743: \end{align*}
1744: From this, together with \lref{ur1cont}, the claim follows.
1745: \end{proof}
1746: \end{theopargself}
1747: \begin{theopargself}
1748: \begin{proof}[of \thref{urqfixed}]
1749: According to \pref{urqainxad}, $\R_q$ maps $P_{\alpha,\delta}$
1750: into itself, and thus, in discrete time, $\vc{p}(t) \in
1751: P_{\alpha,\delta}$ for every $t\in\NNzer$. The analogous
1752: statement is true for continuous time $t\in\RRnn$. To see this,
1753: consider $P_{\alpha,\delta}$ as a closed subset of $\Ll{1}$.
1754: Recall that $\R_q - \Id$ is globally Lipschitz on $\Ll{1}$ by
1755: \pref{recolipschitz}. Moreover, for any $\vc{p} \in
1756: P_{\alpha,\delta}$ and $t\in\ccint{0,1}$, \pref{urgpadcomp} tells
1757: us that
1758: \begin{displaymath}
1759: \vc{p} + t (\R_q(\vc{p}) - \vc{p}) =
1760: (1-t) \vc{p} + t \R_q(\vc{p}) \in P_{\alpha,\delta} \es.
1761: \end{displaymath}
1762: This implies the positive invariance of $P_{\alpha,\delta}$ by
1763: \cite[\thm VI.2.1]{Mar} (see also \cite[\thm 16.5]{Ama}). The
1764: existence of a fixed point once again follows from the
1765: Leray--Schauder--Tychonov theorem \cite[\thm V.19]{ReSi}.
1766: \end{proof}
1767: \end{theopargself}
1768:
1769: On the basis of the above analysis, and further numerical work done
1770: to investigate the fixed point properties \cite{Red,ShAt},
1771: it is plausible that, given the mean copy number $m$, never more than
1772: one fixed point for $\R_q$ exists. Due to the global convergence
1773: results at $q=0$ and $q=1$, any non-uniqueness in the vicinity of
1774: these parameter values could only come from a bifurcation, not from an
1775: independent source. Numerical investigations indicate that no
1776: bifurcation is present, but this needs to be analyzed further.
1777:
1778: Furthermore, the Lipschitz constant for $\Ra_q$ can be expected to be
1779: continuous in the parameter $q$, hence to remain strictly less than 1
1780: on the sets $X_{\alpha,\delta}$ in a neighborhood of $q=1$. So, at
1781: least locally, the contraction property should be preserved.
1782: Nevertheless, we do not expand on this here since it seems possible to
1783: use a rather different approach \cite{Hof03}, which has been used for
1784: similar problems in game theory, to establish a slightly weaker type
1785: of convergence result for all $0 < q < 1$, and probably even on the
1786: larger compact set $\M_{1,m,C}^+$ of \lref{ur0dp}.
1787:
1788:
1789: \section{Concluding remarks}
1790: \label{sec:remarks}
1791:
1792: In this article, we have shown that, for the extreme parameter values $q=0$
1793: (internal UC) and $q=1$ (random UC), any initial configuration satisfying a
1794: specific condition converges to one of the known fixed points, both in
1795: discrete and continuous time. The condition to be met is, for $q=0$, the
1796: existence of the $r$-th moment ($r>1$, see \thref{ur0conv}), respectively, for
1797: $q=1$, that the corresponding generating function has a radius of
1798: convergence $\rho>1$ (\thref{ur1conv}).
1799: Convergence takes place in the total variation
1800: norm in all cases. As argued in the previous section, similar results can be
1801: expected for the intermediate parameter values as well.
1802:
1803: % We have seen in the preceding sections that, definitely for the extreme cases
1804: % $q=0$ and $q=1$ and presumably for the intermediate values as well, the
1805: % deterministic dynamics converges to a unique equilibrium solution in both
1806: % discrete and continuous time. This correspond
1807: These results are valid for deterministic dynamics and thus correspond
1808: to the case of infinite populations. \Wrt biological
1809: relevance, however, we add some arguments that it is reasonable to expect this
1810: to be a good description for large but finite populations as well, \ie for the
1811: underlying (multitype) branching process. For finite state spaces, such as in
1812: the mutation--selection models discussed in \cite{HRWB}, the results by Ethier
1813: and Kurtz \cite[\thm 11.2.1]{EtKu86} and the generalization \cite[\thm
1814: V.7.2]{AtNe72} of the Kesten--Stigum theorem \cite{KeSt66,KLPP97} guarantee
1815: that in the infinite population limit the relative genotype frequencies of the
1816: branching process converge almost surely to the deterministic solution (if the
1817: population does not go to extinction). Since for the UC models considered
1818: here the equilibrium distributions are exponentially small for large copy
1819: numbers (owing to \thref{urqfixed} also for $q\in\ooint{0,1}$), one can expect
1820: these systems to behave very much like ones with finitely many genotypes.
1821: This is also supported by several simulations. Nevertheless, this questions
1822: deserves further attention.
1823:
1824:
1825: \begin{acknowledgement}
1826: We thank Ellen Baake, Joachim Hermisson, and Josef Hofbauer for
1827: their cooperation, and Odo Diekmann and Anton Wakolbinger for
1828: clarifying discussions. O.R. gratefully acknowledges support by a
1829: PhD scholarship of the Studienstiftung des deutschen Volkes. The
1830: authors express their gratitude to the Erwin Schr\"odinger
1831: International Institute for Mathematical Physics in Vienna for
1832: support during a stay in December 2002, where the manuscript for
1833: this article was completed.
1834: \end{acknowledgement}
1835:
1836:
1837: \begin{thebibliography}{10}
1838: \expandafter\ifx\csname url\endcsname\relax
1839: \def\url#1{\texttt{#1}}\fi
1840: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1841:
1842: \bibitem{Ama}
1843: Amann, H.: Ordinary Differential Equations. De Gruyter, Berlin, 1990
1844:
1845: \bibitem{AtNe72}
1846: Athreya, K.B., Ney, P.E.: Branching Processes. Springer, New York, 1972
1847:
1848: \bibitem{BilC}
1849: Billingsley, P.: Convergence of Probability Measures. 2nd ed., Wiley, New York,
1850: 1999
1851:
1852: \bibitem{EtKu86}
1853: Ethier, S.N., Kurtz, T.G.: Markov Processes -- Characterization and
1854: Convergence. Wiley, New York, 1986
1855:
1856: \bibitem{GrLi}
1857: Graur, D., Li, W.-H.: Fundamentals of Molecular Evolution. 2nd ed., Sinauer,
1858: Sunderland, 2000
1859:
1860: \bibitem{HRWB}
1861: Hermisson, J., Redner, O., Wagner, H., Baake, E.: Mutation--selection balance:\
1862: Ancestry, load, and maximum principle. Theor.\ Pop.\ Biol.\ \textbf{62}, 9--46
1863: (2002); \newline \texttt{cond-mat/0202432}
1864:
1865: \bibitem{Hof03}
1866: Hofbauer, J.: In preparation
1867:
1868: \bibitem{HofPriv}
1869: Hofbauer, J.: Private communication (2002)
1870:
1871: \bibitem{Kal}
1872: Kallenberg, O.: Foundations of Modern Probability. Springer, New York, 1997
1873:
1874: \bibitem{KeSt66}
1875: Kesten, H., Stigum, B.P.: A limit theorem for multidimensional
1876: {G}alton--{W}atson processes. Ann.\ Math.\ Statist. \textbf{37}, 1211--1233
1877: (1966)
1878:
1879: \bibitem{KLPP97}
1880: Kurtz, T., Lyons, R., Pemantle, R., Peres, Y.: A conceptual proof of the
1881: {K}esten--{S}tigum theorem for multi-type branching processes. In: Athreya,
1882: K.B., Jagers, P. (eds.), Classical and Modern Branching Processes, pp.
1883: 181--185, Springer, New York, 1997
1884:
1885: \bibitem{Lan}
1886: Lang, S.: Real and Functional Analysis. 3rd ed., Springer, New York, 1993
1887:
1888: \bibitem{LanCA}
1889: Lang, S.: Complex Analysis. 4th ed., Springer, New York, 1999
1890:
1891: \bibitem{Mar}
1892: Martin, R.H.: Nonlinear Operators and Differential Equations in Banach Spaces.
1893: Wiley, New York, 1976
1894:
1895: \bibitem{Oht}
1896: Ohta, T.: On the evolution of multigene families. Theor.\ Pop.\ Biol.
1897: \textbf{23}, 216--240 (1983)
1898:
1899: \bibitem{Ped}
1900: Pedersen, G.K.: Analysis Now. Revised ed., Springer, New York, 1995
1901:
1902: \bibitem{Red}
1903: Redner, O.: Models for Mutation, Selection, and Recombination in
1904: Infinite Populations. PhD Thesis, Univ.\ Greifswald (2003). Shaker,
1905: Aachen, 2003.
1906:
1907: \bibitem{ReSi}
1908: Reed, M., Simon, B.: Methods of Modern Mathematical Physics {I}: Functional
1909: Analysis. Academic Press, San Diego, 1980
1910:
1911: \bibitem{Rem}
1912: Remmert, R.: Classical Topics in Complex Function Theory. Springer, New York,
1913: 1998
1914:
1915: \bibitem{RudRCA}
1916: Rudin, W.: Real and Complex Analysis. 3rd ed., McGraw-Hill, New York, 1986
1917:
1918: \bibitem{Shir}
1919: Shiryaev, A.N.: Probability. 2nd ed., Springer, New York, 1996
1920:
1921: \bibitem{ShAt}
1922: Shpak, M., Atteson, K.: A survey of unequal crossover systems and their
1923: mathematical properties. Bull.\ Math.\ Biol. \textbf{64}, 703--746 (2002)
1924:
1925: \bibitem{ShWa}
1926: Shpak, M., Wagner, G.P.: Asymmetry of configuration space induced by unequal
1927: crossover: Implications for a mathematical theory of evolutionary innovation.
1928: Artificial Life \textbf{6}, 25--43 (2000)
1929:
1930: \bibitem{Tak}
1931: Takahata, N.: A mathematical study on the distribution of the number of
1932: repeated genes per chromosome. Genet.\ Res. \textbf{38}, 97--102 (1981)
1933:
1934: \bibitem{Wsh}
1935: Walsh, J.B.: Persistence of tandem arrays: Implications for satellite and
1936: simple-sequence {DNA}'s. Genetics \textbf{115}, 553--567 (1987)
1937:
1938: \bibitem{WalEn}
1939: Walter, W.: Ordinary Differential Equations. Springer, New York, 1998
1940:
1941: \bibitem{Yos}
1942: Yosida, K.: Functional Analysis. 6th ed., Springer, Berlin, 1980
1943:
1944: \end{thebibliography}
1945:
1946: \end{document}
1947: