q-bio0501028/final.tex
1: \documentstyle[prl,aps,epsfig]{revtex}
2: \renewcommand{\huge}{\large}
3: \renewcommand{\LARGE}{\large}
4: \renewcommand{\Large}{\large}
5: \thispagestyle{empty}
6: \begin{document}
7: \def\be{\begin{equation}}
8: \def\ee{\end{equation}}
9: \def\bea{\begin{eqnarray}}
10: \def\eea{\end{eqnarray}}
11: \def\bml{\begin{mathletters}}
12: \def\eml{\end{mathletters}}
13: \def\l{\label}
14: \def\b{\bullet}
15: \def\eqn#1{(~\ref{eq:#1}~)}
16: \def\no{\nonumber}
17: \def\av#1{{\langle  #1 \rangle}}
18: 
19: %=============================================================================
20: %=============================================================================
21: \title{Evolutionary trajectories in rugged fitness landscapes}
22: 
23: \author{Kavita Jain and Joachim Krug}
24: \address{Institut 
25: f\"ur Theoretische Physik, Universit\"at zu K\"oln, Z\"ulpicher Strasse 77, 
26: 50937 K\"oln, Germany}
27: \maketitle
28: \widetext
29: 
30: \begin{abstract}
31: We consider the evolutionary trajectories traced out by an infinite population
32: undergoing mutation-selection dynamics in static, uncorrelated random
33: fitness landscapes. Starting from the population that consists of a single 
34: genotype, the most populated genotype \textit{jumps} from a 
35: local fitness maximum to another and eventually reaches the global maximum. 
36: We use a strong selection limit, which reduces the dynamics beyond the 
37: first time step to the competition between independent mutant subpopulations,  
38: to study the dynamics of this model and of a simpler one-dimensional model 
39: which ignores the geometry of the sequence space. 
40: We find that the fit genotypes that appear along a trajectory are a subset
41: of suitably defined fitness \textit{records}, and exploit several results
42: from the record theory for non-identically distributed random variables.
43: The genotypes that contribute to the trajectory are those records that are
44: not \textit{bypassed} by superior records arising further away from the
45: initial population. Several conjectures concerning the statistics of 
46: bypassing are extracted from numerical simulations. In particular, for the 
47: one-dimensional model, we propose a simple relation between the bypassing 
48: probability and the dynamic exponent which describes the scaling of the 
49: typical evolution time with genome size. The latter can
50: be determined exactly in terms of the extremal properties of the
51: fitness distribution. 
52: \vskip0.5cm
53: \noindent PACS numbers: {87.10.+e, 87.23.Kg, 05.40.-a}
54: \end{abstract}
55: 
56: %=============================================================================
57: %INTRODUCTION
58: %=============================================================================
59: \section{Introduction}
60: \l{intro}
61: 
62: 
63: The episodic nature of biological evolution has provided motivation for much
64: work on the modeling of evolutionary dynamics in the statistical physics
65: community \cite{Sneppen95,Sibani98}; see 
66: \cite{Peliti97,Baake99,Drossel01} for review. Evolution displays
67: a \textit{punctuated} pattern, with epochs of no or slow change 
68: interspersed with
69: bursts of (relatively) rapid activity, on various levels ranging from
70: the fossil record \cite{Eldredge89,Gould93} to experiments with microbial 
71: populations \cite{Lenski94,Elena96,Burch99,Elena03}. Punctuated behavior is 
72: also seen in simulations of \textit{in vitro} evolution
73: of RNA molecules \cite{Schuster02}, optimization algorithms 
74: \cite{Rujan88,vanNimwegen97}, and artificial life \cite{Adami95}. 
75: It has been recognized for a long time that one possible 
76: scenario that is consistent with punctuated dynamics is the evolution 
77: of a population in a static fitness landscape with many peaks.
78: In this picture, the two modes of evolution correspond to the extended
79: periods of residence of the population at a local fitness maximum
80: and the exploration of a new, higher lying peak, respectively,
81: which entails the rapid crossing of a valley of lower fitness 
82: \cite{Newman85,Lande85}. 
83: In essence, this is the phenomenon of quantum evolution described by 
84: the paleontologist G.G. Simpson sixty
85: years ago \cite{Simpson44}, and more recently referred to in 
86: macroevolutionary theory 
87: as punctuated gradualism \cite{Eldredge89}. 
88: If the population is large, so that the distribution of individuals over the 
89: various phenotypes or 
90: genotypes can be modeled by a continuous field, the transition between 
91: fitness peaks described above displays analogies with physical processes  
92: such as quantum tunneling \cite{Ebeling84}, variable range hopping 
93: \cite{Zhang86} or noise-driven barrier crossing. 
94: 
95: Since this metastable behavior seems ubiquitous in nature, it may be 
96: worthwhile to study a simple model where this can be 
97: analysed in detail. A convenient mathematical framework 
98: to address this issue is the quasispecies model which was originally 
99: introduced to describe large populations of self-replicating 
100: macromolecules \cite{Eigen71,Eigen89}. The quasispecies model is a 
101: mutation-selection model whose steady state has been studied in great detail.  
102: For various choices of fitness landscapes, it exhibits 
103: the phenomenon of error threshold in which beyond a 
104: critical mutation rate, the population delocalises over the whole sequence  
105: space. 
106: 
107: In this paper, we address the question of punctuated \textit{dynamics} in the 
108: quasispecies model. To avoid complications associated
109: with the error threshold phenomenon \cite{Baake99,Eigen89}, we work   
110: in a strong selection limit inspired by the zero temperature limit
111: of the statistical mechanics of disordered systems \cite{Krug02} 
112: (for a different kind of strong selection
113: limit used in population genetics see e.g. \cite{Woodcock96}). 
114: In this limit, the location of the population in the space of genotypes
115: can be identified with the most populated genotype at all times, 
116: and the evolutionary trajectories can be represented in a particularly
117: transparent manner \cite{Krug03}. 
118: Since the population is located at a single genotype at any time, the 
119: evolutionary trajectory changes in a stepwise manner. To generate an 
120: evolutionary 
121: trajectory, a localized population is placed at a randomly chosen point 
122: in the space of all possible genotypes. Due to infinite population, in the 
123: next time step, all genotypes get occupied with nonzero population. 
124: As we describe later in detail, it turns out that 
125: the fit genotypes typically receive small initial 
126: population. Strong selection then reduces the problem to competition 
127: between these fit subpopulations struggling to overcome the poor 
128: initial conditions. 
129: 
130: 
131: In the next section, we describe
132: the quasispecies model and derive the reduced representation 
133: that arises in the strong selection limit. The bulk of the 
134: paper is then devoted to the analysis of the strong selection dynamics, which
135: turns out to be rather closely related to the mathematical theory of records
136: \cite{Glick78,Nevzorov87,Arnold98,Nevzorov01}; a relation between 
137: biological evolution and record statistics has been proposed 
138: previously \cite{Sibani98,Kauffman87}, see also \cite{Krug04}. 
139: 
140: %=============================================================================
141: %Model
142: %=============================================================================
143: \section{Quasispecies dynamics in the strong selection limit}
144: \l{model}
145: 
146: The quasispecies model is defined on the space of genotypes represented by 
147: sequences $\sigma \equiv \{ \sigma_{1},...,\sigma_{N} \} $,  
148: where each of the $N$ letters $\sigma_{i}$ is taken from an alphabet of size
149: $\ell \geq 2$. The number of individuals of genotype $\sigma$ 
150: at time $t$ is represented
151: by a real variable $Z(\sigma,t)$ that obeys the discrete time 
152: evolution equation 
153: \be
154: Z(\sigma,t+1)= \sum_{{\sigma}^{\prime}}
155: p(\sigma^{\prime} \rightarrow \sigma) \; W(\sigma^{\prime}) 
156: \; Z(\sigma^{\prime},t).
157: \l{full}
158: \ee
159: Here the fitness $W(\sigma)$ is defined \cite{Peliti97} as 
160: the expected number of offspring produced by an 
161: individual carrying sequence $\sigma$, 
162: and $p(\sigma^{\prime} \rightarrow \sigma)$ is the probability that
163: genotype $\sigma$ is created as offspring of genotype $\sigma'$ due to 
164: copying errors in the genome. A simple choice for the latter 
165: corresponds to independent point mutations occurring with 
166: probability $\mu$ per generation, so that 
167: $p(\sigma^{\prime} \rightarrow \sigma)= 
168: (\mu/(\ell-1))^{d(\sigma^{\prime},\sigma)} 
169: (1-\mu)^{N-d(\sigma^{\prime},\sigma)}$,  
170: where 
171: \be
172: d(\sigma^{\prime},\sigma) = \sum_{i=1}^N (1 - \delta_{\sigma_i,\sigma_i'})
173: \;\;, \l{Hamming}
174: \ee
175: is the Hamming distance between sequences $\sigma^{\prime}$ and $\sigma$.
176: 
177: A constraint of fixed population size could be enforced by dividing
178: the right hand side of (\ref{full}) by the population averaged fitness
179: $\langle W \rangle$. This introduces an (inessential)
180: nonlinearity into the problem \cite{Peliti97}, which is why we prefer to work 
181: with the
182: unnormalized population variables $Z(\sigma,t)$.   
183: It is clear that within the quasispecies framework, which requires an
184: infinite population from the outset, the actual population size cannot
185: play any role.
186: 
187: To derive the strong selection limit \cite{Krug02}, we 
188: write $Z(\sigma,t)=e^{\kappa E(\sigma,t)}$, 
189: $W(\sigma)=e^{\kappa F(\sigma)}$ 
190: and $\mu=e^{-\kappa}$ where $\kappa$ is the inverse selective 
191: temperature \cite{Peliti97,Franz97} 
192: and $E(\sigma,t)$ and $F(\sigma)$ are logarithmic population and
193: fitness variables, respectively. Throughout this paper we take
194: the fitness landscape to be completely uncorrelated, which implies that
195: the fitnesses $F(\sigma)$ are independent, identically 
196: distributed (i.i.d.) quenched random variables chosen 
197: from a distribution $p(F)$
198: with support on interval $[F_{\mathrm{min}},F_{\mathrm{max}}]$. 
199: The strong selection limit corresponds to $\kappa \rightarrow \infty$, 
200: which yields
201: \be
202: E(\sigma, t+1)= \mbox{max}_{\sigma^\prime} \left[ E(\sigma^\prime,t)+ 
203: F(\sigma^\prime)-d(\sigma,\sigma^\prime) \right] \l{fullss} \;\;.
204: \ee
205: Starting with an initial condition 
206: $Z(\sigma,0)=\delta_{\sigma,\sigma^{(0)}}$ in which only a 
207: single, randomly chosen  
208: sequence $\sigma^{(0)}$ has nonzero population, at the next time step each 
209: sequence gets seeded with the logarithmic population
210: \be
211: \l{sslimit2}
212: E(\sigma,1) = F(\sigma^{(0)}) - d(\sigma,\sigma^{(0)}).
213: \ee
214: In terms of the original variables $Z(\sigma,t)$, this
215: corresponds to a population distribution that decays exponentially
216: with increasing Hamming distance from $\sigma^{(0)}$. 
217: Remarkably, it turns out that for the subsequent time evolution the 
218: mutations are unimportant for genotypes with 
219: high fitness. Since we are primarily interested in such genotypes, the 
220: dynamics of the model can be approximated by allowing each genotype 
221: to reproduce with its intrinsic rate $F(\sigma)$ for $t > 1$. 
222: In terms of the logarithmic variables, this leads to
223: \be
224: E(\sigma,t) = E(\sigma,1)+(t-1) F(\sigma). \l{sslimit1} 
225: \ee
226: Thus, (\ref{fullss}) reduces to a problem of non-interacting sequences whose 
227: population is growing linearly in time. 
228: The approximation of ignoring mutations after the seeding 
229: stage was tested numerically \cite{Krug03} and found to be in 
230: very good agreement with the full model described by (\ref{fullss}). 
231: 
232: A further simplification can be made by noting that 
233: due to (\ref{sslimit2}), the population $E(\sigma,1)$ 
234: is same for all the sequences that lie 
235: on a shell of constant Hamming distance $d(\sigma^{(0)},\sigma)$ away from the 
236: sequence $\sigma^{(0)}$. Since we will be interested in the sequence with 
237: the largest 
238: population at any instant, only the sequence with the largest 
239: fitness within a shell needs to be considered. Labeling a shell by its 
240: Hamming distance $k=0,1,...,N$ from the  sequence $\sigma^{(0)}$, we arrive 
241: at the {\it shell model} \cite{Krug03}
242: in which the population $E(k,t)$ of the fittest sequence 
243: in the shell $k$ with a total number of $\alpha_{k}$ sequences obeys 
244: \be
245: \l{lineq}
246: E(k,t)=-k + F(k) t \l{shell}.
247: \ee
248: Here $F(k)$ is the maximum of $\alpha_k$ 
249: i.i.d. random variables drawn from the fitness
250: distribution $p(F)$; hence the $F(k)$ are independent 
251: but non-identically distributed random variables. 
252: To arrive at the simple form (\ref{shell}), we have redefined the 
253: population as $E(k,t)+F(k)-F(0)$. 
254: The number of sequences in shell $k$ is given by the expression 
255: \be
256: \l{alphak}
257: \alpha_{k}= {N \choose k} \; (\ell-1)^{k},
258: \ee
259: as can be seen by noting that 
260: there are ${N \choose k}$ ways of choosing $k$ letters at which a 
261: sequence $\sigma$ differs from $\sigma^{(0)}$, and each 
262: of these $k$ letters can take $\ell-1$ values.
263: For large $N$, the majority of sequences is contained in a belt of width 
264: $\sim \sqrt{N}$
265: around the distance $k_{\mathrm{max}} = N (\ell -1)/\ell$ 
266: where (\ref{alphak}) is peaked. 
267: 
268: We will also consider an {\it i.i.d. model} for which the population $E(k,t)$ 
269: evolves according to (\ref{shell}) but the fitnesses $F(k)$ are i.i.d. 
270: random variables. This choice corresponds to a one-dimensional sequence space 
271: and we will see that our results are sensitive to the geometry of the 
272: sequence space. The i.i.d. model is closely related to the 
273: zero temperature limit of the problem of
274: directed polymers in the presence of columnar defects studied in 
275: \cite{Krug93} and a version of the parabolic
276: Anderson model \cite{Gaertner04}.   
277: 
278: Figure \ref{lines} illustrates the geometric picture of the evolutionary
279: process (\ref{shell}) 
280: that emerges after these simplifications and approximations. 
281: At a given time $t$, the most populated genotype
282: $k^\ast$ for which $E(k^\ast,t) = \max_k \{ E(k,t) \}$ leads until it is 
283: overtaken by a sequence $k'^\ast$ and so on.  
284: While the last leader is obviously located at the global fitness maximum, the 
285: identity of the previous leaders is non-deterministic. 
286: The inset of 
287: Fig.~\ref{lines} shows the punctuated evolution of the leading 
288: sequence $k^{\ast}$. 
289: Our main interest in the present work is in the statistical properties
290: of these leadership changes or evolutionary jumps. 
291: Initially, the sequence $\sigma^{(0)}$ with population $E(0,t)$ leads.
292: A shell $k'$ can overtake the currently leading shell 
293: $k < k'$ at the crossing time $T({k,k^{\prime}})$ given by 
294: \be
295: T(k',k)= \frac{k'- k} {F(k')-F(k)} \;\;, \l{Tkkprime}
296: \ee
297: which is positive provided $F(k') > F(k)$. In 
298: Fig.~\ref{lines}, only the shells $k=1, 2, 6, 8, 15$ 
299: can overtake $k=0$ (and only $k=2, 6, 8, 15$ can overtake $k=1$, ...)
300: since $F(15) > F(8) > F(6) > F(2) > F(1) > F(0)$. Such a set of fitness in 
301: which each value is progressively 
302: larger than all the previous ones defines a sequence of {\it{records}}. 
303: However, in order 
304: to appear in the evolutionary trajectory, it is not sufficient to be a record. 
305: As shown in Fig.~\ref{lines}, the shell $k=2$ is 
306: bypassed by shell $k=8$ since 
307: $T(8,1)=\mbox{min}_{k > 1} T(k,1)$. In general, if the current 
308: leader is in shell $k$, then the next leader is in the shell $k'$ for which 
309: the crossing time (\ref{Tkkprime}) is \textit{minimized} 
310: \cite{Rujan88}. Thus, in principle, the properties of leadership
311: changes can be formulated in terms of the extremal statistics of the 
312: matrix of crossing times. However, as will be explained in more detail in 
313: Sect.\ref{mft}, this minimization problem is too cumbersome 
314: to handle analytically and is further complicated by the presence of 
315: strong correlations among the crossing times.    
316: 
317: The nontrivial dynamics of the change in leadership described above is 
318: due to the competition between the initial population at a sequence and its 
319: fitness. As discussed above, a potential leader must be a record and hence 
320: must occur at a large Hamming distance away from the sequence $\sigma^{(0)}$
321: but the initial population at such a sequence is small 
322: due to (\ref{sslimit2}). Thus, the disadvantage due to poor initial condition 
323: may be overcome by a better fitness. Further, since 
324: the fitness of the leader is correlated to its Hamming distance from the 
325: sequence $\sigma^{(0)}$, the fitness $F(k^*)$ also shows punctuated evolution.
326:  
327: The rest of the paper is organised as follows.
328: In Sections~\ref{record} and \ref{jump}, we define records and jumps precisely 
329: and study some of their statistical properties, both for the shell model
330: and for the i.i.d. model. In Section~\ref{approach}, we 
331: describe the dynamics of the approach to the global maximum of the fitness 
332: landscape. A conjecture
333: concerning the probability of bypassing in the i.i.d. model, 
334: based on the results obtained in Sections~\ref{jump} and \ref{approach} 
335: along with numerical evidence, is presented in Section~\ref{beta-z}.
336: In Section~\ref{mft}, we introduce and discuss a further simplification of the 
337: problem described by (\ref{shell}), which highlights the strong 
338: interdependence of the crossing times.  
339: Finally, we summarise our results and conclude 
340: with a brief discussion of the possible biological relevance of this work 
341: in Section~\ref{conclude}.
342: 
343: %=============================================================================
344: %RECORDS
345: %=============================================================================
346: \section{Record statistics in sequence space}
347: \l{record}
348: 
349: In a sequence $\{ X_{k} \}$ of random variables, an upper (or lower) 
350: record is said to occur at $m$ if $X_{m} > X_{k}$ (or $X_{m} < X_{k}$) 
351: for all $k < m$ \cite{Glick78,Nevzorov87,Arnold98,Nevzorov01}. Since only the upper 
352: records are pertinent to our study, we shall henceforth refer to them as 
353: records. In the following subsections, we study some characteristics of 
354: these records; in particular, we find the mean number of records and 
355: the typical 
356: spacing between them. The record statistics of i.i.d. variables is 
357: well studied 
358: and we briefly review some of the known results as they are useful for 
359: later discussion. We study the record statistics in the shell model, for which 
360: the random variables are not identically distributed, in some detail. 
361: An appealing general feature of record statistics is that the results
362: are independent of the underlying probability distribution $p(F)$, 
363: which therefore does not need to be specified in this section. 
364: 
365: %=============================================================================
366: \subsection{Number of records}
367: 
368: In this subsection, we calculate the average number 
369: ${\cal{R}}_{\mathrm{shell}}$ of 
370: records in the shell model. 
371: For i.i.d. random variables, it is well known that the average number of records
372: among $N$ variables is ${\cal{R}}_{\mathrm{iid}} \approx \ln N$ for large $N$. To see 
373: this, note that the probability 
374: $\tilde{P}_{\mathrm{iid}}(k)$ that the $k$'th random variable is a record is equal to the 
375: probability 
376: that it is the largest among the first $k$ random variables. 
377: Since the location of the global 
378: maximum is uniformly distributed for i.i.d. variables, it follows that 
379: $\tilde{P}_{\mathrm{iid}}(k)=1/k$ for any choice of $p(F)$. 
380: Summing $\tilde{P}_{\mathrm{iid}}(k)$ up to $k=N$ yields the average number 
381: ${\cal{R}}_{\mathrm{iid}}$ of records to be $\ln N$. In fact, it 
382: is known that the probability distribution of the number of records is 
383: a Poisson distribution with mean $\ln N$ \cite{Sibani98}. 
384: 
385: The record statistics for non-i.i.d. variables is much less studied. 
386: A class of models in which records are obtained from a sequence $\{ X_{k} \}$ 
387: where each $X_{k}$ itself is the maximum of a set of $\alpha_{k}$ 
388: i.i.d. random variables was considered by Nevzorov \cite{Nevzorov84}. 
389: Such models have been used, for instance, in an (unsuccessful) attempt to 
390: account for the frequent occurrence of Olympic records due to an 
391: increasing world 
392: population \cite{Yang75}. The i.i.d. model is a special case  
393: corresponding to $\alpha_{k}=1$ for all $k$ and 
394: the shell model is obtained by using $\alpha_k$ given by (\ref{alphak}).
395: 
396: 
397: To analyze the Nevzorov model, 
398: we define a binary random variable $Y_{k}$ which takes value $1$ 
399: if a record occurs in the $k$'th set and $0$ otherwise. Since the distribution 
400: $p_{k}(F)$ of the maximum of $\alpha_{k}$ i.i.d. random variables is given by \cite{David70}
401: \be
402: \l{pk}
403: p_{k}(F)= \alpha_{k} \; p(F) \; \left( \int^{F}_{F_{\mathrm{min}}} p(x) dx \right)^{\alpha_{k}-1},
404: \ee 
405: we have
406: \be
407: \mbox{Prob}(Y_{k}=1)=\int_{F_{\mathrm{min}}}^{F_{\mathrm{max}}} dF \; 
408: \prod_{i=0}^{k-1} q_{i}(F) \;\; p_{k}(F)= 
409: \frac{\alpha_{k}}{(\alpha_{0}+...+\alpha_{k})} \;\;,\l{PYk1}
410: \ee
411: where $q_{k}(F)$ is the cumulative distribution corresponding to $p_{k}(F)$. 
412: The above equation expresses the fact that the $k$'th record value is a 
413: global maximum amongst $\sum_{j=0}^{k} \alpha_{j}$ variables 
414: and there are $\alpha_{k}$ ways in which it can occur in the $k$'th set. 
415: Further, the joint probability $\mbox{Prob}(Y_{k_{1}}=1,Y_{k_{2}}=1)$ for 
416: $k_{1} < k_{2}$ is given by 
417: \be
418: \mbox{Prob}(Y_{k_{1}}=1,Y_{k_{2}}=1)= \int_{F_{\mathrm{min}}}^{F_{\mathrm{max}}} dF \; 
419: \prod_{i=0}^{k_{1}-1} q_{i}(F) \;\;  p_{k_{1}}(F)
420: \int_{F}^{F_{\mathrm{max}}} dG \;\prod_{j=k_{1}+1}^{k_{2}-1} q_{j}(G) \;\; 
421: p_{k_{2}}(G)= 
422: \frac{\alpha_{k_{1}} \; \alpha_{k_{2}}}{(\alpha_{0}+...+\alpha_{k_{1}}) \;
423: (\alpha_{0}+...+\alpha_{k_{2}})} \;\;. \l{PYk12}
424: \ee
425: In a similar manner, it can be shown that 
426: $\mbox{Prob}(Y_{k_{1}}=1,...,Y_{k_{m}}=1) =
427: \prod_{j=1}^{m} \mbox{Prob}(Y_{k_{j}}=1)$ for any $m \ge 2$. 
428: Thus, the $Y_{k}$'s are 
429: independent, non-identically distributed variables 
430: \cite{Nevzorov87,Arnold98,Nevzorov01}. 
431: 
432: For the shell model, due to (\ref{alphak}) and (\ref{PYk1}), the probability 
433: $\tilde{P}_{\mathrm{shell}}(k,N)$ that a record occurs in the $k$'th shell is given by 
434: \be
435: \tilde{P}_{\mathrm{shell}}(k,N)=\mbox{Prob}(Y_{k}=1) \approx 1- \left( \frac{1}{\ell-1} \right) 
436: \frac{a}{1-a} \;\;\;\;,\;\;\;\; a= \frac{k}{N} < \frac{\ell-1}{\ell} =
437: \frac{k_{\mathrm{max}}}{N} \;\;,
438: \l{Ptildeshell}
439: \ee 
440: where we have used that 
441: ${N \choose k-m} / {N \choose k} \approx [a/(1-a)]^{m}$ for $k, N \gg 1$ with 
442: $k/N$ fixed. Since it is easier to break records in the beginning, the 
443: probability to find a record is near unity for
444: $k \ll N$. However, it vanishes beyond $k_{\mathrm{max}}$ because 
445: the global maximum typically occurs in the shell $k_{\mathrm{max}}$. 
446: The average 
447: number ${\cal{R}}_{\mathrm{shell}}$ of records can be obtained by simply 
448: integrating $\tilde{P}_{\mathrm{shell}}(k,N)$ over $k$ and we find  
449: \be
450: {\cal{R}}_{\mathrm{shell}} \approx \frac{(\ell-\ln\ell-1)}{\ell-1} \; N \;\;.
451: \ee
452: Thus, ${\cal{R}}_{\mathrm{shell}}$ 
453: increases with $\ell$ and as $\ell \rightarrow \infty$, 
454: ${\cal{R}}_{\mathrm{shell}} \rightarrow N$ because the population in each shell
455: becomes infinite. 
456: Since the $Y_{k}$'s are independent random variables with finite mean 
457: and variance, the number of records satisfies the central limit theorem
458: and becomes Gaussian for large $N$. Specifically, for large $N$, 
459: the variance of the number of records is
460: \be
461: \l{variance} {\cal{V}}_{\mathrm{shell}} = \sum_{k=0}^N 
462: \mbox{Prob}(Y_k = 1)[1 - \mbox{Prob}(Y_k = 1)]
463: \approx
464: \frac{N}{\ell - 1} 
465: \left( \frac{\ell + 1}{\ell - 1} \ln \ell - 2 \right) \;\;,
466: \ee
467: which is maximal for $\ell = 5$ and vanishes for large $\ell$.
468: The ratio ${\cal{V}}_{\mathrm{shell}}/{\cal{R}}_{\mathrm{shell}}$ is always 
469: considerably smaller
470: than unity, which implies that the distribution is much sharper than the
471: log-Poisson distribution obtained in the i.i.d. case.
472: 
473: %=============================================================================
474: \subsection{Inter-record spacings}
475: \l{recordspace}
476: 
477: For the discussion of the spacings between records, it is convenient
478: to label the records ``backwards'' in time, with $r_1$ denoting
479: the position of the last record (i.e., the global maximum), $r_2 < r_1$
480: the penultimate record, and so on.
481: In this way the pathologies associated with the fact that
482: the expected waiting time for the next record to occur is
483: infinite in the i.i.d. case \cite{Glick78} can be avoided.  
484: The probability $\tilde{P}(r_{j})$ that the $j$'th record
485: occurs at location $r_{j}$ 
486: can be found for the Nevzorov model in a manner analogous to 
487: (\ref{PYk12}). We obtain 
488: \be
489: \tilde{P}(r_{j})= 
490: \sum_{r_{1},..., r_{j-1}} \;  
491: \prod_{k=0,...,j-1} \frac{\alpha_{r_{k+1}}}
492: {\alpha_{0}+...+\alpha_{{r_{k}-1}}}  \;\;,\l{recloc}
493: \ee 
494: with $N \geq r_{1} >...> r_{j-1} > r_j$ and $r_{0}=N+1$. The factors on the 
495: right hand side of the above equation simply reflect the fact that 
496: the record at location $r_{k+1}$ remains the maximum 
497: until the next record occurs at $r_{k} > r_{k+1}$. 
498: 
499: For the i.i.d. model, using $\alpha_{k}=1$ for all $k$ in (\ref{recloc}),  
500: the average location $\av{r_{j}}=\sum r_{j} \tilde{P}(r_{j})$ can be 
501: calculated. One finds that the average inter-record distance 
502: $\tilde{\Delta}_{\mathrm{iid}}(j)=\av{r_{j}}-\av{r_{j+1}}$ 
503: between the $j$'th and $(j+1)$'th record behaves as \cite{Arnold98,Nevzorov01}
504: \be
505: \tilde{\Delta}_{\mathrm{iid}}(j) \approx \frac{N}{2} \left( \frac{1}{2} \right)^{j} 
506: \;\;,\;\;j=1,2,... \;\;. \l{interreciid}
507: \ee
508: A simple argument, useful for later discussion, can also be employed 
509: to obtain the above equation.
510: The record labeled by $j=1$, being the global maximum, is equally 
511: likely to occur 
512: anywhere between $1$ and $N$. Thus, on average, it is located at $N/2$. 
513: Similarly, the record labeled by $j=2$ is a global maximum in the 
514: range $[1,r_1)$ with 
515: uniformly distributed location, which gives the average location 
516: $\av{r_{2}}=(1/2) \av{r_1} = N/4$. Repeating this argument, we obtain the result in 
517: (\ref{interreciid}).
518: 
519: For the shell model, since the most likely position of the 
520: global maximum is in the shell with the largest number $\alpha_k$ of 
521: sequences, we have $\av{r_{1}} = k_{\mathrm{max}}$. 
522: For sake of simplicity, we consider binary sequences ($\ell=2$)
523: in the following but the scaling behavior obtained below holds 
524: for $\ell > 2$ as well. We find 
525: that for $j \ge 2$, the average location $\av{r_{j}}$ of the 
526: $j$'th record is given by 
527: \be
528: \av{r_{j}}=\av{r_{1}} -\frac{1}{\pi} \sqrt{\frac{N}{2}} 
529: \left( \frac{2}{\sqrt{\pi}} \right)^{j-2} 
530: \int_{-\infty}^{\infty} dx_{j-1} \; 
531: \frac{e^{-x_{j-1}^{2}}}{\mbox{erfc}(x_{j-1})} \; \int_{x_{j-1}}^{\infty} 
532: dx_{j-2} \; \frac{e^{-x_{j-2}^{2}}}{\mbox{erfc}(x_{j-2})}  \; ... \;
533: \int_{x_{2}}^{\infty} 
534: dx_{1} \; \frac{e^{-2 x_{1}^{2}}}{\mbox{erfc}(x_{1})}  \;\;,\;\;j \ge 2 \;\;,
535: \l{rj}
536: \ee
537: where we have used Eqs.(\ref{cnr})-(\ref{avgcnr}) and 
538: performed a Gaussian integral. The average location $\av{r_{2}}$ 
539: of the second record is given by 
540: $\av{r_{2}}=\av{r_{1}}- 2.0064 \sqrt{N}/ \pi \sqrt{2}$. Thus, the second 
541: record (and in fact, $j$'th record for $j$ of order unity) can be found 
542: within ${\cal{O}}(\sqrt{N})$ distance of the global maximum since 
543: $\alpha_{k}$ has width $\sim \sqrt{N}$ about $k_{\mathrm{max}}$. 
544: For $j > 2$, 
545: after repeated integration by parts, (\ref{rj}) can be rewritten as 
546: \be
547: \av{r_{j}}=\av{r_{1}}-\frac{1}{\pi} \sqrt{\frac{N}{2}} \sum_{m=1}^{j-2} 
548: \frac{(\ln 2)^{j-2-m}}{(j-2-m)!} \;\; G(m) \;\;,\;\;j > 2 \;\;,\l{sspacing}
549: \ee
550: where 
551: \be
552: G(m)= (-1)^{m} \int_{-\infty}^{\infty} dx \; 
553: \frac{e^{-2 x^{2}}}{\mbox{erfc}(x)} \; 
554: \frac{\left( \ln {\mbox{erfc(x)}} \right)^{m}- 
555: \left( \ln {2} \right)^{m}}{m!} \;\;.
556: \ee
557: As outlined in Appendix {\ref{appendix1}, the integral $G(m)$ can be 
558: estimated by the saddle point method and we find 
559: $G(m) \approx \sqrt{ \pi m/2}$ for large $m$. Since the leading contribution 
560: to the sum in (\ref{sspacing}) comes from the $m=j-2$ term, we have 
561: \be
562: \tilde{\Delta}_{{\mathrm{shell}}}(j) \approx  
563: \sqrt {\frac{N}{4 \pi j}} \;\;,\;\; j \gg 1 \;\;. \l{interrecshell}
564: \ee
565: Thus, while the inter-record spacing decays exponentially with $j$ for 
566: the i.i.d. model, it falls as a power law for the shell model. 
567: The spacing between the first few records [$j = {\cal{O}}(1)$]
568: is of order $\sqrt{N}$, while for the bulk of the records with $j = 
569: {\cal{O}}(N)$ the spacing is of order unity; this is consistent with
570: the vanishing of the record occurrence probability (\ref{Ptildeshell})
571: near $k=k_{\mathrm{max}}$. 
572: 
573: 
574: %=============================================================================
575: %JUMPS
576: %=============================================================================
577: \section{Jump statistics}
578: \l{jump}
579: 
580: As we described in Sect.\ref{model}, in our model, evolutionary jumps
581: are a subset of records and if a jump occurs at 
582: $k^{\prime}$, the next jump is said to occur 
583: at $k > k^{\prime}$ if (i) $F(k)$ is a record (ii) the overtaking time 
584: $T(k, k^{\prime})= \mbox{min}_{j \geq k^{\prime}}  
585: \{ T(j, k^{\prime}) \}$. By convention, the first jump and the first record
586: occurs at $k=0$. Due to the second condition, some of the records can get 
587: bypassed and fail to appear in the set of jumps. 
588: In this section, we find the mean number of jumps and the 
589: inter-jump spacing for both the i.i.d and the shell model. 
590: In contrast to the properties of records, the statistics of jumps
591: depends explicitly on the fitness distribution $p(F)$ and we will consider 
592: distributions for which the tail behavior corresponds to the 
593: three universality classes of 
594: standard extreme value theory \cite{Sornette00} and which also 
595: appear in a very similar form in the theory of records 
596: \cite{Nevzorov87,Arnold98,Nevzorov01}. 
597: 
598: %=============================================================================
599: \subsection{Mean number of jumps}
600: \l{meanjump}
601: 
602: We begin by discussing numerical results for the average number of 
603: jumps in the i.i.d. model. 
604: It was found in \cite{Krug03} that the average number ${\cal{J}}_{\mathrm{iid}}$ 
605: of jumps grows as $\beta \ln N$ where the 
606: prefactor $\beta < 1$, and was conjectured to be 
607: \be
608: \beta \approx \cases { 1/2   & {$\;,\;\;p(F) \sim e^{-F}$} \cr
609: (\delta-1)/(2 \delta -1) & {$\;,\;\;p(F) \sim F^{-1-\delta}, \;\; \delta \geq 1$}  \cr
610: (2 + \nu)/(3 + 2 \nu)  & {$\;,\;\;p(F) \sim (F_{\mathrm{max}}-F)^{\nu}, \;\; \nu > -1$} 
611: } \;\;. \l{beta1d}
612: \ee
613: Figure~\ref{iidjump} shows ${\cal{J}}_{\mathrm{iid}}$ increasing linearly with 
614: $\ln N$ and slope $\beta$ for 
615: some distributions in accordance with (\ref{beta1d}). 
616: Thus, in the i.i.d. model
617: the jumps can be viewed as ``diluted'' records, in the sense that the mean 
618: number of 
619: records and jumps differ only up to a prefactor and  
620: the probability $P_{\mathrm{iid}}(k,N)$ that a jump occurs at $k$ 
621: is given by $\beta /k$. In this picture, the probability for a given
622: record to be bypassed is simply $1-\beta$. However, bypassing is not
623: completely random, as the variance of the number of jumps is found
624: consistently to be smaller than the mean. This implies a certain amount
625: of ``anti-bunching'' among the jumps, which can also be detected
626: by the direct measurement of correlation 
627: $C_{\mathrm{iid}}(k_{1},k_{2})=P_{\mathrm{iid}}(k_{1},k_{2})-
628: P_{\mathrm{iid}}(k_{1}) P_{\mathrm{iid}}(k_{2})$  
629: where $P_{\mathrm{iid}}(k_{1},k_{2})$ is the joint distribution of 
630: having a jump at $k_{1}$ and $k_{2}$, as shown in the inset 
631: of Fig.~\ref{rspace}. 
632: Some further discussion of the conjecture (\ref{beta1d}) will be provided
633: in Sect.\ref{beta-z}.
634: 
635: Estimates for the mean number of jumps in the shell model were also
636: given in \cite{Krug03}, but the range of sequence lengths was too limited
637: to allow for a definite statement. Here we present the results of our 
638: simulations for large values of $N$ obtained using the approximation 
639: described below. Since the shell fitness $F(k)$ is the 
640: maximum of $\alpha_k$ i.i.d. random variables drawn from the distribution
641: $p(F)$, it can be obtained from a uniform random variable 
642: $u$ using the relation   
643: \be
644: \l{uniform}
645: \int_{F_{\mathrm{min}}}^{F(k)} p(x) dx =u^{1/\alpha_{k}}.
646: \ee 
647: However, the binomial coefficient ${N \choose k}$ increases  
648: exponentially with $N$ for 
649: large $k$ and the distribution of $u^{1/\alpha_{k}}$ approaches a 
650: delta-function centred at unity for $k \gg 1$ making it difficult to determine 
651: the distribution $p_{k}(F)$ of $F(k)$ accurately when $N$ is large. 
652: For the exponential fitness distribution, the relation (\ref{uniform})
653: becomes 
654: \be
655: \l{Fexp}
656: F(k)=-\ln (1- e^{\ln u/\alpha_{k}}) \simeq - \ln [\ln(1/u)]+\ln(\alpha_{k}).
657: \ee
658: Since the last expression only involves the logarithms of binomial 
659: coefficients, $F(k)$ can be easily generated up to large values of $N$.  
660: While a similar approximation can be employed for other 
661: distributions with unbounded tails, we have not been able 
662: to obtain reliable results for bounded distributions.  
663: 
664: In Fig.~\ref{sjump}, we show simulation results for the
665: probability $P_{\mathrm{shell}}(k,N)$ that a jump occurs in 
666: the $k$'th shell, for a binary alphabet  
667: ($\ell=2$) and two different values of $N$. The data collapses 
668: onto the scaling form 
669: \be
670: P_{\mathrm{shell}}(k,N) \approx  N^{-1/2} f(k/N) \;\;\;,\;\; 
671: x < \frac{\ell-1}{\ell} = \frac{k_{\mathrm{max}}}{N} \;\;,
672: \l{scaling_jumps}
673: \ee
674: where the scaling function behaves as $f(x) \sim x^{-1/2}$ for small $x$.
675: This has the interesting consequence that 
676: $P_{\mathrm{shell}} \sim 1/\sqrt{k}$ is independent
677: of $N$, and hence the number of jumps grows as $\sqrt{k}$ 
678: for $k \ll k_{\mathrm{max}}$.
679: The scaling function appears to be independent of the alphabet size $\ell$,
680: which therefore only changes the cutoff at $k = k_{\mathrm{max}}$ \cite{Krug04}.
681: Thus, the average number ${\cal{J}}_{\mathrm{shell}}$ of jumps obtained by 
682: integrating $P_{\mathrm{shell}}(k,N)$ over $k$ grows with $N$ as 
683: \be
684: {\cal{J}}_{\mathrm{shell}} \sim \sqrt{\frac{(\ell -1) N}{\ell}} \;\;\;.
685: \ee
686: Our simulations indicate that the above dependence of 
687: ${\cal{J}}_{\mathrm{shell}}$ on $N$ is also true for 
688: normal-distributed fitness and we expect it to hold 
689: for all distributions decaying more rapidly than 
690: any power law \cite{Krug04}. Further, the distribution of the number of jumps 
691: is a Gaussian (as in the case of records) with both mean and variance 
692: scaling as $\sqrt{N}$.  
693: 
694: For the power law case, we find that ${\cal{J}}_{\mathrm{shell}} \to 1$ for 
695: large $N$. 
696: As we shall see in Section~\ref{approach}, the globally fittest sequence 
697: takes over the leadership in a time of order unity in this case, 
698: which explains the above behavior of ${\cal{J}}_{\mathrm{shell}}$. 
699: In summary, unlike the i.i.d. model, most of the records are 
700: bypassed in the shell model both for exponential and power law distributions. 
701: 
702: 
703: %=============================================================================
704: \subsection{Inter-jump spacings}
705: \l{inter-jump}
706: 
707: We now turn to a discussion of the inter-jump spacings 
708: $\Delta(j)$ 
709: defined in analogy to the inter-record spacings.
710: We denote by $s_j$ the position of the $j$'th jump, with
711: $j=1$ referring to the last jump, which is also the last
712: record (the global maximum), and define the inter-jump spacing
713: as $\Delta(j) = \av{s_{j}} - \av{s_{j+1}}$ where $\av{s_{j}}$ is the average 
714: location of the $j$'th jump. 
715: For the i.i.d. model, an approximate calculation of  
716: $\Delta_{\mathrm{iid}}(j)$ can be carried out by assuming jumps to be 
717: randomly diluted records.
718: The position $s_j$ of the $j$'th jump equals the position
719: $r_k$ of the $k$'th record, where $k \geq j$ because of the possibility of
720: bypassing. If the record $k+1$ is not bypassed, then $s_{j+1} = r_{k+1}$
721: and $s_{j+1} = (1/2) s_j$ on average due to the argument 
722: given after (\ref{interreciid}) for the inter-record spacings.
723: In the diluted record picture,
724: this is true with probability $\beta$. With probability $\beta (1 - \beta)$,
725: record $k+1$ is bypassed and $s_{j+1} = r_{k+2}$ so that 
726: $s_{j+1} = (1/4) s_j$ on the average. Similarly, with
727: probability $\beta (1 - \beta)^l$, $l$ records are bypassed
728: and the ratio $s_{j+1}/s_j = (1/2)^{l+1}$ on average. Summing over
729: all possibilities, we find that $\av{s_{j+1}} = b \av{s_j}$ with
730: \be
731: \l{bgeom}
732: b = \sum_{l = 0}^\infty 2^{-(l+1)} \beta (1 - \beta)^l = \frac{\beta}{1 + \beta}.
733: \ee
734: Taking into account that the sum over all inter-jump spacings adds up to 
735: $\av{s_1} = \av{r_1} = N/2$, we obtain
736: \be
737: \Delta_{\mathrm{iid}}(j) \approx \frac{N}{2} (1- b) b^{j-1}
738: \;\;,\;\;j=1,2,... \;\;.\l{interjumpiid}
739: \ee
740: The numerical data shown in Fig.~\ref{rspace} supports the general
741: form of this expression, but with the coefficient 
742: $b \approx \beta/2$ rather than $\beta/(1+\beta)$. This is another
743: indication of correlations that are not accounted for in the random
744: dilution picture.
745: 
746: For the shell model, as shown in Fig.~\ref{sjump}, the data 
747: for $\Delta_{\mathrm{shell}}(j)$ for various $N$ collapses onto a 
748: monotonically decreasing curve if we assume $\Delta_{\mathrm{shell}}(j)$ 
749: to be of the scaling form 
750: \be
751: \Delta_{\mathrm{shell}}(j) \approx 
752: \sqrt{N} \; h(j/\sqrt{N}) \;\;. \l{interjumpshell}
753: \ee
754: The scaling with $\sqrt{N}$ follows naturally from the scaling form 
755: (\ref{scaling_jumps}) for the jump occurrence probability. The latter
756: implies that the density of jumps on the $k$-axis is of order $1/\sqrt{N}$
757: for finite $k/N$, hence the spacing is of order $\sqrt{N}$, and the
758: argument of the scaling function is $j/\sqrt{N}$ because the total number
759: of jumps is also of order $\sqrt{N}$. However, the tail of the scaling 
760: function $h$ does not obey the scaling form (\ref{interjumpshell}) and 
761: approaches zero with increasing $N$. Thus, while for finite $k/N$, the jumps 
762: are roughly equally spaced with spacing $\sqrt{N}$, the spacing is of 
763: ${\cal{O}} (N^{s})$ with $s < 1/2$ for $k/N \rightarrow 0$.
764: 
765: %=============================================================================
766: %APPROACH
767: %=============================================================================
768: 
769: \section{Approach to the global fitness maximum}
770: \l{approach}
771: 
772: So far, we have discussed the statistics on the $k^{\ast}$ axis of the 
773: inset of Fig.~\ref{lines} and now we focus on the temporal statistics.   
774: As explained in Section~\ref{model}, a fit sequence $k^{\ast}$ leads till 
775: it is 
776: overtaken by an even fitter one and eventually the globally fittest sequence 
777: emerges as the leader at typical time $T^{\ast}$. 
778: In this section, we find this typical time $T^{\ast}$ required 
779: for the population to reach the global maximum of the fitness landscape 
780: and the distribution of the evolution times to jump from one local maximum to 
781: another. 
782: 
783: %=============================================================================
784: \subsection{Dynamic scaling}
785: 
786: The location $k^{\ast}(t)$ of the most populated sequence at time $t$ for 
787: which the logarithmic population is 
788: $E({k^{\ast}},t)=\mbox{max}_{k} \{ E(k,t) \}$ increases with time till 
789: the global maximum is reached. In Appendix~\ref{appendix2}, the 
790: distribution $P_t(k^{\ast})$ is explicitly calculated for the 
791: i.i.d. model in the limit of infinite genome size ($N \to \infty$). 
792: This distribution is usually of the scaling form 
793: \be
794: \l{Ptscale}
795: P_t(k^\ast) = t^{-1/z} \Phi(k^\ast/t^{1/z}) \;, \l{Ptkast}
796: \ee
797: where the scaling function $\Phi$ depends on the underlying fitness 
798: distribution
799: and the dynamic exponent $z$ is given by 
800: \be
801: z= \cases { 1   & {$\;,\;\;p(F) \sim e^{-F}$} \cr
802:   (2 + \nu)/(1 + \nu)  & {$\;,\;\;p(F) \sim (F_{\mathrm{max}}-F)^{\nu}, \;\; \nu > -1$} \cr
803:      (\delta-1)/\delta & {$\;,\;\;p(F) \sim F^{-1-\delta}, \;\; \delta \geq 1$}
804: } \;\;, \l{1dz}
805: \ee
806: for the three classes of fitness distributions introduced in Sect.\ref{jump}. 
807: The corresponding behavior $\overline{k^{\ast}(t)} \sim t^{1/z}$ 
808: of the mean location has been derived previously using Flory-type arguments 
809: \cite{Zhang86,Krug93} and can also be seen using (\ref{Ptkast}). As we have 
810: already discussed, the global maximum is reached
811: when $\overline{k^{\ast}(t)} \approx N/2$, which defines a total evolution 
812: time $T^\ast \sim N^z$. One thus expects a scaling form 
813: \be
814: \overline{k^{\ast} (t,N)} \approx t^{1/z} \varphi (t/T^{\ast}) \;\;,
815: \ee
816: where the scaling function $\varphi(x)$ is a constant for $x \ll 1$ and 
817: decays as $x^{-1/z}$ for $x \gg 1$ \cite{remark}.
818: 
819: An alternative approach \cite{Krug02,Krug03} to estimating 
820: the typical time $T^{\ast}$ required by the population 
821: to reach the global maximum starts from the observation that
822: $T^\ast$ is of the order of the time $T_1$ at which 
823: the globally fittest sequence at typical location $r_{1} = s_1$ and fitness 
824: $F(r_1)$ 
825: overtakes the penultimate leader with respective quantities $s_2$ and 
826: $F(s_2)$ (refer Sect.\ref{inter-jump}). From (\ref{Tkkprime}), we have
827: \be
828: T_1 =  \frac{s_1 - s_2}{F(s_1)-F(s_2)} \;\;,
829: \l{Tast}
830: \ee
831: where the inter-jump spacing in the numerator is given by 
832: (\ref{interjumpiid}). 
833: The estimation of the denominator involves a subtlety -- in previous 
834: works \cite{Krug02,Krug03}, it was assumed that $F(s_1) - F(s_2)$ is of the
835: order of the \textit{fitness gap} $\epsilon$, which was defined as 
836: the difference between the values of the global maximum 
837: and the second largest fitness of the fitness landscape. 
838: Because the second largest fitness does not necessarily appear in the 
839: record sequence,
840: $\epsilon$ is only a lower bound on $F(r_1) - F(r_2)$, which in turn is clearly
841: a lower bound on the fitness difference of interest,
842: $\epsilon \leq F(r_1) - F(r_2) \leq F(s_1) - F(s_2)$. However, our 
843: explicit calculations show that $F(r_1) - F(r_2)$ is of the 
844: same order as $\epsilon$; moreover, at least for the i.i.d. model, we know 
845: that at most a few
846: records are bypassed between $s_1$ and $s_2$, and hence the assumption
847: that $F(s_1) - F(s_2) \sim \epsilon$ seems justified.
848: 
849: The calculation of the distribution of fitness gap $\epsilon$ is a 
850: standard exercise in
851: extreme value statistics \cite{David70}. 
852: In a system with a total number $S$ of sequences, the typical value of the 
853: fitness gap increases as $S^{1/\delta}$ for the unbounded power 
854: law distribution, decreases as $S^{-1/(1+\nu)}$ for the bounded distribution,
855: and is of order unity for the exponential distribution. 
856: For the i.i.d. model, using 
857: $s_{1} - s_{2} \sim N$ [see (\ref{interjumpiid})] and $S=N$ in 
858: (\ref{Tast}), we recover $T_1 \sim T^{\ast} \sim N^{z}$ with $z$ given in 
859: (\ref{1dz}). The other fitness distributions can be treated in a similar 
860: manner. 
861: 
862: For the shell model, $S=\ell^{N}$ and due to (\ref{interjumpshell}), the 
863: numerator is of order $\sqrt{N}$ so that $T^{\ast} \sim \sqrt{N}$ for 
864: the case of exponentially distributed fitness. Presumably, the time $T_{j}$ 
865: at which the $j$'th jump occurs is also of order $\sqrt{N}$ for 
866: $j \sim {\cal{O}}(1)$. Since the total number of jumps is of the same order, 
867: it follows that initially there are many, quick jumps followed by few 
868: jumps that take ${\cal{O}} (\sqrt{N})$ time. 
869: This result agrees qualitatively with that seen 
870: in experiments (discussed later) concerning the pace of evolution which is 
871: initially rapid and later slows down considerably. 
872: For the bounded distributions, $T^{\ast}$ increases exponentially with $N$ 
873: whereas it \textit{decreases} exponentially for the 
874: fat-tailed power law distributions. The latter result implies 
875: that, for large $N$, the global maximum
876: takes over in a single time step, which explains why the mean number of 
877: jumps tends to unity  for the power 
878: law distributions (see Sect.\ref{meanjump}).
879: 
880: 
881: %=============================================================================
882: \subsection{Universal tails of the evolution time distribution}
883: 
884: 
885: In the last subsection, we found the typical time $T^{\ast}$ to 
886: reach the global maximum and now we 
887: consider the distribution $P(T_{1},N)$ of the time $T_{1}$ 
888: at which the final jump occurs. For the i.i.d. model, since the 
889: typical $T_{1}$ also grows 
890: as $N^{z}$, we may expect the normalised distribution 
891: $P(T_{1},N)$ to be of the scaling form 
892: $P(T_{1},N) \approx N^{-z} g_{1}(T_{1}/N^{z})$.
893: In general, for $j \sim {\cal{O}}(1)$, the distribution $P(T_{j},N)$ of the time 
894: $T_{j}$ at which the $j$'th jump occurs is of the scaling form 
895: \be
896: \l{Tjscale}
897: P(T_{j},N) \approx N^{-z} g_{j}(T_{j}/N^{z}) \;\;.
898: \ee
899: Although the dynamic exponent 
900: $z$ depends on the underlying fitness distribution, we shall now show 
901: that the tail of the 
902: distribution $P(T_{j},N)$ is universal. The events contributing to 
903: large $T_{1}$ are the ones for which $F(s_1) - F(s_2)$ is small; 
904: in these cases we expect the general bound $F(s_1) - F(s_2) \geq F(r_1) - F(r_2)$
905: to be saturated, i.e. the second record is not bypassed and $s_2 = r_2$. 
906: Thus, using (\ref{Tast}), we obtain
907: \be
908: \l{PT1}
909: P(T_{1},N) \approx \left| \frac{d \epsilon_1}{d T_1} \right|
910: \mbox{Prob}(\epsilon_1 = \Delta_{\mathrm{iid}}(1)/T_1) \approx
911: \frac{\Delta_{\mathrm{iid}}(1)}{T_{1}^2}  \; 
912: \mbox{Prob}(\epsilon_{1}=0) \;\;, 
913: \ee
914: for large $T_1$,
915: where $\epsilon_{1}=F(r_{1})-F(r_{2})$. 
916: The probability distribution of $\epsilon_1$ can be obtained along
917: the lines of the derivation of the distribution of the fitness gap
918: $\epsilon$ in \cite{Krug02}, and it is found that 
919: the probability for a near-vanishing difference between two successive record 
920: values is nonzero for any $p(F)$.
921: We conclude that $P(T_{1},N)$ has a 
922: power law tail with 
923: exponent $-2$ for any underlying fitness distribution \cite{Krug03}.
924: This is an example of the generation of a power law through a change
925: of variables (from $\epsilon_1$ to $T_1 \sim 1/\epsilon_1$) as described
926: in \cite{Sornette00}. 
927: 
928: Similarly, the events contributing to the tail of $P(T_{2},N)$ 
929: are the ones in which the record at $r_{2}$ is the penultimate leader and 
930: the record at $r_{3}$ is the leader previous to it. Thus, we demand that 
931: none of these two records should be bypassed; in particular,
932: $r_2$ should not be bypassed, which requires 
933: that $T(r_{1}, r_{2}) > T(r_{2},r_{3})$.  
934: This condition can be written as 
935: $\epsilon_{1} < C \epsilon_{2}$, where
936: $\epsilon_2 = F(r_{2})-F(r_{3})$ and $C$ is the average value
937: of $(r_1-r_2)/(r_2-r_3)$, a number of order unity. Thus we obtain 
938: \be
939: P(T_{2},N) \approx  \frac{\Delta_{\mathrm{iid}}(2)}{T_{2}^2}  
940: \int_{0}^{C \Delta_{\mathrm{iid}}(2)/T_2} d \epsilon_{1} \; 
941: \mbox{Prob}(\epsilon_{1}, \epsilon_{2},N)  \;
942: \approx \frac{C \Delta_{\mathrm{iid}}(2)^2}{T_2^3}
943: \mbox{Prob}(\epsilon_{1}=0, \epsilon_{2}=0,N).  
944: \ee
945: Since Prob$(\epsilon_{1}=0, \epsilon_{2}=0,N)$ can be shown to be 
946: nonzero, we conclude  
947: that $P(T_{2},N) \sim T_2^{-3}$ for large $T_2$. 
948: Extending the above arguments in a similar fashion to the next evolution 
949: times, we find that the scaling functions in (\ref{Tjscale}) behave as
950: $g_{j}(x) \sim x^{-1-j}$ for $x \gg 1$. 
951: Interestingly, this implies that the expected time $\av{T_j}$ is finite
952: for $j \geq 2$ and infinite for $j = 1$.
953: In Fig.~\ref{evol}, the prediction 
954: $P(T_{j},N) \sim T^{-1-j}$ for $j=1,2,3$ is compared with data 
955: obtained using Monte Carlo simulations for the i.i.d. model. 
956: The behavior of the universal tails of $P(T_{j})$ discussed above 
957: is true for the shell model as well.
958: 
959: 
960: %=============================================================================
961: %BETA-Z
962: %=============================================================================
963: \section{Bypassing probability and dynamic exponent: a conjecture}
964: \l{beta-z}
965: 
966: As we have seen in the previous sections, the jump statistics are not 
967: analytically tractable due to the constraint of minimal overtaking time. 
968: For the i.i.d. model, the jumps differ from the records only up to a prefactor 
969: $\beta$ conjectured to be given by  (\ref{beta1d}). 
970: Comparing the expressions in Eqs.(\ref{beta1d}) and (\ref{1dz}), we observe 
971: that 
972: the bypassing probability $1- \beta$ appears to be related to the 
973: dynamic exponent $z$ by the following universal relation 
974: \be
975: \beta = (1-\beta) \; z \;\;. 
976: \l{relation}
977: \ee
978: A derivation of this relation (which eludes us so far) would constitute a 
979: proof of the conjecture (\ref{beta1d}). 
980: 
981: Interestingly, the relation (\ref{relation}) can be interpreted in terms of a 
982: kind
983: of \textit{duality} between the $k$- and the $t$-axis of the inset of 
984: Fig.~\ref{lines}. So far we have identified each jump with the position on the
985: $E$-axis where the line that takes over the leadership when the jump
986: occurs originates; but we may just as well identify the jump with the 
987: corresponding crossing time $T(k,k^{\prime})$ at which the leadership 
988: shifts from $k'$ to $k$. 
989: Clearly, there is a one-to-one correspondence between the jumps defined on 
990: two axes and the average number ${\cal{J}}_{\mathrm{iid}}^{\prime}$ of jumps 
991: on the time axis is equal to ${\cal{J}}_{\mathrm{iid}}$ discussed in earlier 
992: sections. 
993: Thus, ${\cal{J}}_{\mathrm{iid}}^{\prime}={\cal{J}}_{\mathrm{iid}} 
994: \approx \beta \ln {N}$ and 
995: since the 
996: typical time $T^{\ast}$ to reach the global maximum scales as $N^{z}$, we have 
997: ${\cal{J}}_{\mathrm{iid}}^{\prime} \approx \beta^{\prime} 
998: \ln {T^{\ast}}$ with $\beta^{\prime}= 1- \beta$ due to the conjecture 
999: (\ref{relation}).  
1000: This leads us to expect that the probability 
1001: $P_{\mathrm{iid}}(t,N)$ that a jump occurs at time $t$ decays as 
1002: $\beta^{\prime}/t$ for $t \ll N^{z}$ and as 
1003: $1/t^{2}$ for $t \gg N^{z}$; the latter behavior is the universal 
1004: $t^{-2}$ tail explained in Section~{\ref{approach}}. 
1005: We conclude that 
1006: the sum of the jump probabilities along the $k$- and the $t$-axes should
1007: sum up to a universal function, 
1008: {\it i.e.}
1009: \be
1010: P_{\mathrm{iid}}(t=X,N)+P_{\mathrm{iid}}(k=X,N) = 1/X \;\;,
1011: \ee
1012: for any choice of fitness distribution. 
1013: The numerical evidence supporting this claim is shown in Fig.~\ref{sum}.
1014: Furthermore, in analogy to the jump spacing $\Delta_{\mathrm{iid}}(j)$ along 
1015: the $k$-axis, 
1016: one can also consider the quantity 
1017: $\Delta_{\mathrm{iid}}^{\prime}(j) = \langle T_j \rangle - 
1018: \langle T_{j+1} \rangle$ 
1019: which is the spacing 
1020: between the successive jumps on the time-axis. Replacing $N$ by $T^{\ast}$ 
1021: and $b = \beta/2$ by $\beta'/2 = (1-\beta)/2$ in (\ref{interjumpiid}), we 
1022: expect
1023: \be
1024: \Delta_{\mathrm{iid}}^{\prime}(j) \sim N^{z} 
1025: \left( \frac{1-\beta}{2} \right)^{j-1} \;\;,\;\;j=1,2,...  \;\;. \l{cnjctre}
1026: \ee
1027: Numerical results consistent with this expression 
1028: are shown in the inset of Fig.~\ref{sum} for some distributions. 
1029: The deviations seen in the data for the first jump ($j=1$) reflect the 
1030: fact that,
1031: because of the $1/T_1^2$-tail derived in Eq.(\ref{PT1}), the average of $T_1$
1032: is not defined and hence grows with the number of disorder realizations. 
1033: 
1034: %=============================================================================
1035: %RECORD TIMES
1036: %=============================================================================
1037: \section{A Model Based on Record Times}
1038: \l{mft}
1039: 
1040: In this section, we introduce a further simplification of the i.i.d. model. 
1041: As we have discussed already, a sequence $k$ may occur in the set of 
1042: jumps provided $F(k)$ is a record. Thus, it is sufficient to consider 
1043: only the subset of sequences whose fitness is a record.
1044: Here it is convenient to label the records forward in time, so we
1045: denote by $R_j$ the location of the $j$'th record with $j=1$ labeling
1046: the first record ($R_1 = 1$ by convention), $R_2 > R_1$ the second record,
1047: and so on.  Note that 
1048: there are two sources of randomness in the problem -- one arising from the
1049: record locations $R_j$ and the other due to record 
1050: values $F(R_j)$. For exponentially distributed fitness, it is known that 
1051: the differences between 
1052: successive record values are independent and exponentially distributed 
1053: random variables \cite{Tata69}. Thus, the fitness of two  
1054: successive records differs by unity on average. 
1055: These considerations allow us to 
1056: eliminate the randomness associated with the record values by replacing
1057: the i.i.d. model of (\ref{shell}}) with exponentially distributed fitness by 
1058: a simpler model for which the population evolves as 
1059: \be
1060: \tilde E(j,t)=-R_{j}+ j t.
1061: \label{mfmodel}
1062: \ee
1063: Like in the original i.i.d. model, we find numerically that the 
1064: average number $\tilde {\cal{J}}$ of jumps grows logarithmically with $N$
1065: with a prefactor $\tilde \beta \approx 0.63$ (see Fig.~\ref{mftjump}). 
1066: This is distinctly different from the value $\beta \approx 1/2$ found for 
1067: the i.i.d. model with exponential fitness distribution, indicating that the 
1068: randomness in the record values is relevant. Somewhat surprisingly, in 
1069: contrast to the conjectured values of $\beta$ for the full i.i.d. problem 
1070: given in (\ref{beta1d}), $\tilde \beta$ does not 
1071: seem to be a simple rational number.
1072: 
1073: Our primary motivation for introducing this simplified model is to gain further
1074: insight into the mathematical structure of bypassing. 
1075: For the model defined by (\ref{mfmodel}), 
1076: the crossing time $\tilde T(j,j^{\prime})$ at which the line associated 
1077: with the $j$'th record overtakes that associated with record $j'$ is given by
1078: \be
1079: \label{mftimes} 
1080: \tilde T(j,j')= \frac{R_{j}-R_{j'}}{j-j'}.  
1081: \ee
1082: Then the probability $\tilde \beta_{2}$ that the second record is not
1083: bypassed can be written as 
1084: \bea
1085: \tilde \beta_{2} &=& 
1086: \mbox{Prob} \left[ R_{2}-1= \mbox{min} \left( R_{2}-1, \frac{R_{3}-1}{2}, 
1087: \frac{R_{4}-1}{3},...\right) \right].  
1088: \eea
1089: The evaluation of the 
1090: condition on the right hand using the joint 
1091: probability distribution for the record 
1092: times \cite{Nevzorov87,Arnold98,Nevzorov01}
1093: \be
1094: \label{jointRn}
1095: \mbox{Prob}(R_2,R_3,...,R_n) = 
1096: \frac{1}{(R_2 - 1)(R_3 - 1)(R_4 - 1)...(R_n - 1)R_n}
1097: \ee
1098: is clearly a difficult task. An upper bound on $\tilde \beta_2$ is obtained by 
1099: requiring that the record at $R_2$ is not bypassed by the one at $R_3$, i.e.
1100: that $\tilde T(2,1) < \tilde T(3,1)$. This gives  
1101: \be
1102: \label{b2bound}
1103: \tilde \beta_2 \leq 
1104: \mbox{Prob}(R_{3} > 2 R_{2} - 1)= \sum_{R_{2}=2}^{\infty} 
1105: \frac{1}{R_{2}-1} \; 
1106: \frac{1}{2 R_{2}-1} = 2 \; (1- \ln 2) \approx 0.613706 \;\;.
1107: \ee
1108: In our simulations on a large system, we find 
1109: $\tilde \beta_{2} \approx 0.600786$, showing that bypassing of $R_2$
1110: by the records beyond $R_3$ is rather unimportant. 
1111: 
1112: Consider next the behavior of the crossing times (\ref{mftimes}) when $j$ and
1113: $j'$ are large. Williams \cite{Williams73} has shown that the sequence of 
1114: record times
1115: can be generated from the recursion relation 
1116: \cite{Glick78,Nevzorov87,Arnold98,Nevzorov01}  
1117: \be
1118: \label{Williams}
1119: R_{j+1} =  [ e^{X_j} R_{j}] + 1,
1120: \ee
1121: where the $X_j$ are independent, exponentially distributed random variables 
1122: with mean one,
1123: and $[ a ]$ is the integer part of $a$. For large $j$, the integer constraint 
1124: can be ignored, and hence 
1125: \be
1126: \label{Tijlarge}
1127: \tilde T(j,j') \approx \frac{R_{j'}}{j - j'} 
1128: \left[ \exp \left(\sum_{i=j'+1}^j X_i \right) - 1
1129: \right].
1130: \ee
1131: Recalling that 
1132: the choice of the next non-bypassed record involves finding the minimum among
1133: all crossing times $\tilde T(j,j')$ with $j > j'$, we see that that the current
1134: location $R_{j'}$ cancels in the comparison between two such crossing times.
1135: The problem thus acquires translation invariance in the 
1136: record space, in the sense that the position of the next non-bypassed
1137: record depends only on $j - j'$ and on the random
1138: variables $X_i$ associated with the records between $j'$ and $j$. It is 
1139: therefore plausible
1140: that the bypassing probability tends to a constant for large $j$, and one is 
1141: tempted
1142: to describe the process by a Markov chain on the
1143: set of records with the transition probability 
1144: \be
1145: \label{Pjk}
1146: P_{j',j} = \mbox{Prob}[ \tilde T(j,j') = \min_{n > j'} \tilde T(n,j')].
1147: \ee     
1148: This is the conditional probability that the next jump occurs at $j$, given 
1149: that
1150: the preceding jump was at $j'$, averaged over all realizations of record times.
1151: Using the representation (\ref{Tijlarge}), the $P_{j,j'}$ are manifestly 
1152: translationally
1153: invariant for large $j,j'$, depending only on $j-j'$.
1154: Even in the asymptotic limit in which the expression (\ref{Tijlarge}) can be 
1155: used,
1156: the evaluation of (\ref{Pjk}) is cumbersome, but an analytic 
1157: upper bound on $ P_{j,j+1}$ can be obtained along the lines of 
1158: (\ref{b2bound}).  
1159: In the limit of large $j$, we can write
1160: \be
1161: \label{Pjj+1bound}
1162: P_{j,j+1} \leq \mbox{Prob}[\tilde T(j+1,j) < \tilde T(j+2,j)] 
1163: = \mbox{Prob}[e^{X_{j+1}} + e^{-X_j} > 2] = \ln 2,
1164: \ee
1165: where $X_j$ and $X_{j+1}$ are the independent exponential random variables 
1166: used in the 
1167: representation (\ref{Williams}). The numerical evalution of (\ref{Pjk}) yields
1168: $P_{j,j+1} \approx 0.669$, $P_{j,j+2} \approx 0.225$ and 
1169: $P_{j,j+3} \approx 0.075$,
1170: indicating a roughly exponential decay of the transition probability. 
1171: From the $P_{j',j}$, the mean density of jumps (non-bypassed records) can be 
1172: computed according to 
1173: \be
1174: \label{betamf2}
1175: \tilde \beta_{\mathrm{Markov}} = 
1176: \left( \sum_{n=1}^{\infty} n P_{j,j+n} \right)^{-1} \approx 0.676,
1177: \ee
1178: which is significantly larger than the direct numerical estimate 
1179: $\tilde \beta \approx 0.63$
1180: (see Fig.~\ref{mftjump}). This shows that the transition probability 
1181: (\ref{Pjk}) is 
1182: \textit{not} an exact representation of the process. The reason is that 
1183: (\ref{Pjk})
1184: is an \textit{annealed} average, whereas in the full problem the record times
1185: (or, equivalently, the exponential random variables in (\ref{Williams}))
1186: must be 
1187: treated as \textit{quenched}: Minimizing the crossing times $\tilde T(j,j')$
1188: for a given $j'$ involves, in principle, \textit{all} $j > j'$, and 
1189: the \textit{same} set of 
1190: random variables is used every time this minimization is repeated for 
1191: different $j'$.
1192: 
1193: We thus have to conclude that the range of attainable analytic results, even 
1194: for the 
1195: simplified problem (\ref{mfmodel}), is very limited. An extension of the bound
1196: (\ref{b2bound}) to the full i.i.d. model should be feasible using the 
1197: representation
1198: of the joint distribution of record times and record values through a Markov  
1199: chain \cite{Arnold98,Nevzorov01}; however, as such a bound is unlikely to 
1200: provide much insight
1201: into the conjectured relation (\ref{relation}), we have not pursued this 
1202: approach.
1203: 
1204: %=============================================================================
1205: %CONCLUSION
1206: %=============================================================================
1207: \section{Conclusions}
1208: \l{conclude}
1209: 
1210: In this article, we characterised the evolutionary trajectories 
1211: traced out by a quasispecies population in an uncorrelated rugged 
1212: fitness landscape. These trajectories approach the global fitness maximum 
1213: through a sequence of \textit {jumps} 
1214: which mark a change in the identity of the most populated genotype. 
1215: The statistics of these evolutionary jumps 
1216: was studied mainly numerically. However, useful insights were provided by 
1217: a study of \textit{record} statistics which could be handled analytically. 
1218: It was found that the jump statistics are qualitatively similar to records, 
1219: but there are quantitative differences because, as shown in 
1220: Fig.~\ref{lines},   
1221: a record breaking genotype can be \textit{bypassed} 
1222: by a superior one before it can acquire dominance 
1223: in the population (i.e. qualify to be a jump). 
1224: 
1225: The statistics of records and jumps depends strongly on the 
1226: geometry of the space of genotypes. The natural setting for genotype 
1227: evolution is the Hamming space of sequences of fixed length $N$. However, 
1228: computational effort could be greatly reduced by lumping together 
1229: the sequences within a shell of constant Hamming distance with respect to 
1230: the initial population \cite{Krug03}. Complementary to this shell model, we 
1231: also considered
1232: a model of i.i.d. shell fitnesses, which corresponds effectively to a 
1233: one-dimensional sequence space. While for the i.i.d. model, the average 
1234: number of jumps differs from the number of records only through the 
1235: prefactor $\beta$ of the logarithm of $N$, for the shell model
1236: the ratio of the two numbers 
1237: ${\cal{J}}_{\mathrm{shell}}/{\cal{R}}_{\mathrm{shell}} \rightarrow 0$ as 
1238: $N \rightarrow \infty$. 
1239: For fat-tailed fitness distributions the evolutionary
1240: trajectories in the shell model may even degenerate, in the sense that the 
1241: global fitness maximum is reached in a single step. For distributions
1242: decaying faster than a power law, like the exponential and normal 
1243: distributions,
1244: we find numerically that ${\cal{J}}_{\mathrm{shell}} \sim \sqrt{N}$; 
1245: an analytic understanding of this result would be very desirable. 
1246: 
1247: A universal feature of the
1248: evolutionary trajectories, which is independent of the geometry of genotype
1249: space as well as of the fitness distribution, is the hierarchy of power 
1250: law tails
1251: for the distributions of the times at which the jumps occur. In particular, 
1252: the $T^{-2}$-tail for the total evolution time implies that the average
1253: of $T$ is infinite. The dependence of the 
1254: \textit{typical} evolution time on the size of the sequence space
1255: can be characterized by a dynamic exponent $z$, which was obtained 
1256: exactly in terms of the extremal properties of the fitness distribution.
1257: On a mathematical level, perhaps the most intriguing result of this work
1258: is the conjectured relation (\ref{relation}) between  $\beta$ and $z$ for the 
1259: i.i.d. model. It would be extremely interesting to understand how such a 
1260: simple relation arises from the properties of the matrix of crossing 
1261: times (\ref{Tkkprime}); however,
1262: in view of the difficulties encountered even in the analysis of the simplified 
1263: problem (\ref{mfmodel}), we do not see at present how further progress in this
1264: direction can be achieved.
1265: 
1266: In this article, we worked in the limit of infinite population and strong 
1267: selection. However, we expect our results to 
1268: hold, at least qualitatively, for finite selective temperature as well. The
1269: jumps appear instantaneous in the strong selection limit;
1270: at finite selective temperature 
1271: they take a finite amount of time in which the peak
1272: associated with the new leader catches up with the currently dominating
1273: peak and the population distribution briefly becomes bimodal 
1274: \cite{Rujan88,Ebeling84,Krug93}.   
1275: Although the quasispecies theory, which works in the infinite population 
1276: limit, is inadequate 
1277: to address the fluctuation effects that become important
1278: when small mutant populations cross a fitness valley \cite{vanNimwegen00},   
1279: for the related problem of episodic behavior in evolutionary
1280: computation \cite{vanNimwegen97}, some of the finite
1281: population behavior has been understood on the basis of the infinite population
1282: limit. Thus we expect our investigation to give some insight into the
1283: behavior of finite populations in rugged fitness landscapes 
1284: \cite{Campos02a,Campos02b} as well. 
1285: 
1286: We close with some remarks on the applicability of the present work to the 
1287: evolution of
1288: biological populations. A realization of asexual mutation-selection dynamics 
1289: in a 
1290: static fitness landscape that is believed to be quite rugged is provided by
1291: the long-term experiments on populations of \textit{Escherichia coli} carried
1292: out by Lenski and collaborators \cite{Lenski94,Elena96,Elena03}. These 
1293: experiments
1294: show evidence for punctuated behavior both in the fitness and in the 
1295: morphological
1296: features (such as cell size) of the evolving populations, which is 
1297: attributed to the
1298: emergence and fixation of beneficial mutations. Fixation implies that a 
1299: mutation which
1300: is initially present only in a single individual is inherited by a growing 
1301: fraction
1302: of the population and eventually acquires dominance. 
1303: 
1304: At least on a qualitative level,
1305: this process corresponds in our model to that in which a subpopulation 
1306: residing in a distant shell
1307: of sequence space takes over the leadership of the population, 
1308: as described by (\ref{lineq}) and illustrated in Fig.~\ref{lines}. 
1309: The bypassing of one subpopulation by another is analogous to the phenomenon of
1310: \textit{clonal interference}, in which one beneficial mutation is superseded by
1311: another one before reaching fixation \cite{Gerrish98}. 
1312: The key difference between our model and the behavior of real asexual 
1313: evolution is that in the latter case beneficial mutants arise through the 
1314: stochastic search of a finite
1315: population in an immensely large sequence space, while in the quasispecies 
1316: model \textit{all possible mutants are present} 
1317: (in extremely small numbers) \textit{after the first time
1318: step}. It is, therefore, mandatory to include the 
1319: stochastic finite population dynamics
1320: into the model. Nevertheless, some features of the competition between 
1321: the mutant populations (however they may have arisen) could 
1322: well survive in a more complete, realistic treatment.   
1323: 
1324: \section*{Acknowledgments} 
1325: 
1326: We are grateful to A. Engel, L. Peliti, P. Eichelsbacher, W. Kirsch, 
1327: T. Kriecherbauer and N. Kumar for useful discussions. This work was 
1328: supported by DFG within 
1329: SFB/TR 12 \textit{Symmetries and Universality in Mesoscopic Systems.} 
1330: 
1331: 
1332: %=============================================================================
1333: %APPENDIX
1334: %=============================================================================
1335: \appendix
1336: 
1337: \section{Inter-record spacing for the shell model}
1338: \l{appendix1}
1339: Using the Stirling's formula $r ! \approx \sqrt{ 2 \pi r} (r/e)^{r}$ in  
1340: the binomial coefficient ${N \choose r}$ for $r, N \gg 1$ with $r/N$ fixed and 
1341: expanding ${N \choose r}$ about its maximum at $r=N/2$, we have 
1342: \be
1343: \frac{1}{2^{N}} \; {N \choose r} \approx \sqrt{\frac{2}{\pi N}} \; 
1344: \mbox{exp} \left[ {-\frac{(N/2-r)}{N/2}}^{2} \right] \;\;. \l{cnr}
1345: \ee
1346: This expression can be used to show that
1347: \bea
1348: \frac{1}{2^{N}} \; \sum_{r=0}^{y}{N \choose r} & \approx & 
1349: \frac{1}{2} \mbox{erfc}(\alpha) \l{sumcnr} \\
1350: \frac{1}{2^{N}} \; \sum_{r=0}^{y} r {N \choose r} & \approx & \frac{N}{4} 
1351: \left( \mbox{erfc}( \alpha) - \sqrt{\frac{2}{\pi N}} \mbox{exp} 
1352: (- \alpha^{2}) \right) \;\;,\l{avgcnr}
1353: \eea
1354: by approximating the sum on the left hand side by an integral. In 
1355: the above expressions, 
1356: $\alpha = (N/2-y)/\sqrt{N/2}$ and 
1357: $\mbox{erfc(x)} = (2/\sqrt{\pi}) \int_{x}^{\infty} dt \; e^{-t^{2}}$ is 
1358: the complementary error function. One can derive (\ref{rj}) for the 
1359: average location $\av{r_{j}}=\sum r_{j} \tilde{P}(r_{j})$ where 
1360: $\tilde{P}(r_{j})$ is given by (\ref{recloc}) using 
1361: Eqs.(\ref{cnr})-(\ref{avgcnr}) and replacing the sums by integrals. 
1362: 
1363: The integral $G(m)$ in Section~\ref{recordspace} can be estimated by saddle 
1364: point method as follows. We have 
1365: \be
1366: G(m) = (-1)^{m} \int_{-\infty}^{\infty} dx \; 
1367: \frac{e^{-2 x^{2}}}{\mbox{erfc}(x)} \; 
1368: \frac{\left(\ln {\mbox{erfc(x)}}\right)^{m}- \left( \ln {2} \right)^{m}}{m!}
1369: \approx  (-1)^{m} 
1370: \int_{-\infty}^{\infty} dx \; \frac{e^{-{\cal{G}}(x)}}{m!} \;\;,
1371: \ee
1372: where ${\cal{G}}(x) \approx x^{2} - \ln x \; (2 m + 1+ x^{-2})$ for 
1373: large $m$. Here we have used 
1374: that $\mbox{erfc}(x) \approx e^{-x^{2}}/ (\sqrt{\pi} x) $ for $x \gg 1$. 
1375: Expanding ${\cal{G}}(x)$ about its minimum $x_{0} \approx m- \ln \sqrt{m}$ 
1376: up to $O(x-x_{0})^{2}$ and doing the Gaussian 
1377: integral, we obtain $G(m) \approx \sqrt{ \pi m/2}$ for large $m$. 
1378: 
1379: %=============================================================================
1380: \section{Distribution of the most populated sequence}
1381: \l{appendix2}
1382: 
1383: Our goal is to compute the probability $P_t(k^\ast)$ that the 
1384: sequence $k^{\ast}$ has 
1385: the maximum population at time $t$ in the i.i.d. model. 
1386: This distribution is given by 
1387: \be  
1388: P_t(k^\ast) = \int_{E_{\mathrm{min}}(k^\ast,t)}^{E_{\mathrm{max}}(k^\ast,t)} 
1389: dE \;\; p_t^{(k^\ast)}(E)  \;\; 
1390: \prod_{k \neq k^{\ast}} 
1391: q_t^{(k)}(E) \;\;,
1392: \ee 
1393: where 
1394: \be
1395: p_t^{(k)}(E) = t^{-1} p[(E+k)/t] 
1396: \ee
1397: is the distribution of $E(k,t)$ obtained from the fitness distribution
1398: $p(F)$ via the variable transformation (\ref{lineq}). The 
1399: limits of the support of $p_t^{(k)}(E)$ are 
1400: $E_{\mathrm{min}}(k,t) = -k+ F_{\mathrm{min}} t$
1401: and $E_{\mathrm{max}}(k,t) = -k+ F_{\mathrm{max}} t $, and 
1402: $q_t^{(k)}(E)= \int_{E_{\mathrm{min}}}^{E} dE \; p_t^{(k)}(E) $
1403: is the corresponding cumulative distribution for $E > E_{\mathrm{min}}$ 
1404: and zero otherwise. In the following, we show 
1405: that the distribution $P_t(k^\ast)$ is of the scaling form 
1406: $P_t(k^\ast) \approx t^{-1/z} \; \Phi (k^{\ast}/{t^{1/z}})$ 
1407: for various choices of fitness distribution. 
1408: 
1409: (i) $\;$ \underline{$p(F)=e^{-F}$:} 
1410: \bea
1411: P_t(k^{\ast}) = \frac{1}{t} 
1412: \int_{0}^{\infty} dx \; e^{-(x+k^{\ast})/t} 
1413: \prod_{k \neq k^{\ast}} (1-e^{-(x+k)/t}) \approx \frac{1}{t} e^{-k^{\ast}/t}  \;\;,
1414: \eea
1415: where the last expression is obtained by exponentiating the product and 
1416: evaluating the resulting sum as an integral for an infinite system.  
1417: Thus in this case $z = 1$ and the scaling function is given by 
1418: $\Phi_1(y) = e^{-y}$.
1419: 
1420: (ii) $\;$ \underline{$p(F)=(1+\nu) (1-F)^{\nu}$, $\nu \geq -1$:}
1421: \bea
1422: P_t(k^\ast) &=& \frac{1+\nu}{t} \int_{0}^{t-k^{\ast}} dx \; 
1423: \left( 1- \frac{x+k^{\ast}}{t} \right)^{\nu} 
1424: \prod_{k \neq k^{\ast}} 
1425: \left[ 1-\left( 1- \frac{x+k}{t} \right)^{1+\nu} \right] \\
1426:  & \approx &  \frac{1+\nu}{t} \int_{0}^{t-k^{\ast}} dx \; 
1427: \left( 1- \frac{x+k^{\ast}}{t} \right)^{\nu} 
1428: \mbox{exp} \left[ - \frac{t}{2+\nu} 
1429: \left( 1- \frac{x}{t} \right)^{2+\nu} \right] \\
1430: & \approx & \frac{1}{t^{1/z}} \; \Phi_{2} \left( y= \frac{k^{\ast}}{t^{1/z}} 
1431: \right) \;\;,
1432: \eea
1433: where $z=(2+\nu)/(1+\nu)$ and the scaling function $\Phi_{2}(y)$ is given by
1434: \be
1435: \Phi_{2}(y) \approx  \frac{1+\nu}{(2+\nu)^{1/z}} 
1436: \int_{y^{2+\nu}/(2+\nu)}^{\infty} dx \; 
1437: e^{-x} x^{-1/z} \left( y+ ((2+\nu) x)^{1/(2+\nu)} \right)^{\nu} \;\;.
1438: \ee
1439: 
1440: 
1441: (iii) $\;$ \underline{$p(F)= \delta \; F^{-1-\delta}$, $\delta > 1$:}
1442: \bea
1443: P_t(k^\ast) &=& \frac{\delta}{t} \int_{t}^{\infty} dx \; 
1444: \left( \frac{t}{x+k^{\ast}} \right)^{1+\delta} 
1445: \prod_{k \neq k^{\ast}} \left[ 1- \left( \frac{t}{x+k} \right)^\delta 
1446: \right] \\
1447: & \approx & \frac{1}{t^{1/z}} \; \Phi_{3} \left( y= \frac{k^{\ast}}{t^{1/z}} 
1448: \right) \;\;,
1449: \eea
1450: where $z=(\delta-1)/\delta$ and the scaling function $\Phi_{3}(y)$ 
1451: is given by 
1452: \be
1453: \l{f3}
1454: \Phi_{3}(y) \approx \delta (\delta -1)^{1/(\delta -1)}
1455: \int_{0}^{\infty} dx 
1456: \frac{e^{-x} x^{1/(\delta-1)}}{(1+((\delta-1) x)^{1/(\delta-1)}y)^{1+\delta}} 
1457: \;\;.
1458: \ee
1459: 
1460: %=============================================================================
1461: %BIBLIOGRAPHY
1462: %=============================================================================
1463: 
1464: \begin{thebibliography}{999}
1465: 
1466: \bibitem{Sneppen95} K. Sneppen, P. Bak, H. Flyvbjerg and M. H. Jensen,
1467: Proc. Natl. Acad. Sci. USA \textbf{92}, 5209 (1995).
1468: 
1469: \bibitem{Sibani98} P. Sibani, M. Brandt and P. Alstr\o m, Int. J. Mod. 
1470: Phys. {\bf 12}, 361 (1998).
1471:  
1472: \bibitem{Peliti97} L. Peliti, {\tt{cond-mat/9712027}}.
1473: 
1474: \bibitem{Baake99} E. Baake and W. Gabriel,  
1475: in {\it Annual Reviews of Computational Physics VII},
1476: ed. by  D. Stauffer (World Scientific, Singapore, 2000), p. 203.
1477: 
1478: \bibitem{Drossel01} B. Drossel, Adv. Phys. \textbf{50}, 209 (2001).
1479: 
1480: \bibitem{Eldredge89} N. Eldredge, \textit{Macroevolutionary Dynamics} 
1481: (McGraw-Hill, New York 1989).
1482: 
1483: \bibitem{Gould93} S.J. Gould and N. Eldredge, Nature {\bf 366}, 223 (1993).
1484: 
1485: 
1486: \bibitem{Lenski94} R.E. Lenski and M. Travisano, 
1487: Proc. Natl. Acad. USA \textbf{91}, 6808 (1994).
1488: 
1489: \bibitem{Elena96} S.F. Elena, V.S. Cooper, and R.E. Lenski, Science
1490: \textbf{272}, 1802 (1996).
1491: 
1492: \bibitem{Burch99} C.L. Burch and L. Chao, Genetics \textbf{151}, 921 (1999).
1493: 
1494: \bibitem{Elena03} S.F. Elena and R.E. Lenski, 
1495: Nature Reviews Genetics \textbf{4}, 457 (2003).
1496: 
1497: \bibitem{Schuster02} P. Schuster, in \textit{Biological Evolution and 
1498: Statistical Physics}, eds. M. L\"assig and A. Valleriani 
1499: (Springer, Berlin, 2002), p.55.
1500: 
1501: \bibitem{Rujan88} P. Ruj\'an, Z. Phys. B \textbf{73}, 391 (1988). 
1502: 
1503: \bibitem{vanNimwegen97} E. van Nimwegen, J.P. Crutchfield, and M. Mitchell,
1504: Phys. Lett. A \textbf{229}, 144 (1997); 
1505: Theor. Comp. Sci. \textbf{229}, 41 (1999). 
1506: 
1507: \bibitem{Adami95} C. Adami, Phys. Lett. A \textbf{203}, 29 (1995).
1508: 
1509: \bibitem{Newman85} C.M. Newman, J.E. Cohen and C. Kipnis, Nature
1510: {\bf 315}, 400 (1985).
1511: 
1512: \bibitem{Lande85} R. Lande, 
1513: Proc. Natl. Acad. Sci. USA \textbf{82}, 7641 (1985).
1514: 
1515: \bibitem{Simpson44} G.G. Simpson, \textit{Tempo and Mode in Evolution} 
1516: (Columbia University Press, New York 1944).
1517: 
1518: \bibitem{Ebeling84} W. Ebeling, A. Engel, B. Esser and R. Feistel,
1519: J. Stat. Phys. {\bf 37}, 369 (1984). 
1520: 
1521: \bibitem{Zhang86} Y. C. Zhang, Phys. Rev. Lett. {\bf 56}, 2113 (1986).
1522: 
1523: \bibitem{Eigen71} M. Eigen, Naturwissenschaften {\bf 58}, 465 (1971). 
1524: 
1525: \bibitem{Eigen89}
1526: M. Eigen, J. McCaskill, and P. Schuster, Adv. Chem. Phys. {\bf 75}, 149 (1989).
1527: 
1528: \bibitem{Krug02} J. Krug in \textit{Biological Evolution and 
1529: Statistical Physics}, 
1530: eds. M. L\"assig and A. Valleriani (Springer, Berlin, 2002), p. 205.
1531: 
1532: \bibitem{Woodcock96} G. Woodcock and P.G. Higgs, 
1533: J. theor. Biol. \textbf{179}, 61 (1996).
1534: 
1535: \bibitem{Krug03} J. Krug and C. Karl, Physica A {\bf 318}, 137 (2003).
1536: 
1537: \bibitem{Glick78} N. Glick, Amer. Math. Monthly {\bf 85}, 2 (1978). 
1538: 
1539: \bibitem{Nevzorov87} V.B. Nevzorov, Theory Probab. Appl., 
1540: {\bf 32}, 201 (1987).   
1541: 
1542: \bibitem{Arnold98} B.C. Arnold, N. Balakrishnan and 
1543: H.N. Nagaraja, \textit{Records}
1544: (Wiley, New York 1998).
1545: 
1546: \bibitem{Nevzorov01} V.B. Nevzorov, 
1547: \textit{Records: Mathematical Theory} (American Mathematical
1548: Society, Providence, 2001).
1549: 
1550: \bibitem{Kauffman87} S.A. Kauffman, S. Levin, J. theor. Biol. 
1551: \textbf{128}, 11 (1987).
1552: 
1553: \bibitem{Krug04} J. Krug and K. Jain in {\it Proceedings of 8th 
1554: ICCMSP/Marrakech}, to be published in Physica A, {{\tt q-bio.PE/0409019}}. 
1555: 
1556: \bibitem{Franz97} S. Franz and L. Peliti, J. Phys. A \textbf{30}, 4481 (1997).
1557: 
1558: \bibitem{Krug93} J. Krug and T. Halpin-Healy, J. Phys. I 
1559: France \textbf{3}, 2179 (1993).
1560: 
1561: \bibitem{Gaertner04} J. G\"artner and W. K\"onig, in 
1562: J.-D. Deuschel and A. Greven (Eds.), \textit{Interacting Stochastic Systems} 
1563: (Springer, Berlin 2005) p 153.
1564: 
1565: \bibitem{Nevzorov84} V. B. Nevzorov, Theory Probab. Appl., 
1566: {\bf 29}, 845 (1984). 
1567: 
1568: \bibitem{Yang75} M. C. K. Yang, J. Appl. Prob. {\bf 12}, 148 (1975).
1569: 
1570: \bibitem{David70} H.A. David, \textit{Order Statistics} 
1571: (Wiley, New York, 1970).
1572: 
1573: \bibitem{Sornette00} D. Sornette, 
1574: \textit{Critical Phenomena in Natural Sciences} (Springer, Berlin 2000).
1575: 
1576: \bibitem{remark} This scaling form is not true for 
1577: power law distributed 
1578: fitness with $1 < \delta \leq 2$, because the scaling function $\Phi$ in
1579: (\ref{Ptscale}) does not possess a first moment in this case (see (\ref{f3})). 
1580: Nevertheless, the time scale 
1581: $T^{\ast} \sim N^{z}$ with $z$ given in (\ref{1dz}) for all $\delta > 1$. 
1582: 
1583: \bibitem{Tata69} M. N. Tata, Z. Wahrsch. Verw. Geb. {\bf 12}, 9 (1969). 
1584: 
1585: 
1586: \bibitem{Williams73} D. Williams, Bull. London Math. Soc. {\bf 5}, 235 (1973). 
1587: 
1588: \bibitem{vanNimwegen00} E. van Nimwegen and J.P. Crutchfield, 
1589: Bull. Math. Biol. \textbf{62}, 799 (2000).
1590: 
1591: \bibitem{Campos02a} P.R.A. Campos, C. Adami and C.O. Wilke, 
1592: Physica A \textbf{304}, 495 (2002).
1593: 
1594: \bibitem{Campos02b} P.R.A. Campos and V.M. de Oliveira, 
1595: Int. J. Mod. Phys. C \textbf{13}, 1003 (2002).
1596: 
1597: 
1598: \bibitem{Gerrish98} P.J. Gerrish and R.E. Lenski, 
1599: Genetica \textbf{102/103}, 127 (1998).
1600: 
1601: \end{thebibliography}
1602: 
1603: %=============================================================================
1604: %FIGURES
1605: %=============================================================================
1606: \begin{figure}
1607: \begin{center}
1608: \psfig{figure=lines.ps,width=9cm,angle=270}
1609: \caption{
1610: Time evolution of the logarithmic population variable $E(k,t) = - k +
1611: F(k)t$. The fitnesses $F(k)$ of the sequences corresponding to thin green
1612: lines are not records whereas those corresponding to bold red lines
1613: are. The lines appearing in the upper envelope define the most populated
1614: sequence $k^\ast$; an evolutionary jump occurs when two such lines cross.
1615: Note that the sequences $k=2$ and $k=6$, although constituting fitness
1616: records, are bypassed as they do not satisfy the minimum crossing time
1617: constraint. The inset shows the punctuated evolution of the most
1618: populated sequence $k^\ast$ as a function of time.}
1619: \label{lines}
1620: \end{center}
1621: \end{figure}
1622: 
1623: \begin{figure}
1624: \begin{center}
1625: \psfig{figure=iidjump.ps,width=9cm,angle=270}
1626: \caption{Mean number ${\cal{J}}_{\mathrm{iid}}$ of jumps for 
1627: the i.i.d. model for 
1628: various fitness distributions. The lines are the best fits to the functional 
1629: form ${\cal{J}}_{\mathrm{iid}}= \beta \ln N + \mbox{constant}$.}
1630: \label{iidjump}
1631: \end{center}
1632: \end{figure}
1633: 
1634: 
1635: \begin{figure}
1636: \begin{center}
1637: \psfig{figure=sjump.ps,width=9cm,angle=270}
1638: \caption{Data collapse for the scaled jump distribution 
1639: $\sqrt{N} P_{\mathrm{shell}}(k,N)$ vs. $k/N$ for 
1640: the shell model with $\ell=2$ and $p(F)=e^{-F}$. Inset: 
1641: Scaled inter-jump spacing $\Delta_{\mathrm{shell}}(j)/\sqrt{N}$ vs. 
1642: $j/\sqrt{N}$ to show that jumps are roughly equally spaced in the shell model.}
1643: \label{sjump}
1644: \end{center}
1645: \end{figure}
1646: 
1647: \begin{figure}
1648: \begin{center}
1649: \psfig{figure=rspace.ps,width=9cm,angle=270}
1650: \caption{Semi-log plot for inter-jump spacing $\Delta_{\mathrm{iid}}(j)$ 
1651: for the i.i.d. model in accordance with (\ref{interjumpiid}) with 
1652: slope $b=\beta/2$. 
1653: Inset: Correlation $C_{\mathrm{iid}}(k_{1},k_{2})$ 
1654: vs. $k_{2}-k_{1}$ for two fixed values of $k_{1}$ 
1655: to show correlations between jumps in the i.i.d. model.}
1656: \label{rspace}
1657: \end{center}
1658: \end{figure}
1659: 
1660: \begin{figure}
1661: \begin{center}
1662: \psfig{figure=evol.ps,width=9cm,angle=270}
1663: \caption{Log-log plot for the ratios $P(T_{2})/P(T_{1})$ and 
1664: $P(T_{3})/P(T_{2})$ of evolution time distributions
1665: for the i.i.d. model with $p(F)=e^{-F}$. The ratios decay as 
1666: $1/T$ for large $T$, as the line with 
1667: slope $-1$ indicates.}
1668: \label{evol}
1669: \end{center}
1670: \end{figure}
1671: 
1672: \begin{figure}
1673: \begin{center}
1674: \psfig{figure=sum.ps,width=9cm,angle=270}
1675: \caption{Distribution 
1676: $P_{\mathrm{iid}}(X,N)=P_{\mathrm{iid}}(t=X,N)+P_{\mathrm{iid}}(k=X,N)$ vs. 
1677: $X$ for various distributions, illustrating the 
1678: duality between the jump processes along the $k$- and $t$-axes. The 
1679: solid line has slope $-1$. Inset: Semi-log plot for the inter-jump spacing 
1680: $\Delta_{\mathrm{iid}}^{\prime}(j)$ along the time axis as a function of 
1681: $j$ for the i.i.d. model. The slope $(1-\beta)/2$ is consistent with the 
1682: conjecture (\ref{cnjctre}).}
1683: \label{sum}
1684: \end{center}
1685: \end{figure}
1686: 
1687: \begin{figure}
1688: \begin{center}
1689: \psfig{figure=mftjump.ps,width=9cm,angle=270}
1690: \caption{Average number $\tilde {\cal{J}}$ of jumps 
1691: vs. $N$ for the record time model defined by (\ref{mfmodel}). The solid line 
1692: $0.63 \ln N-0.13$ is the best fit.}
1693: \label{mftjump}
1694: \end{center}
1695: \end{figure}
1696: 
1697: %=============================================================================
1698: %=============================================================================
1699: 
1700: \end{document}
1701: