- Go Mobile »
- Access by Staats- und Universitaetsbibliothek Bremen
Statistical mechanics of ecological systems: Neutral theory and beyond
Rev. Mod. Phys. 88, 035003 – Published 26 July, 2016
DOI: https://doi.org/10.1103/RevModPhys.88.035003
Abstract
The simplest theories often have much merit and many limitations, and, in this vein, the value of neutral theory (NT) of biodiversity has been the subject of much debate over the past 15 years. NT was proposed at the turn of the century by Stephen Hubbell to explain several patterns observed in the organization of ecosystems. Among ecologists, it had a polarizing effect: There were a few ecologists who were enthusiastic, and there were a larger number who firmly opposed it. Physicists and mathematicians, instead, welcomed the theory with excitement. Indeed, NT spawned several theoretical studies that attempted to explain empirical data and predicted trends of quantities that had not yet been studied. While there are a few reviews of NT oriented toward ecologists, the goal here is to review the quantitative aspects of NT and its extensions for physicists who are interested in learning what NT is, what its successes are, and what important problems remain unresolved. Furthermore, this review could also be of interest to theoretical ecologists because many potentially interesting results are buried in the vast NT literature. It is proposed to make these more accessible by extracting them and presenting them in a logical fashion. The focus of this review is broader than NT: new, more recent approaches for studying ecological systems and how one might introduce realistic non-neutral models are also discussed.
Physics Subject Headings (PhySH)
Article Text
“It is interesting to contemplate an entangled bank, clothed with many plants of many kinds, with birds singing on the bushes, with various insects flitting about, and with worms crawling through the damp earth, and to reflect that these elaborately constructed forms, so different from each other in so complex a manner, have been all produced by laws acting around us.” In this celebrated text from the Origin of Species (), Darwin eloquently conveys his amazement for the underlying laws of nature: despite the striking diversity of shapes and forms, it exhibits deep commonalities that have emerged over wide scales of space, time, and organizational complexity. For more than 50 years now, ecologists have collected census data for several ecosystems around the world from diverse communities such as tropical forests, coral reefs, plankton, etc. However, despite the contrasting biological and environmental conditions in these ecological communities, some macroecological patterns can be detected that reflect strikingly similar characteristics in very different communities (see Table ). This suggests that there are ecological mechanisms that are insensitive to the details of the systems and that can structure general patterns. Although the biological properties of individual species and their interactions retain their importance in many respects, it is likely that the processes that generate such macroecological patterns are common to a variety of ecosystems and they can therefore be considered to be universal. The question then is to understand how these patterns arise from just a few simple key features shared by all ecosystems. Contrary to inanimate matter, living organisms adapt and evolve through the key elements of inheritance, mutation, and selection.
This fascinating intellectual challenge fits perfectly into the way physicists approach scientific problems and their style of inquiry. Statistical physics and thermodynamics have taught us an important lesson that not all microscopic ingredients are equally important if a macroscopic description is all one desires. Consider, for example, a simple system like a gas. In the case of an ideal gas, the assumptions are that the molecules behave as pointlike particles that do not interact and that only exchange energy with the walls of the container in which they are kept at a given temperature. Despite its vast simplifications, the theory yields amazingly accurate predictions of a multitude of phenomena, at least in a low-density regime and/or at not too low temperatures. Just as statistical mechanics provides a framework to relate the microscopic properties of individual atoms and molecules to the macroscopic or bulk properties of materials, ecology needs a theory to relate key biological properties at the individual scale, with macroecological properties at the community scale. Nevertheless, this step is more than a mere generalization of the standard statistical mechanics approach. Indeed, in contrast to inanimate matter, for which particles have a given identity with known interactions that are always at play, in ecosystems we deal with entities that evolve, mutate, and change, and that can turn on or off as well as tune their interactions with partners. Thus the problem at the core of the statistical physics of ecological systems is to identify the key elements one needs to incorporate in models in order to reproduce the known emergent patterns and eventually discover new ones.
Historically, the first models defining the dynamics of interacting ecological species were those of Lotka and Volterra, which describe asymmetrical interactions between predator-prey or resource-consumers systems. The Lotka and Volterra equations have provided much theoretical guidance. For instance, developed a model for studying interactions among consumers which exploit common resources. By making use of different time scales, he showed how resources can be included in the Lotka and Volterra equations. He also derived the formulas for the competition matrix, which provided hints about how it can be measured empirically (). Seemingly, this suggested a viable way to measure niche overlap between species. Soon after, studied the problem of how similar competing species can be and yet coexist in a community. According to Gause’s competitive exclusion principle (), two species cannot occupy the same niche in the same environment for a long time (see Table ). They found that environmental fluctuations limit niche overlap and therefore species’ similarity. All these studies, and many other variations and generalizations, provided a robust theoretical basis for understanding ecosystems. Their limitations were identified as more research was carried out (). These approaches share the idea that species can be well described by deterministic models which are shaped by the fundamental concept of niche, which, however, can be appropriately defined only a posteriori. This theoretical focus is not eminently suited for studying a wide range of empirical patterns. In fact, all such models have several drawbacks: (1) They are mostly deterministic models and often do not take into account stochastic effects in the demographic dynamics (). (2) As the number of species in the system increases, they become analytically intractable and computationally expensive. (3) They have a lot of parameters that are difficult to estimate from ecological data or experiments. (4) It is very difficult to draw generalizations that include spatial degrees of freedom. And (5) while time series of abundance are easily analyzed, it remains challenging to analytically study the macroecological patterns they generate and thus, their universal properties.
A pioneering attempt to explain macroecological patterns as a dynamic equilibrium of basic and universal ecological processes—and that also implicitly introduced the concept of neutrality in ecology—was made by in the famous monograph titled “The theory of island biogeography.” In this work, they proposed that the number of species present on an island (and forming a local community) changes as the result of two opposing forces: on the one hand, species not yet present on the island can reach the island from the mainland (where there is a metacommunity); and on the other hand, the species already present on the island may become extinct. MacArthur and Wilson’s model implies a radical departure from the then main current of thought among contemporary ecologists for at least three reasons: (1) Their theory stresses that demographic and environmental stochasticity can play a role in structuring the community as part of the classical principle of competitive exclusion. (2) The number of coexisting species is the result of a dynamic balance between the rates of immigration and extinction. (3) No matter which species contribute to this dynamic balance between immigration and extinction on the island, all the species are treated as identical. Therefore, they introduced a model that is neutral at the level of species (see Table ), even though they did not think of it as such.
Just a few years later, the ecologist H. Caswell proposed a model in which the species in a community are essentially a collection of noninteracting entities and their abundance is driven solely by random migration and immigration. In contrast to the mainstream vision of niche community assembly, where species persist in the community because they adapt to the habitat, Caswell stressed the importance of random dispersal in shaping ecological communities. Although the model was unable to correctly describe the empirical trends observed in a real ecosystem, it is important because it pictured ecosystems as an open system, within which various species have come together by chance, past history, and random diffusion.
In 2001, greatly inspired by the theory of island biogeography and the dispersal limitation concept (see Table ), Hubbell published an influential monograph titled “The Unified Neutral Theory of Biodiversity and Biogeography” [however, the debate dates back to 1979 ()]. Unlike the niche theory and the approach adopted by Lotka and Volterra, the neutral theory (NT) aims to model only species on the same trophic level (monotrophic communities, see Table ), species that therefore compete with each other because they all feed on the same pool of limited resources. For instance, competition arises among plant species in a forest because all of them place demands on similar resources like carbon, light, or nitrate. Other examples include species of corals, bees, hoverflies, butterflies, birds, and so on. The NT is an ecological theory within which organisms of a community have identical per-capita probabilities of giving birth, dying, migrating, and speciating, regardless of the species they belong to. Thus, from an ecological point of view, the originality of Hubbell’s NT lies in the combination of several factors: (i) it assumes competitive equivalence among interacting species; (ii) it is an individual-based stochastic theory founded on mechanistic assumptions about the processes controlling the origin and interaction of biological populations at the individual level (i.e., speciation, birth, death, and migration); (iii) it can be formulated as a dispersal limited sampling theory; and (iv) it is able to describe several macroecological patterns through just a few fundamental ecological processes, such as birth, death, and migration (; ; ; ). Although the theory has been highly criticized by many ecologists as being unrealistic (; ; ; ; ; ), it does provide very good results when describing observed ecological patterns and it is simple enough to allow analytical treatment (, , ; ; ). However, such precision does not necessarily imply that communities are truly neutral and indeed non-neutral models can also produce similar patterns (; ). Yet the NT does call into question approaches that are either more complex or equally unrealistic (; ). Moreover, NT is not only a useful tool to reveal universal patterns but also it is a framework that provides valuable information when it fails. One of the strengths of NT is that one can, in fact, falsify one or more of its assumptions and thereby actually test the theory. Few models in community ecology meet this gold standard. These features have made NT an important approach in the study of biodiversity (; ; ; ; ; ; ; ).
From a physicist’s perspective, NT is appealing as it represents a “thermodynamic” theory of ecosystems. Similar to the kinetic theory of ideal gases in physics, NT is a basic theory that provides the essential ingredients to further explore theories that involve more complex assumptions. Indeed, NT captures the fundamental approach of physicists, which can be summarized by Einstein’s celebrated quote “Make everything as simple as possible, but not simpler.” Finally, it should be noted that the NT of biodiversity is basically the analog of the theory of neutral evolution in population genetics () and indeed several results obtained in population genetics can be mapped to the corresponding ecological case ().
Statistical physics is contributing decisively to our understanding of biological and ecological systems by providing powerful theoretical tools and innovative steps (, ) to understand empirical data about emerging patterns of biodiversity. The aim of this review is not to present a complete and exhaustive summary of all the contributions to this field in recent years—a goal that would be almost impossible in such an active and broad interdisciplinary field—but rather, we want to introduce this exciting new field to physicists that have no background in ecology and yet are interested in learning about NT. Thus, we focus on what has already been done and what issues must be addressed most urgently in this nascent field, that of statistical physics applied to ecological systems. A nice feature of this field is the availability of ecological data that can be used to falsify models and highlight their limitations. At the same time, we see how the development of a quantitative theoretical framework will enable one to better understand the multiplicity of empirical experiments and ecological data.
This review is organized into six main sections. Section is a review of several important results that have been obtained by solving neutral models at stationarity. In particular, we present the theoretical framework based on Markovian assumptions to model ecological communities, where different models may be seen as the results of different NT ensembles. We also show how NT, despite its simplicity, can describe patterns observed in real ecosystems. In Sec. , we present more recent results on dynamic quantities related to NT. In particular, we discuss the continuum limit approximation of the discrete Markovian framework, paying special attention to boundary conditions, a subtle aspect of the time-dependent solution of the NT. In Sec. , we provide examples of how space plays an essential role in shaping the organization of an ecosystem. We discuss both phenomenological and spatially implicit and explicit NT models. A final subsection is devoted to the modeling of environmental fragmentation and habitat loss. In Secs. and we propose some emerging topics in this fledgling field and present the problems currently being faced. Finally, we close the review with a section dedicated to conclusions.
Neutral theory deals with ecological communities within a single trophic level, i.e., communities whose species compete for the same pool of resources (see Table ). This means that neutral models will generally be tested on data describing species that occupy the same position in the food chain, like trees in a forest, breeding birds in a given region, butterflies in a landscape, plankton, etc. Therefore, ecological food webs with predator-prey-type interactions are not suitable to be studied with standard neutral models.
As explained in the Introduction, ecologists have been studying an array of biodiversity descriptors over the last 60 years (see Table ), including RSA, SAR, and spatial PCF (see Fig. ). For instance, the RSA represents one of the most commonly used static measures to summarize information on ecosystem diversity. The analysis of this pattern reveals that the RSA distributions in tropical forests share similar shapes, regardless of the type of ecosystem, geographical location, or the details of species interactions (see Fig. ). Therefore, the functional form of the RSA [see seminal papers by and for a theoretical explanation of its origins] has been one of the great problems studied by ecologists. Indeed, much attention has been devoted to the precise functional forms of these patterns. A meta-analysis revealed that RSA distributions have basically three shapes: more often unimodal (log-normal like) in fully sampled communities, and either power law or without any mode (log-series like) within incompletely censused regions (). Also, it is difficult to tease apart, from the RSA alone, the nature of the basic processes driving communities (), because neutral and non-neutral mechanisms coexist in nature.
Visual scheme of two important macroecological patterns (see Table ): RSA and SAR. The functional shape of the RSA depends on the spatial scale considered, while the SAR generally displays a triphasic behavior (see Sec. ). There is a growing appreciation that the various descriptors of biodiversity are intrinsically interrelated, and substantial efforts have been devoted to understand the links between them. From .
Tree relative species-abundance data from the Barro Colorado Island (BCI), Yasuni, Pasoh, Lambir, Korup, and Sinharaja plots for trees that are 10 cm in stem diameter at breast height. The frequency distributions are plotted using Preston’s binning method as described by and the bars are the observed number of species binned into
It is instructive to derive some of these functional forms by starting with a very simple but extreme neutral model, which assumes that species are independent and randomly distributed in space. This null model tells us what we should expect when the observed macroecological patterns are driven only by randomness, with no underlying ecological mechanisms. If the density of individuals in a very large region is
The SAR curve can also be calculated within a slightly more accurate model, which still assumes that species are independent and randomly situated in space. Let us now suppose that a region with area
Although this model was originally studied by , we now know that it significantly overestimates species diversity at almost all spatial scales ().
Beta diversity (see Table ) can be estimated under the assumptions we have mentioned. Now, regardless of the spatial distance between two individuals, the probability that two of them belong to the same species
The failure of the random placement model to capture the RSA, SAR, and beta diversity is a clear indication that ecological patterns are driven by nontrivial mechanisms that need to be appropriately identified. Thus, we shall assess to what extent the NT at stationarity can provide predictions in agreement with empirical data.
There are two related but distinct analytical frameworks that have been used to mathematically formulate the NT of biodiversity at stationarity for both local and metacommunities: the first fixes the total population, whereas the second fixes the average total population of a community. From an ecological point of view, a local community is defined as a group of potentially interacting species sharing the same environment and resources. Mathematically, when modeling a local community the total community population abundance remains fixed. Alternatively, a metacommunity can be considered a set of interacting communities that are linked by dispersal and migration phenomena. In this case, it is the average total abundance of the whole metacommunity that is held constant. From the physical point of view, these roughly correspond to the microcanonical (fixed total abundance) and the grand-canonical (fixed average total abundance) ensembles, respectively. The microcanonical ensemble or so-called zero-sum dynamics when death and birth events always occur as a pair originates from the sampling frameworks in population genetics pioneered by Warren Ewens and Ronald Fisher (). Note that even though a fixed size sample is one way to analyze available data, for the majority of cases (apart from very small size samples) the grand-canonical ensemble approach is that used routinely in statistical physics and it provides a precise yet largely simplified description of the system. The ultimate reason for this lies in the surprising accuracy of the asymptotic expansion of the gamma function (the mathematical framework heavily uses combinatorials and factorials, etc.). The Stirling approximation can be used for very large values of the gamma function; nevertheless, it is quite accurate even for values of the arguments of the order of 20. The advantages of the master equation (ME) (see Sec. ) and the grand-canonical ensemble approach stem from their computational simplicity, which make the results more intuitively transparent.
We now introduce the mathematical tools of stochastic processes that will be used extensively in the rest of the article. Let Assuming that the stochastic dynamics are Markovian, the time evolution of where Equation is typically intractable with analytical tools, because it involves a sum of all the configurations. If each term in the summation is zero, i.e., detailed balance is said to hold. A necessary and sufficient condition for the validity of detailed balance is that for all possible cycles in the configuration space, the probability of walking through it in one direction is equal to the probability of walking through it in the opposite direction. Given a cycle This condition evidently corresponds to a time-reversible condition. Now let us apply these mathematical tools to the study of community dynamics that is driven by random demographical drift. Consider a well-mixed local community. This is equivalent to saying that the distribution of species in space is not relevant, which should hold for an ecosystem with a linear size smaller than or of the same order as the seed dispersal range. In this case one can use In this case, where is the birth rate, and the death rate. This particular choice corresponds to a sort of mean-field (ME) approach (; , ; ; ; ; ). Our many-body ecological problem can also be formulated in a language more familiar to statistical physicists, where we consider the distribution of balls into urns. The “urns” are the species and the “balls” are the individuals. Birth and death processes correspond to adding or removing a ball to or from one of the urns using a rule as dictated by Eqs. and . Equation with the choice can be simplified if one assumes that the initial condition is factorized as The stationary solution of Eq. is easily seen to satisfy the detailed balance (; ). First we note that because of the neutrality hypothesis species are assumed to be demographically identical and therefore, we can drop the where When there are Depending on the functional form of The random walk and Bose-Einstein distribution.: If one chooses where The Fermi-Dirac distribution: If Boltzmann counting: If This is the familiar grand-canonical ensemble Boltzmann counting in physics, where Density independent dynamics: We now consider the dynamic rules of birth, death, and speciation that govern the population of an individual species. The most simple ecologically meaningful case is to consider where The RSA This is the celebrated Fisher log-series distribution, i.e., the distribution Fisher proposed as that describing the empirical RSA in real ecosystems (). The parameter In 1948, F. W. Preston, published a paper () challenging Fisher’s point of view. He showed that the log series is not a good description for the data from a large sample of birds. In fact, he observed an internal mode in the RSA that was absent in a log-series distribution. In particular, Preston introduced a way to plot the experimental RSA data by octave abundance classes (i.e., [ Different neutral models of community ecology. In all these models, There are several ecological meaningful mechanisms that can generate a bell-shaped Preston-like RSA. The first of these involves density-dependent effects on birth and death rates. The second involves considering a Fisher log series as the RSA of a metacommunity acting as a source of immigrants to a local community embedded within it. The dynamics of the local community are governed by births, deaths, and immigration, whereas the metacommunity is characterized by births, deaths, and speciation. This leads to a local community RSA with an internal mode (; ; , ). A third way is to incrementally aggregate several local communities (see Appendix ). Local dynamics with density-dependent rates: One of the most fundamental and long debated questions in community ecology is to understand what maintains species diversity within communities (; ). Mechanisms which are able to promote diversity include competitive trade-offs among species because of species-specific traits, balance between speciation and extinction, frequency or density dependence, and environmental variability. For instance, processes that hold the abundance of a common species in check inevitably lead to rare-species advantages, given that the space or resources freed up by density-dependent death can be exploited by less-common species. Therefore, interspecies frequency dependence is the community-level consequence of intraspecies density dependence, and, thus, they may be thought of as two different manifestations of the same phenomenon (). Because these processes are capable of driving community-level patterns such as SAR or RSA, one might hope to identify mechanisms by delving into observed patterns. However, communities with different governing processes can unexpectedly show similar patterns, thus suggesting that their shape cannot be, in general, used to identify specific mechanisms (; ). In this review we focus on two of the most prominent hypotheses that explain species coexistence through frequency and density dependence: the Janzen-Connell (; ) and the Chesson-Warner hypotheses (). These mechanisms generally predict the reproductive advantage of a rare species due to ecological factors and they can be readily captured in a common mathematical framework that is presented later. The Janzen-Connell hypothesis postulates that host-specific pathogens or predators act in the vicinity of the maternal parent. Thus, seeds that disperse farther away from the mother are more likely to escape mortality. This spatially structured mortality effect suppresses the uncontrolled population growth of locally abundant species relative to uncommon species, thereby producing a reproductive advantage to a rare species. The Chesson-Warner storage hypothesis explores the consequences of a variable external environment and it relies on three empirically validated observations: species respond in a species-specific manner to the fluctuating environment, there is a covariance between the environment and intraspecies and interspecies competition, or life history stages buffer the growth of population against unfavorable conditions. Such conditions prevail when species have similar per-capita rates of mortality but they reproduce asynchronously and there are overlapping generations. We begin by noting that the mean number of species with The density dependence arising from the Janzen-Connell effect is spatially explicit, in plant communities at least, because it is caused by interactions among neighboring individuals. Here instead we formulate a density dependence which is spatially implicit and only abundance driven. We therefore introduce a modified symmetric theory that incorporates a rare-species advantage or common species disadvantage by making and for and where Strikingly, any relative abundance data can be considered as arising from effective density-dependent processes in which the birth and death rates are given by these expressions. Thus, one would expect that the per-capita birth rate or fecundity drops as the abundance increases, whereas mortality ought to increase with abundance. Indeed, the per-capita death rate can be arranged to be an increasing function of The mathematical formulation of the density dependence may seem unusual to ecologists familiar with the logistic or Lotka-Volterra systems of equations, wherein the density dependence is typically described as a polynomial expansion of powers of one readily arrives at the following relative species-abundance relationship: where Various analytical solutions for the Markovian (ME) approach have been suggested in the literature. Based on pioneering work by , an analytical solution was proposed for the ME when the size of both the local community and the metacommunity was fixed (; ; ). Here we refer to the work of , in whose model the birth and death rates in the metacommunity are given by During a unit of time, the population ( Let where Therefore the density of species of relative abundance where Let us now consider a local community (of size where The stationary solution where Finally, the average number of species with abundance which displays an internal mode for appropriate values of the model parameters. A generalization of this result to a system of two communities of arbitrary yet fixed sizes that are subject to both speciation and migration () has also been carried out. suggested that the log-series solution for the species-abundance distribution in the metacommunity is applicable only for species-rich communities, and that it does not adequately describe species-poor metacommunities. Application of the ME approach to the asymmetric species case was considered () and it was further demonstrated that the species-abundance distribution has exactly the same sampling formula for both zero-sum and nonzero-sum models within the neutral approximation (; ). The simplest mode of speciation, a point mutation, has been the one most commonly used to derive the species-abundance distribution of the metacommunity. A Markovian approach incorporating various modes of speciation such as random fission was also presented by , . An innovative way to model speciation was used by . They introduced protracted speciation, i.e., a gradual process whereby new species are created with a few individuals, instead of being an instantaneous process starting with exactly one individual. This gradual speciation improves predictions of species lifetimes, speciation rates, and the number of rare species. An alternative approach to compute the distribution of individuals in a community is based on coalescent theory. The idea of this approach consists of tracing each member of the community back to their ancestors that first immigrated into the community (; ). A succinct comparison between the coalescent approach and other methods is presented by and . Within this framework, the local community consists of a fixed number of individuals Two observations are crucial to derive the sampling formula for the species-abundance distribution. First, because there is no speciation in the local community, each individual is either an immigrant or a descendent of an immigrant. Thus, the information pertaining to the species-abundance distribution is specified by considering “the ancestral tree” (tracing back each individual to its immigrant ancestor, which is somewhat similar to the phylogenetic tree construction in genetics) and the species composition of the set of all the ancestors. Second, this set of all ancestors can be considered as a random sample from the metacomunity and its species-abundance distribution is provided by the Ewens’ sampling formula (), corresponding to the Fisher’s log series in the limit of large sample size that describes neutral population genetics models with speciation and no immigration. As we do not know the ancestor of each individual, the final formula for the species-abundance composition of the data comes from summing the probabilities of all possible combinations of ancestors that give rise to the observed data. In a tour de force calculation, the resulting sampling formula is found to be () where where These expressions are simplified versions of the original results presented and as expected they provide a comparable fit to the ME approach for the data from the tropical forests. It is heartening that equivalent results can be derived using very different approaches.
So far we have focused on the implications of neutral theory when models describing neutral patterns reach a steady state. The stationary condition allows one to take advantage of a variety of different mathematical techniques to obtain analytical expressions of ecological patterns. However, stationarity is not always a good approximation, either because the ecosystem is still in a state of flux or because the assumption may hide different and important processes that lead to the same final steady state. Furthermore, one can calculate time-dependent correlation functions (); see, for example, Eq. . It is therefore essential to understand the temporal behavior of ecosystems in order to discriminate between ecological processes that would otherwise be indistinguishable at stationarity. Statistical comparisons between time-dependent patterns are usually more difficult than those between stationary patterns because they require more data and long empirical time series that are rarely available. In addition, although time-dependent solutions facilitate stronger tests when confronted with data, they are more difficult to obtain and this is the reason why only a few studies have investigated the temporal behavior of neutral models.
An important method to study the time dependence is van Kampen’s system size expansion. Assuming that the total population of the system
Some neutral models can also be formulated in terms of discrete-time Markov processes that have the form
Others have focused on specific temporal patterns that can sometimes be calculated more easily. For instance, a measure that has been used to quantify the diversity of a community is the Simpson index ()
Although it is not mandatory from the neutral assumption, models often assume that demographic stochasticity is the main source of fluctuations in stochastic neutral dynamics (; ). Demographic randomness originates from the intrinsic stochastic nature of birth and death events within a discrete population of individuals. However, other sources may be important, such as environmental stochasticity that, by contrast, encompasses effects of abiotic and biotic environmental variables on all individuals [strictly speaking, we should more correctly refer to symmetric, instead of neutral, models (see Table ) when these incorporate environmental fluctuations, because the per-capita vital rates vary across individuals; however, many authors do not appreciate this distinction]. There is theoretical and empirical evidence that forest dynamics exhibit signatures of environmental variability (; ; ). Usually, neutral models based only on demographic stochasticity tend to overestimate the expected times to extinction for abundant species (), whose temporal fluctuations will also be underestimated (). Incorporating an environmental source of randomness as well as more realistic forms of speciation have made some dynamic aspects of NT more realistic ().
Here we focus on the dynamic aspects of the SAD () which under appropriate assumptions can be studied in detail (; ) and whose predictions can be benchmarked against empirical data. We also review recent progress in modeling dynamic patterns, including the species turnover distribution () as well as species’ persistence-times (or lifetimes) distributions (; ; ; ); see Table . The microscopic description of a system starts by correctly identifying the variables that define all the possible configurations of the system. Having decided how to describe the states of the system, the next step is to consider the transition rates among different states. The configurations of an ecological community can be described by different variables according to different levels of coarsening of the spatiotemporal scales. However, for simplicity, we focus on ecosystems comprising If the ecological community at time Assuming that where where the time has been rescaled ( and the entries of the matrix The matrix whose solution, for initial condition So far we have assumed that the system has a fixed and finite total population where time has not been rescaled and Starting from the discrete formulation of the birth and death rates that we have seen previously, one can derive the expressions of the corresponding continuous rates. Following our earlier discussion, we can write the following general expansion: where The link established between the FP equation and the ME provides a useful interpretation of the coefficients: As expected, different choices of Further insights into the nature of stochasticity can be achieved by writing down Eq. in the equivalent Langevin form, which is an equation for the state variables themselves. Within the Itô prescription (), Eq. is equivalent to the following Langevin equation: where For the present model, the Fokker-Planck equation is equivalent to where we have instead assumed that the time correlation is The stochastic dynamics described by Eq. governs the evolution of the population when it is strictly positive. Assuming that • We use reflecting boundary conditions at • A second possibility arises when a community can lose individuals without any replacement or a net emigration flux of individuals from the ecosystem exists. One can then describe the system by introducing absorbing boundary conditions at It is possible to define other kinds of boundaries, according to different ecological behaviors when the population consists of just a few individuals. However, the reflecting and absorbing boundaries capture the most interesting situations in ecology. As we have alluded, the reflecting and absorbing solutions of the Fokker-Planck equation have rather unusual properties. This is essentially due to the fact that the diffusion term An ecosystem described by Eq. can reach a nontrivial steady state when setting reflecting boundary conditions at where One can analytically solve Eq. at any time with arbitrary initial conditions, so that the evolution of the population can be traced even far from stationary conditions. The time-dependent solution with reflecting boundary conditions ( where From the Fokker-Planck equation , it is easy to derive the evolution of the mean population per species. It is simply where where According to the NT, the turnover of ecological communities reflects their continuous reassembly through immigration and emigration and local extinction and speciation. Species’ histories overlap by chance due to stochasticity, yet their lifetimes are finite and distributed according to the underlying governing process. Nontrivial stationary communities are reached because old species are continuously replaced by new species, bringing about a turnover of species that can be studied and modeled within our framework. To measure species turnover one usually considers the population of a given species at different times, then studying the temporal evolution of their ratios. For an ecosystem close to stationarity, one can look at the distribution Here In Eq. , one can use the time-dependent reflecting or absorbing solution according to different ecological dynamics. We should use the reflecting solution when not concerned with the extinction of the species present at the initial time point and especially when accounting for any new species introduced through immigration or speciation. The expression for the reflecting STD can be found in . One can show that it has the following power-law asymptotic behavior for a fixed where the functions STD for the interval 1990–1995 in the BCI forest. The main panel shows the results for individuals of more than 10 cm dbh, and the inset the results for individuals of more than 1 cm dbh (Center for Tropical Forest Science website). The black line represents the analytical solution given by Eq. . From . These fits not only provide direct information about the time scale of evolution but also they underline the importance of rare species in the STD. A theoretical framework to study and analyze persistence or extinctions of species in ecosystems allows one to understand the link between environmental changes [like habitat destruction or climate change (; ; ; ; )] and the increasing number of threatened species. The persistence or lifetime The simplest baseline model to study the persistence-time distribution is a random walk in the species abundance A further step in modeling lifetime distributions can be made by taking into account birth, death, and speciation (; ; ; ) through a mean-field scheme of the voter model with speciation (; ) (see description in Sec. ), i.e., ME with birth and death rates given by Eqs. and in the large for We have seen in Sec. that the RSA pattern does not depend on the biological details of the ecosystem under analysis. Thus, one may wonder if the persistence-time distribution is also a universal macroecological pattern. Indeed, it has been shown (; ) that the power law with an exponential cutoff shape predicted for the persistence-time distribution by the NT (see Appendix ) is common to very different types of ecosystems (see Fig. ). Other exact formulas for species ages and species lifetimes have been proposed in neutral () and non-neutral models (; ). Comparison between persistence empirical distributions for North American breeding birds, Kansas grasslands, New Jersey BSS forest, and an estuarine fish community and the corresponding theoretical species persistence-times pdfs. The circles and solid lines show the observational distributions and fits, respectively. The finiteness of the time window Another interesting and related quantity is the survival distribution In the continuum limit, a crucial distribution for the analysis of species’ extinction is the time-dependent solution of Eq. with absorbing boundaries at Note that Eq. is finite at The lifetime distribution can be calculated analytically using Eq. , i.e., which depends only on
So far we have considered models, which assume that all individuals, at a given time, experience essentially the same conditions. They live in well-mixed habitats in which environmental heterogeneity and spatial distance are not important. These models are conceptually and mathematically simple and this is the basic reason why they have been widely used. However, the spatial structure of ecosystems is able to control the shape of many patterns and it is also a critical factor for understanding species’ coexistence (). In fact, models with well-mixed populations predict that the best competitor is able to displace a pool of species competing for the same limiting resource. There is empirical and theoretical evidence that space plays a crucial role in maintaining species diversity in communities with a single limiting resource, as demonstrated in the grasslands of the Cedar Creek Natural History Area (USA) ().
Space is therefore an essential element for understanding the organization of an ecosystem and most empirical observations are spatial. Often, the dynamics and composition of a community cannot be disentangled from its spatial aspects. Unfortunately, it is very difficult to derive analytical predictions for spatial (stochastic) models. The main difficulty is due to the fact that these are out-of-equilibrium models ().
Spatial effects can be incorporated into the theoretical framework, with the ME of Eq. remaining formally the same by considering the index
At present, no coherent spatial neutral theory exists but rather there is a collection of models and techniques that can explain some spatial patterns. In this section, we review some of those approaches.
The relationship between the number of species and the area sampled is probably one of the oldest quantities studied in ecology (). referred to it as “one of community ecology’s few genuine laws.” The SAR is defined as the average number of species
postulated a power-law relationship
Triphasic shape of the species-area relationship. The left panel shows the three behaviors on different scales. At a local scale the relationship is linear, becoming a power-law relationship at the regional scale and returning to linear at very large intercontinental scales. The right panel shows empirical data of species diversity at the regional scale (from the BCI forest). From .
Despite some notable exceptions (), the value of the exponent
Several models tried to reproduce the empirical behavior and a simple but useful assumption involves considering different species as independent realizations of the same process (). It is important to note that this assumption is stronger than neutrality, because neutrality does not imply independence. Under these assumptions the SAR is given by Eq. .
The EAR is defined as the number of species that are completely contained (i.e., endemic) in a given area (see Table ). It is not generally simply related to the SAR (). If the species are considered as independent realizations of a unique process, then
where
The
where
that can be rewritten as
If the spatial positions of different species are independent, then
Most theoretical work consists of attempts to relate these ecological quantities with other spatial and nonspatial observable factors. For instance, a typical problem is to calculate the SAR knowing the Phenomenological models do not assume any microscopic dynamics but they are rather based on a given phenomenological distribution of individuals in space. The simplest assumption is to consider individuals at random positions in space (; ). This null model, usually known as a “random placement” model, can be used to obtain predictions for the SAR and EAR having the RSA or the SAD as an input. Even though this assumption is not realistic, the random placement model turns out to be very useful to capture the relevant aspects of the relationship between RSA and SAR, and it also gives reasonable predictions that can be benchmarked against empirical data. In addition, this assumption also allows direct connections between SAR and EAR to be formulated (). Consider where The simple framework of the random placement model also allows the EAR to be calculated. Using Eq. we obtain Despite the simplicity of the approximation, the EAR evaluated using random placement captures the quantitative behavior of several observed ecosystems (). Under random placement assumptions, one can obtain the EAR from the SAR and vice versa. To calculate the EAR in Eq. , we calculated the number of species with zero individuals in the area complementary to that of interest. This number is equal to the difference between the number of species in the whole area and the number of species in the complementary area. The complementary area has a nontrivial shape and under general assumptions this quantity is not easy to calculate. Under the random placement assumption, the number of species in the complementary area is the SAR of the complementary area. The EAR is therefore presented a careful analysis of the reliability of this extrapolation to predict the empirical EAR and showed that the random placement approximation describes the empirical data well. One can obtain a closed form expression for the EAR and SAR by starting with a RSA distribution. In the case of a Fisher log series [see Eq. ], the SAR reads which follows from the observation that In real ecosystems, individuals are not distributed uniformly in space but rather, due to dispersal limitation, individuals of the same species tend to be clustered. This is confirmed by empirical There are two possible ways to include space in a neutral stochastic model, either implicitly or explicitly. Spatially implicit models are based on the observation that one can relate the sample area In a metacommunity the number of species where This result corresponds exactly to that found using the random placement in Eq. . At small sample sizes, the number of species Real ecosystems are of course spatially explicit, but one might wonder how spatially implicit models or, more generally, models that do not consider space explicitly, are predictive and how their parameters are related to spatially explicit ones. A way to assess this is to measure the efficacy of nonspatial models in predicting the behavior of spatially explicit models (). As expected, nonspatial models have a good predictive power when the dispersal lengths are sufficiently large and they are particularly good in predicting nonspatial patterns such as the RSA. Spatially explicit models are typically defined as birth-death-diffusion processes. A model is fully specified given a ME and can be obtained in several ways, i.e., it is possible to write several different MEs that include space in a neutral model. The ME is not tractable analytically and one has to introduce approximations in order to get analytical results. Spatially explicit models are particularly difficult to solve because of the lack of detailed balance (see Sec. ). The voter model () was originally introduced to describe opinion formation, whereby voters are located in a network and each one has one opinion among Microscopic moves of the multispecies voter model in a 2D lattice and with a nonregular network (). In the 2D structure, each site is occupied by one and only one individual, whose color represents the species it belongs to. At each time step, one random individual is replaced by a daughter of one of its neighbors with probability Consider a lattice of We want to write an equation for where The stationary solution of Eq. can be obtained from a Fourier series in the continuum limit, by taking the limit of where where with It is possible to obtain interesting results on the SAR via extensive numerical simulations. In the case of NT, one can take advantage of the coalescent approach (; ). Instead of simulating the stochastic dynamics directly, one can reconstruct the genealogy of the individuals in the sample area by regressing in time. The main value of this method is that there is no need to wait for any transient state to decay and, therefore, this approach is much faster than forward dynamics () as well as allowing infinite landscapes to be simulated. The coalescent approach has been applied to the MVM with a different dispersal kernel (). Instead of a nearest-neighbor diffusion, once an individual is removed it is replaced by the offspring of another individual with a probability that depends on its distance from the individual removed. In an infinite landscape, the SAR shows the characteristic inverted-S shape. The model depends only on two parameters (up to a choice of the functional form of the dispersal kernel): the speciation rate where the exponent where An attempt to connect spatial and temporal patterns can be found (), where the persistence-time distribution was studied in MVM (see Sec. ). It was shown that the empirical persistence-time distribution is consistent with that predicted by MVM and where Here we assumed that Resources are not equally distributed and, even over small scales, their distribution in space is far from uniform. This spatial heterogeneity affects the distribution of individuals and species in space and has clear implications for the conservation of ecosystems. The space within which ecosystems are embedded is often fragmented and a species may be present in only part of the landscape. Moreover, different patches are not independent but they are rather connected via immigration. In the absence of speciation, the voter model () predicts monodominance in dimension where where a small mutation rate [ NT has also been applied to predict the extinction rate of species after habitat loss (). Habitat loss corresponds to a reduction of the area available and therefore to the total number of individuals. When this area is destroyed, the endemic species suddenly disappear and what follows is a delayed series of extinctions due to habitat loss. The community was modeled as a neutral assembly of species and a typical time scale of extinctions was obtained, along with its dependence on the number of species and the habitat destroyed. The predictions obtained in this way reproduced the available data of avifaunal extinctions well ().
The concept of a niche is central in classical ecology (; ; ). An ecological niche is “the requirements of a species for existence in a given environment and its impacts on that environment” () and it describes how an organism or a population responds to changes in resources, competitors, and predators. A possible mathematical realization of the concept of a niche is the Hutchinsonian niche, which involves a
The mathematical representation of an ecological community that includes niche aspects typically coincides with the Lotka-Volterra equations. In this case, the focus is on the properties of the fixed points (or other dynamical attractors) of these systems of equations and the typical problem that is analyzed is their stability, in relation to the parameters and the species present in the system.
The main difference between neutral and niche theory therefore depends on which mechanism plays the main role in shaping ecosystems (). Neutral theory assumes that random processes, such as dispersal, demographic stochasticity, speciation, and ecological drift, have a stronger impact on many of the observed patterns than niche differences. Niche theory assumes the opposite, that the quantitatively important processes are related to differences in species and their interdependence.
As we might expect in a real ecosystem both stochasticity and niche differences play a role, and it is natural to try to quantify how neutral behavior emerges from a niche model. In many cases, niche-based and neutral models yield compatible fits of biodiversity patterns (; ; ; ), and it is impossible to distinguish between the mechanisms by looking at those patterns. As pointed out by neutrality emerges when species have the same or very similar fitness.
Neutrality has often been proposed to emerge under some conditions from models considering niche differences (; ; ), and neutrality and niche theory were proposed to be the extremes of a more general model (; ). In both cases, a Lotka-Volterra equation is introduced to describe community dynamics, while population dynamics is modeled by also taking into account demographic stochasticity and immigration. By considering different values of parameters, one can move from a scenario where species differences matter a lot and the stable configuration is very close to the solution of the deterministic Lotka-Volterra equation to a scenario where demographic stochasticity is more important and the community behaves like a neutral community. A slightly different approach has also been considered (), analyzing a stochastic version of Lotka-Volterra dynamics and quantifying, when stochasticity is varied, the difference between the prediction of the neutral model with the full model. In this case, instead of a continuum of strategies, neutral and niche regimes are two macroscopic phases separated by a phase transition. Others have also incorporated neutral and non-neutral features within the same framework (; ): species with comparable sizes were considered functionally equivalent, thereby entailing a neutral dynamics; and parameters such as speciation, dispersal, and populations of organisms of distinct sizes were obtained from allometric scaling laws. With this approach, were able to explain why species richness reaches a maximum at intermediate body size.
A different approach was based on effectively considering niche theory as a model where per-capita death and birth rates are species dependent (). In this case, one might expect the difference between species abundances to reflect the differences between these parameters. More precisely, in a neutral scenario, all the species fluctuate around a given abundance value, while when niche characteristics play a role, each species fluctuates around its own distinct value of abundance. A third scenario was proposed (), wherein the per-capita birth and death rates do not have a monotonic effect on species abundance. The symmetry between species, due to the neutrality of the process, can yet be spontaneously broken. In this case, the stable states are not symmetric in terms of species abundance.
So far macroevolutionary patterns such as phylogenetic trees have not been a major focus in NT. However, there is evidence that NT is not able to satisfactorily capture phylogenetic diversity (). Multiple patterns of evolutionary history in bacterial communities seem to deviate from the predictions of NT (). However, recent models which incorporate either neutral population dynamics into the cladogenesis () or a mild selection into an otherwise neutral model () are able to slow down the diversification process as well as match phylogenies observed in nature. These represent promising developments, which can improve and broaden the spectrum of predictions of NT.
Ecologists have highly criticized NT because of its unrealistic assumptions. The patterns that we have studied so far can in fact be explained without introducing species differences, and this has led some ecologists to oppose NT because it assumes that nature is actually governed by neutral processes, whereas it is not. Clearly, no one believes that nature is truly neutral. The patterns of community ecology are actually generated by a cocktail of processes, and it is both inappropriate and dangerous to consider processes in isolation from a macroecological pattern or empirical data set.
However, it is informative to study whether, how, and which ecological processes can drive a community toward a state in which demographic stochasticity and immigration play a crucial role in the face of strong species’ differences that are dictated by classical competitive exclusion. Such a state should allow similar species to emerge in the niche space with the ability to coexist for sufficient time. Indeed, when this problem was studied, it was shown that species can evolve into groups of relatively more similar species that coexist for very long times (). A large number of species were placed at random along a hypothetical niche axis, which represents a specific trait, assuming that interspecific competition can be calculated through niche overlap. Running a classical Lotka-Volterra competition model and studying evolution, groups of multiple species were evident that aggregated around similar values in the niche axis and they could coexist for many generations before the majority of them head toward an inexorable extinction. Eventually, only one species survives from each group, producing the expected pattern of single species equally spaced in the niche space.
In other words, the niche similarity of species prevents competitive exclusion from swiftly selecting the best competitor among a group of similar species, allowing their coexistence for very long times even though only the superior species will ultimately persist.
This model may be considered as one of the possible steps toward a reconciliation of niche and neutral theories. Species that are initially ecologically nonequivalent, and that therefore behave in a non-neutral fashion, are driven by community and evolutionary processes toward states in which the dynamics may well be better approximated by neutral models over appropriate spatial and temporal scales. More recently, further support for this approach came from showing that the model can produce multimodal RSAs (). Immigration may also be an important component in the neutral-like behavior of communities (; ). Parasitoids competing for a common species () have been used to show that clusters of species separated by gaps emerge along the niche axis, confirming—using a quite different approach—that processes exist that can lead community dynamics to be effectively neutral ().
The basic Lotka-Volterra model has been extended to investigate the possibility that some processes decrease the risk of competitive exclusion so that species lumps are not only transient, but ultimately permanent (). Density-dependent regulation was introduced that stabilizes the coexistence of species within a group. This approach unfortunately has a drawback that the mechanism introduces a discontinuity in the competition strength of the species (), which means that unmodeled species differences may be responsible for coexistence in the community. However, it has been shown that more realistic density-dependent terms, or even other mechanisms [e.g., migration ()], can eliminate this problem, making the approach more robust ().
The maximum entropy principle is a useful method to obtain the least biased information from empirical measurements (). This powerful tool, borrowed from statistical mechanics, has a wide range of applications (), including ecology (; ). In its ecological application, the Max Ent principle is an inference method () used to evaluate the effective strength of interactions among species based on either species-abundance data () or simply the presence or absence of the species (). This methodology provides a way to systematically incorporate the most important species interactions into the development of a theory beyond the purely noninteracting case. In addition, the Max Ent principle was implemented as a method to predict biodiversity patterns across different spatial scales using only the information on local interactions (; ). Here we discuss the extension of Max Ent to the study of spatial biodiversity patterns. Consider an ecosystem in which
where
In this section, we consider how one might advance theory in terms of adding essential details without making the system unnecessarily complex or having an explosion of species-specific parameters. What is the relative importance of species traits, their interactions, and spatial and environmental effects? Throughout this review, we have treated the total number of species and the total number of individuals as parameters which are provided as input. How would one determine and predict these parameters? A mechanistic explanation is needed for why one region may be more biodiverse than another, what sustains biodiversity, and how evolutionary pressures sculpt ecological communities. At the moment, we do not have a spatially explicit theory that can analytically predict the most important ecological patterns. This would be important, because it would allow us to understand what drives biodiversity across spatial and temporal scales. Indeed, it is likely that biological processes are not equally important across scales. As evident in particle physics, we might eventually find that ecosystems will need to be described by different effective theories according to the range of scales in which we are interested. In a pioneering paper, which received great interest, a ME was proposed that seemed to permit analytical calculations of the SAR (). Indeed, an analytical expression for the SAR was obtained. However, the promised solution was neither correct, nor was it an approximation of the actual SAR (). The main technical difficulty preventing an analytical solution was that the model was inherently out of equilibrium. Detailed balance, which is the condition that makes the calculation of stationary probabilities possible, did not hold (). The time is ripe for explorations of this kind. The challenge is to develop new and powerful techniques in nonequilibrium statistical mechanics. Others have applied NT to study biodiversity patterns on river networks. By exploiting the network connectivity of such dendritic landscapes (), they were able to explain several empirical large-scale spatial biodiversity patterns (, ; ) and show how changes in riverine ecosystems may impact the spread of species and local species richness (; ). Over the last few decades, ecologists have come to appreciate the importance of spatial patterns and processes, and the explicit introduction of space has the potential to revolutionize what we know about natural populations and communities (; ). It has become apparent that key ecological patterns, such as SAR, RSA, and spatial patterns of species distributions and turnover, are intimately intertwined and scale dependent (see Fig. ). However, despite a plethora of models that address spatial patterns, only a few practical methods have been proposed to link them across different scales. Specifically, spatial approaches (; ) lack the needed analytical machinery, whereas most theoretical approaches are not spatially explicit or sufficiently flexible (; ). Even neutral theory was conceived to reflect the idealized behavior of natural systems at equilibrium, rather than to reflect nonpristine landscapes produced by environmental change or management (). Therefore, a general methodology is required to predict and link these ecological patterns across scales that is sufficiently robust and flexible to allow its application to a range of natural or managed systems. One possible way to tackle this challenging problem is through a theoretical framework inspired by ideas coming from phenomenological renormalization (). The fundamental assumption at the core of this theoretical setting is that the functional form of the RSA remains the same across all spatial scales, even though the parameters of the curve are likely to vary. Because of this assumption, the spatial dependence of the abundance distribution can be obtained by making the RSA parameters suitable functions of scale. Together with the functional shape of the RSA, the other model input is the spatial PCF, which describes the correlation in species’ abundances between pairs of samples as a function of the distance between them (; ). If populations were randomly distributed in space, distinct communities would on average share the same fraction of species regardless of their spatial separation, and therefore the PCF would not depend on distance. In contrast, in highly aggregated communities correlations in abundance would fall off steeply with increasing distance. The PCF not only measures the rate of turnover in species composition but it also reflects the variation of population clustering across scales, given that the variance in species abundances at any particular scale can be calculated directly from the PCF (). Therefore, the PCF is related to the spatial species-abundance distribution. Thus, the PCF can link the effects of aggregation, similarity decay, species richness, and species abundances across scales. Building on the intrinsic relationship among these patterns, while accounting for spatial correlations with fidelity, is critical for predicting the biodiversity profiles across scales when information on a limited number of fine-scale scattered samples is available (; ). The explosion of publicly available large-scale biodiversity data for paradigmatic ecosystems such as the Amazonian forest () and ocean plankton () makes this problem one of the most exciting scientific challenges in this field. Most neutral (symmetric) models are solely governed by the underlying stochastic birth-death process and do not take into account any environmental stochasticity (). Recently it was shown that environmentally induced variations of the demographic rates dominate the long-term dynamics and have an important impact on some dynamic properties [such as age-size relationships and species extinction time ()], while not affecting the already good accord of neutral models with ecological static patterns (such as the RSA) (). However, analytical results on neutral (symmetric) models for ecological communities subject to correlated environmental noise are mainly missing [although some recent results have been obtained for white noise (; )], incentivizing physicists to explore these paths. Although we are progressing in understanding the suitability and limits of noninteracting and nonspatial models, most neutral models still assume that species interact randomly with each other. However, a network approach to modeling ecological systems provides a powerful representation of the interactions among species (; ; ). Ecological networks may be viewed as a set of different species (nodes) and connections or links (edges) that represent interspecific interactions (e.g., competition, predation, parasitism, and mutualism). The architecture of ecological interaction networks has become a bubbling area of research, and it seems to be a critical feature in shaping and regulating community dynamics and structure diversity patterns (; ). An important step relevant for multitrophic systems will be to obtain a general framework within which a network of preferences or disfavors modifies the birth and death rates of different species and can be superposed on neutral models (like the voter model presented in this review). Recent studies of ecological networks have considered the exciting task of anticipating critical transitions in such systems and to design structures that are less vulnerable to collapse (; ). Connecting stochastic quasineutral models, ecological networks, and critical transitions within a unified theoretical framework is an important challenge as it will enhance our capacity to understand and thus manage the crucial interplay between ecological dynamics and species interactions.
In this paper, we have attempted to describe some of the theoretical frameworks that can be used to understand key issues related to biodiversity and that will serve to address important questions. These frameworks are necessarily elementary and incomplete, yet they have the advantage of being tractable and related to the central issues. Unlike standard approaches to more traditional physics, here the Hamiltonian function or the interactions among the components are completely unknown and even identifying the state variables may sometimes be a nontrivial task. We have a fleeting picture of what an ecosystem is and it is not necessarily in equilibrium. We have little knowledge of the myriad degrees of freedom and their interactions. The real challenge is to discern the most essential degrees of freedom and to develop a framework to understand and predict the emergent ecological behavior.
The basic message of this review is that to resolve these challenging problems, ideas and techniques must be recruited from different disciplines. We are still at the beginning of this adventure. Moving forward is not only important but it is also urgent. The pressures of habitat destruction, pollution, and climate change are having highly undesirable consequences on the health of ecological communities. To address practical issues related to conservation biology, we need models that can be used across scales in order to extrapolate information on biodiversity from accessible regions to inaccessible yet important scales. There is plenty of room for ideas that matter, and community ecology can greatly benefit from the contribution of other disciplines, including physics.
We have greatly benefited from discussions with Stephen Hubbell, Stefano Allesina, David Alonso, Enrico Bertuzzo, Ryan Chisholm, Stephen J. Cornell, Paolo D’Odorico, William F. Fagan, Andrea Giometto, William E. Kunin, Egbert Leigh, Alan J. McKane, Miguel A. Muñoz, Andrew Noble, James P. O’Dwyer, Andrea Rinaldo, Ignacio Rodriguez-Iturbe, James L. Rosindell, and Nadav Shnerb. S. S. acknowledges the University of Padova Physics and Astronomy Department Senior Grant No. 129/2013 Prot. 1634.
In this Appendix, we present an alternative method to introduce density dependence and how ecosystems emerge by assembling local communities.
First, one can set
The steady-state solution of the ME for
with a mean
The number of species containing
where
Note that a nontrivial
where
considered earlier with the advantage of having ecologically meaningful parameters.
The average number of species observed in the local community is
If the sample considered has
that is a compound multinomial Dirichlet distribution where
Metacommunity composition: Now let us gradually assemble the metacommunity of coral reefs by considering it as an assemblage of local communities. Let us start by considering the joint RSA of two local communities
where the actual spatial locations of the local reef communities have been neglected (all local communities are well mixed in the metacommunity, i.e., a mean-field approximation). The elegant result of the
Extending the calculation of the joint RSA distribution to more and more local communities, it can be shown that the RSA of the metacommunity is characterized by an effective immigration parameter
A Fisher log series is observed in two limiting cases—in the metacommunity in which there are no immigration events and in the very small local community that has a high immigration rate from the metacommunity characterized by a Fisher log-series RSA.
In this Appendix, we present the derivation for the persistence (or lifetime) distribution that leads to the empirical choice made in Eq. . The master equation we want to solve is
where the birth rate is
Thus, we introduce the generating function
whose radius of convergence is
with the initial condition
with a “final” condition
implying that
The solution of Eq. , with the chosen final condition gives
and thus
Finally, we get the survival probability
In the scaling, if we consider the limit of fixed
with
The lifetime distribution is simply given by
where the last equality holds in the scaling limit as
which for
In this Appendix, we present another temporal pattern predicted by the model defined by Eq. or . For an ecosystem at or near stationarity, the model can provide the exact expression for the STD when using the time-dependent absorbing solution of Eq. : the so-called species turnover distribution with absorbing boundary conditions. Unlike the reflecting species turnover distribution, this distribution corresponds to a new kind of measure that only accounts for the species present at the initial time, and it does not take into account any new species introduced by immigration or speciation or any old species that reappear after their apparent extinction until
where
It is noteworthy that the STD’s with absorbing and reflecting boundaries are indistinguishable whenever
References (193)
- Adler, P. B., J. HilleRisLambers, and J. M. Levine, 2007, Ecol. Lett. 10, 95.
- Adorisio, M., J. Grilli, S. Suweis, S. Azaele, J. R. Banavar, and A. Maritan, 2014, arXiv:1407.2425.
- Allen, A. P., and V. M. Savage, 2007, Ecol. Lett. 10, 637.
- Allesina, S., J. Grilli, G. Barabás, S. Tang, J. Aljadeff, and A. Maritan, 2015, Nat. Commun. 6, 7842.
- Allesina, S., and S. Tang, 2012, Nature (London) 483, 205.
- Alonso, D., R. S. Etienne, and A. J. McKane, 2006, Trends Ecol. Evol. 21, 451.
- Alonso, D., and A. J. McKane, 2004, Ecol. Lett. 7, 901.
- Alonso, D., A. Ostling, and R. S. Etienne, 2008, Ecol. Lett. 11, 93.
- Arrhenius, O., 1921, J. Ecol. 9, 95.
- Azaele, S., A. Maritan, S. J. Cornell, S. Suweis, J. R. Banavar, D. Gabriel, and W. E. Kunin, 2015, Meth. Ecol. Evol. 6, 324.
- Azaele, S., R. Muneepeerakul, A. Maritan, A. Rinaldo, and I. Rodriguez-Iturbe, 2009, Proc. Natl. Acad. Sci. U.S.A. 106, 7058.
- Azaele, S., R. Muneepeerakul, A. Rinaldo, and I. Rodriguez-Iturbe, 2010, J. Theor. Biol. 262, 323.
- Azaele, S., S. Pigolotti, J. R. Banavar, and A. Maritan, 2006, Nature (London) 444, 926.
- Babak, P., 2006, Phys. Rev. E 74, 021902.
- Babak, P., and F. He, 2009, Theor. Popul. Biol. 75, 76.
- Banavar, J. R., A. Maritan, and I. Volkov, 2010, J. Phys. Condens. Matter 22, 063101.
- Barabás, G., R. D’Andrea, R. Rael, G. Meszéna, and A. Ostling, 2013, Oikos 122, 1565.
- Bastolla, U., M. a. Fortuna, A. Pascual-García, A. Ferrera, B. Luque, and J. Bascompte, 2009, Nature (London) 458, 1018.
- Bell, G., 2000, Am. Nat. 155, 606.
- Bertuzzo, E., S. Suweis, L. Mari, A. Maritan, I. Rodríguez-Iturbe, and A. Rinaldo, 2011, Proc. Natl. Acad. Sci. U.S.A. 108, 4346.
- Black, A. J., and A. J. McKane, 2012, Trends Ecol. Evol. 27, 337.
- Blythe, R. A., and A. J. McKane, 2007, J. Stat. Mech. 07, P07018.
- Bonsall, M. B., V. A. A. Jansen, and M. P. Hassell, 2004, Science 306, 111.
- Borile, C., A. Maritan, and M. A. Muñoz, 2013, J. Stat. Mech. 04, P04032.
- Borile, C., M. A. Muñoz, S. Azaele, J. R. Banavar, and A. Maritan, 2012, Phys. Rev. Lett. 109, 038102.
- Bork, P., C. Bowler, C. de Vargas, G. Gorsky, E. Karsenti, and P. Wincker, 2015, Science 348, 873.
- Brown, J. H., 1995, Macroecology (The University of Chicago Press, Chicago).
- Butler, T., and N. Goldenfeld, 2009, Phys. Rev. E 80, 030902.
- Carrara, F., F. Altermatt, I. Rodriguez-Iturbe, and A. Rinaldo, 2012, Proc. Natl. Acad. Sci. U.S.A. 109, 5761.
- Carrara, F., A. Rinaldo, A. Giometto, and F. Altermatt, 2014, Am. Nat. 183, 13.
- Cencini, M., S. Pigolotti, and M. A. Muñoz, 2012, PLoS One 7, e38232.
- Chandrasekhar, S., 1943, Rev. Mod. Phys. 15, 1.
- Chase, J. M., and M. A. Leibold, 2003, Ecological niches: Linking classical and contemporary approaches (University of Chicago Press, Chicago).
- Chave, J., 2004, Ecol. Lett. 7, 241.
- Chave, J., D. Alonso, and R. S. Etienne, 2006, Nature (London) 441, E1.
- Chave, J., and E. G. Leigh, 2002, Theor. Popul. Biol. 62, 153.
- Chave, J., H. C. Muller-Landau, and S. A. Levin, 2002, Am. Nat. 159, 1.
- Chayes, J. T., L. Chayes, and E. H. Lieb, 1984, Commun. Math. Phys. 93, 57.
- Chesson, P., 2000, Annu. Rev. Ecol. Syst. 31, 343.
- Chesson, P. L., and R. R. Warner, 1981, Am. Nat. 117, 923.
- Chisholm, R. A., 2011, Theor. Popul. Biol. 80, 71.
- Chisholm, R. A., and J. P. O’Dwyer, 2014, Theor. Popul. Biol. 93, 85.
- Chisholm, R. A., et al., 2014, Ecol. Lett. 17, 855.
- Clark, J. S., 2009, Trends Ecol. Evol. 24, 8.
- Coleman, B. D., 1981, Math. Biosci. 54, 191.
- Coleman, B. D., M. A. Mares, and Y.-H. Hsieh, 1982, Ecology 63, 1121.
- Connell, J. H., 1970, Proceedings of the Advanced Study Institute on Dynamics of Numbers in Populations, edited by P. J. Den Boer and G. R. Gradwell (Centre for Agricultural Publishing and Documentation, Wageningen) pp. 298–312.
- Darwin, C. R., 1909, The Origin of Species (The Harvard Classics, New York), Vol. 11.
- Davies, T. J., A. P. Allen, L. Borda-de Água, J. Regetz, and C. J. Melián, 2011, Evolution 65, 1841.
- Diamond, J. M., 1989, Phil. Trans. R. Soc. B 325, 469.
- Du, X., S. Zhou, and R. S. Etienne, 2011, J. Theor. Biol. 278, 127.
- Durrett, R., and S. Levin, 1996, J. Theor. Biol. 179, 119.
- Etienne, R. S., 2005, Ecol. Lett. 8, 253.
- Etienne, R. S., and D. Alonso, 2005, Ecol. Lett. 8, 1147.
- Etienne, R. S., D. Alonso, and A. J. McKane, 2007, J. Theor. Biol. 248, 522.
- Etienne, R. S., and H. Olff, 2004a, Am. Nat. 163, 69.
- Etienne, R. S., and H. Olff, 2004b, Ecol. Lett. 7, 170.
- Etienne, R. S., and J. Rosindell, 2011, PLoS One 6, e14717.
- Ewens, W., 1972, Theor. Popul. Biol. 3, 87.
- Feller, W., 1951, Ann. Math. 54, 173.
- Fisher, A., R. Corbet, and C. Williams, 1943, J. Anim. Ecol. 12, 42.
- Fisher, C. K., and P. Mehta, 2014, Proc. Natl. Acad. Sci. U.S.A. 111, 13111.
- Gardiner, C., 2009, Stochastic Methods: A Handbook for the Natural and Social Sciences (Springer, New York).
- Gause, G. F., 1934, The Struggle for Existence (Williams & Wilkins Company, Baltimore).
- Goldenfeld, N., and C. Woese, 2007, Nature (London) 445, 369.
- Goldenfeld, N., and C. Woese, 2010, arXiv:1011.4125.
- Gould, S. J., 1979, Am. Nat. 114, 335.
- Gravel, D., C. D. Canham, M. Beaudet, and C. Messier, 2006, Ecol. Lett. 9, 399.
- Grilli, J., S. Azaele, J. R. Banavar, and A. Maritan, 2012a, Europhys. Lett. 100, 38002.
- Grilli, J., S. Azaele, J. R. Banavar, and A. Maritan, 2012b, J. Theor. Biol. 313, 87.
- Haegeman, B., and R. S. Etienne, 2008, J. Theor. Biol. 252, 288.
- Haegeman, B., and R. S. Etienne, 2009, Bull. Math. Biol. 71, 1507.
- Haegeman, B., and R. S. Etienne, 2010, Phys. Rev. E 81, 031911.
- Haegeman, B., and M. Loreau, 2011, J. Theor. Biol. 269, 150.
- Halley, J. M., and Y. Iwasa, 2011, Proc. Natl. Acad. Sci. U.S.A. 108, 2316.
- Harte, J., 2003, Nature (London) 424, 1006.
- Harte, J., 2011, Maximum Entropy and Ecology: A Theory of Abundance, Distribution, and Energetics (Oxford University Press, New York).
- Harte, J., T. Zillio, E. Conlisk, and A. Smith, 2008, Ecology 89, 2700.
- Hashitsume, N., M. Toda, R. Kubo, and N. Saitō, 1991, Statistical physics II: Nonequilibrium Statistical Mechanics (Springer Science & Business Media), Vol. 30.
- He, F., and S. P. Hubbell, 2011, Nature (London) 473, 368.
- Holley, R. A., and T. M. Liggett, 1975, Ann. Probab. 3, 643.
- Holt, R. D., 2006, Trends Ecol. Evol. 21, 531.
- Hu, X. S., F. L. He, and S. P. Hubbell, 2006, Oikos 113, 548.
- Huang, K., 2001, Introduction to Statistical Physics (CRC Press, Boca Raton, FL).
- Hubbell, S., 2001, The Unified Theory of Biodiversity and Biogeography (Princeton University Press, Princeton, NJ).
- Hubbell, S. P., 1979, Science 203, 1299.
- Hubbell, S. P., 2005, Funct. Ecol. 19, 166.
- Hubbell, S. P., 2006, Ecology 87, 1387.
- Hutchinson, G. E., 1957, in Cold Spring Harbor Symposia on Quantitative Biology, Vol. 22 (Cold Spring Harbor Laboratory Press), pp. 415–427.
- Hutchinson, G. E., 1959, Am. Nat. 93, 145.
- Janzen, D. H., 1970, Am. Nat. 104, 501.
- Jaynes, E. T., 2003, Probability Theory: The Logic of Science (Cambridge University Press, Cambridge, England).
- Jeraldo, P., et al., 2012, Proc. Natl. Acad. Sci. U.S.A. 109, 9692.
- Kalyuzhny, M., R. Kadmon, and N. M. Shnerb, 2015, Ecol. Lett. 18, 572.
- Kalyuzhny, M., E. Seri, R. Chocron, C. H. Flather, R. Kadmon, and N. M. Shnerb, 2014, Am. Nat. 184, 439.
- Kampen, N. V., 2007, Stochastic Processes in Physics and Chemistry (North-Holland, Amsterdam), 3rd ed.
- Keil, P., T. Herben, J. Rosindell, and D. Storch, 2010, J. Theor. Biol. 265, 78.
- Keitt, T. H., and H. E. Stanley, 1998, Nature (London) 393, 257.
- Kessler, D. A., and N. M. Shnerb, 2015, Phys. Rev. E 91, 042705.
- Kessler, D. A., S. Suweis, M. Formentin, and N. M. Shnerb, 2015, Phys. Rev. E 92, 022722.
- Kimura, M., 1985, The Neutral Theory of Molecular Evolution (Cambridge University Press, Cambridge).
- Kingman, J. F. C., 1982, Stoch. Proc. Appl. 13, 235.
- Krishna, A., P. R. Guimara, P. Jordano, and J. Bascompte, 2008, Oikos 117, 1609.
- Lande, R., S. Engen, and B.-E. Saether, 2003, Stochastic Population Dynamics in Ecology and Conservation (Oxford University Press, Oxford).
- Lebedev, N. N., 1972, Special functions and their applications (Courier Dover Publications, New York).
- Lehnigk, S. H., 1993, The generalized Feller equations and related topics (Longman, New York).
- Leigh, E. G., 2007, J. Evol. Biol. 20, 2075.
- Lynch, H. J., E. H. Campbell Grant, R. Muneepeerakul, M. Arunachalam, I. Rodriguez-Iturbe, and W. F. Fagan, 2011, Water Resour. Res. 47, W05531.
- MacArthur, R., 1970, Theor. Popul. Biol. 1, 1.
- MacArthur, R. H., and E. O. Wilson, 1967, The Theory of Island Biogeography (Princeton University Press, Princeton, NJ).
- Manceau, M., A. Lambert, and H. Morlon, 2015, Ecol. Lett. 18, 347.
- Mari, L., R. Casagrandi, E. Bertuzzo, A. Rinaldo, and M. Gatto, 2014, Ecol. Lett. 17, 426.
- Martín, H. G., and N. Goldenfeld, 2006, Proc. Natl. Acad. Sci. U.S.A. 103, 10310.
- Martin, T. E., 1981, Am. Nat. 118, 823.
- Masoliver, J., and J. Perelló, 2012, Phys. Rev. E 86, 041116; 86, 049906(E) (2012); 89, 029902 (2014).
- May, R. M., 2001, Stability and Complexity in Model Ecosystems (Princeton University Press, Princeton, NJ), Vol. 6.
- May, R. M., 2010, Phil. Trans. R. Soc. B 365, 41.
- May, R. M., and R. H. Mac Arthur, 1972, Proc. Natl. Acad. Sci. U.S.A. 69, 1109.
- McGill, B., 2003, Oikos 102, 679.
- McGill, B., and C. Collins, 2003, Evol. Ecol. Res. 5, 469.
- McGill, B. J., B. A. Maurer, and M. D. Weiser, 2006, Ecology 87, 1411.
- McKane, A., D. Alonso, and R. V. Solé, 2000, Phys. Rev. E 62, 8466.
- McKane, A. J., D. Alonso, and R. V. Solé, 2004, Theor. Popul. Biol. 65, 67.
- Melbinger, A., and M. Vergassola, 2015, Sci. Rep. 5, 15211.
- Morlon, H., G. Chuyong, R. Condit, S. Hubbell, D. Kenfack, D. Thomas, R. Valencia, and J. L. Green, 2008, Ecol. Lett. 11, 904.
- Mouquet, N., and M. Loreau, 2003, Am. Nat. 162, 544.
- Muneepeerakul, R., S. Azaele, S. Levin, A. Rinaldo, and I. Rodriguez-Iturbe, 2011, J. Theor. Biol. 269, 256.
- Muneepeerakul, R., E. Bertuzzo, H. J. Lynch, W. F. Fagan, A. Rinaldo, and I. Rodriguez-Iturbe, 2008, Nature (London) 453, 220.
- Mutshinda, C. M., R. B. O’Hara, and I. P. Woiwod, 2008, Funct. Ecol. 22, 340.
- Nee, S., 2005, Funct. Ecol. 19, 173.
- Noble, A. E., and W. F. Fagan, 2015, Theoret. Ecol. 8, 149.
- Noble, A. E., A. Hastings, and W. F. Fagan, 2011, Phys. Rev. Lett. 107, 228101.
- Noble, A. E., N. M. Temme, W. F. Fagan, and T. H. Keitt, 2011, J. Theor. Biol. 273, 1.
- O’Dwyer, J. P., and R. Chisholm, 2014, Ecol. Lett. 17, 961.
- O’Dwyer, J. P., and J. L. Green, 2010, Ecol. Lett. 13, 87.
- O’Dwyer, J. P., S. W. Kembel, and T. J. Sharpton, 2015, Proc. Natl. Acad. Sci. U.S.A. 112, 8356.
- Paine, Robert T., 1966, Am. Nat. 100, 65.
- Pigolotti, S., and M. Cencini, 2009, J. Theor. Biol. 260, 83.
- Pigolotti, S., A. Flammini, and A. Maritan, 2004, Phys. Rev. E 70, 011916.
- Pigolotti, S., A. Flammini, M. Marsili, and A. Maritan, 2005, Proc. Natl. Acad. Sci. U.S.A. 102, 15747.
- Pimm, S. L., G. J. Russell, J. L. Gittleman, and T. M. Brooks, 1995, Science 269, 347.
- Pinto, O. A., and M. A. Muñoz, 2011, PLoS One 6, e21946.
- Plotkin, J. B., M. D. Potts, N. Leslie, N. Manokaran, J. LaFrankie, and P. S. Ashton, 2000, J. Theor. Biol. 207, 81.
- Power, D. M., 1972, Evolution 26, 451.
- Preston, F. W., 1948, Ecology 29, 254.
- Preston, F. W., 1960, Ecology 41, 611.
- Pueyo, S., 2006, Oikos 112, 392.
- Purves, D. W., and L. A. Turnbull, 2010, J. Anim. Ecol. 79, 1215.
- Purves, D. W., and S. W. Pacala, 2005, in “Biotic interactions in the tropics: Their role in the maintenance of species diversity,” edited by David Burslem, Michelle Pinard, and Sue Hartley (Cambridge University Press, Cambridge, England).
- Ricklefs, R. E., 2003, Oikos 100, 185.
- Ricklefs, R. E., 2006, Ecology 87, 1424.
- Ricklefs, R. E., and S. S. Renner, 2012, Science 335, 464.
- Rosindell, J., and S. J. Cornell, 2007, Ecol. Lett. 10, 586.
- Rosindell, J., and S. J. Cornell, 2009, Ecology 90, 1743.
- Rosindell, J., S. J. Cornell, S. P. Hubbell, and R. S. Etienne, 2010, Ecol. Lett. 13, 716.
- Rosindell, J., L. J. Harmon, and R. S. Etienne, 2015, Ecol. Lett. 18, 472.
- Rosindell, J., S. P. Hubbell, and R. S. Etienne, 2011, Trends Ecol. Evol. 26, 340.
- Rosindell, J., S. P. Hubbell, F. L. He, L. J. Harmon, and R. S. Etienne, 2012, Trends Ecol. Evol. 27, 203.
- Rosindell, J., Y. Wong, and R. S. Etienne, 2008, Ecological Informatics 3, 259.
- Roughgarden, J., R. M. May, and S. A. Levin, 1989, Eds., Perspectives in Ecological Theory (Princeton University Press, Princeton, NJ), Vol. 203, pp. 203–226.
- Scheffer, M., and E. H. van Nes, 2006, Proc. Natl. Acad. Sci. U.S.A. 103, 6230.
- Scheffer, M., et al., 2012, Science 338, 344.
- Schoener, T. W., 1976, in Proceedings of the XVI International Ornithological Congress, Vol. 6 (Australian Academy of Science, Canberra, ACT), pp. 629–642.
- Simpson, E. H., 1949, Nature (London) 163, 688.
- Slik, J. F., et al., 2015, Proc. Natl. Acad. Sci. U.S.A. 112, 7472.
- Storch, D., et al., 2007, Scaling Biodiversity (Cambridge University Press, Cambridge, UK).
- Suweis, S., E. Bertuzzo, L. Mari, I. Rodriguez-Iturbe, A. Maritan, and A. Rinaldo, 2012, J. Theor. Biol. 303, 15.
- Suweis, S., and P. D’Odorico, 2014, PLoS One 9, e101851.
- Suweis, S., A. Rinaldo, and A. Maritan, 2012, J. Stat. Mech. 07, P07017.
- Suweis, S., F. Simini, J. R. Banavar, and A. Maritan, 2013, Nature (London) 500, 449.
- Svenning, J.-C., and R. Condit, 2008, Science 322, 206.
- ter Steege, H., et al., 2013, Science 342, 1243092.
- ter Steege, H., et al., 2015, Sci. Adv. 1, e1500936.
- Thomas, C. D., et al., 2004, Nature (London) 427, 145.
- Tilman, D., 1994, Ecology 75, 2.
- Tilman, D., and P. M. Kareiva, 1997, Spatial ecology: The role of space in population dynamics and interspecific interactions (Princeton University Press, Princeton, NJ), Vol. 30.
- Ulrich, W., M. Ollik, and K. I. Ugland, 2010, Oikos 119, 1149.
- Vallade, M., and B. Houchmandzadeh, 2003, Phys. Rev. E 68, 061902.
- Vallade, M., and B. Houchmandzadeh, 2006, Phys. Rev. E 74, 051914.
- Vanpeteghem, D., O. Zemb, and B. Haegeman, 2008, Math. Biosci. 212, 88.
- Veech, J. A., 2006, J. Biogeography 33, 2145.
- Vergnon, R., R. Leijs, E. H. van Nes, and M. Scheffer, 2013, Am. Nat. 182, 67.
- Vergnon, R., E. H. van Nes, and M. Scheffer, 2012, Nat. Commun. 3, 663.
- Vergnon, R., E. H. van Nes, and M. Scheffer, 2013, Oikos 122, 1573.
- Volkov, I., J. Banavar, F. He, S. Hubbell, and A. Maritan, 2005, Nature (London) 438, 658.
- Volkov, I., J. R. Banavar, S. P. Hubbell, and A. Maritan, 2003, Nature (London) 424, 1035.
- Volkov, I., J. R. Banavar, S. P. Hubbell, and A. Maritan, 2007, Nature (London) 450, 45.
- Volkov, I., J. R. Banavar, S. P. Hubbell, and A. Maritan, 2009, Proc. Natl. Acad. Sci. U.S.A. 106, 13854.
- Volkov, I., J. R. Banavar, and A. Maritan, 2006, J. Stat. Phys. 123, 167.
- Walker, S. C., 2007, Theor. Popul. Biol. 71, 318.
- Watson, H. C., 1835, Remarks on the Geographical Distribution of British Plants;: Chiefly in Connection with Latitude, Elevation, and Climate (Longman, Rees, Orme, Brown, Green, and Longman, Paternoster-Row, London).
- Zillio, T., J. R. Banavar, J. L. Green, J. Harte, and A. Maritan, 2008, Proc. Natl. Acad. Sci. U.S.A. 105, 18714.
- Zillio, T., I. Volkov, J. R. Banavar, S. P. Hubbell, and A. Maritan, 2005, Phys. Rev. Lett. 95, 098101.