1: %2345678901234567890123456789012345678901234567890123456789012345678901234567890
2: \documentclass[12pt]{iopart}
3: %\usepackage{lscape}
4: \eqnobysec
5: \usepackage{setstack}
6: \usepackage{iopams}
7: \usepackage{graphicx}% Include figure files
8: %\usepackage{bm}% bold math
9: %\usepackage{epsf}
10: %\pagestyle{plain}
11:
12: \begin{document}
13:
14: \title[Fluctuations and correlations in biological coevolution]{Fluctuations
15: and correlations in an individual-based model of biological coevolution}
16: \author{R K P Zia\dag\ddag \; and Per Arne Rikvold\P}
17: \address{\dag\ Center for Stochastic Processes in Science and Engineering,
18: Department
19: of Physics, Virginia Polytechnic Institute and State University, Blacksburg,
20: Virginia 24061-0435, USA}
21: \address{\ddag\ Fachbereit Physik, Universit\"{a}t
22: Duisburg-Essen, 45117 Essen, Germany}
23: \address{\P\ School of Computational Science
24: and Information Technology, Center for Materials Research and Technology,
25: and Department of Physics, Florida State University, Tallahassee, Florida
26: 32306-4120, USA}
27:
28: \ead{rkpzia@vt.edu, rikvold@csit.fsu.edu}
29:
30: \date{\today }
31:
32: \begin{abstract}
33: We extend our study of a simple model of biological coevolution to its
34: statistical properties. Staring with a complete description in terms of a
35: master equation, we provide its relation to the deterministic evolution
36: equations used in previous investigations. The stationary states of the
37: mutationless model are generally well approximated by Gaussian
38: distributions, so that the fluctuations and correlations of the populations
39: can be computed analytically. Several specific cases are studied by Monte
40: Carlo simulations, and there is excellent agreement between the data and the
41: theoretical predictions.
42: \end{abstract}
43:
44: \submitto{JPA}
45: \pacs{
46: 05.40.-a %Fluctuation phenomena, random processes, noise, and Brownian motion
47: 87.23.Kg %Dynamics of evolution (In 80: Interdisciplinary physics.)
48: 05.65.+b %Self-organized systems
49: }
50:
51: \maketitle
52:
53:
54:
55: \section{Introduction}
56: \label{sec:Int}
57:
58: The dynamics of populations and species in the context of biological or
59: ecological systems have attracted considerable attention in the community of
60: statistical physicists in the last decade. Though there has been much
61: progress \cite{DROS01,LASS02}, it is rare that the behavior of macroscopic
62: populations can be predicted, even for ``simple models,'' from a set of
63: {\em stochastic} rules for an {\em individual's} propensity to survive and/or
64: reproduce. The difficulties can be traced not only to the presence of many
65: degrees of freedom, with complex internal interactions (e.g., mutualistic or
66: predator-prey), but also to the non-trivial couplings to external reservoirs
67: (such as energy or food). As a result, even if we only focus on systems in
68: stationary states, we must recognize that these are {\em non-equilibrium
69: steady states}, so that the well-known methods of equilibrium thermodynamics
70: should not be blindly applied. At present, given the absence of a
71: universally applicable framework of non-equilibrium statistical mechanics,
72: progress is made through understanding ``one system at a time.'' Within this
73: context, we recently studied a model of coevolution \cite{RIKV03A,RIKZIA03},
74: based on the one (``tangled nature'') introduced by Hall, Christensen, and
75: collaborators \cite{HALL02,CHRI02,COLL03}. The motivations behind these
76: studies are varied. An early model for coevolution and speciation,
77: introduced by Bak and Sneppen \cite{BAK93}, consists of competing species,
78: according to a preassigned notion of ``fitness.'' Speciation arises by
79: imposing a crude version of ``mutation,'' i.e., letting the least fit
80: species (as well as some of their ``neighbors'') be replaced by new species
81: with different, randomly chosen fitness. Despite their simplicity, such models
82: appear to settle into a steady state which exhibits avalanches of
83: extinctions. Though they are hailed as showing a link between Darwin's
84: principle of ``survival of the fittest'' and Gould's notion of ``punctuated
85: equilibria" \cite{GOUL77,GOUL93,NEWM85}, these models are thought to be
86: too simplistic in at least two aspects. Firstly, mutation and selection act
87: on individuals of a populations, rather than on entire species at once.
88: Secondly, whether a species is ``fit'' is not a static notion, but rather
89: depends on what other species are present in the ecosystem.
90:
91: To address both issues, Hall, {\em et.al.} \cite{HALL02,CHRI02,COLL03}
92: recently introduced an individual-based model with a dynamically evolving
93: ``fitness landscape.'' The ``ecosystem'' consists of individuals, each
94: of which is
95: said to belong to a ``species'' identified by a ``genome''
96: represented by a string of bits. Mutations are built in through the random
97: flipping of bits, at a constant slow rate, in the genomes of newborn
98: individuals.
99: In earlier models of speciation, an individual's fitness, i.e., its
100: reproduction
101: probability, is purely a function of the bit string and fixed for all time.
102: Here, the reproduction probability
103: depends on the relative abundance of all other species,
104: so that ``fitness'' becomes a dynamic concept. The ``interspecies
105: interactions'' are, by contrast, comparatively static in nature and, for
106: simplicity, are introduced via fixed quantities associated with {\em pairs\/}
107: of genotypes. Thus, the model accounts only for whether (an individual of) a
108: species is beneficial or detrimental for another species, regardless of the
109: presence of any other species. Again for simplicity, all species reproduce
110: asexually, with identical fecundities. Simulations reveal several
111: interesting behaviours, including the presence of long-lived states separated
112: by bursts of high activity -- ``punctuated equilibria.'' Unfortunately, this
113: model proved too complex for analytic understanding, and only
114: phenomenological descriptions have been advanced to date.
115:
116: In an effort to gain some insight into how its remarkable properties arise,
117: we considered a variation of this model, with simpler interspecies
118: interactions, which enables us both to carry out much longer simulations and
119: to perform linear stability analysis~\cite{RIKV03A,RIKZIA03}. In addition to
120: observing a self-similar picture of intermittency or ``punctuated
121: equilibria'' over several decades of generations, we found that standard
122: measures of diversity display $1/f$ noise in their power spectral densities.
123: This property is quite consistent with another finding: The life-time
124: distribution of the long-lived states approximately
125: follows a inverse-square power law \cite{Krug03}.
126: More detailed investigations of the very long-lived states, which typically
127: consist of a community of a
128: handful of dominant species along with a ``cloud of mutants,''
129: reveal the reason behind their longevity. Essentially none of the closest
130: mutants of the main species in such a community are ``dangerous,'' in that
131: their interactions with the parent species inhibit their exponential growth.
132: In other words, the original community is (linearly) stable against invasion
133: by its most closely related mutants.
134: Only a small fraction of the next-closest mutants (with
135: genes differing from a dominant species by two bits) are dangerous, leading
136: to the eventual demise of the ``quasi-steady'' state. Despite the
137: discovery of these connections, full analytic understanding is still beyond
138: our grasp. In particular, since the theoretical analyses were based entirely
139: on a heuristic, deterministic (``mean-field'') equation of motion, it is
140: unlikely that they can account for the most intriguing behaviour stemming
141: from a {\em stochastic} dynamics.
142:
143: In the present paper, we address some of the issues associated with a fully
144: stochastic description. Starting from a master equation which governs the
145: evolution of all the details (``microscopics'') of the model, we first
146: demonstrate that the deterministic equation in \cite{RIKZIA03} emerges,
147: provided all correlations are ignored. However, it is no easy task to find a
148: quantitative understanding describing the quasi-steady states (QSS),
149: in which the system appears to be stationary but
150: actually has long, finite lifetimes.
151: One complication lies with the inherent metastability aspect of a QSS.
152: Another is that, due to mutations, the populations in a QSS consist of
153: {\em two} components: a handful of dominant species (with at least
154: several hundred
155: individuals, in the specific simulations we ran) and a larger number of
156: minor species (with much less than one hundred individuals).
157: Our approach to the solution is
158: a two-step process. First, for each QSS, we develop a full understanding of
159: a corresponding ``truly stationary state'' (TSS), i.e., one with only the
160: dominant species. Each TSS is associated with an ${\cal N}$-species fixed
161: point discussed in \cite{RIKZIA03} and can be accessed by setting
162: the mutation rate, $\mu $, to zero. The second step is to account for
163: $O\left(\mu \right) $ effects, by including a limited ``cloud of mutants.''
164: Needless
165: to say, a careful definition of such a community will be necessary before
166: any analytic progress is possible. Thus, we will only take the first step
167: here, deferring the more complex problem to a later publication. Beyond
168: that, our eventual goal is to predict the more fascinating phenomena, such
169: as power-law distribution of QSS lifetimes, ``punctuated equilibria'' (or
170: intermittency), and $1/f$ noise.
171:
172: In the next section, for completeness and the readers' convenience, we
173: briefly review the specifications of the model. The following section is
174: devoted to the master equation and the derivation of the deterministic
175: evolution equation used in \cite{RIKZIA03}. Fluctuations and
176: correlations in a TSS, as well as comparisons to simulations in specific
177: cases, are the focus of sections~\ref{sec:TSS} and~\ref{sec:dyn}.
178: Conclusions and an outlook can be found in section~\ref{sec:Concl}. The
179: Appendix is devoted to some of the technical details.
180:
181: \section{Model specification and algorithm}
182:
183: \label{sec:Mod}
184:
185: The model we considered \cite{RIKZIA03} is a simplified version of the one
186: introduced in \cite{HALL02,CHRI02,COLL03}. It consists of a population of
187: individuals, each of which is associated with a string $L$ of bits (0 or 1),
188: representing a ``genome'' of $L$ ``genes.'' The bit-strings, or genotypes,
189: are labeled by integers $I$ , which lie between $1$ and
190: ${\cal N}_{{\rm max}}=2^L$.
191: For simplicity, we will use the term ``species'' to distinguish
192: individuals with different bit-strings or genotypes. The population evolves
193: asexually in discrete time steps ($t=0,1,...$), which may be thought of as
194: ``years'' or ``generations.'' In our model, all individuals of a generation
195: die when those of the next generation are ``born'' (as in, e.g., aphids).
196: Thus, for any particular run (or ``history''), the system is fully specified
197: by the set of integers $n_I(t)$ ($I=1,...,{\cal N}_{{\rm max}}$)
198: representing the number of individuals of genotype $I$ in generation $t$. In
199: cases where the explicit index $I$ is not necessary, we will use the
200: ``vector'' notation:
201: \begin{equation}
202: \vec{n}(t)\equiv \{n_1(t),...,n_{{\cal N}_{{\rm max}}}(t)\}\,\,.
203: \label{eq:n^arrow}
204: \end{equation}
205:
206: To model competition and interspecies interactions, we let an individual die
207: with some non-vanishing probability {\em before} it reproduces, e.g., salmon
208: that die in the oceans. All survivors then give birth to $F$ offspring,
209: which constitute the next generation. Competition for resources (e.g.,
210: space, energy, food) is often introduced via a Verhulst \cite{VERH1838}
211: factor, which also plays the role of preventing unlimited growth and enters
212: typically via a ratio $N_{{\rm tot}}(t)/N_0$. Here, $N_0$ is a parameter
213: representing the ``carrying capacity'' of the ``ecosystem,'' and
214: \begin{equation}
215: N_{{\rm tot}}(t)\equiv \sum_In_I(t) \label{eq:N_totDef}
216: \end{equation}
217: is the total population at time $t$. With a ``healthy'' fecundity ($F$),
218: the population is unlikely to ``collapse'' ($N_{{\rm tot}}=0$). Instead,
219: $N_{{\rm tot}}(t)$ is rarely far from $N_0$. Interspecies interactions are
220: modeled by a matrix ${\bf M}$, the element $M_I^J$ being the effect of the
221: species $J$ on species $I$. For reasons provided in \cite{RIKZIA03}, we set
222: $M_I^I=0$ and choose random off-diagonal elements from a uniform distribution
223: over $[-1,1]$. In all our simulations, ${\bf M}$ is fixed at $t=0$ and does
224: {\em not} evolve in time. If both $M_I^J$ and $M_J^I$ are positive, the
225: two species are said to be ``mutualistic,'' and if they are of opposite
226: signs, we have a predator-prey relationship. Not surprisingly, populations
227: with both elements being negative are extremely unstable. Finally, the
228: probability of an individual of species $I$ to survive to reproduction is
229: specified by \cite{RIKZIA03}
230: \begin{equation}
231: P(I;\vec{n}(t))=\frac 1{1+\exp \left[ -M_I^Jn_J(t)/N_{{\rm tot}}(t)+
232: N_{{\rm tot}}(t)/N_0\right] }\;,
233: \label{eq:P}
234: \end{equation}
235: which is often shortened to just $P(I)$. Here, as in the rest of the paper,
236: we use the Einstein summation convention, i.e., a repeated index (e.g., $J$
237: in $M_I^Jn_J$) is summed over (from $1$ to ${\cal N}_{{\rm max}}$, in
238: general). As will be clear below, where we deal with systems with large
239: $N_0$ (``macroscopic,'' though not necessarily in the sense of typical
240: thermodynamic systems), quantities with a single subscripted index (e.g.,
241: $n_J$) are generally of order $N_0$; those with a single superscript, of
242: order $1/N_0$; those with both (e.g., $M_I^J$) or none (e.g., $P(I)$), of
243: $O\left( 1\right) $; etc. Exceptions are noted with a caret (or ``hat'').
244: For example, we denote the ``normalized'' covariance matrix for the
245: populations (i.e., $\left( \left\langle n_In_J\right\rangle -\left\langle
246: n_I\right\rangle \left\langle n_J\right\rangle \right) /N_0$) by
247: $\hat{G}_{IJ}$, which is a quantity of $O\left( 1\right) $ rather than
248: $O\left(N_0^2\right) $. Similarly, its inverse ($\hat{\Gamma}^{IJ}$) is
249: also of $O\left( 1\right) $ , as opposed to of $O\left( 1/N_0^2\right) $.
250: To avoid confusion, we will remind the readers of such exceptions at the
251: appropriate points.
252:
253: The last ingredient in our model is mutation. In the absence of mutations,
254: the diversity of the population never increases with time. Indeed, the only
255: {\em rigorously} stationary state is the collapsed one, $n_I = 0$ for all $I$.
256: Nevertheless,
257: non-trivial states (with extremely long life times and called TSS's here)
258: are typically reached, for all {\em relevant} time scales.
259: While the rigorously stationary state is independent of the intitial
260: condition $\vec{n}(0)$, the TSS
261: states are entirely dependent on the initial population,
262: and they display the dominant characteristics of the corresponding QSS's.
263: Identified by a fixed point (FP) of the deterministic evolution equation,
264: \begin{equation}
265: n_I(t+1)=n_I(t)FP(I;\vec{n}(t)\})\;, \label{eq:EOM-no mu}
266: \end{equation}
267: a TSS provides the basis for linear stability analysis and for future
268: investigations of the associated QSS.
269:
270: Returning to mutations, they are important, not only to promote diversity
271: and to model ``speciation,'' but also to provide the main ingredient for the
272: interesting phenomenon of, say, intermittency (``punctuated equilibria'')
273: \cite{RIKZIA03}. As in all bit-string models involving asexual reproduction,
274: we allow each offspring to carry the genes of its parent, except for a
275: probability of $\mu /L$ that each bit be changed. As a result, on the
276: average and for small $\mu $, a survivor produces $\mu F$ offspring with
277: different genetic material. To keep track of the ``biodiversity,'' we define
278: the species richness ${\cal N}(t)$ as the number of populated species at $t$
279: (i.e., only ${\cal N}$ $I$'s are present at $t$). Another common measure
280: which characterizes the relative abundance of the species better (but will
281: not be the focus here) is the Shannon-Wiener index, $\sum_J\rho (J)\ln \rho
282: (J)$, where
283: \begin{equation}
284: \rho (J)\equiv n_J(t)/N_{{\rm tot}}(t) \label{eq:rhoDef}
285: \end{equation}
286: is just the fraction of species $J$ in the system.
287:
288: Let us briefly summarize the algorithm used to simulate this model,
289: referring to \cite{RIKZIA03} for the details. There are three layers of
290: nested loops: (1) over generations $t$, (2) over ${\cal N}(t)$, and (3) over
291: $n_I(t)$. In the innermost (last) loop, each individual produces, with
292: probability $P(I)$, $F$ offspring, each of which is allowed to mutate before
293: $\vec{n}(t+1)$ is recorded. In most of our studies, we chose $L=13$,
294: $N_0=2000$, $F=4$, $\mu =10^{-3}$, and $N_{{\rm tot}}\left( 0\right) =100$,
295: with random initial $\vec{n}(0)$. Thus, ${\cal N}_{{\rm max}}=8192$, though
296: far fewer species are typically present in the system at any given time,
297: i.e., ${\cal N}\ll {\cal N}_{{\rm max}}$.
298:
299: \section{Master equation and mean-field theory}
300:
301: \label{sec:MEMFT}
302:
303: In our recent investigations, the Monte Carlo studies generate stochastic
304: sequences of configurations (i.e., non-negative integers, or just
305: ``points,'' in the ${\cal N}_{{\rm max}}$-dimensional space) but the
306: theoretical analysis was based on deterministic, heuristic equations of
307: motion for the averages of the populations (``mean-field'' theory). To
308: understand the full stochastic process, we need a complete description,
309: involving the probability that the system is found with a specific number of
310: individuals at time $t$, namely, ${\cal P}\left( \vec{n},t\right) $. Its
311: evolution is governed by the master equation, given in an appendix of \cite
312: {RIKZIA03}. Before turning to this equation, let us emphasize the difference
313: between ${\cal P}\left( \vec{n},t\right) $ here and the $\vec{n}(t)$ above.
314: In the former, $\vec{n}$ is a co-ordinate (in ${\cal N}_{{\rm max}}$-space)
315: and ${\cal P}$ an evolving {\em function} in this space. By contrast,
316: $\vec{n}(t)$ is just the trajectory of a single point in this space. A
317: single Monte Carlo run generates a particular trajectory (or ``history''):
318: ${n}_I\left( t\right) $, and can be represented as a (Kronecker) delta
319: function jumping from point to point:
320: ${\cal P}_{\mathrm{a\;MC\;run}}\left( \vec{n},t \right) =
321: \prod_I\delta \left( n_I,{n}_I\left( t\right) \right)$.
322: The full dynamics of ${\cal P}\left( \vec{n},t\right) $ is simulated by
323: averaging over many runs, and thus difficult to access. When we turn out
324: attention to the TSS's below, the process simplifies, since they are
325: characterized by {\em static } distributions:
326: ${\cal P}^{*}\left( \vec{n}\right) $. Then, it is sufficient to perform a
327: single, long run during which the system rarely wanders far (say,
328: $O\left( \sqrt{N_0}\right) $) from the neighborhood of a specific point
329: (typically, fixed points of the mean-field evolution equations).
330:
331: Returning to the issue at hand, finding an equation for
332: ${\cal P}\left( \vec{n},t\right) $,
333: we recapitulate the probability for an individual of species $I$ to survive:
334: \begin{equation}
335: P\left( I\right) =\left\{ 1+\exp \left[ \frac{N_{{\rm tot}}}{N_0}
336: - M_I^J \frac{n_J}{N_{{\rm tot}}}\right] \right\} ^{-1} \; .
337: \label{eq:B1}
338: \end{equation}
339: Again, note the different interpretation we give for this expression versus
340: equation
341: (\ref{eq:P}). Here, $P\left( I\right) =P\left( I;\vec{n}\right) $ denote
342: ${\cal N}_{{\rm max}}$ functions defined in ${\cal N}_{{\rm max}}$-space,
343: independent of $t$. In contrast, for a particular MC run at a particular
344: time $t$, we need only ${\cal N}\left( t\right) $ functions (for a specific
345: point $\{n_J(t)\}$, in a much smaller, ${\cal N}$-space). Next, we must keep
346: track of {\em all} of the possible number of survivors. Since all of these
347: individuals reproduce, we will call them ``parents.'' Defining the symbol
348: ${n_I \atopwithdelims[] m_I}$
349: %${\QATOPD[ ] {n_I}{m_I}}$
350: as the rate for $m_I$ individuals to survive from
351: the original $n_I$, we simply write a binomial distribution:
352: \begin{equation}
353: {n_I \atopwithdelims[] m_I}
354: %{\QATOPD[ ] {n_I}{m_I}}
355: =\frac{n_I!}{m_I!\left( n_I-m_I\right) !}\left[
356: P\left( I\right) \right] ^{m_I}\left[ 1-P\left( I\right) \right] ^{n_I-m_I}.
357: \label{eq:B2}
358: \end{equation}
359: Next, each parent gives rise to $F$ offspring, not every one of which is of
360: the same species. In the simulations, it is possible to have a mutant
361: whose genome differs from the parent by two or more bits. However, this is
362: quite rare, being less than
363: $O\left( \mu ^2N_{{\rm tot}}\right) =O\left( \mu^2N_0\right) \sim 10^{-3}$.
364: Thus, we will keep our analysis simple by
365: restricting our analysis here to mutations which flip only a single bit.
366: Then, there can be only $L+1$ possible varieties of offspring for each
367: parent. To account for these, let us introduce the notation
368: \begin{eqnarray*}
369: &&b_{J,0}\quad \mathrm{for\;the\;number\;of\;offspring\;from\;parent}\;J\;
370: \mathrm{with\;no\;mutations} \\
371: &&b_{J,\alpha }\quad \mathrm{for\;the\;number\;of\;offspring\;from\;parent
372: \;}J \;\mathrm{with\;the\;}\alpha \mathrm{th\;bit\;flipped}
373: \end{eqnarray*}
374: and define the multinomial-like symbol
375: \begin{equation}
376: {Fm_J \atopwithdelims[] b_{J,0},b_{J,1},...,b_{J,L}}
377: %{\ {\QATOPD[ ] {Fm_J}{b_{J,0},b_{J,1},...,b_{J,L}}}}
378: =\frac{\left(Fm_J\right) !}{\left( b_{J,0}\right) !}
379: \left( 1-\mu \right)^{b_{J,0}}
380: \prod_{\alpha =1}^L\frac 1{\left( b_{J,\alpha }\right) !}
381: \left(\frac \mu L\right) ^{b_{J,\alpha }} \; .
382: \label{eq:B3}
383: \end{equation}
384: This is the probability that the $Fm_J$ offspring are distributed into the
385: specific set $\left\{ b_{J,0},b_{J,1},...,b_{J,L}\right\} $. The last
386: ingredient needed is the connection matrix
387: \begin{equation}
388: \Delta _K^{J,\alpha }=\left\{
389: \begin{array}{cc}
390: 1 & \mathrm{if\;genotype\;}K\;\mathrm{is\;}J\mathrm{\;with\;the\;}\alpha
391: \mathrm{th\;bit\;flipped} \\
392: 0 & \mathrm{otherwise}
393: \end{array}
394: \right. \label{eq:B4}
395: \end{equation}
396: so that the number of offspring born {\em into} species $K$ due to
397: mutations is
398: \begin{equation}
399: B_K\equiv \sum_{J,\alpha >0}\Delta _K^{J,\alpha }b_{J,\alpha }\;.
400: \label{eq:B5}
401: \end{equation}
402: With these ingredients, we arrive at the master equation
403: \footnote{
404: Many master equations are written for continuous time, in the form of
405: $\partial _t{\cal P}(C,t)=\sum_{C^{\prime }}L\left( C,C^{\prime }\right)
406: {\cal P}(C^{\prime },t)$. For discrete-time processes, we find it more
407: convenient to express the evolution in the form used here.}:
408: \begin{equation}
409: \fl
410: {\cal P}\left( \vec{n}^{\prime },t+1\right)
411: =\sum_{ \vec{n},\vec{m},\left\{b\right\} }
412: \prod_K
413: \delta \left( n_K^{\prime },b_{K,0}+B_K \right)
414: \prod_J
415: {Fm_J \atopwithdelims[] b_{J,0},b_{J,1},...,b_{J,L}}
416: %\ {\QATOPD[ ] {Fm_J}{b_{J,0},b_{J,1},...,b_{J,L}}}}
417: \prod_I
418: {n_I \atopwithdelims[] m_I}
419: %\QATOPD[ ]{n_I}{m_I}}}
420: {\cal P}\left( \vec{n},t \right) \;,
421: \label{eq:B6}
422: \end{equation}
423: where $\delta \left( n^{\prime },n\right) $ is the Kronecker delta.
424: Given a particular initial configuration $\vec{n}_0$,
425: ${\cal P}\left( \vec{n},t\right) $ can be found, in principle, by recursion
426: with ${\cal P}\left(\vec{n},0\right)=\delta \left( \vec{n},\vec{n}_0\right)$.
427: In that sense, we note that the more precise notation is
428: ${\cal P}\left( \vec{n},t|\vec{n}_0,0\right) $,
429: explicitly showing that it is the probability to find our
430: system in state $\vec{n}$ at time $t$, {\em conditioned} on a specific
431: initial condition. However, this notation seems unnecessarily cumbersome, so
432: that we will just use ${\cal P}\left( \vec{n},t\right) $ in its place.
433:
434: Note that equation (\ref{eq:B6}) is just a special example of the general
435: evolution of conditional probabilities in a Markov process, i.e.,
436: \begin{equation}
437: {\cal P}\left( \vec{n}^{\prime },t+1\right) =\sum_{\vec{n}}R\left(
438: \vec{n}^{\prime }|\vec{n}\right) {\cal P}\left( \vec{n},t\right) \;,
439: \label{eq:MEgen}
440: \end{equation}
441: where $R\left( \vec{n}^{\prime }|\vec{n}\right) $ is the conditional
442: probability for finding the system in $\vec{n}^{\prime }$ given that it was
443: in $\vec{n}$, also known as the transition rate. For our case, $R$ is
444: explicitly
445: \begin{equation}
446: R\left( \vec{n}^{\prime }|\vec{n}\right) =\sum_{\vec{m},\left\{ b\right\}}
447: \prod_K \delta \left( n_K' , b_K+B_K \right) \prod_J
448: {Fm_J \atopwithdelims[] b_{J,0},b_{J,1},...,b_{J,L}}
449: %{\QATOPD[ ]{Fm_J}{b_{J,0},b_{J,1},...,b_{J,L}}}}
450: \prod_I
451: {n_I \atopwithdelims[] m_I}
452: %{\QATOPD[ ] {n_I}{m_I}}}
453: \;.
454: \label{eq:R}
455: \end{equation}
456:
457: Once ${\cal P}\left( \vec{n},t\right) $ is found, the time dependence of the
458: expectation value of any quantity can be obtained via
459: \begin{equation}
460: \left\langle \bullet \right\rangle _t\equiv \sum_{\vec{n}}\bullet {\cal P}
461: \left( \vec{n},t\right) \,\,,
462: \label{eq:B7}
463: \end{equation}
464: in principle. Specifically, our main interest here will be $\left\langle
465: n_I\right\rangle _t$, the average number of individuals of species $I$ at
466: time $t$, as well as quantities quadratic in $n$ (e.g., covariances
467: and correlations). For
468: example, to see how $\left\langle n_K\right\rangle _t$ evolves in time, we
469: multiply equation (\ref{eq:MEgen}) by $n_K^{\prime }$, and sum over
470: $\vec{n}^{\prime }$
471: \begin{equation}
472: \left\langle n_K\right\rangle _{t+1}=\left\langle f_K\left( \vec{n}\right)
473: \right\rangle _t\,\,,
474: \label{eq:nEvolve}
475: \end{equation}
476: where
477: \begin{equation}
478: f_K\left( \vec{n}\right) \equiv \sum_{\vec{n}^{\prime }}n_K^{\prime }R\left(
479: \vec{n}^{\prime }|\vec{n}\right) \,\,.
480: \label{eq:fDef}
481: \end{equation}
482: Exploiting the explicit form of $R$ above and
483: \begin{equation}
484: \sum_mm\frac{n!}{m!\left( n-m\right) !}q^m\left( 1-q\right) ^{n-m}=mq\,\,,
485: \end{equation}
486: it is straightforward to find $f$. Reminding the readers of the $\vec{n}$
487: dependence in $P$, we write explicitly
488: \begin{equation}
489: f_K\left( \vec{n}\right) =F\left[ \left( 1-\mu \right) n_KP\left( K;\vec{n}
490: \right) +\left( \mu /L\right) \sum_J\sum_{\alpha =1}^L\Delta _K^{J,\alpha
491: }n_JP\left( J;\vec{n}\right) \right] \,\,. \label{eq:f}
492: \end{equation}
493: The interpretation of various terms in this expression is clear: $F$ offspring
494: are born to $n_JP\left( J\right) $ survivors of species $J$, with
495: rearrangements into the new generations due to mutations. Inserting it into
496: equation (\ref{eq:nEvolve}), we arrive at an {\em exact} equation
497: \begin{equation}
498: \left\langle n_K\right\rangle _{t+1}=F\left[ \left( 1-\mu \right)
499: \left\langle n_KP\left( K\right) \right\rangle _t+\left( \mu /L\right)
500: \sum_J\sum_{\alpha =1}^L\Delta _K^{J,\alpha }\left\langle n_JP\left(
501: J\right) \right\rangle _t\right] \,\,. \label{eq:nExact}
502: \end{equation}
503: Its simplicity is deceptive, since the $P$'s are, from equation (\ref{eq:B1}),
504: non-trivial functions of $\vec{n}$. If a formal expansion in powers of
505: $\vec{n}$ were inserted for these functions, then averages of all powers,
506: $\left\langle n_Jn_K...\right\rangle _t$, will appear on the right-hand side.
507: Of course, equations for these new averages can be written formally, but the
508: result would be more complex than the BBGKY hierarchy \cite{BBGKY}. Needless
509: to say, good approximation schemes are crucial for further progress.
510:
511: One such scheme, also known as the ``mean-field'' approximation, is to
512: ignore all correlations in order to produce a closed equation for
513: $\left\langle n_K\right\rangle $. Thus, we replace all averages of the
514: products by products of the averages:
515: \begin{equation}
516: \left\langle n_J n_K... \right\rangle \rightarrow \left\langle
517: n_J\right\rangle \left\langle n_K\right\rangle ...
518: \end{equation}
519: so that, e.g.,
520: \begin{equation}
521: \left\langle n_KP \left( K ; \vec{n} \right) \right\rangle \rightarrow
522: \left\langle n_K\right\rangle P\left( K;\left\langle \vec{n}\right\rangle
523: \right) \;.
524: \end{equation}
525: In terms of the less cumbersome notation $n_K\left( t\right) \equiv
526: \left\langle n_K\right\rangle _t$, equation (\ref{eq:nEvolve}) reduces to
527: \begin{equation}
528: n_K(t+1)=f_K\left( \{n_J\left( t\right) \}\right) \,\,.
529: \label{eq:MFT}
530: \end{equation}
531: Apart from the change of notation ($P_K(\{n_J(t)\})$
532: $\rightarrow P(K;\vec{n}(t))$),
533: this equation is precisely the starting point of our earlier
534: analysis, equation~(2) in \cite{RIKZIA03}:
535: \begin{equation}
536: \fl
537: n_K(t+1)=n_K(t)FP_K(\{n_J(t)\})[1-\mu ]
538: +
539: (\mu/L)F\sum_{I(K)}n_{I(K)}(t)P_{I(K)}(\{n_J(t)\})\;,
540: \label{eq:EOM}
541: \end{equation}
542: where $I\left( K\right) $ runs over all values of $I$ that differ from $K$
543: by one bit.
544:
545: In the low-mutation-rate ($\mu \ll 1$) regime, much of the behaviour of a
546: QSS is well approximated by a TSS ($\mu =0$), which can be understood in
547: terms of the fixed points of the mutationless version of this mean-field
548: theory. Here, let us briefly summarize the predictions of this theory. With
549: $\mu =0$, the number of populated species in the system, ${\cal N}$, never
550: increases. Therefore, we can restrict our attention to a small subspace
551: (${\cal N}$-dimensional, with ${\cal N}<5$ in typical simulation runs) of
552: the full $2^L$-dimensional space. In our previous work \cite{RIKZIA03}, we
553: use tildes (e.g., ${\bf \tilde{M}}$) to emphasize this aspect. To keep the
554: notation simple, we will drop the tildes here and keep in mind that, for
555: example, ${\bf M}$ is an ${\cal N}\times {\cal N}$ matrix. Also, with
556: ${\cal N}$ being an $O\left( 1\right) $ quantity (i.e., small compared to
557: $N_0$), the population of every species is generically $O\left( N_0\right)$.
558:
559: Next, $f_I$ simplifies to
560: \begin{equation}
561: f_I\left( \vec{n}\right)
562: {\underset {\mu =0} \rightarrow }
563: Fn_IP\left( I;\vec{%
564: n}\right) \,\,.
565: \label{eq:f(n)}
566: \end{equation}
567: Since the fixed point, $\vec{n}^{*}$, obeys
568: \begin{equation}
569: \vec{n}^{*}=\vec{f}\left( \vec{n}^{*}\right) \,\,,
570: \label{eq:n*eqn}
571: \end{equation}
572: we have $FP\left( I;\vec{n}^{*}\right) =1$ for all $I$. Defining the inverse
573: of ${\bf M}$ by
574: \begin{equation}
575: {\bf W}\equiv {\bf M}^{-1}\,\,,
576: \label{eq:W}
577: \end{equation}
578: (with elements $W_I^J$) and the sum
579: \begin{equation}
580: \sigma \equiv \sum_{IJ}W_I^J\,\,, \label{eq:Sigma}
581: \end{equation}
582: we found that the fractions of each species are given by
583: \begin{equation}
584: \rho ^{*}\left( I\right) \equiv n_I^{*}/N_{{\rm tot}}^{*}=\sum_JW_I^J/\sigma
585: \,\,, \label{eq:fix3}
586: \end{equation}
587: and the total $N_{{\rm tot}}^{*}$, by
588: \begin{equation}
589: N_{{\rm tot}}^{*}/N_0=\ln (F-1)+1/\sigma \;. \label{eq:fix4}
590: \end{equation}
591: Finally, the elements of ${\bf S}$, the stability matrix, is
592: \begin{equation}
593: S_I^J\equiv f_I{}^J\left( \vec{n}^{*}\right) \equiv \left. \partial
594: f_I/\partial n_J\right| _{\vec{n}^{*}}=\delta _I^J+\Lambda _I^J\,\,,
595: \label{eq:S}
596: \end{equation}
597: where $\delta _I^J$ is the unit matrix. Here,
598: \begin{equation}
599: \Lambda _I^J=\left( 1-\frac 1F\right) \left( M_I^J-\ln (F-1)-2/\sigma
600: \right) \rho ^{*}\left( I\right) \;.
601: \label{eq:A2}
602: \end{equation}
603: (denoted by $\tilde{\Lambda}_{IJ}$ in \cite{RIKZIA03} and referred to as the
604: community matrix in biological literature) plays the role of a ``restoring''
605: force, driving the population back towards the (stable) fixed point.
606:
607: \section{Fluctuations and correlations of populations}
608: \label{sec:TSS}
609:
610: It is clear that all stochastic aspects of the system are lost in the
611: mean-field approximation. In this section, we study the simplest aspect,
612: namely, an approximate description of the long-lived, quasi-steady
613: states (QSS). Now, as we have seen in simulations \cite{RIKZIA03}, these
614: states are dominated by a few mutually supportive species, along with a few
615: individuals of closely related ``benign mutants.'' Thus, we restrict
616: ourselves in this paper to systems {\em without mutation}, so that several
617: simplifications apply. First, there is the obvious reduction of $R$ from
618: equation (\ref{eq:R}) to
619: \begin{equation}
620: R \left( \vec{n}^{\prime } | \vec{n} \right)
621: {\underset {\mu =0} \rightarrow }
622: %{\rightarrow }_{\mu =0}
623: \sum_{\vec{m}} \prod_K \delta \left( n_K^{\prime },Fm_K\right)
624: \prod_I
625: {n_I \atopwithdelims[] m_I}
626: %{\QATOPD[ ] {n_I}{m_I}
627: \,\,.
628: \label{eq:R0}
629: \end{equation}
630: Next, since there can be no new species, we can focus on populations with
631: ${\cal N}\sim O\left( 1\right) $ species, each of which having $O\left(
632: N_0\right) $ individuals. Finally, though it is possible for a fluctuation
633: to collapse the entire population (which is, rigorously, {\em the unique}
634: stationary state associated with equations~(\ref{eq:MEgen},\ref{eq:R0})), such
635: events are so extreme that the life times of a typical non-trivial state
636: should be $O\left( \rme^{N_0}\right) $. So, for all practical purposes, we may
637: regard such states as ``truly'' stationary (the TSS's).
638: \footnote{
639: It is possible to force such states to be rigorously stationary, by slight
640: modifications of the rates. The simplest is to let the survival
641: probabilities $P\left( I\right) $ be unity when $n_I=1$.}
642:
643: A complete description of a TSS is provided by a stationary
644: {\em distribution} ${\cal P}^{*}\left( \vec{n}\right) $, which satisfies
645: \begin{equation}
646: {\cal P}^{*}\left( \vec{n}^{\prime }\right) =\sum_{\vec{n}}R\left( \vec{n}
647: ^{\prime }|\vec{n}\right) {\cal P}^{*}\left( \vec{n}\right) \,\,.
648: \label{eq:P*}
649: \end{equation}
650: Given ${\cal P}^{*}\left( \vec{n}\right) $, we can find stationary averages
651: of any quantity:
652: \begin{equation}
653: \left\langle \bullet \right\rangle ^{*}\equiv \sum_{\vec{n}}\left( \bullet
654: \right) {\cal P}^{*}\left( \vec{n}\right) \,\,. \label{eq:<dot>}
655: \end{equation}
656: Here, our main interest will be mean populations and their correlations:
657: \begin{equation}
658: \left\langle n_I\right\rangle ^{*}\quad \mathrm{and}\quad \left\langle
659: n_In_J\right\rangle ^{*}-\left\langle n_I\right\rangle ^{*}\left\langle
660: n_J\right\rangle ^{*}\,\,.
661: \label{eq:n+C}
662: \end{equation}
663:
664: Even in the absence of mutations and with simplified $R$'s, there are
665: substantial non-linearities (through $P\left( \vec{n}\right) $) in the
666: problem that prevent us from finding solutions to equation (\ref{eq:P*}) in
667: general. On the other hand, observations in simulations (and lessons learned
668: from the central limit theorem) show that our distributions are well
669: approximated by Gaussians. Indeed, a systematic expansion for
670: ${\cal P}^{*}$ can be formulated, starting from the Gaussian form:
671: \begin{equation}
672: \fl
673: {\cal P}_{\rm G} \left( \vec{n}\right)
674: =
675: \left( 2\pi N_0\right) ^{-{\cal N}/2}\left(
676: \det {\bf \hat{\Gamma}}\right) ^{1/2}\exp \left[ -\frac 1{2N_0}\left( n_I-
677: \bar{n}_I\right) \hat{\Gamma}^{IJ}\left( n_J-\bar{n}_J\right) \right] \,\,.
678: \label{eq:PG}
679: \end{equation}
680: where $\bar{n}_I$ and $\hat{\Gamma}^{IJ}$ are parameters to be determined.
681: Of course, the first of these is of $O\left( N_0\right) $. As for the
682: latter, we expect the fluctuations in our problem to be
683: $O\left( \sqrt{N_0}\right) $, i.e., covariances of $O\left( N_0\right) $.
684: To incorporate this
685: expectation, we have chosen to write ${\cal P}_{\rm G}$ in a form such that
686: the matrix ${\bf \hat{\Gamma}}$ has elements of $O\left( 1\right) $:
687: \begin{equation}
688: \hat{\Gamma}^{IJ}\sim O\left( 1\right) \,\,, \label{eq:Gam}
689: \end{equation}
690: {\em despite} the presence of two superscripts.
691:
692: Within the context of such a scheme, it is possible to compute these
693: quantities from the microscopic rates. Before proceeding to this
694: computation, we remark that this approximation for ${\cal P}^{*}$ will lead
695: to the predictions
696: \begin{equation}
697: \left\langle n_I\right\rangle ^{*}=\bar{n}_I\quad \mathrm{and}\quad
698: \left\langle n_In_J\right\rangle ^{*}-\left\langle n_I\right\rangle
699: ^{*}\left\langle n_J\right\rangle ^{*}=N_0\hat{G}_{IJ}\,,
700: \label{eq:nAv+cov}
701: \end{equation}
702: where ${\bf \hat{G}}$ is the inverse of ${\bf \hat{\Gamma}}$, i.e.,
703: \begin{equation}
704: \quad \hat{G}_{IJ}\hat{\Gamma}^{JK}=\delta _I^K\,\,.
705: \label{eq:GGam=1}
706: \end{equation}
707: Of course, we expect
708: \begin{equation}
709: \hat{G}_{IJ}\sim O\left( 1\right) \,\,,
710: \label{eq:G}
711: \end{equation}
712: despite its two subscripts.
713:
714: Though such approaches are well known \cite{vanKampen}, we provide a few
715: details in the Appendix, both for the sake of completeness and for the
716: convenience of readers unfamiliar with these methods. Here, we present only
717: the highlights of the analysis and how they apply to our case. As shown
718: below, the agreement between our predictions and simulation data in
719: three specific
720: cases are excellent and validates this entire approach.
721:
722: Turning to the computation of $\bar{n}_I$ and ${\bf \hat{\Gamma}}$ (or
723: ${\bf \hat{G}}$), we may expect the former to be simply related to the
724: $n_I^{*}$ of mean-field theory (equation (\ref{eq:n*eqn})). As shown in the
725: appendix, we have
726: \begin{equation}
727: \bar{n}_I=n_I^{*}\left[ 1+O\left( 1/N_0\right) \right]
728: \label{eq:nbar=n*}
729: \end{equation}
730: (provided none of the eigenvalues of ${\bf S}$ are close to unity). Indeed,
731: the first correction can be computed explicitly. Since these corrections are
732: not necessary for finding the fluctuations and correlations, we will not
733: quote the results here, but refer the reader to equation (\ref{eq:n-G}) in the
734: Appendix. To find ${\bf \hat{G}}$, we need not only
735: $f_I\left( \vec{n}\right) $ and the stability matrix ${\bf S}$
736: (given by equations (\ref{eq:f(n)},\ref{eq:S},\ref{eq:A2})), but also
737: \begin{equation}
738: H_{IJ}\left( \vec{n}\right) \equiv \sum_{\vec{n}^{\prime }}n_I^{\prime
739: }n_J^{\prime }R\left( \vec{n}^{\prime }|\vec{n}\right) \,\,.
740: \end{equation}
741: In the limit $\mu \rightarrow 0$, $R$ reduces to equation (\ref{eq:R0}),
742: so that
743: \begin{eqnarray}
744: H_{IJ}\left( \vec{n}\right) &\rightarrow &\sum_{\vec{m}}\left( Fm_I\right)
745: \left( Fm_J\right) \prod_I
746: {n_I \atopwithdelims[] m_I}
747: %{\QATOPD[ ] {n_I}{m_I}}}
748: \\
749: &=&F^2\left[ n_In_JP\left( I\right) P\left( J\right)
750: + \delta_I^J n_IP\left( I\right) \left( 1-P\left( I\right) \right) \right]
751: \,\,,
752: \label{eq:H(n)}
753: \end{eqnarray}
754: which is indeed of the form in equation (\ref{eq:h_IJ}):
755: $H_{IJ}=f_If_J+N_0\hat{H}_{IJ}$. Thus, we readily identify $\hat{H}_{IJ}$
756: of equation (\ref{eq:Htilde}):
757: \begin{equation}
758: \hat{H}_{IJ}=\delta_I^J F^2n_I^{*}P\left( I;\vec{n}^{*}\right)
759: \left[ 1-P\left( I;\vec{n}^{*}\right) \right] /N_0\,\,.
760: \label{eq:Htilde-here}
761: \end{equation}
762: But, $FP\left( I;\vec{n}^{*}\right) =1$ for all $I$, so that
763: \begin{equation}
764: \hat{H}_{IJ}=\delta_I^J \left( F-1\right) n_I^{*}/N_0\,\,.
765: \end{equation}
766: Referring to the Appendix again for the details, ${\bf \hat{G}}$ can be
767: obtained from ${\bf S}$ and ${\bf \hat{H}}$ via the linear relationship
768: \begin{equation}
769: {\bf \hat{G}-S\hat{G}S}^T={\bf \hat{H}\,\,.} \label{eq:GSH}
770: \end{equation}
771: Inserting the explicit form for ${\bf \hat{H}}$ into
772: equations (\ref{eq:Hab}, \ref{eq:explicit}),
773: we arrive at a complete solution for the covariance matrix
774: in terms of $\lambda _a,u_a^K$, and $v_I^a$ (respectively, the eigenvalues
775: and the left and right eigenvectors of $S_K^I$, normalized by $%
776: u_a^Iv_I^b=\delta _a^b$) :
777: \begin{equation}
778: \hat{G}_{IJ}=\left( F-1\right) \sum_{a,b,K}v_I^av_J^bu_a^Ku_b^K
779: \frac{n_K^{*}/N_0}{1-\lambda _a\lambda _b}+O\left( 1/N_0\right) \,\,.
780: \label{eq:explicitG}
781: \end{equation}
782:
783: \begin{figure}[t]
784: \includegraphics[angle=0,width=.47\textwidth]{n1n2hist.eps}
785: \includegraphics[angle=0,width=.53\textwidth]{pops_hist.eps}
786: \caption[]{
787: Theoretical Gaussian (thick, grey curves in the background) and simulated
788: (thin, black curves in the foreground)
789: TSS probability densities for ${\cal N} = 2$, $F=4$, and
790: $N_0 = 2000$, corresponding to the second column in table~\ref{tab:I}.
791: The Monte Carlo
792: simulation was performed over 524,290 generations with zero probability
793: of mutations.
794: ({\it a})
795: Contour plot of the
796: joint distribution for $n_1$ and $n_2$,
797: equation (\protect\ref{eq:PG}).
798: The negative slope of the long
799: axis indicates that $n_1$ and $n_2$ are negatively correlated.
800: ({\it b})
801: Marginal distributions for $n_1$ (right) and $n_2$ (left),
802: equation (\protect\ref{eq:Pproj}).
803: }
804: \label{fig:pops}
805: \end{figure}
806:
807: Before comparing these predictions to simulations, we provide an intuitive
808: picture for these fluctuations and correlations. Since we are concerned with
809: distributions well approximated by Gaussians, the underlying process is just
810: an Ornstein-Uhlenbeck one \cite{vanKampen}.
811: Such a process can be described by a Langevin
812: equation:
813: \begin{equation}
814: \vec{\varepsilon}\left( t+1\right) -\vec{\varepsilon}\left( t\right) =
815: {\bf \Lambda }\vec{\varepsilon}\left( t\right) +\vec{\eta}\left( t\right) \;,
816: \label{eq:Langevin}
817: \end{equation}
818: where ${\bf \Lambda }$ plays the role of restoring forces which drive
819: $\vec{\varepsilon}$ towards zero, and $\vec{\eta}$ is a Gaussian noise with
820: \begin{equation}
821: \left\langle \vec{\eta}\left( t\right) \right\rangle =0;\,\,\left\langle
822: \vec{\eta}\left( t\right) \vec{\eta}\left( t^{\prime }\right) \right\rangle
823: = {\bf \hat{H}}\delta \left( t,t^{\prime }\right) \,\,.
824: \end{equation}
825: The notation for the two matrices (${\bf \Lambda }$, ${\bf \hat{H}}$) is
826: chosen deliberately. For this problem, they are precisely those given above:
827: equations (\ref{eq:A2},\ref{eq:Htilde-here}), while
828: $\vec{\varepsilon}\left( t\right) $ is just the deviation from the average:
829: \begin{equation}
830: \vec{\varepsilon}\left( t\right) =\vec{n}\left( t\right) -\bar{n}\,\,.
831: \label{eq:epsilon}
832: \end{equation}
833: The deterministic part of equation (\ref{eq:Langevin}) is intuitively clear,
834: being the same as in the mean-field evolution equation, linearized about the
835: fixed point. Now, $\vec{\eta}$ can be seen as the (microscopic) noise in the
836: stochastic process. In our case, this is clearly due to the uncertainties
837: associated with survival. Since we have a simple two-state (dying or
838: surviving) random process, we can hardly be surprised by the presence of the
839: factors $nP\left( 1-P\right) $. Also, since this randomness is imposed on
840: each individual, the diagonal form of ${\bf \hat{H}}$ can be anticipated.
841: Perhaps only the factor $F^2$ cannot be easily surmised.
842:
843: \Table{
844: %\fulltable{
845: \label{tab:I}
846: Theoretical results
847: for two TSS communities with ${\cal N} = 2$ and~3, respectively, compared
848: with corresponding quantities from Monte Carlo simulations with $N_0 = 2000$
849: over 524,290 generations. The communities are defined by the interaction
850: matrices $\bf M$.
851: The quantities shown are the normalized fixed-point population vector,
852: $\vec{n}^*/N_0$ [equations (\ref{eq:fix3}) and~(\ref{eq:fix4})]
853: and the corresponding Monte Carlo average
854: $\langle \vec{n} \rangle_{\rm MC} /N_0$,
855: the normalized population covariance matrix $\bf \hat{G}$
856: [defined by equation (\ref{eq:nAv+cov}) and calculated from equation
857: (\protect\ref{eq:recrel}) with 20 iterations for ${\cal N} =2$ and 30
858: iterations for ${\cal N} =3$]
859: and ${\bf \hat{G}}_{\rm MC}$,
860: the step-covariance matrix $\bf \hat{g} $ [equation (\ref{eq:g})] and
861: ${\bf \hat{g}}_{\rm MC}$,
862: and the normalized
863: correlation matrix between steps $\vec{s}(t)$ and the deviation from the
864: average population $\vec{\varepsilon}(t) = \vec{n}(t) - \vec{n}^*$,
865: $\bf \hat{C}$ [equation (\ref{eq:se})] and ${\bf \hat{C}}_{\rm MC}$.
866: All numbers are given to four significant digits.
867: Results for ${\cal N} = 4$ are shown in table~\protect\ref{tab:II}.
868: }
869: \\
870: \br
871: %\hline
872: %\begin{footnotesize}
873: %\begin{tabular}{cccc}
874: \begin{tabular}{ccc}
875: ${\cal N}$ & 2 & 3
876: %& 4
877: \\
878: \mr
879: %\hline
880: \ms
881: $\bf M$
882: &
883: $\left(
884: \begin{array}{cc}
885: 0 & 0.9448 \\
886: 0.8563 & 0
887: \end{array}
888: \right) $
889: &
890: $\left(
891: \begin{array}{ccc}
892: 0 & 0.7497 & 0.9450 \\
893: 0.8935 & 0 & 0.6212 \\
894: 0.6474 & 0.9881 & 0
895: \end{array}
896: \right) $
897: %&
898: \\
899: \ms
900: \hline
901: \ms
902: $\vec{n}^*/N_0$
903: &
904: $\left(
905: \begin{array}{c}
906: 0.8119 \\
907: 0.7359
908: \end{array}
909: \right) $
910: &
911: $\left(
912: \begin{array}{c}
913: 0.6062 \\
914: 0.4897 \\
915: 0.5388
916: \end{array}
917: \right) $
918: %&
919: \\
920: \ms
921: \hline
922: \ms
923: $\langle \vec{n} \rangle_{\rm MC} /N_0$
924: &
925: $\left(
926: \begin{array}{c}
927: 0.8112 \\
928: 0.7350
929: \end{array}
930: \right) $
931: &
932: $\left(
933: \begin{array}{c}
934: 0.6055 \\
935: 0.4892 \\
936: 0.5378
937: \end{array}
938: \right) $
939: %&
940: \\
941: \ms
942: \hline
943: \ms
944: $\bf \hat{G}$
945: &
946: $\left(
947: \begin{array}{cc}
948: 3.294& -0.8722 \\
949: -0.8722 & 3.224
950: \end{array}
951: \right) $
952: &
953: $\left(
954: \begin{array}{ccc}
955: 3.667 & -0.9323 & -0.9558 \\
956: -0.9323 & 3.551 & -0.9249 \\
957: -0.9558 & -0.9249 & 3.590
958: \end{array}
959: \right) $
960: %&
961: \\
962: \ms
963: \hline
964: \ms
965: ${\bf \hat{G}}_{\rm MC}$
966: &
967: $\left(
968: \begin{array}{cc}
969: 3.295& -0.8784 \\
970: -0.8784 & 3.227
971: \end{array}
972: \right) $
973: &
974: $\left(
975: \begin{array}{ccc}
976: 3.690 & -0.9440 & -0.9584 \\
977: -0.9440 & 3.573 & -0.9289 \\
978: -0.9584 & -0.9289 & 3.592
979: \end{array}
980: \right) $
981: %&
982: \\
983: \ms
984: \hline
985: \ms
986: $\bf \hat{g} $
987: &
988: $\left(
989: \begin{array}{cc}
990: 4.455 & 1.368 \\
991: 1.368 & 3.882
992: \end{array}
993: \right) $
994: &
995: $\left(
996: \begin{array}{ccc}
997: 3.039 & 0.7902 & 0.8768 \\
998: 0.7902 & 2.285 & 0.7152 \\
999: 0.8768 & 0.7152 & 2.592
1000: \end{array}
1001: \right) $
1002: %&
1003: \\
1004: \ms
1005: \hline
1006: \ms
1007: ${\bf \hat{g}}_{\rm MC}$
1008: &
1009: $\left(
1010: \begin{array}{cc}
1011: 4.442 & 1.352 \\
1012: 1.352 & 3.865
1013: \end{array}
1014: \right) $
1015: &
1016: $\left(
1017: \begin{array}{ccc}
1018: 3.050 & 0.7865 & 0.8716 \\
1019: 0.7865 & 2.287 & 0.7098 \\
1020: 0.8716 & 0.7098 & 2.586
1021: \end{array}
1022: \right) $
1023: %&
1024: \\
1025: \ms
1026: \hline
1027: \ms
1028: $\bf \hat{C}$
1029: &
1030: $\left(
1031: \begin{array}{cc}
1032: -2.227 & -0.6494 \\
1033: -0.7189 & -1.941
1034: \end{array}
1035: \right) $
1036: &
1037: $\left(
1038: \begin{array}{ccc}
1039: -1.520 & -0.5253 & -0.2812\\
1040: -0.2649 & -1.143 & -0.5244\\
1041: -0.5956 & -0.1908 & -1.296
1042: \end{array}
1043: \right) $
1044: %&
1045: \\
1046: \ms
1047: \hline
1048: \ms
1049: ${\bf \hat{C}}_{\rm MC}$
1050: &
1051: $\left(
1052: \begin{array}{cc}
1053: -2.221 & -0.6385 \\
1054: -0.7139 & -1.933
1055: \end{array}
1056: \right) $
1057: &
1058: $\left(
1059: \begin{array}{ccc}
1060: -1.524 & -0.5221 & -0.2814\\
1061: -0.2643 & -1.143 & -0.5224\\
1062: -0.5902 & -0.1873 & -1.293
1063: \end{array}
1064: \right) $
1065: %&
1066: \\
1067: \ms
1068: \br
1069: %\hline
1070: \end{tabular}
1071: %\end{footnotesize}
1072: %\endfulltable
1073: \endTable
1074:
1075: \Table{
1076: %\fulltable{
1077: \label{tab:II}
1078: Theoretical results
1079: for a TSS community with ${\cal N} = 4$, compared
1080: with corresponding quantities from a Monte Carlo simulation with
1081: $N_0 = 10,000$ over 524,290 generations.
1082: The quantities shown are the same as in table~\protect\ref{tab:I}.
1083: The eigenvalues of $\bf S$ are relatively close to unity,
1084: so evaluation of $\bf G$ from equation (\protect\ref{eq:recrel}) required
1085: 180 iterations.
1086: }
1087: \\
1088: \br
1089: \ms
1090: \begin{tabular}{cc}
1091: ${\cal N}$ & 4
1092: \\
1093: \ms
1094: \mr
1095: \ms
1096: $\bf M$
1097: &
1098: $\left(
1099: \begin{array}{cccc}
1100: 0 & 0.5507 & 0.5101 & 0.9847 \\
1101: 0.9543 & 0 & 0.8437 & 0.9508 \\
1102: 0.4724 & 0.6190 & 0 & 0.7371 \\
1103: 0.9734 & 0.6808 & 0.1862 & 0
1104: \end{array}
1105: \right) $
1106: \\
1107: \ms
1108: \hline
1109: \ms
1110: $\vec{n}^*/N_0$
1111: &
1112: $\left(
1113: \begin{array}{c}
1114: 0.3355 \\
1115: 0.7034 \\
1116: 0.2366 \\
1117: 0.3468
1118: \end{array}
1119: \right) $
1120: %&
1121: \\
1122: \ms
1123: \hline
1124: \ms
1125: $\langle \vec{n} \rangle_{\rm MC} /N_0$
1126: &
1127: $\left(
1128: \begin{array}{c}
1129: 0.3356 \\
1130: 0.7035 \\
1131: 0.2354 \\
1132: 0.3472
1133: \end{array}
1134: \right) $
1135: \\
1136: \ms
1137: \hline
1138: \ms
1139: $\bf \hat{G}$
1140: &
1141: $\left(
1142: \begin{array}{cccc}
1143: 4.215 & -1.043 & -2.583 & 1.132 \\
1144: -1.043 & 3.882 & -0.1463 & -1.006 \\
1145: -2.583 & -0.1463 & 6.509 & -3.791 \\
1146: 1.132 & -1.006 & -3.791 & 5.653
1147: \end{array}
1148: \right) $
1149: \\
1150: \ms
1151: \hline
1152: \ms
1153: ${\bf \hat{G}}_{\rm MC}$
1154: &
1155: $\left(
1156: \begin{array}{cccc}
1157: 4.252 & -1.055 & -2.620 & 1.152 \\
1158: -1.055 & 3.870 & -0.1366 & -1.003 \\
1159: -2.620 & -0.1366 & 6.612 & -3.879 \\
1160: 1.151 & -1.003 & -3.879 & 5.718
1161: \end{array}
1162: \right) $
1163: \\
1164: \ms
1165: \hline
1166: \ms
1167: $\bf \hat{g} $
1168: &
1169: $\left(
1170: \begin{array}{cccc}
1171: 1.387 & 0.6468 & 0.2384 & 0.3094 \\
1172: 0.6468 & 3.704 & 0.4482 & 0.6518 \\
1173: 0.2384 & 0.4482 & 0.8927 & 0.2414 \\
1174: 0.3094 & 0.6518& 0.2414 & 1.461
1175: \end{array}
1176: \right) $
1177: \\
1178: \ms
1179: \hline
1180: \ms
1181: ${\bf \hat{g}}_{\rm MC}$
1182: &
1183: $\left(
1184: \begin{array}{cccc}
1185: 1.388 & 0.6417 & 0.2368 & 0.3090 \\
1186: 0.6417 & 3.694 & 0.4440 & 0.6501 \\
1187: 0.2368 & 0.4440 & 0.8870 & 0.2381 \\
1188: 0.3090 & 0.6501& 0.2381 & 1.462
1189: \end{array}
1190: \right) $
1191: \\
1192: \ms
1193: \hline
1194: \ms
1195: $\bf \hat{C}$
1196: &
1197: $\left(
1198: \begin{array}{cccc}
1199: -0.6934 & -0.3951 & -0.07285 & -0.1841 \\
1200: -0.2517 & -1.852 & -0.1801 & -0.3281 \\
1201: -0.1655 & -0.2681 & -0.4464 & -0.02048 \\
1202: -0.1253 & -0.3237 & -0.2209 & -0.7304
1203: \end{array}
1204: \right) $
1205: \\
1206: \ms
1207: \hline
1208: \ms
1209: ${\bf \hat{C}}_{\rm MC}$
1210: &
1211: $\left(
1212: \begin{array}{cccc}
1213: -0.6942 & -0.3907 & -0.07252 & -0.1851 \\
1214: -0.2510 & -1.847 & -0.1779 & -0.3259 \\
1215: -0.1642 & -0.2661 & -0.4435 & -0.01783 \\
1216: -0.1239 & -0.3242 & -0.2202 & -0.7310
1217: \end{array}
1218: \right) $
1219: \\
1220: \ms
1221: \br
1222: \end{tabular}
1223: \endTable
1224:
1225: We have carried out simulations for ten cases, with ${\cal N}=2,3$, and
1226: $4$. For simplicity, in each case we chose a set of species which served
1227: as the dominant ones in one of the ten QSS communities
1228: included in Table~I of \cite{RIKZIA03}. Of
1229: course, since $\mu $ was set to zero in the simulations reported here,
1230: the communities are TSS's.
1231: The runs were carried out for 524,290 generations each. From the recorded
1232: $n_I\left( t\right) $, we computed the averages
1233: $\left\langle n_I\right\rangle_{\rm MC}$ and the covariance matrix
1234: ${\bf \hat{G}}_{\rm MC}$. For the ${\cal N}=2$ case, we can easily display a
1235: histogram of all the populations
1236: (figure~\ref{fig:pops}(a)).
1237: In addition, in figure~\ref{fig:pops}(b), we show how good the Gaussian
1238: approximation is by plotting the projections of both this histogram (onto
1239: one or the other axis) and the theoretical predictions, e.g.,
1240: \begin{eqnarray}
1241: {\cal P}_{\rm proj}\left( n_1\right) &=&\int \rmd n_2{\cal P}_{\rm G}
1242: \left( n_1,n_2\right)
1243: \\
1244: &=&\left( \frac{\det {\bf \hat{\Gamma}}}{2\pi N_0\hat{\Gamma}_{22}}\right)
1245: ^{1/2}\exp \left[ -\frac{\det {\bf \hat{\Gamma}}}{2N_0\hat{\Gamma}_{22}}
1246: \left( n_1-\bar{n}_1\right) ^2\right] \,\,.
1247: \label{eq:Pproj}
1248: \end{eqnarray}
1249: We emphasize that the theoretical curves are produced with {\em no} fitting
1250: parameters -- all quantities were computed from the model specifications
1251: (i.e., $N_0$, $F$, and ${\bf M}$). For the ${\cal N}>2$ cases, it is difficult
1252: to display full histograms. Instead, we only provide the comparison for the
1253: averages and covariance matrices for three particular TSS's
1254: in tables~\ref{tab:I} and~\ref{tab:II}.
1255: As we see, the agreement is
1256: excellent, well within the expected accuracy of the approximation, $O\left(
1257: 1/N_0\right) $, and the statistical errors,
1258: $O\left( 1/\sqrt{524,290}\right) $.
1259:
1260: \section{Distributions containing dynamical information}
1261: \label{sec:dyn}
1262:
1263: In the previous section, we focused on the static distribution of the
1264: populations in a TSS. In other words, we can compute (within the Gaussian
1265: approximation) correlation functions that involve {\em any number} of
1266: species, all ``at the same time.'' Here, we investigate the information
1267: contained in the dynamics of the stochastic process, i.e., time-dependent
1268: correlations. For an {\em evolving} population, a natural question is how
1269: the system changes from one generation to the next. To probe this issue at a
1270: quantitative level, let us consider two examples, the statistics of
1271: ``steps,''
1272: \begin{equation}
1273: \vec{s}\left( t\right) \equiv \vec{n}\left( t+1\right) -\vec{n}\left(
1274: t\right) \;,
1275: \label{eq:stepdef}
1276: \end{equation}
1277: and the correlation of these steps with the deviations from the average:
1278: $\vec{\varepsilon}\left( t\right) $.
1279:
1280: In a stationary state, the mean-field prediction for the step size is
1281: clearly zero ($\vec{s}^{\,{\rm MF}}\equiv 0$), as must also be the case for
1282: the average $\left\langle \vec{s} \, \right\rangle ^{*}$. Nevertheless, we
1283: expect a typical step size to be $O\left( \sqrt{N_0}\right) $. For a more
1284: detailed picture, we may seek the step-size distribution (SSD) in the steady
1285: state: ${\cal P}_s^{*}\left( \vec{s} \, \right) $. A precise definition is
1286: \begin{equation}
1287: {\cal P}_s^{*}\left( \vec{s} \, \right) \equiv
1288: {\underset {t \rightarrow \infty } \lim }
1289: \sum_{\vec{n}^{\prime },\vec{n}}\delta \left( \vec{s},\vec{n}^{\prime
1290: }-\vec{n}\right) {\cal P}\left( \vec{n}^{\prime },t+1;\vec{n},t\right) \,\,,
1291: \label{eq:PsDef}
1292: \end{equation}
1293: where ${\cal P}\left( \vec{n}^{\prime },t+1;\vec{n},t\right) $ is the
1294: {\em joint } probability for finding the system with population $\vec{n}$
1295: at time $t$ {\em and} with $\vec{n}^{\prime }$ in the next step. From the
1296: master equation (\ref{eq:MEgen}), it is clear that this is just
1297: $R\left( \vec{n}^{\prime }|\vec{n}\right) {\cal P}\left( \vec{n},t\right) $,
1298: so that
1299: \begin{equation}
1300: {\underset {t \rightarrow \infty } \lim }
1301: {\cal P}\left( \vec{n}^{\prime },t+1;
1302: \vec{n},t\right) =R\left( \vec{n}^{\prime }|\vec{n}\right)
1303: {\cal P}^{*}\left( \vec{n}\right) \,\,.
1304: \end{equation}
1305: Thus, once the steady-state distribution ${\cal P}^{*}\left( \vec{n}\right) $
1306: is known, the SSD can be obtained from:
1307: \begin{equation}
1308: {\cal P}_s^{*}\left( \vec{s} \, \right) =\sum_{\vec{n}}R\left(\vec{n}
1309: +\vec{s}| \vec{n}\right) {\cal P}^{*}\left( \vec{n}\right) \,\,.
1310: \label{eq:Ps*}
1311: \end{equation}
1312:
1313: Applying this formalism to our case of coevolving species {\em without}
1314: mutations, we can exploit all the approximations detailed in the previous
1315: section, namely, a Gaussian for the stationary state
1316: (${\cal P}^{*}\left( \vec{n}\right) \cong {\cal P}_{\rm G}
1317: \left( \vec{n}\right) $
1318: of equation (\ref{eq:PG})), continuous variables for $\vec{n}$, and integrals
1319: instead of sums. Given equation (\ref{eq:P*}), the success of that scheme is
1320: implicitly dependent on the fact that
1321: $R\left( \vec{n}^{\prime }|\vec{n}\right) $ is also well approximated by
1322: a Gaussian
1323: \footnote{
1324: In the same manner as for the stationary distributions, this property can be
1325: derived from the definition of $R$ using straightforward, but tedious,
1326: manipulations with the binomials. Here, we can treat this as an assumption,
1327: the justification of which will be the agreement with simulation data.}.
1328: As a result, we need not carry out another lengthy analysis to conclude that
1329: ${\cal P}_s^{*}\left( \vec{s}\right) $ should also be of the form
1330: \begin{equation}
1331: {\cal P}_s^{*}\left( \vec{s} \, \right) \cong \left( 2\pi N_0\right)
1332: ^{-{\cal N}/2}\left( \det {\bf \hat{\gamma}}\right) ^{1/2}\exp \left[
1333: -\frac 1{2N_0}
1334: %\sum_{I,J}
1335: s_I\hat{\gamma}^{IJ}s_J\right] \,\,.
1336: \label{eq:PsG}
1337: \end{equation}
1338: However, it is clear that the matrix ${\bf \hat{\gamma}}$ is distinct from
1339: ${\bf \hat{\Gamma}}$, since the former must contain some ``dynamic''
1340: information. (Note that the caret is to remind us that, despite the presence
1341: of two superscripts, $\hat{\gamma}^{IJ}$ is of $O\left( 1\right) $.) Now,
1342: within the context of simple Gaussians, ${\bf \hat{\gamma}}$ can be found by
1343: computing its inverse
1344: \begin{equation}
1345: {\bf \hat{g}\equiv \hat{\gamma}}^{-1}
1346: \end{equation}
1347: which is just the second moment
1348: \begin{equation}
1349: \left\langle s_Is_J\right\rangle _s\equiv \sum_{\vec{s}}s_Is_J{\cal P}%
1350: _s^{*}\left( \vec{s}\right) \,\,.
1351: \end{equation}
1352: Let us first derive an exact formula for this quantity in terms of
1353: $\left\langle \bullet \right\rangle ^{*}$. Starting with
1354: \begin{eqnarray}
1355: \left\langle s_Is_J\right\rangle _s &=&\sum_{\vec{s},\vec{n}}s_Is_JR\left(
1356: \vec{n}+\vec{s}|\vec{n}\right) {\cal P}^{*}\left( \vec{n}\right)
1357: \label{eq:ss-def} \\
1358: &=&\sum_{\vec{n}^{\prime },\vec{n}}\left( n_I^{\prime }-n_I\right) \left(
1359: n_J^{\prime }-n_J\right) R\left( \vec{n}^{\prime }|\vec{n}\right)
1360: {\cal P}^{*}\left( \vec{n}\right) \,\,,
1361: \end{eqnarray}
1362: we use equations (\ref{eq:P*}) and (\ref{eq:fDef}) and arrive at the
1363: {\em exact} relation:
1364: \begin{equation}
1365: \left\langle s_Is_J\right\rangle _s=2\left\langle n_In_J\right\rangle
1366: ^{*}-\left\langle f_In_J\right\rangle ^{*}-\left\langle n_If_J\right\rangle
1367: ^{*}\,\,. \label{eq:ss-exact}
1368: \end{equation}
1369:
1370: Next, we exploit
1371: ${\cal P}^{*}\left( \vec{n}\right) \cong {\cal P}_{\rm G}
1372: \left( \vec{n}\right) $
1373: for computing $\left\langle \bullet \right\rangle ^{*}$ and
1374: apply equation (\ref{eq:Q*}). Thus,
1375: \begin{equation}
1376: \left\langle n_In_J\right\rangle ^{*}\cong \bar{n}_I\bar{n}_J+N_0\hat{G}%
1377: _{IJ}
1378: \label{eq:scov}
1379: \end{equation}
1380: and
1381: \begin{equation}
1382: \left\langle f_In_J\right\rangle ^{*}\cong f_I\left(\bar{n}\right)\bar{n}_J
1383: +N_0f_I^K\left( \bar{n}\right) \hat{G}_{KJ}
1384: + \frac{N_0}2f_I^{KM}\left( \bar{n}\right) \bar{n}_J\hat{G}_{KM}\,\,.
1385: \label{eq:mess}
1386: \end{equation}
1387: Thanks to equation (\ref{eq:nbar}), the first and the last terms in
1388: equation (\ref{eq:mess}) combine, so that
1389: \begin{equation}
1390: \left\langle f_In_J\right\rangle ^{*}\cong \bar{n}_I\bar{n}_J
1391: + N_0f_I^K\left( \bar{n}\right) \hat{G}_{KJ}\,\,.
1392: \label{eq:<fn>}
1393: \end{equation}
1394: To the order kept here, $f_I^K\left( \bar{n}\right) $ is just
1395: $f_I^K\left( \vec{n}^{*}\right) =S_I^K$. So, collecting various items
1396: and using equation (\ref{eq:S}), we arrive at
1397: \begin{eqnarray}
1398: \left\langle s_Is_J\right\rangle _s &\cong &N_0\left( 2\hat{G}_{IJ}
1399: - S_I^K\hat{G}_{KJ}-\hat{G}_{IK}S_J^K\right) \\
1400: &=&\,-N_0\left( \Lambda _I^K\hat{G}_{KJ}+\hat{G}_{IK}\Lambda _J^K\right) \,.
1401: \end{eqnarray}
1402: Finally, we relate this result to the Gaussian approximation
1403: equation (\ref{eq:PsG}) intended for the SSD, which provides
1404: \begin{equation}
1405: \left\langle s_Is_J\right\rangle _s\cong N_0\hat{g}_{IJ}\,\,,
1406: \end{equation}
1407: and obtain a simple equation for ${\bf \hat{g}}$:
1408: \begin{equation}
1409: {\bf \hat{g}}=-\left( {\bf \Lambda \hat{G}+\hat{G}\Lambda }^T\right) \,\,.
1410: \label{eq:g}
1411: \end{equation}
1412: As in the previous section, these predictions are well borne out in
1413: simulations. For the same ${\cal N}=2$ case as above,
1414: we display a two-dimensional
1415: histogram of the step sizes in figure~\ref{fig:diffs}.
1416: %as well as its projection onto one of the axes (figure~2b).
1417: Note that this distribution is indeed quite different from
1418: that for the populations (figure~\ref{fig:pops}).
1419: For the ${\cal N}>2$ cases, full
1420: histograms are difficult to display and we only show the correlation
1421: matrices ${\bf \hat{g}}$ and their Monte Carlo counterparts
1422: ${\bf \hat{g}}_{\rm MC}$ in tables~\ref{tab:I} and~\ref{tab:II}.
1423: \begin{figure}[t]
1424: \includegraphics[angle=0,width=.47\textwidth]{s1s2hist.eps}
1425: \caption[]{
1426: Contour plot of the
1427: theoretical Gaussian and simulated joint probability densities for the
1428: steps, $s_1$ and $s_2$, equation (\protect\ref{eq:PsG}).
1429: The line types and parameters are the same as in
1430: figure~\protect\ref{fig:pops},
1431: and the simulated histogram was obtained from the same
1432: simulation run as in that figure.
1433: The positive slope of the long axis indicates that the steps are
1434: positively correlated.
1435: }
1436: \label{fig:diffs}
1437: \end{figure}
1438:
1439:
1440: Although the SSD probes the underlying dynamics, it does not contain {\em all}
1441: information of the stochastic (Ornstein-Uhlenbeck) process. In particular,
1442: equation
1443: (\ref{eq:g}) shows that the antisymmetric part of ${\bf \Lambda \hat{G}}$
1444: is ``missing.'' To remedy this shortcoming, we turn to another question
1445: which naturally comes to mind: How are the steps ($\vec{s} \,$) correlated with
1446: the deviations from the average ($\vec{\varepsilon}\equiv \vec{n}-\bar{n}$)?
1447: Note that, since two quantities are involved ($\vec{s}$ and
1448: $\vec{\varepsilon}$), a full distribution associated with this question is
1449: slightly more involved than ${\cal P}^{*}\left( \vec{n}\right) $ or
1450: ${\cal P}_s^{*}\left( \vec{s}\right) $. Nevertheless, within the
1451: approximation scheme we use, it would be a (generalized) Gaussian in the
1452: steady state. For simplicity, let us focus on, instead of the full
1453: distribution,
1454: only the normalized correlation of $\vec{s}$ with $\vec{\varepsilon}$:
1455: \begin{equation}
1456: \hat{C}_{IJ}\equiv \left\langle s_I\varepsilon _J\right\rangle /N_0\,\,.
1457: \end{equation}
1458:
1459: By using this notation, we are displaying again our expectation that
1460: $\left\langle s_I\varepsilon _J\right\rangle $ is also of
1461: $O\left( N_0\right)$. Starting at the same point as equation (\ref{eq:ss-def}),
1462: we write
1463: \begin{eqnarray}
1464: \left\langle s_I\varepsilon _J\right\rangle &=&\sum_{\vec{s},\vec{n}}
1465: s_I\varepsilon _JR\left( \vec{n}+\vec{s}|\vec{n}\right) {\cal P}^{*}\left(
1466: \vec{n}\right)
1467: \label{eq:se-def} \\
1468: &=&\sum_{\vec{n}^{\prime },\vec{n}}\left( n_I^{\prime }-n_I\right) \left(
1469: n_J-\bar{n}_J\right) R\left( \vec{n}^{\prime }|\vec{n}\right) {\cal P}^{*}
1470: \left( \vec{n}\right) \,\,.
1471: \end{eqnarray}
1472: Since
1473: $\left\langle \vec{n}^{\prime }\right\rangle ^{*}
1474: =\left\langle \vec{n}\right\rangle ^{*}$, this reduces to
1475: \begin{equation}
1476: \left\langle s_I\varepsilon _J\right\rangle =\left\langle
1477: f_In_J\right\rangle ^{*}-\left\langle n_In_J\right\rangle ^{*}\,\,,
1478: \label{eq:se-exact}
1479: \end{equation}
1480: which is again an {\em exact} relationship. Repeating the analysis in the
1481: previous paragraph, we find, in the Gaussian approximation, a simple result:
1482: $\left\langle s_I\varepsilon _J\right\rangle =\,N_0\Lambda _I^K\hat{G}_{KJ}$%
1483: , or
1484: \begin{equation}
1485: \,{\bf \hat{C}=\Lambda \hat{G}}\,. \label{eq:se}
1486: \end{equation}
1487: Unlike the step-size covariance, this matrix is not symmetric in general and
1488: contains the full information of the dynamics (to this order of our
1489: approximation). Note also that, since ${\bf \Lambda }$ is negative and
1490: ${\bf \hat{G}}$ is positive, this correlation is typically negative. This
1491: simply reflects a restoring dynamics: steps and deviations from the mean
1492: tend to be opposite. In tables~\ref{tab:I} and~\ref{tab:II},
1493: we see that there is excellent agreement
1494: between the predicted ${\bf \hat{C}}$'s and their counterparts from
1495: simulations.
1496:
1497: Another commonly studied correlation is the connected two-time function,
1498: $\left( \langle n_I(t+\tau) n_J(t) \rangle
1499: - \langle n_I \rangle \langle n_J \rangle \right) / N_0$,
1500: which, in the stationary state, is just
1501: \begin{equation}
1502: \hat{\Xi}_{IJ}(\tau)
1503: = \langle \varepsilon_I(t+\tau) \varepsilon_I(t) \rangle / N_0
1504: \;.
1505: \label{eq:xi1}
1506: \end{equation}
1507: Within the Gaussian approximation, equation (\ref{eq:Langevin})
1508: can be used repeatedly to
1509: express $\varepsilon (t+\tau)$ in terms of $\varepsilon(t)$ and the noise,
1510: $\eta_I(t+t')$ at each time step.
1511: Since $\langle \eta_I(t+t') \varepsilon(t) \rangle
1512: = \langle \eta_I(t+t') \rangle \langle \varepsilon(t) \rangle = 0$
1513: for all $t'$, we find
1514: \begin{equation}
1515: {\bf \hat{\Xi}}(\tau) = {\bf {S}}^\tau {\bf \hat{G}}
1516: \;.
1517: \label{eq:xi2}
1518: \end{equation}
1519:
1520:
1521: \section{Concluding remarks}
1522:
1523: \label{sec:Concl}
1524:
1525: Since we have presented a considerable amount of mathematical details above,
1526: it is worthwhile to provide a short summary. In the model presented in
1527: \cite {RIKZIA03}, the entire evolution depends only on the parameters,
1528: %\begin{equation}
1529: $\mu$, $F$ , $N_0$ , and $M_{IJ}$.
1530: %\label{eq:parameters}
1531: %\end{equation}
1532: Strictly speaking, the only stationary state corresponds to total extinction
1533: ($\vec{n}=0$). But, if mutation is suppressed ($\mu =0$), then, not only does
1534: the configuration space break up into many sectors, but there will also be
1535: non-trivial long-lived ($O\left( \rme^{N_0}\right) $ generations) states.
1536: Referred to as ``truly stationary states,'' these communities consist of a
1537: fixed number (${\cal N}$ $\geq 1$) of species. Typically, the population of
1538: each species is $O\left( N_0\right) $. Provided we are not near a ``critical
1539: point,'' the full distribution of these populations can be studied by a
1540: systematic approximation scheme, starting with a multivariate Gaussian,
1541: parametrized by its mean and covariance matrix. For each TSS, our theory
1542: predicts these parameters and thus, the full distribution. Specifically,
1543: from the set of parameters,
1544: $\mu$, $F$ , $N_0$ , and $M_{IJ}$,
1545: %the set given above (\ref{eq:parameters}),
1546: we can compute $\vec{n}^{*}$,
1547: ${\bf \hat{H}}$, and ${\bf S=1-\Lambda }$
1548: (equations \ref{eq:fix3}-\ref{eq:A2}).
1549: Then, at the lowest order in our approximation scheme, the mean $\bar{n}$
1550: and the covariance matrix ${\bf \hat{G}}$ are given explicitly
1551: (equations \ref{eq:nbar=n*} and \ref{eq:GSH}-\ref{eq:explicitG}).
1552: In addition to
1553: this ``static'' aspect of the steady state, we also presented two ways of
1554: characterizing the ``dynamic'' aspect. One is the full distribution of sizes
1555: of steps (changes in the populations in a single time step, denoted by
1556: $\vec{s} \, $). Well approximated by a Gaussian also, this distribution
1557: has zero mean and covariance ${\bf \hat{g}}$ , which is given explicitly by
1558: equation (\ref{eq:g}). The other is the correlation of the steps with the
1559: populations just before the step (specified by the deviations from the
1560: average, $\vec{\varepsilon} \, $). Denoted by ${\bf \hat{C}}$, this correlation
1561: is given explicitly by equation (\ref{eq:se}). Finally, we have shown that
1562: there is excellent
1563: agreement between these analytical predictions and Monte Carlo simulations
1564: in ten typical TSS communities (data for three of which are shown in
1565: tables~\ref{tab:I} and~\ref{tab:II}
1566: and figures~\ref{fig:pops} and~\ref{fig:diffs}).
1567:
1568: With a solid understanding of the steady-state properties of stable
1569: communities in the absence of mutations, our next goal is a study of
1570: quasi-steady states of populations with low mutation rates. To carry out
1571: a serious analysis with $\mu >0$, we must define the restrictions on $R$
1572: carefully. Otherwise, it would be impossible to find a satisfactory solution
1573: to the equation
1574: \begin{equation}
1575: {\cal P}^{*}\left( \vec{n}^{\prime }\right) = \sum
1576: R\left( \vec{n}^{\prime }|\vec{n}\right)
1577: {\cal P}^{*}\left( \vec{n}\right) \;.
1578: \end{equation}
1579: As in the case for $\mu =0$, we may seek an approximate solution, in the
1580: form of a product of a Gaussian distribution for the dominant species and
1581: exponential distributions for the minority mutants \cite{ZIA04}. Provided
1582: this program is successful, the next step would be to study the probability
1583: of this type of QSS having ``dangerous'' mutants at the higher order of
1584: $\mu$, which hopefully will lead to some understanding of the distribution of
1585: QSS lifetimes \cite{Krug03}
1586: and the presence of $1/f$ noise in power spectral densities.
1587: Beyond this step, perhaps sophisticated renormalization-group techniques can
1588: be marshalled to account for the self-similar patterns displayed, as well as
1589: to identify universality classes for such behaviour. Needless to say, even
1590: for such a simple model of coevolution, much remains to be explored.
1591:
1592: \ack
1593: %\section{Acknowledgements}
1594:
1595: We acknowledge enlightening discussions with many colleagues, in particular,
1596: HW Diehl, BU Felderhof, K~Jain, J~Krug, AJ McKane, and L~Sch\"{a}fer.
1597: One of us (RKPZ) thanks HW Diehl for his hospitality at the
1598: University of Duisburg-Essen (Germany) where some of this work was carried
1599: out. This research is supported in part by the Alexander von Humboldt
1600: Foundation (Germany), Florida State University through the Center for
1601: Materials Research and Technology and the School of Computational Science
1602: and Information Technology, and grants from the US National Science
1603: Foundation (DMR-0088451, DMR-0120310, and DMR-0240078).
1604:
1605:
1606: \appendix
1607: \section{Mathematical Detail}
1608:
1609: Starting from the general expression for a discrete-time Markov process,
1610: \begin{equation}
1611: {\cal P}\left( \vec{n}^{\prime },t+1\right) =\sum_{\vec{n}}
1612: R\left( \vec{n}^{\prime }|\vec{n}\right)
1613: {\cal P}\left( \vec{n},t\right) \;,
1614: \label{eq:A1}
1615: \end{equation}
1616: we write the mean-field approximation for the evolution equation as
1617: \begin{equation}
1618: \vec{n}^{{\rm MF}}\left( t+1\right) =
1619: \vec{f}\left( \vec{n}^{{\rm MF}}\left(t\right) \right) \;,
1620: \label{eq:A2'}
1621: \end{equation}
1622: where the {\em functional form} of $\vec{f}$ is given by
1623: \begin{equation}
1624: \vec{f}\left( \vec{n}\right) =\sum_{\vec{n}^{\prime }}\vec{n}^{\prime}
1625: R\left( \vec{n}^{\prime }|\vec{n}\right) \;.
1626: \label{eq:A3}
1627: \end{equation}
1628: Note that the components of $\vec{n}$ play the role of coordinates in equation
1629: (\ref{eq:A1}) and
1630: take only non-negative integer values in models of population dynamics.
1631: However, there is no guarantee that $\vec{f}$ will be an integer in
1632: equation (\ref{eq:A2'}),
1633: so that we must allow $\vec{n}^{{\rm MF}}$ to be
1634: continuous variables (functions of $t$). As we will see below, our analysis
1635: is much simplified if we also assume $\vec{n}$ to be continuous.
1636:
1637: Focusing only on simple fixed points (as opposed to fixed cycles involving
1638: two or more points) of the mean-field theory, we denote a FP by $\vec{n}^{*}$.
1639: (To keep the notation from being too cumbersome, we suppress the superscript
1640: $^{{\rm MF}}$ and only keep in mind that $\vec{n}^{*}$ represents just
1641: ${\cal N}$ real numbers.) The equation for determining $\vec{n}^{*}$ is
1642: \begin{equation}
1643: \vec{n}^{*}=\vec{f}\left( \vec{n}^{*}\right) \,\,.
1644: \label{eq:n*}
1645: \end{equation}
1646: Anticipating the need for the associated stability matrix, we write it as
1647: \begin{equation}
1648: {\bf S\equiv }\left. \vec{\nabla}\vec{f} \, \right| _{\vec{n}^{*}}\,\,,
1649: \label{eq:Sdef}
1650: \end{equation}
1651: where $\vec{\nabla}$ stands for derivatives with respect to $\vec{n}$ .
1652: Explicitly, the matrix elements of ${\bf S}$ are
1653: \begin{equation}
1654: S_I^J=\left. \partial f_I/\partial n_J\right| _{\vec{n}^{*}} \label{eq:S'}
1655: \;,
1656: \end{equation}
1657: clearly asymmetric in general. As in our earlier study, we restrict
1658: ourselves to stable fixed points, so that the real parts of the eigenvalues
1659: of ${\bf S}$ (denoted by ${\bf 1}+{\bf \hat{\Lambda}}$ in \cite{RIKZIA03})
1660: should lie in the interval $(-1,1)$. Indeed, we restrict our attention
1661: here to isolated FP's far from others, so that no (real part of any)
1662: eigenvalue is close to unity.
1663:
1664: Turning to the full stochastic problem, we consider a stationary
1665: distribution ${\cal P}^{*}\left( \vec{n}\right) $, which satisfies
1666: \begin{equation}
1667: {\cal P}^{*}\left( \vec{n}^{\prime }\right) =
1668: \sum_{\vec{n}} R\left( \vec{n}^{\prime }|\vec{n}\right)
1669: {\cal P}^{*}\left( \vec{n}\right) \,\,.
1670: \label{eq:P*ap}
1671: \end{equation}
1672: Therefore, it is a right eigenvector of the matrix
1673: $R\left( \vec{n}^{\prime}|\vec{n}\right) $ with unit eigenvalue. Its
1674: existence is guaranteed (though not necessarily its uniqueness, in general)
1675: by the presence of a left eigenvector of unit eigenvalue
1676: (i.e., $u\left( \vec{n}\right) =1 \,\, \forall \vec{n}$),
1677: thanks to having probability conserving rates. If ${\cal P}^{*}$
1678: is known, then we can compute the stationary averages of all quantities via
1679: equation (\ref{eq:<dot>}). A different approach is to study the equations
1680: satisfied by these averages. Multiplying equation (\ref{eq:P*ap}) by a quantity
1681: and summing, we find, in general, no closed equation. Instead, a tangle of
1682: infinitely many coupled equations appear. Examples are
1683: \begin{equation}
1684: \left\langle n_I\right\rangle ^{*}
1685: =\left\langle f_I\left( \vec{n}\right)\right\rangle ^{*}
1686: \quad \mathrm{and}\quad
1687: \left\langle n_In_J\right\rangle^{*}
1688: =\left\langle H_{IJ}\left( \vec{n}\right) \right\rangle ^{*}\,\,,
1689: \label{eq:f+H}
1690: \end{equation}
1691: where $f_I$ is an element of $\vec{f}$, given by equation (\ref{eq:A3}), and
1692: \begin{equation}
1693: H_{IJ}\left( \vec{n}\right) \equiv \sum_{\vec{n}^{\prime }}
1694: n_I^{\prime}n_J^{\prime }R\left( \vec{n}^{\prime }|\vec{n}\right) \,\,.
1695: \label{eq:Hdef}
1696: \end{equation}
1697: In both cases, the right-hand sides contain expectation values of all powers of
1698: $\vec{n}$, so that infinitely many equations are needed to close the set.
1699:
1700: In practice, for generic situations such as those described in the main
1701: text, we may approximate ${\cal P}^{*}$ by a Gaussian:
1702: \begin{equation}
1703: \fl
1704: {\cal P}^{*}\left( \vec{n}\right) \cong {\cal P}_{\rm G}
1705: \left( \vec{n}\right)
1706: =\left( 2\pi N_0\right)^{-{\cal N}/2}
1707: \left( \det {\bf \hat{\Gamma}}\right)^{1/2}
1708: \exp \left[ -\frac 1{2N_0}\left( n_I-\bar{n}_I\right) \hat{\Gamma}^{IJ}
1709: \left( n_J-\bar{n}_J\right) \right] \,\,,
1710: \label{eq:PGap}
1711: \end{equation}
1712: where the unknowns $\bar{n}_I$ and $\hat{\Gamma}^{IJ}$ are to be determined.
1713: Naturally, we expect that the center of the
1714: Gaussian lies close to the mean-field FP,
1715: $\vec{n}^{*}$. This scenario is corroborated by our case studies
1716: above and our analysis below. Indeed, this approximation can be used as the
1717: starting point of a systematic expansion, which relies on having a large
1718: parameter ($N_0$) controlling the fluctuations to
1719: $O\left( \sqrt{N_0}\right)$. In this context, $\bar{n}$ and
1720: ${\bf \hat{\Gamma}}$ are assumed to be $O\left( N_0\right) $ and
1721: $O\left( 1\right) $, respectively. In addition to playing the role of
1722: ordering a systematic expansion, a large $N_0$ (and so, large $\vec{n}^{*}$)
1723: provides two further simplifications: (i) $\vec{n}$ may be promoted to
1724: continuous variables so that sums can be replaced by integrals, and
1725: (ii) integration can be extended to $\left[ -\infty ,\infty\right] $
1726: while only incurring errors of $O\left( \rme^{-\sqrt{N_0}}\right) $.
1727:
1728: Once $\bar{n}_I$ and $\hat{\Gamma}^{IJ}$ are determined, then we can set up
1729: expansions for the averages of any quantity $Q$:
1730: \begin{equation}
1731: \left\langle Q(\vec{n})\right\rangle ^{*}\equiv
1732: \sum_{\vec{n}} Q(\vec{n}){\cal P}^{*}\left( \vec{n}\right) \,\,,
1733: \label{eq:Q1}
1734: \end{equation}
1735: starting with
1736: \begin{equation}
1737: \int Q(\vec{n}){\cal P}_{\rm G}
1738: \left( \vec{n}\right) {\rm d} \vec{n} \,\,.
1739: \label{eq:Q2}
1740: \end{equation}
1741: Given that $n_I$ does not deviate far from $\bar{n}_I$, such integrals can
1742: be handled by first expanding $Q$ around $\bar{n}_I$:
1743: \begin{equation}
1744: Q(\vec{n}) = \bar{Q} +
1745: Q^I\left( n_I-\bar{n}_I\right) +
1746: \frac{1}{2} Q^{IJ}\left( n_I-\bar{n}_I\right) \left( n_J-\bar{n}_J\right)
1747: + \ldots \;,
1748: \label{eq:Qexpand}
1749: \end{equation}
1750: where the summation convention is used and
1751: \begin{eqnarray}
1752: \bar{Q} &\equiv &\left. Q\left(\vec{n}\right) \right| _{\vec{n}=\bar{n}}\;,
1753: \label{eq:Q3a} \\
1754: Q^I &\equiv &\left. \frac{\partial Q\left( \vec{n}\right) }{\partial n_I}
1755: \right| _{\vec{n}=\bar{n}} \;,
1756: \label{eq:Q3b} \\
1757: Q^{IJ} &\equiv &\left. \frac{\partial ^2 Q\left( \vec{n}\right) }
1758: {\partial n_I\partial n_J}\right| _{\vec{n}=\bar{n}} \;.
1759: \label{eq:Q3c}
1760: \end{eqnarray}
1761: The integral can now be performed, so that
1762: \begin{equation}
1763: \left\langle Q(\vec{n})\right\rangle ^{*}\cong \int Q(\vec{n})
1764: {\cal P}_{\rm G}\left( \vec{n}\right) {\rm d} \vec{n}
1765: =\bar{Q}\,+\frac{N_0}2Q^{IJ}\hat{G}_{IJ}+\ldots \;,
1766: \label{eq:Q*}
1767: \end{equation}
1768: where ${\bf \hat{G}}={\bf \hat{\Gamma}}^{-1}$. There is no need to be
1769: alarmed at the factor $N_0$ in the second term. Since $Q^{IJ}$ involves two
1770: derivatives with respect to $\vec{n}$,
1771: it is typically of the order of $1/N_0^2$
1772: compared to $\bar{Q}$. Thus, the second term in this expansion is
1773: $O\left(1/N_0\right) $ compared to the first.
1774:
1775: With this machinery, we seek equations to fix $\bar{n}_I$ and ${\bf \hat{G}}$.
1776: From equation (\ref{eq:n+C}), we have, with $Q=f_I$,
1777: \begin{equation}
1778: \bar{n}_I=\left\langle f_I\left( \vec{n}\right) \right\rangle ^{*}\cong
1779: f_I\left( \bar{n}\right) +\frac{N_0}2f_I{}^{JK}\hat{G}_{JK}+\ldots \,\,.
1780: \label{eq:nbar}
1781: \end{equation}
1782: Again, we expect the second term to be $O\left( 1/N_0\right) $ compared to
1783: the first, so that we may write an expansion for $\bar{n}_I$:
1784: \begin{equation}
1785: \bar{n}_I=\bar{n}_I^{(0)}+\bar{n}_I^{(1)}+\ldots \,\,.
1786: \label{eq:n-exp}
1787: \end{equation}
1788: Inserting this back into equation
1789: (\ref{eq:nbar}) and noting equation (\ref{eq:n*}), we find the expected
1790: \begin{equation}
1791: \bar{n}_I^{(0)}=n_I^{*}\,\,.
1792: \label{eq:n0}
1793: \end{equation}
1794: Proceeding to the next order, we find
1795: \begin{equation}
1796: \bar{n}_I^{(1)}=f_I^J\left( \bar{n}_I^{(0)}\right) \bar{n}_I^{(1)}
1797: +\frac{N_0}{2} f_I^{JK}\hat{G}_{JK}\,\,,
1798: \label{eq:n1G}
1799: \end{equation}
1800: where all functions of $\vec{n}$ in the last term can be evaluated at
1801: $\bar{n}_I^{(0)}$. To save notation, we will just write $\hat{G}_{JK}$
1802: in lieu of the more explicit form: $\hat{G}_{JK}^{(0)}$. We recognize that
1803: $f_I^J\left(\bar{n}_I^{(0)}\right) $ is just the stability matrix $S_I^J$
1804: and write
1805: \begin{equation}
1806: \bar{n}_I^{(1)}=\frac{N_0}2V_I^Mf_M{}^{JK}\hat{G}_{JK} \;,
1807: \label{eq:n-G}
1808: \end{equation}
1809: where
1810: \begin{equation}
1811: {\bf V\equiv }\left( {\bf 1-S}\right) ^{-1}
1812: \label{eq:Vdef}
1813: \end{equation}
1814: (the inverse of $-{\bf \tilde{\Lambda}}$ in \cite{RIKZIA03}). Let us
1815: emphasize that $V_I^M$ and $\hat{G}_{JK}$ are supposedly $O\left( 1\right) $,
1816: while $f_M{}^{JK}$ is $O\left( 1/N_0\right) $ so that the right-hand side of
1817: equation
1818: (\ref{eq:n-G}) is a quantity of $O\left( 1\right) $. This expression also
1819: provides a precise meaning to the phrase ``no eigenvalue (of ${\bf S}$) is
1820: close to unity,'' which appeared as a restriction on the FP's we study.
1821:
1822: Extending this technique to $n_In_J$ , we consider
1823: \begin{equation}
1824: N_0\hat{G}_{IJ}+\left\langle n_I\right\rangle ^{*}\left\langle
1825: n_J\right\rangle ^{*}=\left\langle n_In_J\right\rangle ^{*}=\left\langle
1826: H_{IJ}\left( \vec{n}\right) \right\rangle ^{*}\,\,,
1827: \label{eq:nn-exact}
1828: \end{equation}
1829: from equations (\ref{eq:n+C}) and (\ref{eq:f+H}). Again, using the Gaussian
1830: approximation for both sides, we arrive at
1831: \begin{equation}
1832: N_0\hat{G}_{IJ}+\bar{n}_I\bar{n}_J\cong H_{IJ}\left( \bar{n}\right)
1833: +\frac{N_0}2H_{IJ}{}^{KM}\hat{G}_{KM}+\ldots \,\,.
1834: \label{eq:G-H}
1835: \end{equation}
1836: At the lowest order, $O\left( N_0^2\right) $, of this equation, internal
1837: consistency will ensure that
1838: $H_{IJ}\left( \bar{n}^{(0)}\right) \cong \bar{n}_I^{(0)}\bar{n}_J^{(0)}$
1839: and should provide a check for tedious, error-prone
1840: computations. Furthermore, we can consider the {\em difference} :
1841: \begin{equation}
1842: H_{IJ}\left( \vec{n}\right) -
1843: f_I\left( \vec{n}\right) f_J\left( \vec{n}\right)
1844: \label{eq:h_IJ}
1845: \end{equation}
1846: and find that it is often one order lower (as explicitly shown in the main
1847: text for our model). In other words, we have
1848: \begin{equation}
1849: H_{IJ}\left( \vec{n}^{*}\right) -
1850: f_I\left( \vec{n}^{*}\right) f_J\left( \vec{n}^{*}\right)
1851: \sim O\left( N_0\right) \,\,.
1852: \end{equation}
1853: To emphasize this property, we explicitly extract a factor $N_0$ and
1854: define
1855: \begin{equation}
1856: \hat{H}_{IJ}\left( \vec{n}\right) \equiv \left[ H_{IJ}\left( \vec{n}\right)
1857: -f_I\left( \vec{n}\right) f_J\left( \vec{n}\right) \right] /N_0\,\,,
1858: \label{eq:Htilde0}
1859: \end{equation}
1860: so that
1861: \begin{equation}
1862: H_{IJ}\left( \bar{n}\right)
1863: =f_I\left( \bar{n}\right) f_J\left( \bar{n}\right)
1864: +N_0\hat{H}_{IJ}\left( \bar{n}\right) \,\,.
1865: \label{eq:Hff}
1866: \end{equation}
1867: Since we keep only terms of orders $N_0^2$ and $N_0$, we can evaluate
1868: $\hat{H}_{IJ}\left( \bar{n}\right) $ at $\vec{n}^{*}$ and define,
1869: for simplicity,
1870: \begin{equation}
1871: \hat{H}_{IJ}\equiv \hat{H}_{IJ}\left( \vec{n}^{*}\right) \,\,.
1872: \label{eq:Htilde}
1873: \end{equation}
1874: As a reminder, this quantity is of $O\left( 1\right) $. Next, to lowest
1875: order needed (i.e., $O\left( 1\right) $), we find
1876: \begin{equation}
1877: H_{IJ}{}^{KM}=S_I^KS_J^M+S_I^MS_J^K+f_If_J{}^{KM}+f_I{}^{KM}f_J{}\,\,,
1878: \label{eq:H_IJKM}
1879: \end{equation}
1880: all evaluated at the FP. Inserting these expressions into
1881: equation (\ref{eq:G-H}), we obtain
1882: \begin{eqnarray}
1883: \fl
1884: N_0G_{IJ}+\bar{n}_I\bar{n}_J
1885: &\cong&
1886: f_I\left( \bar{n}\right) f_J\left( \bar{n}\right)
1887: + N_0\hat{H}_{IJ} \nonumber\\
1888: && +\frac{N_0}2 \left[S_I^KS_J^M+S_I^MS_J^K
1889: + f_If_J{}^{KM}+f_I{}^{KM}f_J\right] G_{KM}+\ldots
1890: \,\,.
1891: \nonumber\\
1892: \end{eqnarray}
1893: Exploiting equation (\ref{eq:nbar}), we arrive at an equation for
1894: ${\bf \hat{G}}$ :
1895: \begin{equation}
1896: \hat{G}_{IJ}
1897: =\hat{H}_{IJ}+\frac 12\left[ S_I^KS_J^M+S_I^MS_J^K\right] \hat{G}_{KM} \;.
1898: \end{equation}
1899: But, ${\bf G}$ is symmetric, so that the final equation, in matrix form,
1900: reads
1901: \begin{equation}
1902: {\bf \hat{G}-S\hat{G}S}^T={\bf \hat{H}\,\,.} \label{eq:ansG}
1903: \end{equation}
1904: Here, ${\bf S}^T$ denotes the transpose of ${\bf S}$, and, from
1905: equation (\ref{eq:Hdef}), ${\bf \hat{H}}$ is necessarily symmetric. Again, let
1906: us emphasize that all quantities in this equation are $O\left( 1\right) $.
1907: Unfortunately, there is no simple way
1908: to express ${\bf \hat{G}}$
1909: in terms of the known matrices: ${\bf S}$ and ${\bf \hat{H}}$.
1910: A formal series can be written as
1911: \begin{equation}
1912: {\bf \hat{G}}
1913: =
1914: {\bf \hat{H}+S\,\hat{H}} \thinspace {\bf S}^T+{\bf SS\,\hat{H}}
1915: \thinspace {\bf S}^T{\bf S}^T \thinspace +...
1916: \;,
1917: \label{eq:formser}
1918: \end{equation}
1919: equivalent to a recursion relation,
1920: \begin{equation}
1921: {\bf \hat{G}}_n = {\bf \hat{H}} + {\bf {S}}{\bf \hat{G}}_{n-1}{\bf {S}}^T
1922: \label{eq:recrel}
1923: \end{equation}
1924: with ${\bf \hat{G}}_0 = {\bf \hat{H}}$.
1925: It was this recursion method that was used to obtain numerical results for
1926: comparison with the Monte Carlo results in the figures and tables.
1927: Convergence to four significant digits was obtained with
1928: $n=20$ for ${\cal N} =2$, $n=30$ for ${\cal N} =3$,
1929: and $n=140$ for ${\cal N} =4$.
1930:
1931: An explicit form for ${\bf \hat{G}}$
1932: can be found when we examine this equation in the frame where
1933: ${\bf S}$ is diagonal. To be explicit, take matrix elements of
1934: equation (\ref{eq:ansG}) with $u_a^I$, the left eigenvectors of ${\bf S}$,
1935: which satisfy
1936: \begin{equation}
1937: u_a^IS_I^K=\lambda _au_a^I\qquad \mathrm{(no\;sum\;on\;}a) \;.
1938: \end{equation}
1939: The result is
1940: \begin{equation}
1941: \hat{G}_{ab}=\hat{H}_{ab}/\left( 1-\lambda _a\lambda _b\right)
1942: \qquad \mathrm{(no\;sum)} \;,
1943: \label{final}
1944: \end{equation}
1945: where
1946: \begin{equation}
1947: \hat{G}_{ab}=\hat{G}_{IJ}u_a^Iu_b^J\qquad \mathrm{and}\qquad \hat{H}_{ab}
1948: =\hat{H}_{IJ}u_a^Iu_b^J\,\,. \label{eq:Hab}
1949: \end{equation}
1950: To find the original matrix elements, we apply the dual set $v_I^a$ (i.e.,
1951: the right eigenvectors of ${\bf S}$, normalized by $v_I^au_b^I=\delta _b^a$)
1952: and obtain
1953: \begin{equation}
1954: \hat{G}_{IJ}=v_I^a v_J^b \hat{G}_{ab}
1955: =\sum_{a,b}v_I^av_J^b\hat{H}_{ab}/\left(1-\lambda _a\lambda _b\right) \;.
1956: \label{eq:explicit}
1957: \end{equation}
1958:
1959: Finally, this expression again shows the importance of insisting that ``no
1960: eigenvalue (of ${\bf S}$) is close to unity.'' In case one or more $\lambda $
1961: approaches unity, the system would be labelled ``critical'' or
1962: ``multi-critical,'' in the language of phase transitions.
1963: Away from such
1964: points, the Gaussian approximation, along with its associated
1965: Ornstein-Uhlenbeck process, is adequate. For further details, see, for
1966: example, the books by van Kampen~\cite{vanKampen} and Risken~\cite{Risken}.
1967:
1968:
1969: \section*{References}
1970:
1971: \begin{thebibliography}{99}
1972: \bibitem{DROS01} Drossel B 2001 {\it Adv. Phys.} {\bf 50} 209
1973:
1974: \bibitem{LASS02} L{\"a}ssig M and Valleriani A 2002
1975: {\em Biological Evolution and Statistical Physics} (Berlin: Springer-Verlag)
1976:
1977: \bibitem{RIKV03A} Rikvold P~A and Zia R~K~P 2003, in {\em Computer
1978: Simulation Studies in Condensed Matter Physics XVI}, ed D~P Landau,
1979: S~P Lewis and H-B Sch{\"{u}}ttler (Berlin: Springer-Verlag, in press)
1980: {\it Preprint\/} arXiv:nlin.AO/0303010
1981:
1982: \bibitem{RIKZIA03} Rikvold P~A and Zia R~K~P 2003 {\it Phys. Rev. E}
1983: {\bf 68} 031913
1984:
1985: \bibitem{HALL02} Hall M, Christensen K, di~Collobiano S~A and
1986: Jensen H~J 2002 {\it Phys. Rev. E} {\bf 66} 011904
1987:
1988: \bibitem{CHRI02} Christensen K, di~Collobiano S~A, Hall M and
1989: Jensen H~J 2002 {\it J. theor. Biol.} {\bf 216} 73
1990:
1991: \bibitem{COLL03} di~Collobiano S~A, Christensen K and Jensen H~J 2003
1992: {\it J. Phys. A} {\bf 36} 883
1993:
1994: \bibitem{BAK93} Bak P and Sneppen K 1993 {\it Phys. Rev. Lett.}
1995: {\bf 71} 4083
1996:
1997: \bibitem{GOUL77} Gould S~J and Eldredge N 1977 {\it Paleobiology}
1998: {\bf 3} 115
1999:
2000: \bibitem{GOUL93} Gould S~J and Eldredge N 1993 {\it Nature} {\bf 366} 223
2001:
2002: \bibitem{NEWM85} Newman C~M, Cohen J~E and Kipnis C 1985
2003: {\it Nature} {\bf 315} 400
2004:
2005: \bibitem{Krug03} A similar distribution has been reported in Krug J and
2006: Karl C 2003 {\em Physica} {\bf A318} 137. Recast as a distribution of
2007: inverse lifetimes (i.e., ``decay rates''), $p\left( \tau \rightarrow
2008: \infty
2009: \right) \,\sim \,1/\tau ^2$ becomes
2010: $p \left( r\rightarrow 0\right) \,\sim \,const.$, a property which
2011: appears not only plausible but also ``generic'' or ``universal.''
2012:
2013: \bibitem{VERH1838} Verhulst P~F 1838 {\it Corres. Math. et Physique}
2014: {\bf 10} 113
2015:
2016: \bibitem{BBGKY} See, for example, Mohling F 1982 {\em Statistical Mechanics,
2017: Method and Application} (New York: Wiley);
2018: or Dorfman J~R 1999 {\em An introduction to Chaos in
2019: Nonequilibrium Statistical Mechanics} (Cambridge: Cambridge Univ.\ Press)
2020:
2021: \bibitem{vanKampen} See, for example, Chapter IX in van Kampen N 1981
2022: {\em Stochastic Processes in Physics and Chemistry}
2023: (Amsterdam: North Holland).
2024: A treatment which closely parallels the one given here can be found in Howard M
2025: and Zia R~K~P 2001 {\it Int. J. Mod. Phys. B} {\bf 15} 391
2026:
2027: \bibitem{ZIA04} Zia R~K~P and Rikvold P~A, in preparation
2028:
2029: \bibitem{Risken} Risken H 1984 {\em The Fokker-Planck Equation}
2030: (Berlin: Springer-Verlag)
2031:
2032: \end{thebibliography}
2033:
2034: \end{document}
2035: