- Featured in Physics
- Open Access
Interactions and Migration Rescuing Ecological Diversity
PRX Life 2, 013014 – Published 18 March, 2024
Abstract
How diversity is maintained in natural ecosystems is a long-standing question in Theoretical Ecology. By studying a system that combines ecological dynamics, heterogeneous interactions, and spatial structure, we uncover a new mechanism for the survival of diversity-rich ecosystems in the presence of demographic fluctuations. For a single species, one finds a continuous phase transition between an extinction and a survival state, that falls into the universality class of Directed Percolation. Here we show that the case of many species with heterogeneous interactions is different and richer. By merging theory and simulations, we demonstrate that with sufficiently strong demographic noise, the system exhibits behavior akin to the single-species case, undergoing a continuous transition. Conversely, at low demographic noise, we observe unique features indicative of the ecosystem's complexity. The combined effects of the heterogeneity in the interaction network and migration enable the community to thrive, even in situations where demographic noise would lead to the extinction of isolated species. The emergence of mutualism induces the development of global bistability, accompanied by sudden tipping points. We present a way to predict the catastrophic shift from high diversity to extinction by probing responses to perturbations as an early warning signal.
Physics Subject Headings (PhySH)
Viewpoint
How Migration May Stabilize the Diversity of Ecosystems
A model based on statistical physics suggests that the combination of species migration and interspecies interactions may allow a complex ecological system to maintain its diversity.
See more in Physics
Article Text
Community ecology explores how the interactions between different species shape the diversity-rich ecosystems that characterize the natural world. Understanding the main mechanisms at play is a challenge that spans different scientific fields and it is relevant for human health .
There are three salient facts that one has to take into account in this endeavor. Many ecosystems of interest are species-rich. The interactions between these large sets of species, and the induced ecological dynamics, can lead to complex dynamical behaviors such as chaos and a very large number of possible equilibria . Many ecosystems are spatially extended: the ecological dynamics takes place at some local scale, but individuals can then explore different spatial locations through migration . This can lead to the appearance of complex ecological phenomena, such as traveling activity fronts, pattern formation, and persistent chaotic dynamics . Ecosystems are subject to noise, particularly environmental and demographic noise (due to stochasticity in births and deaths). Both noises induce fluctuations that are a key factor in determining abundance distributions, and their time-dependence . Understanding the interplay between these three properties of ecosystems is essential for answering many central questions in community ecology.
In this work, we consider spatially extended species-rich ecosystems subject to demographic noise. We will consider populations that are large but spatially structured, so that demographic fluctuations globally average out, but they have an important effect on the local dynamics. This is the case, for example, in semiarid ecosystems: the total number of plants is such that global fluctuations are negligible, but at the local level stochasticity can play a fundamental role . Our aim is to understand how in these cases interactions and spatial migration can allow for large diversity and finite abundances despite the adversarial role of demographic noise. In fact, in an isolated community, demographic noise leads to extinctions, irreversibly reducing the ecosystem's diversity until there are no species left .
Previous works, following the classical theory of Island Biogeography by MacArthur and Wilson , proposed as a rescuing mechanism the immigration from a static reservoir (or “mainland,” when thinking of an island-mainland system) . Nevertheless, this approach simply shifts the question from how diversity is maintained on the island to its maintenance on the mainland. Here we use a different approach. We consider ecosystems as a network of ecological communities (a metacommunity) coupled by passive dispersal. In this case, the immigration rates are not externally imposed, but they are the result of the internal dynamics. If a species goes locally extinct in one of the communities, immigrants from the neighboring ones can reinvade, providing an “insurance” (or “storage”) effect . This makes the possibility of a global extinction much more unlikely, and it can allow the ecosystem to self-sustain at finite abundances and diversity. The stabilization of high-diversity states by spatial structure is a very general phenomenon: it can arise in the presence of spatial heterogeneity of environmental conditions or when abundances in different spatial locations exhibit unsynchronized fluctuations . Providing a theory for this mechanism for species-rich ecosystems subject to demographic noise, and assessing the role of interactions, is the main contribution of this work.
The situation is well understood in the case of a few species, in which, depending on the competition between migration and death-birth rates, the system is found to be either in a survival or in an extinct state. A transition separates the two regimes . This phase transition falls in the universality class of Directed Percolation, a second-order out-of-equilibrium transition studied in statistical physics and widely used to describe spreading phenomena, from forest fires to epidemics .
In a many-species metacommunity with constant competitive interactions, it was recently shown that a similar second-order phase transition takes place and that it also belongs to the Directed Percolation universality class . Because the transition is continuous with vanishing abundances, interactions, which are quadratic in the abundances, are subleading at the critical point. As a consequence, the main mechanism at play in this case is still the competition between migration and death-birth rates. We shall show that the scenario for heterogeneous interactions is different and goes beyond the directed percolation paradigm. The transition can become discontinuous. The ecosystem can exhibit global bistability and tipping points between drastically different alternative states. Upon small changes in the environmental condition, the system can therefore undergo catastrophic shifts from a state with large diversity and finite abundances to one in which all species are extinct. As in many other dynamical systems, from coral reefs to arid ecosystems and from Earth's climate to financial markets , it is important to find early warning signals of these kinds of transitions in order to prevent them. We have identified a specific probe, which is based on the response of the ecosystem to perturbations, and that can be monitored in experiments. Our analytical framework shows that interactions play a key role both in the overall scenario and in promoting a self-sustained survival state, in agreement with results obtained for constant mutualistic interactions . Remarkably, in our case, heterogeneous interactions of the pool of species are not necessarily mutualistic on average. It is the ecological dynamics that shapes the ecosystem in a self-sustained phase characterized by emergent mutualistic behavior among the nonextinct species.
In our work, we make use of several methods developed in statistical physics that are particularly well suited for species-rich ecosystems, which are complex systems formed by many interacting degrees of freedom undergoing stochastic dynamics. To model the heterogeneity in the interactions, we sample the coupling coefficients from a random ensemble. We have thus to deal with “disordered” ecosystems, which can be analyzed by transferring methods from spin-glass theory . This disorder approach, which dates back to May's seminal paper , has recently inspired a growing body of work and also received positive experimental confirmations . Previous works have explored within this framework the effect of heterogeneous interactions , demographic fluctuations , and spatial structure , but the analysis we present here is, to our knowledge, the first analytical study in which the three ingredients are combined.
The model we focus on is a disordered Generalized Lotka Volterra (GLV) system of metacommunity subject to demographic noise. For one community, the disordered GLV has been shown to have a rich phase diagram, and to display several dynamical regimes: single equilibrium, multistability, and chaos . We expect this complex behavior also in the case of spatially structured ecosystems . In this work, we focus on the moderate-heterogeneity regime in which there is a single stable equilibrium. This allows us to disentangle the multistability due to the fragmentation of the basins of attraction of the ecological dynamics at strong heterogeneity from the bistability of the feedback mechanism between abundance and immigration. Our analysis is performed using a mean-field approximation on the spatial fluctuations, which is equivalent to considering that the community network is a fully connected graph.
Note that because of their generality, Lotka-Volterra equations have been applied to a variety of fields besides their original ecological interpretation, from immunology to economics and game theory . Our results could therefore find applications beyond ecology, notably for the study of global bistability and economic crashes.
We consider a metacommunity of
which corresponds to Lotka-Volterra dynamics, with constant growth rate
To model the heterogeneity in the interactions of species-rich ecosystems, we follow and consider the disordered LV model. As already discussed in the Introduction, the disorder approach has attracted recently a lot of attention and also received positive experimental confirmations . In this framework, the interaction coefficients are random variables, with mean
We will restrict the choice of
In the model defined by Eq. , individuals can migrate on the patches network through diffusion, with a constant diffusion coefficient
Each species is subject to a white demographic noise
The autocorrelation of the demographic noise defines the noise strength
Some further insights into the effect of the demographic noise can be obtained considering it in the absence of all the other terms. In this case, an exact solution to the associated Fokker-Planck equation is available, showing that starting from any initial condition the population goes to 0 abundance with some finite rate . Therefore, also in the continuous model, extinction is possible over finite times, and not only asymptotically, as would be the case, for example, with environmental noise.
The dynamics of species in the presence of birth and death has important connections with the celebrated directed percolation problem studied in out-of-equilibrium physics and statistical field theory . Directed percolation is a model of particles that hop on a network and are subjected to births and deaths; a graphical illustration of the process can be found in Fig. for a one-dimensional network. Directed percolation was originally introduced to model spreading phenomena, from forest fires to epidemics . In our case, the sites of the network represent spatial locations, or patches, on which (or from which) species can migrate; the particles indicate which sites are colonized by species. At each time step the particles can produce an offspring in a neighboring site, die, or just survive. In our case, this corresponds to colonization or extinction. Depending on the competition between death and birth rates, the activity can spread to the entire system and lead to a finite density of particles (active, self-sustaining state) or die out (absorbing, inactive state). Between these two phases, there is a continuous phase transition, characterized by universal critical behavior . We show in Fig. the phase diagram in the mean-field approximation (discussed in the next section). A direct link between DP and GLV is obtained by coarse-graining . In this way, the discrete DP occupation variable becomes a continuous quantity that represents the mean occupation, the competition between birth and death rates gives rise to a logistic growth, hopping is replaced by diffusion, and the stochastic fluctuations generate the demographic noise. This leads to a set of independent GLV Eqs. in the absence of interactions, one for each species. Each equation corresponds to an independent directed percolation process.
Phase diagram for Directed Percolation in the mean-field approximation: in green the active phase, in which at long times there is a finite density of particles, in white the inactive phase, in which all particles eventually die.
The directed percolation transition can therefore be interpreted as a transition between a self-sustained phase where migration enables a finite abundance of species to persist, to a regime, characteristic of small (or zero) dispersal, where species go extinct due to demographic noise. The aim of this work is to develop a theory for these phenomena for species-rich ecosystems in the presence of heterogeneous interactions. Upon increasing the number of species in the pool and considering heterogeneous interactions, the set of directed percolation processes is no longer independent, and the complexity of the model increases considerably. In fact, the system becomes equivalent to the collection of an infinite number of directed percolation processes, coupled by random interactions—an interesting and open statistical physics problem.
In this work, we aim to study systems in which both the number of species and the number of patches are very large. To obtain analytical results, we follow the statistical physics “way” and take the limit of an infinite number of species and an infinite number of patches. In this double limit (whose order is irrelevant) the macroscopic properties of the system do not depend on the particular realization of the demographic noise and of the interactions: the macroscopic properties are self-averaging in the jargon of disordered systems .
The large-
Our derivation follows the one developed in Ref. for LV models, and it can be found in Appendix for generic values
Since this allows us to decouple different species, for simplicity we will omit the species index
The first term represents the average interaction with all other species. It is given by the product of the mean of the interaction strength and the mean abundance,
The second term represents the fluctuation of the interaction with all other species. It is given by the product of the standard deviation of the interaction coefficients and Gaussian noise with zero mean and correlation matching the time autocorrelation of the single species abundances:
The noise
To distinguish the roles of fluctuating and static noises in the GLV equation, we introduce two kinds of averages:
The last term in the dynamical mean-field treatment of the interactions is due to a feedback mechanism: a fluctuation of the abundance of species
In the
The DMFT closure consists then in replacing the empirical averages over species
Equation can also be interpreted as the Langevin equation associated with a Directed Percolation (DP) process, with the addition of a memory term (that is absent in the special case
Interestingly, whereas a system of a few species interacting and diffusing on a network was established to boil down to a standard DP problem , the case of many species is fundamentally different and belongs to a different class. Indeed, a system of many species is equivalent to a family of many DP processes, characterized by different values of static and fluctuating noises and coupled through the common self-consistently determined mean, correlation, and response functions. Understanding the behavior of this self-consistent DP problem is an open challenge. In this work, we study whether the DP transition can fundamentally change nature due to this self-consistent coupling. Even if the transition remained qualitatively DP-like (continuous and from an absorbing state to a fluctuating one), critical properties could change. In fact, although an environmental noise can be shown to be an irrelevant perturbation of the associated field theory , within DMFT the environmental noise inherits the time dependence of the correlation function through the self-consistency. It can therefore develop long-range correlations in time at the critical point, possibly altering the critical behavior and leading to a new universality class.
Studying the coupled field theories is a formidable task. In the following, we simplify the problem by doing a mean-field approximation, which allows us to obtain a general theory independent of the underlying network of patches.
We replace the term
By substituting
This substitution allows us to decouple stochastic processes for the abundance in different patches. Omitting for simplicity the index
Since all patches are equivalent on a fully connected lattice, the
In the case of symmetric interactions,
The memory term and
where
The self-consistent equations can be expressed as averages with respect to the Boltzmann distribution:
and analogously for
These equations can be solved iteratively: starting from a suitable initial condition for
In conclusion, within the
In the case of a generic value of the spatial heterogeneity of the interactions
The mapping to an equilibrium distribution requires symmetry in the interactions: nonsymmetric interactions correspond to nonconservative forces, which explicitly break time reversal and lead to nonequilibrium steady states. To show that our results hold also in this case, at least if the asymmetry is not too strong, we have analyzed the case of small asymmetry in perturbation theory. The analysis of the Martin–Siggia–Rose–De Dominicis–Janssen action
In the following, we present our analytical results focusing on ecosystems with parameters By solving the DMFT equations described in the previous section, one finds that when the diffusion constant is large enough, the system is in a self-sustained phase (active phase in the directed percolation jargon) in which a nonzero abundance is maintained despite the presence of demographic fluctuations. In this regime, although some species go globally extinct on all patches, others survive thanks to the migration from neighboring patches. This mechanism is sufficient to prevent extinctions due to demographic stochasticity and leads to a self-sustained metacommunity. In the following, we discuss the salient properties of this phase, focusing on two ecologically relevant observables: the average abundance, As expected, demographic noise is detrimental to survival: the fraction of surviving species, or diversity, and the average abundance decrease with the strength of demographic fluctuations; see the bottom panels of Fig. . On the contrary, dispersal is beneficial, as shown in the top panels of Fig. . The behavior of the diversity for species-rich ecosystems with heterogeneous interactions in the presence of demographic noise is a novel result of our approach: in the case of fixed external immigration, previously often considered in the literature, all species are kept alive by the immigration, albeit some at very small abundances, it is therefore not possible to rigorously define the ecosystem diversity . We find that the species that go extinct are those whose growth is on average more affected by the interactions with the rest of the ecosystem, as quantified by the static part of the environmental noise The limits As for the distribution of the abundances, we find an exponential decay (see Appendix ), as is the case in other models with random fully connected interactions . When demographic fluctuations are sufficiently strong, decreasing the diffusion constant leads to a continuous phase transition from an active phase in which some species are able to self-sustain, to an inactive phase in which they are all extinct. The critical value of the diffusion constant is the same as would be obtained in the absence of interactions, where the system directly maps to directed percolation, or in the case of constant interactions ; see Fig. and Appendix . This is to be expected: upon approaching the transition, the abundances tend to zero, and therefore the interactions, which have a quadratic dependence on the abundances, become irrelevant. The critical exponents indeed match the ones falling in the Directed Percolation universality class; in particular, the abundance goes to zero linearly [Fig. ]. Interestingly, approaching the transition the diversity does not go to zero and instead tends to a finite value [Fig. ]. The average abundance goes to zero not because more and more species are going extinct, but because all surviving species are simultaneously decreasing their abundances. This homogenization in the behavior of species is yet another consequence of the irrelevance of the interactions, the only trait distinguishing one species from another in our model. (a) The phase diagram for constant interactions across patches ( At smaller demographic noise this picture changes drastically and interactions play a major role, as shown in the phase diagram in Fig. . The ecosystem is able to self-sustain at values of the diffusion constant for which in the absence of interactions it would be in the inactive phase. Upon further lowering It was recently shown that a metacommunity subject to demographic noise and constant mutualistic interactions exhibits a similar discontinuous phase transition . The authors of also performed numerical simulations with random (patch-independent) interactions, showing that the surviving species have more mutualistic interactions than the total species pool. We find that a similar mechanism is at play in our case: it is an emergent phenomenon due to ecological dynamics which is present even though interactions are not on average mutualistic (in fact they are competitive, Thermal averaged interaction term, In Fig. we also show the phase diagram in the case of independent ( The phase diagram for independent interactions across patches ( One can also analytically show that the phase diagrams remains qualitatively unchanged considering a small asymmetry in the interactions (
To confirm the generality of our results, we now consider different variations of the model studied in the previous section. The aim is to show that our results hold in a broader setting. We shall be particularly interested in considering the case of a large but finite number of species, a large but finite number of patches, a small but finite asymmetry of interactions, as well as intermediate values of Generically, for moderate system sizes ( In Fig. we show the behavior of the average abundance as a function of the diffusion constant for two different values of the temperature, starting from two different initial conditions. To probe the existence of hysteresis, and therefore a discontinuous transition and metastability, we numerically simulate systems with different initial conditions. For the green curves, the initial abundances were uniformly sampled between 0 and 1, for the red curves between 0 and 0.1. The former should therefore be more prone to evolve toward the self-sustained solution, if it exists, whereas the latter is prone to evolve toward the “all-extinct” solution. Average abundance We find that indeed at higher temperatures, Instead, at The heterogeneity in the interaction network is essential to allow the ecosystem to self-sustain below the single DP critical point: indeed if we consider the same parameters but take We are now interested in focusing on cases in which the interactions between species are not fully symmetric, and the interaction matrices are partially correlated between patches, i.e., As we have already discussed, we have analytically established that a very small asymmetry is not a singular perturbation. Thus, our results should qualitatively hold also for a finite, at least not too large, asymmetry. To confirm this finding and study intermediate values of Average abundance In conclusion, combining all these numerical tests, we conclude that the scenario obtained from the analytical solution is robust and holds broadly. We will come back to this point in the Conclusion to suggest other extensions and tests.
In the previous section, we have shown that dispersal can rescue complex and large ecosystems from extinction due to demographic noise. Depending on the strength of the demographic noise, the transition from the self-sustained to the extinct phase can be either continuous or discontinuous. The latter takes place for low demographic noise and low dispersal. In this regime, we have found that the transition is accompanied by a metastable regime and hysteresis. Such a transition is what is called in ecology, and in environmental and social sciences, a tipping point or regime shift , and in physics it is called a spinodal. Tipping points are often catastrophic events, as the abrupt rapid shifts almost always lead to negative consequences and a less favorable state of the system. Our case is no exception, as the system's transition is from a self-sustained state with high diversity to one in which all species are extinct. As was done for several other tipping points , it is therefore important to find early signs or precursors that can allow one to detect the closeness of the system to the tipping point before the catastrophic shift actually takes place.
In our case, following intuition that comes from the physics of spinodal points, we focus on responses to perturbations as a probe of closeness to the tipping point. We can show analytically (see Appendix ) that the instability of the self-sustained state is accompanied by a diverging response to perturbations. This phenomenon is strongly linked to the saddle-node bifurcation of the mean-field equations that governs the transition.
In particular, we have studied the change of the average abundance due to a change in the carrying capacity. Such a response, which can be measured in controlled laboratory experiments, does diverge approaching the discontinuous transition; see Fig. for the
We uncovered a rich phase diagram for many-species Lotka-Volterra metacommunities subject to heterogeneous symmetric interactions, demographic noise, and diffusion. If the demographic fluctuations are too strong, they drive all species to extinctions, but when the diffusion constant is large enough these extinctions can be compensated by recolonizations from neighboring sites, and the ecosystem is able to self-sustain at finite abundance and diversity. The system exhibits a phase transition between an extinction and a survival phase. The transition can be either continuous or discontinuous, depending on whether the behavior of the system is dominated by the demographic fluctuations or the heterogeneous interaction network.
When the demographic fluctuations are strong, the transition is continuous and interactions play a secondary role. In fact, the transition is completely analogous to what one would have in the absence of the interactions (even the critical values of the diffusion constant coincide). This is because when the abundances tend to zero, the interactions become subdominant and the system falls in the standard Directed Percolation universality class.
The situation is drastically different at lower demographic noise. In this case, the transition becomes discontinuous and the system exhibits novel features, which are a signature of the complexity of the ecosystem and the major role played by the interactions. There is an extended range of parameters in which without interactions, i.e., for single species, the system would be driven to extinction, but the metacommunity is instead able to self-sustain at finite abundances. This is possible because strongly competing species are eliminated from the community, while surviving species cooperate to self-sustain in such harsh conditions. For small demographic noise and lowering the diffusion constant, the ecosystem reaches a tipping point at which all surviving species go extinct; close to this point the ecosystem is subject to collapses upon small perturbations, and its dynamics exhibits hysteresis. We therefore find that mutualism naturally emerges from an (on average) competitive pool of species when conditions become harsher. This has a double effect: it allows the ecosystem to survive in conditions in which all species in isolation would go extinct, but it also makes it fragile to perturbations. In this regime, it is not possible to predict the vicinity of the catastrophic shift of the ecosystem by looking at the average abundance. As an early warning sign, we propose to monitor the response of the system to perturbations. We have shown that this is a suitable probe, as it diverges approaching the discontinuous transition.
We confirm and complement our analytical approach with numerical simulations, which show that our results are quite robust to modifications of the model, in particular to the introduction of a small asymmetry in the interactions, to various degrees of correlation of the interaction network between different spatial locations, and for a system with a finite number of species and patches.
There are several directions that warrant future investigations. We focused on a fully connected spatial system, which provides a mean-field analysis for generic spatial lattices. On the other hand, our DMFT treatment of the interactions is directly generalizable to any other spatial network, including finite-dimensional ones. It would be very interesting to study cases in which the patches are located in a finite-dimensional lattice or on random structures. In particular, it would be interesting to find out (1) whether the discontinuous transition is also present in this case or if finite-dimensional fluctuations destroy the metastable region, and (2) whether the continuous transition can still be described in terms of directed percolation, or if interactions, although secondary, can alter its universality class. It would also be worth analyzing stronger asymmetries in the interactions, e.g., lowering the value of
Finally, the species-rich LV model with heterogeneous and strong interactions displays multiple equilibria and chaotic dynamics . The possibility of different patches to converge to different stationary states could strongly modify the behavior of the system, in particular allowing the system to experience higher values of global diversity, possibly violating May's bound .
Note added. During the preparation of this manuscript, we became aware of the work of Denk and Hallatscheck on tipping points in mutualistic Lotka-Volterra communities . Their results are complementary and agree with ours.
We would like to thank J. Denk and O. Hallatscheck for sharing their results and constructive interactions. We also thank Joseph Baron, M. Barbier, J. F. Arnoldi, and L. F. Cugliandolo for stimulating discussions. This work was supported by the Simons Foundation Grant on Cracking the Glass Problem (No. 454935) (G.B.).
Here we outline the derivation, adapted from Ref. , of the Dynamical Mean Field Theory for our system, for a generic value of the spatial correlation of the interactions
We consider
to which we have added a perturbation to the carrying capacity
We have introduced the response function
The dynamics of species 0 will depend on these new trajectories:
Because the correlations between interaction coefficients in any two patches are the same, these Gaussian variables can generically be decomposed into a common random contribution, identical in all patches and proportional to the correlation
We want to describe its statistical properties in the limit
By similarly evaluating all terms in , we obtain
where
Species 0 is statistically equivalent to all the others, therefore we can replace the averages over the
Since species have been effectively decoupled, we can suppress the species index.
In the single equilibrium phase, we expect the process to reach a time translation invariant regime, in which the one-time averages are time-independent, and two-times observables only depend on the times difference. This was shown in for a single community with demographic noise and fixed immigration, and it is known to be the case for Directed Percolation and in a many-species metacommunity with constant interactions . It is also confirmed by our numerical results, which show a quick relaxation of one-time observables to an asymptotic value (see Appendix ), at least away from phase transitions. Since the autocorrelation of the abundance of one species does not tend to zero at large times, we can decompose
where
While the derivation is so far valid for any spatial network, we will now restrict ourselves to a fully connected network, in which the empirical average over neighbors can be replaced by its thermal average. In the large-
The same is true for
We separate
invariance, we obtain
where we have summed the random variables that had the same behavior of the correlations (
The equations simplify in the extreme cases
In the case of symmetric interactions (
We can integrate by parts the term with the memory kernel:
We have obtained an additional quadratic term in
Using the Martin–Siggia–Rose–De Dominicis–Janssen (MSRDJ) formalism, we can show that the stationary probability distribution associated with the stochastic differential equation
is the Boltzmann distribution with the effective Hamiltonian:
where we have reintroduced the perturbations
Its equilibrium distribution is given by
where
The MSRDJ action can be written in terms of a deterministic and a dissipative part ,
The time reversal transformation for the two fields is given by
The deterministic and dissipative part of the action are independently invariant under this transformation:
The action is invariant under the time-reversal transformation using
In the following, we restrict ourselves to the
At equilibrium we can rewrite the integrated disorder-dependent responses to a perturbation of the carrying capacity and of the immigration rate in terms of connected correlation functions of
When the time dependence is not present, we are considering a time-independent perturbation.
Adding a perturbation in site
In the third line we used the fact that the correlations between different patches are subleading to take separately the thermal averages. Solving for
We can then average over
The MSRDJ action with nonsymmetrical interactions is given by
where we have defined
An average
where
We want to estimate the scaling of
to show that it is not singular approaching a phase transition. In the simple equilibrium phase, the connected correlation function decays exponentially, with a typical relaxation time
The correlation function
Inserting these scalings in Eq. , we obtain
Considering a small asymmetry in the interactions, observables are shifted by a correction of order
The self-consistency condition for
We can separate the effective Hamiltonian into a quadratic and a logarithmic part:
For
The integral is now finite for
The term of order
As noted in Ref. , this is the same condition that would determine the criticality of the Directed Percolation process with corresponding growth rate and growth factor:
The diversity, given by the fraction of nonextinct species, can be obtained as
At the continuous transition, all moments of
This condition has no dependence on the distribution of the interactions; indeed, it is the same one that would be obtained with zero or constant interactions .
For
The reason for this behavior is that at low demographic noise, the abundances of a species with carrying capacity
By a careful (and cumbersome) expansion of the self-consistent equations, we can show that approaching the continuous transition
As noted before, two types of stochasticity contribute to the distribution of abundances. Each species is subjected to demographic and environmental noise, making their abundance a time-dependent random variable. For each species, the abundance is distributed according to the Boltzmann distribution with Hamiltonian
with
We could also be interested in the distribution across species of the abundance averaged over patches or time, given by
Examples of these abundance probability distributions are shown in Fig. .
(a) Probability distribution of the abundance
The divergence of response functions when approaching a tipping point is a generic feature of saddle node bifurcations . Let us consider a generic dynamical system, described by
The stationary point is stable if the Jacobian of
We show below how this mechanism is at play in our case when approaching the stability limit of the self-sustained phase. We study the response of the system to a perturbation in the environmental conditions in the case of independent interaction coefficients (
The response of the order parameters to a variation of
Thanks to the fact that
Substituting
We collect the three order parameters in a vector
where
The response to a variation of
We expect the same qualitative behavior of the response to perturbations for generic values of
In the case of fixed interaction matrices, a finite fraction of the species goes extinct; the interaction matrix restricted to the surviving species has a smaller mean than the starting one. The statistics of the reduced interaction matrix can be computed at 0 temperature :
Since we are not at 0 temperature, in our case this formula is only an approximation, but it provides a useful estimate of the variation of the mean interaction. We find that the interaction mean decreases (more mutualistic) when decreasing the diffusion coefficient [Fig. ]; it is negative in the entire metastability region. In Fig. we show the distribution of the interaction coefficients considering all species or only surviving ones in numerical simulations. The distribution of the interaction coefficients is slightly shifted to more negative values, and indeed
Left: Analytical estimate of the mean of the reduced interaction matrix using Eq. for
To compute the average interaction term, we can again use the cavity method and imagine to add a species (with index 0) to the community. Using Eq. ,
We can now average it over all species (all values of
Note that we will always find
In the case of independent interaction matrices, all species survive, so that the interaction matrix is not modified.
The numerical simulation of demographic noise poses some technical challenges. Naively sampling it as a Gaussian variable can result in negative species abundances, an unphysical result that makes the scheme numerically unstable. A clever solution was found in Ref. , and improved in . The idea is to separate the process in a deterministic part:
and a stochastic one:
At each time step we numerically integrate the two in sequence. For the stochastic part, an exact solution of the associated Fokker-Planck equation is available for any initial condition, and it can be efficiently sampled using Gamma and Poisson variables:
For the deterministic part, we rely on the Euler method,
Some of the challenges encountered in numerical simulations become clear examining the time evolution of the average abundances (Figs. and ). At high temperature (top) the average abundance fluctuates significantly even with a large number of species and patches (
Time evolution of the average abundance for two different temperatures and two average values of the initial conditions. Note the different time ranges in the top and bottom figures: at high temperature the abundances have converged to their asymptotic values at
Time evolution of the average abundance with partial correlation between patches (
References (80)
- C. A. Lozupone, J. I. Stombaugh, J. I. Gordon, J. K. Jansson, and R. Knight, Diversity, stability and resilience of the human gut microbiota, Nature (London) 489, 220 (2012).
- D. A. Kessler and N. M. Shnerb, Generalized model of island biodiversity, Phys. Rev. E 91, 042705 (2015).
- G. Bunin, Ecological communities with Lotka-Volterra dynamics, Phys. Rev. E 95, 042414 (2017).
- G. Biroli, G. Bunin, and C. Cammarota, Marginally stable equilibria in critical ecosystems, New J. Phys. 20, 083051 (2018).
- A. Altieri, F. Roy, C. Cammarota, and G. Biroli, Properties of equilibria and glassy phases of the random Lotka-Volterra model with demographic noise, Phys. Rev. Lett. 126, 258301 (2021).
- T. Galla, Dynamically evolved community size and stability of random Lotka-Volterra ecosystems(a), Europhys. Lett. 123, 48004 (2018).
- T. L. Rogers, B. J. Johnson, and S. B. Munch, Chaos is not rare in natural ecosystems, Nat. Ecol. Evol. 6, 1105 (2022).
- T. Gross, W. Ebenhöh, and U. Feudel, Long food chains are in general chaotic, Oikos 109, 135 (2005).
- M. A. Leibold, M. Holyoak, N. Mouquet, P. Amarasekare, J. M. Chase, M. F. Hoopes, R. D. Holt, J. B. Shurin, R. Law, D. Tilman, M. Loreau, and A. Gonzalez, The metacommunity concept: A framework for multi-scale community ecology, Ecol. Lett. 7, 601 (2004).
- M. P. Hassell, H. N. Comins, and R. M. Mayt, Spatial structure and chaos in insect population dynamics, Nature (London) 353, 255 (1991).
- M. Mobilia, I. T. Georgiev, and U. C. Täuber, Phase transitions and spatio-temporal fluctuations in stochastic lattice Lotka–Volterra models, J. Stat. Phys. 128, 447 (2007).
- F. Olmeda and S. Rulands, Long-range interactions and disorder facilitate pattern formation in spatial complex systems, arXiv:2303.12611.
- U. Dobramysl, M. Mobilia, M. Pleimling, and U. C. Tauber, Stochastic population dynamics in spatially extended predator-prey systems, J. Phys. A 51, 063001 (2018).
- F. Roy, M. Barbier, G. Biroli, and G. Bunin, Complex interactions can create persistent fluctuations in high-diversity ecosystems, PLoS Comput. Biol. 16, e1007827 (2020).
- M. T. Pearce, A. Agarwala, and D. S. Fisher, Stabilization of extensive fine-scale diversity by ecologically driven spatiotemporal chaos, Proc. Natl. Acad. Sci. USA 117, 14572 (2020).
- J. Denk and O. Hallatschek, Self-consistent dispersal puts tight constraints on the spatiotemporal organization of species-rich metacommunities, Proc. Natl. Acad. Sci. USA 119, e2200390119 (2022).
- J. W. Baron and T. Galla, Dispersal-induced instability in complex ecosystems, Nat. Commun. 11, 6032 (2020).
- I. A. van de Leemput, E. H. van Nes, and M. Scheffer, Resilience of alternative states in spatially extended ecosystems, PLoS ONE 10, e0116859 (2015).
- R. M. May, Stability and Complexity in Model Ecosystems (Princeton University Press, Princeton, NJ, 1974), Vol. 1.
- J. Grilli, Macroecological laws describe variation and diversity in microbial communities, Nat. Commun. 11, 4743 (2020).
- S. Azaele, S. Suweis, J. Grilli, I. Volkov, J. R. Banavar, and A. Maritan, Statistical mechanics of ecological systems: Neutral theory and beyond, Rev. Mod. Phys. 88, 035003 (2016).
- A. Kamenev, B. Meerson, and B. Shklovskii, How colored environmental noise affects population extinction, Phys. Rev. Lett. 101, 268103 (2008).
- F. Larroya and T. Galla, Demographic noise in complex ecological communities, J. Phys.: Complex. 4, 025012 (2023).
- D. A. Vasseur and P. Yodzis, The color of environmental noise, Ecology 85, 1146 (2004).
- O. L. Petchey, A. Gonzalez, and H. B. Wilson, Effects on population persistence: The interaction between environmental noise colour, intraspecific competition and space, Proc. R. Soc. London, Ser. B 264, 1841 (1997).
- J. Realpe-Gomez, M. Baudena, T. Galla, A. J. McKane, and M. Rietkerk, Demographic noise and resilience in a semi-arid ecosystem model, Ecol. Complex. 15, 97 (2013).
- G. Bell and Associate Editor: Dolph Schluter, The distribution of abundance in neutral communities, Am. Natural. 155, 606 (2000).
- L. G. Shoemaker, L. L. Sullivan, I. Donohue, J. S. Cabral, R. J. Williams, M. M. Mayfield, J. M. Chase, C. Chu, W. Stanley Harpole, A. Huth, J. HilleRisLambers, A. R. M. James, N. J. B. Kraft, F. May, R. Muthukrishnan, S. Satterlee, F. Taubert, X. Wang, T. Wiegand, Q. Yang, and K. C. Abbott, Integrating the underlying structure of stochasticity into community ecology, Ecology 101, e02922 (2020).
- F. Peruzzo, M. Mobilia, and S. Azaele, Spatial patterns emerging from a stochastic process near criticality, Phys. Rev. X 10, 011032 (2020).
- R. H. MacArthur and E. O. Wilson, The Theory of Island Biogeography (Princeton University Press, Princeton, NJ, 1967).
- J. Hu, D. R. Amor, M. Barbier, G. Bunin, and J. Gore, Emergent phases of ecological diversity and dynamics mapped in microcosms, Science 378, 85 (2022).
- G. G. Lorenzana and A. Altieri, Well-mixed Lotka-Volterra model with random strongly competitive interactions, Phys. Rev. E 105, 024307 (2022).
- M. Loreau, N. Mouquet, and A. Gonzalez, Biodiversity as spatial insurance in heterogeneous landscapes, Proc. Natl. Acad. Sci. USA 100, 12765 (2003).
- P. Chesson, General theory of competitive coexistence in spatially-varying environments, Theor. Popul. Biol. 58, 211 (2000).
- D. Gravel, F. Massol, and M. A. Leibold, Stability and complexity in model meta-ecosystems, Nat. Commun. 7, 12457 (2016).
- S. Pettersson and M. N. Jacobi, Spatial heterogeneity enhance robustness of large multi-species ecosystems, PLoS Comput. Biol. 17, e1008899 (2021).
- A. Mahadevan, M. T. Pearce, and D. S. Fisher, Spatiotemporal ecological chaos enables gradual evolutionary diversification without niches or tradeoffs, eLife 12, e82734 (2023).
- S. R. Broadbent and J. M. Hammersley, Percolation processes: I. Crystals and mazes, Math. Proc. Cambridge Philos. Soc. 53, 629 (1957).
- H. K. Janssen, Spontaneous symmetry breaking in directed percolation with many colors: Differentiation of species in the gribov process, Phys. Rev. Lett. 78, 2890 (1997).
- H. Hinrichsen, Non-equilibrium critical phenomena and phase transitions into absorbing states, Adv. Phys. 49, 815 (2000).
- M. Scheffer, S. Carpenter, J. A. Foley, C. Folke, and B. Walker, Catastrophic shifts in ecosystems, Nature (London) 413, 591 (2001).
- S. Kéfi, M. Rietkerk, M. van Baalen, and M. Loreau, Local facilitation, bistability and transitions in arid ecosystems, Theor. Popul. Biol. 71, 367 (2007).
- T. M. Lenton, H. Held, E. Kriegler, J. W. Hall, W. Lucht, S. Rahmstorf, and H. J. Schellnhuber, Tipping elements in the Earth's climate system, Proc. Natl. Acad. Sci. USA 105, 1786 (2008).
- J.-P. Bouchaud, Crises and collective socio-economic phenomena: Simple models and challenges, J. Stat. Phys. 151, 567 (2013).
- J. Denk and O. Hallatschek, Tipping points emerge from weak mutualism in metacommunities, Ecology (to be published).
- M. Mezard, G. Parisi, and M. Virasoro, Spin Glass Theory and Beyond: An Introduction to the Replica Method and Its Applications, Volume 9 of World Scientific Lecture Notes in Physics (World Scientific, Singapore, 1986).
- R. M. May, Will a large complex system be stable? Nature (London) 238, 413 (1972).
- A. Altieri and G. Biroli, Effects of intraspecific cooperative interactions in large ecosystems, SciPost Phys. 12, 013 (2022).
- C. K. Fisher and P. Mehta, The transition between the niche and neutral regimes in ecology, Proc. Natl. Acad. Sci. USA 111, 13111 (2014).
- E. Pigani, D. Sgarbossa, S. Suweis, A. Maritan, and S. Azaele, Delay effects on the stability of large ecosystems, Proc. Natl. Acad. Sci. USA 119, e2211449119 (2022).
- S. Suweis, F. Ferraro, S. Azaele, and A. Maritan, Generalized lotka-volterra systems with time correlated stochastic interactions, arXiv:2307.02851.
- M. Barbier, J. F. Arnoldi, G. Bunin, and M. Loreau, Generic assembly patterns in complex ecological communities, Proc. Natl. Acad. Sci. USA 115, 2156 (2018).
- U. Behn, J. L. van Hemmen, and B. Sulzer, Memory B cells stabilize cycles in a repressive network, in Theoretical and Experimental Insights into Immunology (Springer, Berlin, 1992), pp. 249–260.
- J. Moran and J.-P. Bouchaud, May's instability in large economies, Phys. Rev. E 100, 032307 (2019).
- R. M. Goodwin, Chaotic Economic Dynamics (Oxford University Press, Oxford, 1990).
- I. M. Bomze, Lotka-Volterra equation and replicator dynamics: New issues in classification, Biol. Cybern. 72, 447 (1995).
- G. Bunin, Interaction patterns and diversity in assembled ecological communities, arXiv:1607.04734.
- W. Feller, Two singular diffusion problems, Ann. Math. 54, 173 (1951).
- I. Dornic, H. Chaté, and M. A. Muñoz, Integration of langevin equations with multiplicative noise and the viability of field theories for absorbing phase transitions, Phys. Rev. Lett. 94, 100601 (2005).
- J. L. Cardy and R. L. Sugar, Directed percolation and Reggeon field theory, J. Phys. A 13, L423 (1980).
- F. Roy, G. Biroli, G. Bunin, and C. Cammarota, Numerical implementation of dynamical mean field theory for disordered systems: Application to the Lotka–Volterra model of ecosystems, J. Phys. A 52, 484001 (2019).
- R. Zwanzig, Nonequilibrium Statistical Mechanics (Oxford University Press, Oxford, 2001).
- A. Georges, G. Kotliar, W. Krauth, and M. J. Rozenberg, Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions, Rev. Mod. Phys. 68, 13 (1996).
- L. F. Cugliandolo, Recent applications of dynamical mean-field methods, arXiv:2305.01229.
- G. B. Arous and A. Guionnet, Symmetric Langevin spin glass dynamics, Ann. Probab. 25, 1367 (1997).
- S. Chen and U. C. Täuber, Non-equilibrium relaxation in a stochastic lattice Lotka–Volterra model, Phys. Biol. 13, 025005 (2016).
- P. C. Martin, E. D. Siggia, and H. A. Rose, Statistical dynamics of classical systems, Phys. Rev. A 8, 423 (1973).
- H.-K. Janssen, On a Lagrangian for classical field dynamics and renormalization group calculations of dynamical critical properties, Z. Phys. B 23, 377 (1976).
- C. De Dominicis, Dynamics as a substitute for replicas in systems with quenched random impurities, Phys. Rev. B 18, 4913 (1978).
- C. Aron, G. Biroli, and L. F. Cugliandolo, Symmetries of generating functionals of Langevin processes with colored multiplicative noise, J. Stat. Mech. (2010) P11018.
- J. Wu, P. Mehta, and D. Schwab, Understanding species abundance distributions in complex ecosystems of interacting species, arXiv:2103.02081.
- J. W. Baron, T. J. Jewell, C. Ryder, and T. Galla, Breakdown of random-matrix universality in persistent lotka-volterra communities, Phys. Rev. Lett. 130, 137401 (2023).
- The discontinuous transition takes place slightly before the analytical prediction. Besides finite size effects, we note that this phenomenon is to be expected for this kind of transition. In fact, when the red curve (low initial condition) jumps to high abundance, this does not necessarily indicate that the
solution has become locally unstable, but rather that its basin of attraction has shrunk and does not include the considered initial condition anymore. It is therefore to be expected that this occurs for . A similar phenomenon takes place for spinodal transition in physics. - At
the dynamics is so slow (especially close to the tipping points) that at some of the abundances have not yet converged to their asymptotic values. This leads to an apparent smoothing of the discontinuous transition, whose existence is nevertheless ensured by the abrupt change of behavior of the evolution of the abundance (see Fig. 16), analogous to the one observed for . - T. M. Lenton, Environmental tipping points, Ann. Rev. Environ. Resourc. 38, 1 (2013).
- M. Scheffer, S. R. Carpenter, T. M. Lenton, J. Bascompte, W. Brock, V. Dakos, J. Van de Koppel, I. A. Van de Leemput, S. A. Levin, E. H. Van Nes et al., Anticipating critical transitions, Science 338, 344 (2012).
- V. Dakos, C. A. Boulton, J. E. Buxton, J. F. Abrams, D. I. Armstrong McKay, S. Bathiany, L. Blaschke, N. Boers, D. Dylewsky, C. López-Martínez, I. Parry, P. Ritchie, B. van der Bolt, L. van der Laan, E. Weinans, and S. Kéfi, Tipping point detection and early-warnings in climate, ecological, and human systems, EGUsphere (2023).
- A. Altieri, G. Biroli, and C. Cammarota, Dynamical mean-field theory and aging dynamics, J. Phys. A 53, 375006 (2020).
- C. Aron, D. G. Barci, L. F. Cugliandolo, Z. G. Arenas, and G. S. Lozano, Dynamical symmetries of Markov processes with multiplicative white noise, J. Stat. Mech. (2016) 053207.
- H. Weissmann, N. M. Shnerb, and D. A. Kessler, Simulation of spatial systems with demographic noise, Phys. Rev. E 98, 022131 (2018).