1: \documentclass{article}
2: %\documentclass{revtex4}
3: \newcommand\rcsinfo{ \$Source: /home/axel/paper/web/RCS/web.tex,v $ $
4: -- \$Revision: 2.33 $ $ -- \$Date: 2005/04/28 07:17:34 $ $ }
5:
6: \usepackage{natbib}
7: \usepackage{rotating}
8: %% \usepackage{bibendnote}
9: %% \let\footnote=\bibendnote
10: \let\bibendnote=\footnote
11:
12: %\pagestyle{myheadings}
13: %\markright{\footnotesize \rcsinfo}
14:
15: % The following parameters seem to provide a reasonable page setup.
16: \topmargin 0.0cm
17: \oddsidemargin 0.7cm
18: \textwidth 15cm
19: \textheight 21cm
20: \footskip 1.0cm
21:
22: %% \makeatletter
23: %% \renewcommand*{\@seccntformat}[1]{%
24: %% \csname the#1\endcsname.\quad
25: %% }
26: %% \makeatother
27:
28: %% \usepackage{mydoublespace}
29: %% %\def\baselinestretch{2}
30: %% \addtolength{\footnotesep}{1ex}
31:
32: \setlength{\rightskip}{0pt plus 1fil} % Flatterrand
33:
34: %The next command sets up an environment for the abstract to your
35: %paper.
36: \usepackage{version}
37: \newcommand{\cm}[1]{({\small \sf #1})} % use for comments
38: \usepackage{amsmath}
39: \usepackage{url}
40: %\usepackage{srcltx}
41: \usepackage{graphicx}
42: \usepackage{threeparttable}
43: %\usepackage{showkeys}
44: \newlength{\imgwidth}\setlength{\imgwidth}{0.7\textwidth}
45: \includeversion{onlysubmission}\excludeversion{onlypreprint}
46: %\excludeversion{onlysubmission}\includeversion{onlypreprint}
47: \includeversion{nowordcount}
48:
49: \title{An explanatory model for food-web structure and evolution}
50: \author{A. G. Rossberg$^\ast$, H. Matsuda, T. Amemiya, K. Itoh\\
51: \normalsize{Yokohama National University, Graduate School of Environment}\\
52: \normalsize{and Information Sciences, Yokohama 240-8501, Japan}\\
53: \normalsize{$^\ast$Corresponding author. Tel.:
54: +81-45-339-4369, fax: +81-45-339-4353}\\
55: \normalsize{E-mail addresses: rossberg@ynu.ac.jp (A.G.R.),
56: matsuda2@ynu.ac.jp (H.M.),
57: }\\\normalsize{amemiyat@ynu.ac.jp (T.A.), itohkimi@ynu.ac.jp (K.I.)}}
58: \date{}
59:
60:
61: \begin{document}
62:
63:
64: % Double-space the manuscript.
65: %\baselineskip24pt
66: % Make the title.
67: \maketitle
68:
69: \thispagestyle{headings}
70:
71: \begin{abstract}
72: \setlength{\rightskip}{0pt plus 1fil} % Flatterrand
73: Food webs are networks describing who is eating whom in an
74: ecological community. By now it is clear that many aspects of
75: food-web structure are reproducible across diverse habitats, yet
76: little is known about the driving force behind this structure.
77: Evolutionary and population dynamical mechanisms have been
78: considered. We propose a model for the evolutionary dynamics of
79: food-web topology and show that it accurately reproduces observed
80: food-web characteristic in the steady state. It is based on the
81: observation that most consumers are larger than their resource
82: species and the hypothesis that speciation and extinction rates
83: decrease with increasing body mass. Results give strong support to
84: the evolutionary hypothesis.
85: \end{abstract}
86:
87: \textbf{Keywords:} food-webs, evolution, networks, model validation
88: \newpage{}
89:
90:
91: %% \cm{Everything that looks like this is a comment and will not appear
92: %% in the submitted version.}
93:
94:
95: \section{Introduction}
96: \label{sec:intro}
97:
98:
99: The complex networks of feeding relations in ecological communities,
100: food webs, have fascinated researchers for over thirty years
101: \citep{cohen90:_commun_food_webs_AND_THEREIN}. Using
102: carefully collected empirical data%
103: %% \citep{polis91:_compl_des_web,%
104: %% havens92:_scale_webs,%
105: %% warren89:_spatial_freshw_web,%
106: %% hall91:_food_rich_web,%
107: %% martinez91:_artif_attr,%
108: %% goldwasser93:_const_carib_web,%
109: %% baird89:_chesap_bay}%
110: , previous indications
111: \citep{cohen90:_commun_food_webs_AND_THEREIN,havens92:_scale_webs}
112: could be confirmed that several aspects of the structure of empirical
113: food webs are reproducible across habitats as diverse as Caribbean
114: islands, deserts, and lakes
115: \citep{camacho02:_robus_patter_food_web_struc,
116: %%stat-analysis,%
117: milo02:_networ_motif,%
118: garlaschelli03:_univer,%
119: williams00:_simpl,%
120: cattin04:_phylog%
121: }. Most striking, perhaps, were findings that certain descriptive
122: model food webs, constructed by following just a few simple rules for
123: connecting the elements of an abstract species pool, can
124: quantitatively reproduce large sets of basic statistics of food-web
125: topology \citep{
126: cohen90:_commun_food_webs_AND_THEREIN,%
127: williams00:_simpl,%
128: cattin04:_phylog%
129: }. These models do not yet expose the dynamical processes or
130: mechanisms by which the structures they describe are formed. But they
131: set a standard by which the accuracy of any dynamical explanation
132: should be measured \citep{williams00:_simpl}.
133:
134: Two major mechanisms for shaping food webs have been considered. One
135: is the selection of those food webs that support stable population
136: dynamics, because the degree of stability observed in nature is hard
137: to achieve with simple, randomly assembled food-web models
138: \citep{%web-stability%
139: may73:_stabil_complex_eco,yodzis81:stability_real_ecosystems%
140: }. The other mechanism is evolutionary dynamics, which is suggested
141: by the observation that the similarities in the trophic role of
142: species and their phylogenetic kinship are highly correlated
143: \citep{cattin04:_phylog}. Even though there has been plenty of work
144: on dynamical food-web models incorporating population dynamics,
145: evolutionary dynamics, or combinations of both \citep[for a review,
146: see][]{drossel03:_model_food_webs}, non of these models could so far
147: be shown to reproduce food-web structure to the accuracy of the more
148: recent descriptive models, namely the \emph{niche model}
149: \citep{williams00:_simpl} and the \emph{nested hierarchy model}
150: \citep{cattin04:_phylog}. Purely evolutionary models
151: \citep[e.g.,][]{amaral99:_envir_chang_coext_patter_fossil_recor} have
152: been developed mainly for understanding the dynamics of the total
153: number of species, and not the detailed food-web structure. Work on
154: models combining evolutionary and population dynamics [e.g.
155: \citet{%pop-dyn%
156: caldarelli98:_model_multispec_commun,drossel01:_influen_predat_prey_popul_dynam,yoshida03:_evolut_web_sys%
157: }] has not yet arrived at conclusive results, one problem being high
158: model complexity. A recent systematic tests of a plain population
159: dynamical model by \citet{montoya03:_topol_webs_real_to_assemb} turned
160: out clearly negative.
161:
162: Here we introduce a new model (called \emph{speciation model} below)
163: for the evolutionary dynamics of food webs, and show that its accuracy
164: is comparable to the descriptive models. For an in-depth mathematical
165: analysis of the model we refer to \citet{rossberg05:spec}.
166:
167: \section{Model definition}
168: \label{sec:model}
169:
170:
171: We model the evolutionary dynamics of an abstract pool of species
172: belonging to a given habitat, and the topology of the network of
173: feeding relations between them.
174: %
175: In order to separate the plain macroevolutionary dynamics from the
176: complex conditions that determine fitness, evolution is modeled as
177: being undirected. This simplification is here not meant to be a
178: statement about nature, but is used as a technique to isolate the
179: effect of particular processes.
180: %
181: Population dynamics, for example, does not enter the model.
182:
183: The speciation model is most naturally described in terms of a
184: continuous-time stochastic process. (See Appendix~\ref{sec:algorithm}
185: for a description of our computational implementation.) Its dynamics
186: consists of two parts, the dynamics of the community and the dynamics
187: of the food web connecting the species. Table~\ref{tab:par-list}
188: lists the model parameters.
189:
190: \subsection{Evolution of the species pool}
191: \label{sec:pool}
192:
193: Only three kinds of macro-evolutionary processes are taken into
194: account: speciations and extinctions within a habitat and adaptations
195: of new species to the habitat. Anagenesis (evolution without
196: speciation) is omitted for simplicity.
197: %
198: Since changes in the environment can be fast on evolutionary time
199: scales, it is for this model more appropriate to conceive a habitat as
200: a particular set of environmental conditions supporting an ecological
201: community, rather than just a specific location.
202:
203: %% The speciation model is characterized by the rate constants for the
204: %% adaptation $r_1$ of new species to a habitat, for speciation $r_+$,
205: %% and extinction $r_-$; the range $\exp(R)$ of evolution rates spanned;
206: %% the evolutionary variability of rates $D$; the looping tolerance
207: %% $\lambda$; the connection probability $C_0$; and the link persistence
208: %% $\Pi$.
209: %
210:
211:
212: Species at higher trophic levels tend to be larger than at lower
213: levels \citep{memmott00:_predat_size_web,leaper02:_size_in_web}. Due
214: to this correlation, a possible correlation between body size and
215: speciation- and extinction rates leads to correlations between
216: evolution rates and trophic level. Arguments in favor of both,
217: evolution rates increasing \citep{fenchel93:_small_large_spec} and
218: decreasing \citep{bush93:_santa_rosal} with body size have been put
219: forward. Empirically, the question seems to be undecided. Generally
220: one would assume that evolution rates are on the average proportional
221: to some power $M^\alpha$ of the adult body mass $M$. We note that
222: $\alpha=0$ is just as difficult to motivate as any other choice.
223: %
224: Based on the comparison of the observed distributions of the number of
225: links to resources and consumers
226: \citep{camacho02:_robus_patter_food_web_struc}, it can be argued that
227: $\alpha<0$ is more plausible in the context of the model proposed here
228: \citep{rossberg05:spec}.
229:
230: We associate
231: each species $i$ with a ``speed parameter'' $s_i$ (triangles in
232: Fig.~\ref{fig:speciation}) that determines the magnitude
233: of its evolutionary rates $\sim \exp(s_i)$. %% We assume that evolution
234: %% rates decrease with increasing body mass $M$. Even though a scaling
235: %% law relating body mass and evolution rates is unknown, it is
236: %% reasonable to assume that such a relation exists: typical time-scales
237: %% of life history, such as generation times, are known to scale as
238: %% $M^{1/4}$ \citep{peters83:eco_body_size} and maximum population
239: %% densities scale approximately as $M^{-3/4}$
240: %% \citep{%pop-scaling%
241: %% enquist98:_allom_plant,carbone02:_common_rule_scalin_carniv_densit,%
242: %% marquet02:_of_predat_prey_power_laws%
243: %% }. Assuming naively a constant
244: %% number of speciations per individual and generation and a homogeneous
245: %% population densities, this would lead to speciation rates scaling as
246: %% $M^{-1}$. But the model just depends on the assumption that the
247: %% exponent is negative, not on its precise value.
248: %% %% However, our model does not depend on the actual value of the
249: %% %% exponent, but just on the assumption that it is negative.
250: %
251: The community dynamics is characterized by the values of the rate
252: constants for new adaptations $r_1$, speciations $r_+$, and
253: extinctions $r_-$, the range $[0,R]$ of allowed values for $s$, and
254: the evolutionary speed/size dispersion $D$.
255: %
256: The complexity of the processes driving evolution is accounted for by
257: modeling extinctions, speciations, and adaptations of new species to
258: the habitat as independent random events: For any short time interval
259: $[t, t+dt]$, the probability that a species $i$ becomes extinct is
260: $r_- \exp(s_i) \, dt$ and the probability that it speciates is $r_+
261: \exp(s_i) \, dt$. When $i$ speciates, a new species $j$ with slightly
262: different speed parameter $s_j=s_i+\delta$ is introduced into the
263: community, where $\delta$ is a zero-mean Gaussian random variable with
264: $\mathop{\mathrm{var}}\delta=D$. If $s_i+\delta$ exceeds the range
265: $[0,R]$, $s_j=-(s_i+\delta)$ or $s_j=2 R-(s_i+\delta)$ are used
266: instead.
267: %
268: The probability that a new species with a speed parameter in a small
269: range $[s, s+ds]$ adapts to the habitat and joins the community is
270: $r_1 \exp(s)\, ds\, dt$ for all $s$ between $0$ and $R$.
271: %
272: %% As a result, the average equilibrium density of species
273: %% along the $s$ axis has a constant value $r_1/(r_--r_+)$ when $D$ is
274: %% not too large.
275:
276:
277: \subsection{Dynamics of the trophic network}
278: \label{sec:web}
279:
280:
281: The food web is described by the topology of the network of directed
282: trophic links connecting species. Parameters determining its dynamics
283: are the loopiness $\lambda$, the raw connectivity $C_0$, and the
284: reconnection probability $\beta$.
285: %
286: Following \citet{cohen90:_commun_food_webs_AND_THEREIN},
287: \citet{williams00:_simpl}, and \citet{cattin04:_phylog}, we take the
288: fact into account that consumers are typically larger than their
289: resource species. Empirical data show that they are, parasites
290: excluded, at least not much smaller in terms of adult body mass
291: \citep{memmott00:_predat_size_web,leaper02:_size_in_web}. We assume
292: that there is an upper limit $h$ for the ratio of the adult body sizes
293: of resource and consumer species. Resource species are at most $h$
294: times larger than their consumers. Making use of the assumption that
295: evolution rates $\sim\exp(s)$ scale as $M^{\alpha}$ with $\alpha<0$,
296: the difference of the speed parameter $s$ of a resource and its
297: consumer has an upper limit approximately given by $\Lambda:= \left|
298: \alpha \right|\times\ln h$. The degree by which the trophic ordering
299: is violated within the species pool is expressed by the loopiness
300: parameter $\lambda=\Lambda/R$. Summarizing, size constraints restrict
301: the set of \emph{possible consumers} of a species $i$ to the species
302: $l$ with $s_l<s_i+\lambda\,R$ and the \emph{possible resources} to the
303: species $h$ with $s_h>s_i-\lambda\,R$ where $0\le\lambda\le1$ (shaded
304: area in Fig.~\ref{fig:speciation}). Thus, cannibalism and trophic
305: loops \citep{polis91:_compl_des_web} can occur.
306:
307: When a new species $k$ is adapting to the habitat, a possible consumer
308: of $k$ becomes its consumer with probability $C_0$ and, with the same
309: probability, a possible resource of $k$ becomes its resource. When a
310: species $i$ speciates (Fig.~\ref{fig:speciation}), the connections of
311: the descendant species $j$ are determined in three steps. First, all
312: possible resources and consumers of $i$ that are also possible
313: resources and consumers of $j$ are linked to $j$. Possible resources
314: and consumers of $j$ that are not possible consumers or resources of
315: $i$ are linked with probability $C_0$. Then, for each possible
316: consumer or resource of $j$, the information whether it is connected
317: is discarded with probability $\beta$. Finally, a link is established
318: with probability $C_0$ to any consumer or resource for which the
319: information was discarded. Species that are both possible resources
320: and possible consumers of $j$ are treated like two species: a consumer
321: and a resource. This procedure ensures that the average ratio of
322: possible links to realized links remains at $C_0$, in line with the
323: premise that evolution is essentially undirected.
324: %
325:
326: %
327: The steady-state properties of the model are independent of the
328: initial conditions.
329: %
330: Approximations for some basic properties can be obtained analytically
331: \citep{rossberg05:spec}. For example, the expected density of species
332: has a constant value $r_1/(r_--r_+)$ along the $s$ axis. Thus, the
333: observed broad distributions of logarithmic body masses
334: \citep[e.g.][]{fenchel93:_small_large_spec} are here modeled by a
335: simple boxcar function. The number of species $S$ is on the average
336: $\overline{S}=R r_1/(r_--r_+)$ and, with $L$ denoting the number or
337: links, the average directed food-web connectivity defined as $C=L/S^2$
338: is approximately $\overline{C} \approx C_0\, (1+2
339: \lambda-\lambda^2)/2$. The fraction of species that entered the
340: habitat by speciation is $r_+/r_-$, the rest ($1-r_+/r_-$) entered as
341: new adaptations.
342:
343:
344: \section{Model verification}
345: \label{sec:empirical}
346:
347: The speciation model was compared with seven of the best available
348: empirical webs (see Table~\ref{tab:parameters})%
349: %% \citep{polis91:_compl_des_web,%
350: %% havens92:_scale_webs,%
351: %% warren89:_spatial_freshw_web,%
352: %% hall91:_food_rich_web,martinez91:_artif_attr,%
353: %% goldwasser93:_const_carib_web,%
354: %% braid89:_chesap_bay}
355: , all of which exclude parasites, in line with our model.
356: %% Similar datasets have been used in previous studies.
357: %% \citep{williams00:_simpl,%
358: %% camacho02:_robus_patter_food_web_struc,cattin04:_phylog} (CITE MORE)
359: Collecting food-web data is not at all an easy task, and the data is
360: sometimes criticised for being incomplete or inconsistent. But it was
361: for this kind of data that the regularities mentioned in
362: Section~\ref{sec:intro} have been found, which suggests that the data
363: does carries substantial information characteristic for food webs.
364:
365: A simple, visual comparison of the connection matrix of the food web
366: of Little Rock Lake \citep{martinez91:_artif_attr} with a model sample
367: (Fig.~\ref{fig:matrix}) shows that the model reproduces the
368: characteristic mixture of randomly scattered blocks and isolated links
369: found in empirical data. The empirical web has several omnivorous
370: species of high trophic level, which leads to a high density of links
371: near the left edge of the connection matrix. Such isolated, dense
372: blocks are only occasionally found in simulations. They are not
373: typical for empirical webs either.
374:
375: A systematic model verification was done based on twelve commonly used
376: food-web statistics (see Appendix~\ref{sec:properties}). We kept the
377: parameters $R=\ln 10^4$ and $D=0.0025$ fixed, since their precise
378: value has little effect ({doubling either $D$ or $R$ changes the
379: $\chi^2$ values computed below by less than one.}), and set $r_-=1$
380: without loss of generality. Depending on $r_+$, $\lambda$, and
381: $\beta$ (see below), parameters $r_1$ and $C_0$ were chosen so that
382: estimates of $\overline{S}$ and $\overline{C}$ after data
383: standardization (Appendix~\ref{sec:properties}) match with the species
384: abundance $S_\text{e}$ and connectivity $C_\text{e}$ of the empirical
385: webs. Maximum likelihood estimates (see Appendix~\ref{sec:stats})
386: assuming a multivariate Gaussian distribution for the twelve food-web
387: statistics were used for the remaining model parameters $r_+$,
388: $\lambda$, and $\beta$ (Table~\ref{tab:parameters}). As a benchmark,
389: food-web statistics were also computed for the niche model
390: \citep{williams00:_simpl} and for the nested hierarchy model
391: \citep{cattin04:_phylog}. Neither has free parameters. As is shown
392: in Fig.~\ref{fig:properties}, there is good overall agreement between
393: empirical and model data for all models.
394:
395: %% For comparison, the same statistics were calculated for the
396:
397: As a systematic measure for the goodness of fit, we computed the
398: $\chi^2$ statistics given by
399: \begin{align}
400: \label{chi2}
401: \chi^2=(\mathbf{v}_\text{e}-\overline{\mathbf{v}})^T \mathbf{C}^{-1}
402: (\mathbf{v}_\text{e}-\overline{\mathbf{v}}),
403: \end{align}
404: where $\mathbf{v}_\text{e}$ denotes the vector of the twelve
405: food-web statistics for an empirical web, and $\overline{\mathbf{v}}$
406: and $\mathbf{C}$ are the corresponding model mean and covariance
407: matrix (Appendix~\ref{sec:stats}).
408: %% obtained from $1000$ samples \footnote{At
409: %% this sample size, the difference between the $\chi^2$ distribution
410: %% and Hotelling's $T^2$ (valid for small samples) is negligible.}
411: In terms of $\chi^2$ statistics (Table~\ref{tab:parameters}) the
412: speciation model ($\chi^2_\text{S}$) is more accurate than the niche
413: ($\chi^2_\text{N}$) and the nested hierarchy model
414: ($\chi^2_\text{H}$), even when differences in the number of
415: statistical degrees of freedom (DOF $=$ number components of
416: $\mathbf{v}$ minus number of fitting parameters) are taken into
417: account. For example, the overall $p$ value for either niche or nested
418: hierarchy model is less than $10^{-24}$ times that of the speciation
419: model. In all but one case, the speciation model yields the lowest
420: $\chi^2$. This is partially due to inherently larger model variances
421: and partially due to a better accuracy of the model averages: in 46
422: out of 84 cases the average of the speciation model is closest to the
423: empirical value ($p<0.001$). Smaller webs are modeled particularly
424: well. For the five food webs with $S_e$ up to 35 (after species
425: lumping) all $\chi^2_\text{S}$ values lie within the $5\%$ confidence
426: interval $\chi^2_\text{S}<16.9$; and so does their sum $\sum
427: \chi^2_\text{S}=50.5 < 61.7$.
428: %% The two larger food webs
429: %% contain additional structure not captured by any of the models
430: %% (problems with consistent sampling of food webs might also play a
431: %% role here).
432:
433:
434: \section{Discussion}
435:
436: In order to exclude the fallacies that the surprisingly good
437: quantitative fit of the speciation model (1) does not actually depend
438: on evolutionary processes or (2) is merely a consequence of the data
439: standardization and evaluation procedures deployed, the model was also
440: evaluated with $r_+=0$ fixed, i.e.\ with all speciation processes
441: inhibited. In this case the parameter $\beta$ also does not have any
442: effect, so that the only remaining fitting parameter is $\lambda$.
443: The model fits worsened considerably. We obtained $\chi^2$ values
444: ranging from $65$ (Chesapeake Bay) to over 1000 (Ythian Estuary,
445: Little Rock Lake).
446:
447: Of course this does not prove that there is no other explanation for
448: the observations, but it shows that, for the good fit of the model,
449: the speciation events are crucial. Two effects of the evolutionary
450: dynamics are probably particularly important: (1) It leads to
451: approximately exponential distributions in the number of resources of
452: a species \citep{rossberg05:spec}, distributions that are similar to
453: those obtained with the niche model by prescribing a beta distribution
454: for the ``niche width''
455: \cite{camacho02:_analytic_food_webs,camacho02:_robus_patter_food_web_struc}.
456: (2) Related species have similar sets of consumers and resources, an
457: effect that the algorithm of the nested hierarchy model
458: \citep{cattin04:_phylog} was set up to mimic.
459:
460: %
461: The speciation model is at least as accurate in its
462: predictions as the niche model, which was itself a big improvement
463: over the, by now historical, cascade model
464: \citep{cohen90:_commun_food_webs_AND_THEREIN}. The speciation model is
465: the first that involves a mechanistic explanation, confirming the
466: earlier hypothesis \citep{cattin04:_phylog} that the structure of food
467: webs can be understood as the effect of a sequence of evolutionary
468: events.
469: %
470: Recent theoretical and empirical results
471: \citep{mccann00:_diversity_stability_review} show that the population
472: dynamics in complex food webs can be stabilized by mechanisms such as
473: weak links and adaptive foraging
474: \citep{kondoh03:_forag_adapt_relat_between_food}. This indicates that
475: population-dynamical restrictions on food-web structure are less
476: severe than one might first assume, making room for evolutionary
477: dynamics to leave its traces.
478:
479:
480: \section{Acknowledgements}
481: \label{sec:thanks}
482:
483: The authors express their gratitude to N.~D.~Martinez and
484: his group for making their food-web database available and The 21st
485: Century COE Program ``Bio-Eco Environmental Risk Management'' of the
486: Ministry of Education, Culture, Sports, Science and Technology of
487: Japan for generous support.
488:
489:
490:
491:
492: \clearpage{}
493:
494: \newpage
495:
496: \begin{nowordcount}
497:
498: %% \begin{center}
499: %% $\star$ \quad Supporting Online Material \quad $\star$ \\ \bigskip
500: %% \huge \bigskip
501: %% Materials and Methods
502: %% \bigskip
503: %% \end{center}
504: \appendix
505:
506: \makeatletter{}
507: \@addtoreset{equation}{section}
508: \makeatother{}
509: \renewcommand{\theequation}{\thesection.\arabic{equation}}
510:
511:
512: \section*{Appendices}
513: \label{appendices}
514:
515: \section{An Efficient Implementation of the Speciation Model}
516: \label{sec:algorithm}
517:
518: \noindent An evaluation of the speciation model based on
519: a straightforward discretization of speed parameters $s$ and time
520: $t$ would require excessive amounts of computation time. Here the
521: data structures and algorithms that were used in our, more efficient,
522: implementation are described.
523:
524:
525: \subsection{From Poisson processes to event sequences}
526:
527: The main loop of our algorithm advances the dynamics of the food webs
528: by one evolutionary event at each iteration. The question which event
529: comes next is decided based on the following considerations:
530: %
531: Call the quantity $\sigma_i:=\exp(s_i)$ the \emph{raw evolution rate},
532: or simply \emph{evolution rate} of species $i$. We first note that
533: the probability for new species with an evolution rate in the range
534: $[\sigma,\sigma+d\sigma]$ to adapt is
535: \begin{align}
536: \label{sigma}
537: r_1\,\exp(s)\,ds\,dt=r_1\,\sigma\, \frac{ds}{d\sigma}\,
538: d\sigma\,dt=r_1 d\sigma\,dt.
539: \end{align}
540: Thus, the total rate at which species adapt to the habitat is simply
541: \begin{align}
542: \label{R1}
543: \nu_1:=\int_{\sigma_\text{min}}^{\sigma_\text{max}} r_1 d\sigma =
544: r_1\,({\sigma_\text{max}}-{\sigma_\text{min}})
545: \end{align}
546: with $\sigma_\text{min}:=\exp(0)=1$ and $\sigma_\text{max}:=\exp(R)$.
547: Since all possible events are assumed to be statistically independent,
548: the probability that \emph{any} new-adaptation, extinction, or
549: speciation event occurs in the infinitesimal time interval $[t,t+dt]$
550: is $\nu_\text{tot} dt$ with
551: \begin{align}
552: \label{Rtot}
553: \nu_\text{tot}:= \nu_1 + \left(r_- +
554: r_+\vphantom{\sum}\right)\,\sum_{i} \sigma_i,
555: \end{align}
556: where the sum runs over the indices of all species currently in the
557: species pool. The time $\tau$ to the next event follows the
558: distribution
559: \begin{align}
560: \label{between-time}
561: P(\tau)=\nu_\text{tot}\,\exp(-\nu_\text{tot} \tau).
562: \end{align}
563:
564: If the probability for some event $X$
565: to occur in an interval $[t,t+dt]$ is $\nu_X\,dt$, then the
566: probability for event $X$ to occur as the next event is
567: \begin{align}
568: \label{pnext}
569: \frac{\nu_X}{\nu_\text{tot}}.
570: \end{align}
571: To see this, consider first the situation with only two kinds of
572: events: event $A$ occurring at rate $\nu_A$ and event $B$ at rate
573: $\nu_B$ (both are Poisson processes). The joint probability
574: distribution for the time $\tau_A$ to the next event $A$ and the time
575: $\tau_B$ to the next event $B$ is
576: \begin{align}
577: \label{Pab}
578: P_{AB}(\tau_A,\tau_B)=\nu_A\,\exp(-\nu_A\tau_A)\times
579: \nu_B\,\exp(-\nu_B\tau_B).
580: \end{align}
581: The probability that $A$ occurs before $B$ is
582: \begin{align}
583: P \left[ \tau_A < \tau_B \right]=&\int_0^\infty
584: d\tau_A\int_{\tau_A}^\infty d\tau_B
585: P_{AB}(\tau_A,\tau_B)\notag\\
586: \label{AbeforeB}
587: =&\int_0^\infty d\tau_A \nu_A \exp(-\nu_A\tau_A-\nu_B\tau_A)\\
588: =&\frac{\nu_A}{\nu_A+\nu_B}.\notag
589: \end{align}
590: The probability (\ref{pnext}) is obtained by letting event $A$ be
591: event $X$ and event $B$ be any other event.
592:
593: For the sake of formal simplicity the range of allowed values for the
594: speed parameter $s$ is reduced from $[0,R]$ to the half open interval
595: $[0,R[$ without affecting numerical results. In order to decide which
596: event comes next in the sequence of events, a random number $\rho$
597: equally distributed in the range $[0,\nu_\text{tot}[$ is generated.
598: If $\rho<\nu_1$, the next event is taken to be a new adaptation and
599: in view of Eq.~(\ref{sigma}) the evolution rate of the new species is
600: chosen as a random number equally distributed in the range
601: $[\sigma_\text{min},\sigma_\text{max}[$. If $ \nu_1\le\rho$, the next
602: event is taken to be either the extinction or a speciation of the
603: species $i$ that satisfies
604: \begin{align}
605: \label{chose-i}
606: \left(r_- +
607: r_+\vphantom{\sum}\right)\,\sum_{\sigma_j>\sigma_i}\sigma_j \le
608: \rho -\nu_1 < \sigma_i+\left(r_- +
609: r_+\vphantom{\sum}\right)\,\sum_{\sigma_j>\sigma_i}\sigma_j.
610: \end{align}
611: %% The species of the species pool are stored in a sorted list with
612: %% decreasing $\sigma$. Finding the species $i$
613: %% satisfying~(\ref{chose-i}) usually requires only the summation over
614: %% the first few elements of the list. The total sum $\sum_i\sigma_i$,
615: %% that is required to compute $\nu_\text{tot}$ according
616: %% to~(\ref{Rtot}), is kept in memory and updated with every event by
617: %% adding or subtracting the corresponding $\sigma_i$.
618: %
619: Another random number is used to decide if the species $i$ goes
620: extinct [probability $r_-/(r_-+r_+)$] or speciates.
621:
622:
623: \subsection{Data structures for keeping track of the network}
624:
625: The connectivity between species is stored in a matrix of bits
626: $\{b_{ij}\}$ with $b_{ij}=1$ is $j$ eats $i$ and $b_{ij}=0$ otherwise.
627: In order to avoid having to re-arrange the full matrix $b_{ij}$ with
628: every event, the number of rows and columns of $\{b_{ij}\}$ is chose
629: much larger than the actual number of species is the pool.
630: Indices are dynamically allocated with every new adaptation or
631: speciation, and deallocated upon extinction. Unused indices are kept
632: in a stack data-structure.
633: %
634: The determination of the connectivity follows directly the model
635: definition.
636: %% , although some use can be made of the fact that most events
637: %% involve species with high evolution rates $\sigma$ and large $s$.
638: %% Since these have few resources but many consumers, $\{b_{ij}\}$ is
639: %% arranged in such a way that copying all consumers from the parent
640: %% species to the descendant species in a speciation is a fast operation.
641:
642:
643: %% \subsection*{Random numbers}
644:
645: %% \cm{include this only if including the code}Most of the computation
646: %% time of our implementation is used to generate random numbers. We
647: %% reduce this load by making use of the fact that $C_0$ and $\beta$, the
648: %% probabilities of connecting and re-connecting two species, are usually
649: %% small. Using a unit random number $\rho$ in the range $[0,1[$, an
650: %% random decision with small probability $p$ is obtained by verifying if
651: %% the condition $\rho < p$ holds. In the unlikely case that $\rho < p$
652: %% is satisfied, $\rho$ is discarded and a new random number is used for
653: %% the next decision. If $\rho \ge p$, the quantity
654: %% \begin{align}
655: %% \label{newrho}
656: %% \frac{\rho-p}{1-p}
657: %% \end{align}
658: %% is used as the next unit random number $\rho$. This ``recycling'' of
659: %% random number imposes high demands on the random number generator. We
660: %% are using the ``Mersenne Twister'' of Matsumoto and Nishimura
661: %% (\emph{ACM Trans. on Modeling and Computer Simulation} \textbf{8}, pp
662: %% 3-30, 1998).
663:
664:
665: \subsection{Sampling the steady state}
666:
667: It is important to notice that the main loop of the algorithm, which
668: simulates one evolutionary event per iteration, does not proceed with
669: constant speed along the time axis $t$. Episodes with many species
670: require more iterations per unit time than episodes with few species.
671: In order to sample the steady state evenly in $t$, the an estimate $\hat t$
672: of $t$ has to be used, since the true time between events is left
673: undetermined with the method described above. Such an estimate can be
674: obtained by starting with $\hat t=0$ and incrementing $\hat t$ by the
675: current expectation value of the time between events
676: $\tau_\text{tot}:=\nu_\text{tot}^{-1}$ at each iteration.
677:
678: The portion of the food web that involves small species with large
679: $\sigma$ reaches the steady state much faster than the portion that
680: involves only large, slow species (small $\sigma$). In the initial
681: stage of the simulation, it is therefore sufficient to account for the
682: portion involving only slow species. In practice, this is achieved by
683: setting the effective value $R'$ of the parameter $R$ to a low value
684: $R_\text{start}$ (e.g., $R_\text{start}=1$) at the beginning of the
685: simulation and increasing $R'$ after each iteration by a small amount
686: proportional to $\tau_\text{tot}$. When $R'$ has reached $R$, the
687: simulation is continued for some time, and then a steady-state
688: food web is sampled. For the next steady-state sample, the complete
689: food web is discarded, and the simulation is restarted with a
690: $R'=R_\text{start}$. We started each run with an empty species pool.
691:
692: In order to check if the relaxation time is long enough to reach the
693: steady state, we verified that, for the steady-state samples taken,
694: the density of species along the $s$ axis has the theoretically
695: expected constant value $r_1/(r_--r_+)$ for $D=0$.
696:
697:
698: \section{Food-web properties}
699: \label{sec:properties}
700:
701: The food-webs properties that were used to characterize and compare
702: empirical and model webs were: the clustering coefficient
703: \citep{camacho02:_robus_patter_food_web_struc,
704: dorogovtsev02:_evolut_networ} (\textit{Clust} in
705: Figure~\ref{fig:properties}); the fractions of cannibalistic species
706: \citep{williams00:_simpl} (\textit{Cannib}) and species without
707: consumers \citep{cohen90:_commun_food_webs_AND_THEREIN} (\textit{T},
708: top predators); the relative standard deviation in the number of
709: resource species \citep{schoener89:_food_webs} (\textit{GenSD},
710: generality s.d.) and consumers \citep{schoener89:_food_webs}
711: (\textit{VulSD}, vulnerability s.d.); the web average of the maximum
712: of a species' Jaccard similarity \citep{jaccard08:_nouvel_florale}
713: with any other species \citep{williams00:_simpl} (\textit{MxSim}); the
714: fraction of triples of species with two or more resources, which have
715: sets of resources that cannot be ordered to be all contiguous on a
716: line \citep{cattin04:_phylog} (\textit{Ddiet}); the average
717: \citep{cohen90:_commun_food_webs_AND_THEREIN} (\textit{aChnLg}),
718: standard deviation \citep{martinez91:_artif_attr} (\textit{aChnSD}),
719: and average per-species standard deviation
720: \citep{goldwasser93:_const_carib_web} (\textit{aOmniv}, omnivory) of
721: the length of food chains, as well as the $\log_{10}$ of their total
722: number \citep{martinez91:_artif_attr} (\textit{aChnNo}), with the prefix
723: \textit{a} indicating that these quantities were computed using the
724: fast, ``deterministic'' Berger-Shor approximation
725: \citep{berger90:_approx_subgraph} of the maximum acyclic subgraph
726: (MAS) of the food web. The number of non-cannibal trophic links not
727: included in the MAS was measured as \textit{aLoop}. When the output
728: MAS of the Berger-Shor algorithm was not uniquely defined, the average
729: over all possible outputs was used.
730:
731: The statistics were computed after standardization of food webs to
732: allow a consistent comparison of data: Since in many records of
733: empirical food webs the lowest trophic level is poorly resolved, all
734: species without resource species were lumped to a single ``trophic
735: species'' after dropping disconnected species and before the usual
736: lumping of trophically equivalent species
737: \citep{cohen90:_commun_food_webs_AND_THEREIN}, for both empirical and
738: model data.
739:
740: \section{Maximum Likelihood Estimates,
741: Goodness of Fit, and Confidence Intervals}
742: \label{sec:stats}
743:
744: Maximum likelihood estimation of model parameters and a systematic
745: characterization of the goodness of fit ideally require knowledge of
746: the joint probability distribution of the food-web statistics in the
747: steady state for any given set of model parameters. We used the
748: methods described below to obtain satisfactory approximation in a
749: computationally feasible way.
750:
751: Since species number $S$ and directed connectivity $C$ tend to
752: fluctuate strongly in the steady state of the speciation model, it is
753: impracticable to restrict sampling to those webs with $S$ and $C$ in
754: close vicinity of the corresponding empirical values $S_\text{e}$ and
755: $C_\text{e}$ as done by \citet{williams00:_simpl,cattin04:_phylog}.
756: Instead, all webs with values of $S$ and $C$ differing by less than
757: $30\%$ from $S_\text{e}$ and $C_\text{e}$ were used for the
758: statistical analysis. In order to estimate the steady-state mean
759: $\overline{\mathbf{v}}$ and covariance matrix $\mathbf{C}$ of the
760: twelve food-web statistics \emph{conditional to} $S=S_\text{e}$,
761: $C=C_\text{e}$, we first computed steady-state mean
762: $\overline{\mathbf{u}}$ and covariance matrix $\mathbf{D}$ of the
763: vector
764: \begin{align}
765: \mathbf{u}=
766: \left(
767: \begin{matrix}
768: \mathbf{v}\\
769: \mathbf{w}
770: \end{matrix}
771: \right),
772: \end{align}
773: consisting of the twelve food web statistics $\mathbf{v}$ and the
774: vector $\mathbf{w}=(S,C)^T$ from $N=1000$ samples within the
775: $\pm30\%$~range.
776: %
777: Using $\mathbf{w}_\text{e}=(S_\text{e},C_\text{e})^T$ and components
778: of $\overline{\mathbf{u}}$ and $\mathbf{D}^{-1}$ written in the block
779: matrix forms
780: \begin{align}
781: \label{cp}
782: \overline{\mathbf{u}}=
783: \left(
784: \begin{matrix}
785: \overline{\mathbf{u}}_\mathbf{v}\\
786: \overline{\mathbf{u}}_\mathbf{w}
787: \end{matrix}
788: \right),\quad
789: \mathbf{D}^{-1}=
790: \left(
791: \begin{matrix}
792: \mathbf{M}_{\mathbf{v}\mathbf{v}} &
793: \mathbf{M}_{\mathbf{v}\mathbf{w}} \\
794: \mathbf{M}_{\mathbf{w}\mathbf{v}} &
795: \mathbf{M}_{\mathbf{w}\mathbf{w}} &
796: \end{matrix}
797: \right),
798: \end{align}
799: estimates for
800: $\overline{\mathbf{v}}$ and $\mathbf{C}$ can be obtained as
801: \begin{subequations}
802: \label{projection}
803: \begin{align}
804: \overline{\mathbf{v}}=&\overline{\mathbf{u}}_\mathbf{v}+
805: \mathbf{M}_{\mathbf{v}\mathbf{v}}^{-1}\,\mathbf{M}_{\mathbf{v}\mathbf{w}}\,
806: \left( \overline{\mathbf{u}}_\mathbf{w}-\mathbf{w_\text{e}}
807: \right),\\
808: \mathbf{C}=&\mathbf{M}_{\mathbf{v}\mathbf{v}}^{-1}.
809: \end{align}
810: \end{subequations}
811: The linear projection~(\ref{projection}) is straightforwardly verified
812: to be exact when $\mathbf{u}$ and, as a result, $\mathbf{v}$ have
813: multivariate Gaussian distributions and \hbox{$N\to\infty$}. Maximizing the
814: corresponding likelihood function for the empirical statistics
815: conditional to $S=S_\text{e}$, $C=C_\text{e}$,
816: \begin{align}
817: \label{lh}
818: \frac{1}{\sqrt{(2\pi)^{12}
819: \left|
820: \mathbf{C}
821: \right|}}\exp\left[-\frac{1}{2}
822: (\mathbf{v}_\text{e}-\overline{\mathbf{v}})^T \mathbf{C}^{-1}
823: (\mathbf{v}_\text{e}-\overline{\mathbf{v}})\right],
824: \end{align}
825: is equivalent to minimizing
826: \begin{align}
827: \label{chi-det}
828: \chi^2+\mathop{\mathrm{ln}}
829: \left|
830: \mathbf{C}
831: \right|
832: \end{align}
833: with $\chi^2$ as given by Eq.~(\ref{chi2}) in the main text. The
834: maximum likelihood estimates for the model parameters $r_+$,
835: $\lambda$, and $\beta$ were obtained by minimizing~(\ref{chi-det})
836: using the global optimization algorithm of
837: \citet{schonlau98:_global_versus_local_searc_const}.
838:
839: %% Schonlau, Welch, and Jones (see Schonlau, M., Welch, W.,
840: %% Jones, D., ``Global Versus Local Search in Constrained Optimization of
841: %% Computer Models,'' in \textit{New Developments and Applications in
842: %% Experimental Design}, N.~Flournoy, W.~F.~Rosenberger, and W.~K.~Wong
843: %% (ed.), Inst.\ of Math.\ Stat., Hayward, California, Vol.\ 34, 11-25,
844: %% 1998).
845:
846: When the number of samples $N$ used for estimating
847: $\overline{\mathbf{v}}$ and $\mathbf{C}$ in the formula for $\chi^2$
848: (main text) is small, statistical errors of these estimates broaden
849: the probability distribution of $\chi^2$ beyond the usual
850: $\chi^2$-distribution. In the simplest case $\chi^2$ (precisely
851: $N\chi^2$) then follows Hotelling's $T^2$ distribution, for which
852: analytic expressions are known. We decided to use the infinite
853: sample-size confidence intervals (in $95\%$ of all cases $\chi^2<16.9$
854: for $9$ DOF and $<61.7$ for $9\times 5$ DOF) instead of the finite
855: sample size intervals for $N=1000$ ($<17.1$ and $<65.1$).
856: %
857: This is the more conservative choice and, since our procedure for
858: estimating $\overline{\mathbf{v}}$ and $\mathbf{C}$ is different from
859: the one assumed for Hotelling's $T^2$ distribution, it is not clear if
860: these finite-sample-size corrections would still apply.
861:
862: \bibliographystyle{elsart-harv}
863: \bibliography{/home/axel/bib/bibview}
864:
865: \end{nowordcount}
866:
867:
868: \begin{onlysubmission}
869: \newpage{}
870: \end{onlysubmission}
871: \begin{onlypreprint}
872: \thispagestyle{empty} \pagestyle{empty}
873: \end{onlypreprint}
874:
875: % fixme: No single legend should be longer than 200 words. (read
876: % detailed info)
877:
878: \newpage
879:
880: \begin{table*}[h]
881: % \sf \small
882: \begin{threeparttable}
883: \caption{Empirical food webs, fitting parameters, and goodness of fit}
884: % \smallskip
885: \begin{tabular}{llrrrrrrrrrr}
886: Food Web & key &\multicolumn{1}{c}{ $S_\text{e}$\tnote{a}} & \multicolumn{1}{c}{$C_\text{e}\tnote{a}$} & \multicolumn{1}{c}{$\displaystyle\frac{r_+}{r_-}$} & \multicolumn{1}{c}{$r_1$} & \multicolumn{1}{c}{$\lambda$ }&\multicolumn{1}{c}{ $C_0$} &
887: \multicolumn{1}{c}{$\beta$} & \multicolumn{1}{c}{ $\chi^2_\text{N}$\tnote{b}} & \multicolumn{1}{c}{ $\chi^2_\text{S}\tnote{c}$} & \multicolumn{1}{c}{ $\chi^2_\text{H}\tnote{d}$} \\
888: \hline
889: \multicolumn{10}{r}{{%\small
890: number of statistical DOF:} 11\tnote{e}} & 9 & 12 \\
891: \hline
892: Bridge Brook Lake & BB& 15 & 0.28 & 0.91 & 0.17 & 0.12\phantom{5} & 0.37 & 0.059 & 43 & 13.9 & 44 \\
893: \multicolumn{12}{l}{\citep{havens92:_scale_webs}}\\
894: Skipwith Pond &Sk& 25 & 0.32 & 0.93 & 0.21 & 0.009 & 0.53 & 0.012 & 83 & 11.4 & 64 \\
895: \multicolumn{12}{l}{\citep{warren89:_spatial_freshw_web}}\\
896: Coachella Desert &Co& 27 & 0.34 & 0.96 & 0.13 & 0.006 & 0.58 & 0.014 & 37 & 3.9 & 152 \\
897: \multicolumn{12}{l}{\citep{polis91:_compl_des_web}}\\
898: Chesapeake Bay &Ch & 27 & 0.08 & 0.96 & 0.21 & 0.25\phantom{3} & 0.06 & 0.029 & 14 & 7.5 & 23 \\
899: \multicolumn{12}{l}{\citep{baird89:_chesap_bay}}\\
900: St.~Martin Island & SM& 35 & 0.14 & 0.80 & 0.92 & 0.000 & 0.23 & 0.034 & 17 & 13.8 & 12 \\
901: \multicolumn{12}{l}{\citep{goldwasser93:_const_carib_web}}\\
902: Ythan Estuary &Yth& 78 & 0.06 & 0.95 & 0.67 & 0.001 & 0.08 & 0.040 & 58 & 36.8 & 82 \\
903: \multicolumn{12}{l}{\citep{hall91:_food_rich_web}}\\
904: Little Rock Lake & LR& 80 & 0.15 & 0.99 & 0.13 & 0.025 & 0.16 & 0.006 & 52 & 24.2 & 165 \\
905: \multicolumn{12}{l}{\citep{martinez91:_artif_attr}}\\
906: \end{tabular}
907: \begin{tablenotes}
908: \item [a] after species lumping (Appendix~\ref{sec:properties})
909: \item [b] niche model \citep{williams00:_simpl}
910: \item [c] speciation model (this work)
911: \item [d] nested hierarchy model \citep{cattin04:_phylog}
912: \item [e] the statistic \textit{Ddiet}
913: (Appendix~\ref{sec:properties}), which is always zero for the
914: niche model, was excluded
915: \end{tablenotes}
916: \end{threeparttable}
917: \label{tab:parameters}
918: \end{table*}
919:
920: \begin{table}[h]
921: \caption{Model parameters and their theoretical range}
922: \begin{tabular}{lcc}
923: model parameter & & range\\
924: \hline
925: rate constant for new adaptations & $r_1$ & $r_1>0$ \\
926: rate constant for speciations& $r_+$ & $r_+\ge 0$\\
927: rate constant for extinctions& $r_-$ & $r_- > r_+$\\
928: total range of evolution rates & $R$ & $R \ge 0$\\
929: rate/size dispersion constant & $D$ & $D\ge0$ \\
930: loopiness & $\lambda$ & $0\le \lambda \le 1$ \\
931: raw connectivity & $C_0$ & $0\le C_0 \le 1$ \\
932: re-connection probability & $\beta$ & $0\le \beta \le 1$
933: \end{tabular}\centering
934: \label{tab:par-list}
935: \end{table}
936:
937: \clearpage{}
938:
939: \newpage{}
940:
941: \begin{figure}[h]
942: \centering
943: \begin{onlypreprint}
944: \includegraphics[width=0.7\columnwith,keepaspectratio]{newSpeciation2}
945: \end{onlypreprint}
946: \begin{onlysubmission}
947: \includegraphics[width=\textwidth,keepaspectratio]{newSpeciation2}\\
948: \texttt{[newSpeciation2.eps]}
949: \end{onlysubmission}
950: %\begin{nowordcount}
951: \caption{\baselineskip24pt
952: Illustration of the speciation model\newline The food web is
953: represented by a connection matrix (large square). Horizontal
954: lines represent species as resources, vertical lines the same set of
955: species as consumers. Trophic links are indicated by circles at
956: the intersection points. The $s$-axis marks the evolution rate of
957: each species. The process shown is a speciation. Species $j$
958: evolves from species $i$. Most trophic links of $j$ are copied
959: from $i$, but some are modified.}
960: \label{fig:speciation}
961: %\end{nowordcount}
962: \end{figure}
963:
964: \begin{figure}[h]
965: \centering
966: \begin{onlypreprint}
967: \includegraphics[width=0.9\textwidth,keepaspectratio]{samples}
968: \end{onlypreprint}
969: \begin{onlysubmission}
970: \includegraphics[width=0.8\textwidth,keepaspectratio]{samples}\\
971: \texttt{[samples.eps]}
972: \end{onlysubmission}
973: \caption{Examples for connection matrices of food webs\newline
974: The connection matrix of the food webs of Little Rock Lake
975: \citep{martinez91:_artif_attr} and a simulation sample, both after
976: species lumping (see Appendix~\ref{sec:properties} for lumping
977: procedure). Model parameters are as in
978: Table~\ref{tab:parameters}, Little Rock. Each 'X' indicates that
979: the species corresponding to the column eats the species
980: corresponding to the row. The ordering of the species is such as
981: to minimize the number of upper diagonal links and otherwise
982: random.}
983: \label{fig:matrix}
984: \end{figure}
985: %% \begin{figure}[h]
986: %% \centering
987: %% \begin{onlypreprint}
988: %% \includegraphics[width=0.8\columnwith,keepaspectratio]{RYsmall}
989: %% \end{onlypreprint}
990: %% \begin{onlysubmission}
991: %% \includegraphics[width=3cm,keepaspectratio]{RYsmall}
992: %% \end{onlysubmission}
993: %% \begin{nowordcount}
994: %% \caption{\small \sf Cumulative degree distributions for
995: %% Chesapeake Bay (after species lumping). Shaded areas indicate the
996: %% range within one standard deviation of the speciation model
997: %% results.}
998: %% \label{fig:degree}
999: %% \end{nowordcount}
1000: %% \end{figure}
1001:
1002: \begin{figure*}[h]
1003: \centering
1004: \begin{onlypreprint}
1005: \includegraphics[width=\textwidth,keepaspectratio]{graph}
1006: \end{onlypreprint}
1007: \begin{onlysubmission}
1008: \includegraphics[width=\textwidth,keepaspectratio]{graph}\\
1009: \texttt{[graph.eps]}
1010: \end{onlysubmission}
1011: % \begin{nowordcount}
1012: \caption{\baselineskip24pt
1013: Comparison of model and empirical data\newline The graphs
1014: display the values of twelve food-web statistics
1015: (Appendix~\ref{sec:properties}) of seven food webs (key in
1016: Table~\ref{tab:parameters}) obtained from the speciation model
1017: (red stars), in comparison with the empirical data (horizontal
1018: lines), the niche model \citep{williams00:_simpl} (blue
1019: squares), and the nested hierarchy model
1020: \citep{cattin04:_phylog} (blue circles). Vertical lines
1021: indicate model standard deviations.}\
1022: \label{fig:properties}
1023: % \end{nowordcount}
1024: \end{figure*}
1025:
1026: \clearpage
1027:
1028: \end{document}
1029:
1030: %%% Local Variables:
1031: %%% mode: latex
1032: %%% mode: flyspell
1033: %%% TeX-master: t
1034: %%% End:
1035:
1036: % LocalWords: garlaschelli univer milo networ camacho robus struc cohen commun
1037: % LocalWords: williams simpl cattin phylog speciation phylogensis Poissonian
1038: % LocalWords: Matsuda Itoh mccann newman funct compl montoya topol assemb evol
1039: % LocalWords: sneppen extrem slaniana GenSD VulSD MxSim Clust Ddiet aChnSD DOF
1040: % LocalWords: aChnNo aLoop aOmniv Cannib aChnLg paleontologic Neo Jaccard Shor
1041: % LocalWords: lrrrrrrrrrr Skipwith Sk Coachella Ythan Yth dt ds speciates ij
1042: % LocalWords: biologial llrrrrrrrrrr freshw martinez artif attr const McKane
1043: % LocalWords: goldwasser carib chesap dynam evolv dorogovtsev evolut loopiness
1044: % LocalWords: genericity genericities trophically autotrophs Eq schoener ln
1045: % LocalWords: jaccard nouvel florale omnivory schonlau searc Hotelling's baird
1046: % LocalWords: lcc newSpeciation eps
1047: