- You're Mobile Enabled
- Access by Universitaet Linz
Diverse communities behave like typical random ecosystems
Phys. Rev. E 104, 034416 – Published 27 September, 2021
Abstract
In 1972, Robert May triggered a worldwide research program studying ecological communities using random matrix theory. Yet, it remains unclear if and when we can treat real communities as random ecosystems. Here, we draw on recent progress in random matrix theory and statistical physics to extend May's approach to generalized consumer-resource models. We show that in diverse ecosystems adding even modest amounts of noise to consumer preferences results in a transition to “typicality,” where macroscopic ecological properties of communities are indistinguishable from those of random ecosystems, even when resource preferences have prominent designed structures. We test these ideas using numerical simulations on a wide variety of ecological models. Our work offers an explanation for the success of random consumer resource models in reproducing experimentally observed ecological patterns in microbial communities and highlights the difficulty of scaling up bottom-up approaches in synthetic ecology to diverse communities.
Physics Subject Headings (PhySH)
Article Text
One of the most stunning aspects of the natural world is the immense diversity of ecological communities, ranging from rainforests to human microbiomes. Ecological communities are critical for numerous processes ranging from global water cycling processes to animal development and host health . For this reason, understanding the principles governing community assembly and function in diverse communities has wide ranging applications from conservation efforts to pharmaceutical engineering and bioremediation .
Many traditional ecological models focus on ecosystems consisting of a few species and resources. In such low dimensional models, it is often possible to characterize the ecological traits of all the species and resources and then use this information to make predictions about community-level properties . However, many natural communities are extremely diverse and the models and parameters are naturally high dimensional. This problem is especially pronounced in the context of microbial ecology where hundreds of species can coexist in a single location. In this case, a comprehensive parametrization of species and resource traits is no longer feasible, suggesting that new ideas and concepts are required to understand diverse communities.
A similar problem is encountered in statistical physics. For example, an ideal gas is characterized by the unit mole, which has the order of
In 1972, Robert May suggested that large complex ecosystems can also be modeled as random systems . May considered a diverse ecosystem composed of
In May's model, all ecosystem properties are encoded in the species-species interaction matrix. A major limitation of these models is that they neglect resource dynamics, making it difficult to understand how ecosystem properties depend on both the external environment and species consumer preferences. For this reason, community assembly is often analyzed using generalized consumer resource models (CRMs) . In these models, species are modeled as consumer that can consume resources, and sometimes also produce resources . Recently, we have shown that such models, initialized with random parameters, can predict laboratory experiments on complex microbial communities and reproduce large-scale ecological patterns observed in field surveys, including the Earth and Human Microbiome Projects . This suggests that the large-scale, reproducible patterns we see across microbiomes are emergent features of random ecosystems.
Yet, it remains unclear why random ecosystems can accurately describe real ecological communities. To answer these questions, in this paper we exploit ideas from random matrix theory and statistical physics to analyze generalized consumer-resource models in spirit of May's original analysis. We show that the macroscopic ecological properties of diverse ecosystems can be described using random ecosystems, much like thermodynamic quantities like pressure and average energy of the ideal gas can be described by considering particles to be random and independent. To explore these ideas, we devised a more concrete version of May's original thought experiment describing an ecosystem consisting of The original MacArthur consumer resource model consists of The consumption rate of species In the beginning, we focus primarily on a popular variant of the original CRM introduced by Tilman with slightly different resource dynamics : From an ecological perspective, there are significant differences between this model variant and the original CRM. First, the resource supply rate Both the models in Eqs. and make very specific assumptions about resource dynamics. To check the generality of our results, we also numerically analyzed generalizations of the CRM, including linear resource dynamics where resources are supplied externally, and a model of microbial ecology with trophic feedbacks where organisms can feed each other via metabolic byproducts . This analysis can be found in Appendix . Furthermore, for simplicity, in most of this work we assume that In CRMs, the identity of each species is specified by its consumption preferences. In real ecosystems, it is well established that organisms can exhibit strong consumer preferences for particular resources. However, recent work has shown that consumer resource models with random consumer preferences can reproduce experimental observations in field surveys and laboratory experiments . To understand this phenomena, we asked how adding noise to consumer preferences changes macroscopic ecosystem level properties like diversity and average productivity. To do so, we considered a thought experiment where we started with predesigned consumer resource preference, and then added “noise” to the consumer resource preferences. Mathematically, we can decompose the consumer matrix where For simplicity, we started with noninteracting species where each species consumes its own resource. A set of noninteracting species can be constructed by engineering each species to consume a different resource type, with no overlap between consumption preferences. For example, one can imagine designing strains of Escherichia coli, where each strain expresses transporters only for a single carbon source with all other transporters edited out of the genome: i.e., a strain that can only transport lactose, another strain that can only transport sucrose, etc. An ecosystem with such consumer preference structure is shown in Fig. . In such an experiment, horizontal gene transfer would eventually begin distributing transporter genes from one strain to another, so a realistic model would have to allow for some amount of unintended, “off-target” resource consumption. In line with May, we can model the consumer preferences Random interactions destabilize an ecosystem of specialist consumers. (a) Left: an ecosystem with system size
Figure shows how the number of surviving species at steady-state changes as one adds more and more nonspecific resource preferences to an ecosystem initially composed of noninteracting species. As in May's analysis, the appropriate measure of the importance of the random component is the root-mean-squared off-target consumption
Remarkably, the fraction of surviving species converges to the same value as for a completely random consumer preference matrix and remains finite as
Community properties for structured and random ecosystems. (a) Examples of designed interactions Top: the identity matrix; Middle: a Gaussian-type circulant matrix; Bottom: a block matrix (see Appendix for details). Simulations of designed and random ecosystems where the random component of the the consumer preferences
This convergence to random ecosystem behavior is quite robust, and holds for other choices of designed consumer preferences beyond the identity matrix considered above. Figure shows numerical simulations of the diversity
The character of the self-organized state is also robust to changes in the sampling scheme for the random component of the consumer preferences. Gaussian noise in consumer preferences simplifies the analytic calculations but also sometimes results in nonphysical negative values for consumer preferences. We therefore tested two sampling schemes that always produce positive values for consumer preferences: uniformly sampling the random component of preferences
To better understand why mass extinctions happen at
with the species-species interaction matrix given by
(see Fig. and Appendix for details). This matrix is related to May's community matrix governing stability
Effect of random interactions on ecosystem sensitivity. (a) The bipartite interactions
As shown in Fig. , the behavior broadly falls into one of three different regimes depending on the amount of noise introduced in the consumer preferences: a low-noise regime when In the low-noise regime, the engineered structure in the consumer preference controls large scale ecological properties. Furthermore, the eigenvalue spectrum of the LV-interaction matrix With increasing In this regime, we observe two new phenomena that were not accessible in May's original framework. First, the spectrum of the species-species interaction matrix where Second, as As one increases The above analysis suggests that The spectrum of Figure shows that this quantity is initially constant as Note that our results are not restricted to Gaussian noise but also apply to the other cases where the noise in consumer preferences is binary or uniform (see Figs. and ). This is because the central limit theorem guarantees that the statistics of eigenvalues of large random matrices converge to the statistics in Gaussian random matrices for many biologically plausible choices of consumer preferences.
Thus far we have focused on a CRM without resource extinctions specified by Eqs. . As discussed extensively in Appendix , if we allow for resource extinction [Eqs. ] and write
instead of Eq. , then, somewhat surprisingly, our cavity method predicts a second-order phase transition to typicality rather than a cross-over as is the case without resource extinction. The signature of such a second-order transition is the divergence of the susceptibility matrix
Effect of resource extinction on an ecosystem. A schematic for the consumer preference matrix with (a) and without (b) resource extinction for specialist consumers that each eat independent resources. The left schematic corresponds to the initial consumer matrix, and the right schematic to the consumer matrix after species and resource extinctions. Notice that resource extinctions can result in singular consumer matrices. (c) Spectra of
The existence of zero modes can be understood by noting that resource extinction and species extinction correspond to the column and row deletion in the consumption matrix [shown in Fig. ]. Such deletions can change the engineered component of the effective consumer preferences for surviving species and resources, resulting in large fluctuations in the interaction matrix
It is common practice in theoretical ecology to model ecosystems using random matrices. Yet it remains unclear if and when we can treat real communities as random ecosystems. Here, we investigated this question by generalizing May's analysis to consumer resource models and asking when the macroscopic, community-level properties can be accurately predicted using random parameters. We found that introducing even modest amount of stochasticity into consumer preferences ensures that the macroscopic properties of diverse ecosystems will be indistinguishable from those of a completely random ecosystem. Our calculations and numerics suggest that transition to typicality occurs when the total amount of off-target resource consumption becomes comparable to the consumption rate of targeted resources.
We confirmed our analytic calculations using numerical simulations on CRMs with different types of resource dynamics and different classes of nonspecific interactions. We emphasize that despite the fact that random ecosystems can make accurate predictions about macroscopic properties like the average diversity or productivity, they will in general fail to capture species level details. This phenomena is well understood in the context of statistical physics where it is possible to predict thermodynamic quantities such as pressure and temperature even though one cannot accurately predict microstates.
These observations may help explain the surprising success of consumer resource models with random parameters in predicting the behavior of microbial ecosystems in the laboratory and natural environments . They also suggest that maybe possible to predict macroscopic ecosystem level properties like diversity or total biomass even when ecosystems are poorly characterized or have lots of missing data.
The foregoing analysis has several other interesting implications. First, it suggests that bottom-up engineering of complex ecosystems may prove to be very difficult. As the number of components increases, small uncertainties in each of the interaction parameters may eventually overwhelm the designed interactions, and destabilize the intended steady state. Instead, such system are much more likely end up in a typical state which our theory suggests is much more stable than the intended designed state as ecosystems become more diverse.
Our work also suggests that in ecosystems well described by consumer resource models, crossing a May-like transition generically gives rise to typical random ecosystems rather than a marginal stable phase as was found in a recent analysis of the Generalized Lotka-Volterra model (an important caveat to this statement is that adding nonresource based interactions to consumer resource models can restore complicated behavior reminiscent of the marginally stable phase ). This suggests that even when cumulative parameter uncertainties preclude a detailed characterization of an ecosystem, methods from statistical physics and Random Matrix Theory can be employed to predict system-level properties . It will be interesting to explore if and how these insights can be exploited to design top-down control strategies for ecosystems and identify assembly rules for microbial communities with many species .
In this paper, we only consider white noise, which is independently and identically added to all interaction components. In the future, it will be interesting to ask how other specialized noises, resulting from demographic stochasticity, phenotypic variation, can affect our results. Based on our experience, we expect that, even in these more complicated ecosystems, our conclusion will hold quite generally in the thermodynamics limit. But much more work needs to be done to confirm if this intuition is really correct.
We thank Josh Goldford, Zhenyu Liao, Jason Rocks, Guangwei Si, Jean Vila, and Yu Hu for helpful discussions. We also especially appreciate numerous valuable comments from Stefano Allesina. The work was supported by NIH NIGMS Grant No. 1R35GM119461, Simons Investigator in the Mathematical Modeling of Living Systems (MMLS). The authors are pleased to acknowledge that the computational work reported on in this paper was performed on the Shared Computing Cluster which is administered by Boston University Research Computing Services.
We primarily analyze CRMs of the form given by Eqs. and . To do so, we decompose the consumer matrix
with
For all simulations, unless otherwise specified, the default choices for parameters are: To test the generality of our results, we also simulated more complicated variants of the consumer resource model (see Fig. and Appendix ). First, we simulated a consumer resource model with linear resource dynamics : In this model resources are supplied externally at a rate rather than described by a logistic growth. This small change in resource dynamics can significantly change the ecosystem properties because it prevents resources from going extinct in the steady state. In the simulations, we set Second, we simulated a generalization of the MacArthur's Consumer Resource model to a model we call the microbial consumer resource model (MicroCRM). The MicroCRM was introduced in Ref. and refined in Ref. to simulate microbial communities. In this model, in addition to consuming resources species can produce new resources through cross-feeding. This dramatically changes the resource dynamics through the introduction of trophic feedbacks. Unlike the original CRM and the CRM with linear resource dynamics, the MicroCRM possesses no Lyapunov function. Full details of the model are available in the appendices of Refs. . In particular, the dynamics we use are described in Eq. (17) of Ref. with the leakage rate
We begin by defining four susceptibility matrices that measure how the steady-state resource and species abundances respond to changes in the resource supply and species death (growth) rates:
where the bar
For the extinct species and resources, by definition the susceptibilities are zero. For this reason, we focus only on the surviving resources and species. At steady-state, Eq. gives
where
Substituting in for the partial derivatives using the susceptibility matrices defined above, we have
These two equations can be written as single matrix equation for block matrices:
To solve this equation, we define a
When the designed component of the consumer preferences is the identity [i.e.,
-
(i)
,𝑀 * 𝑀 = 𝜙 𝑅 , and⟨ 𝑅 ⟩ = 1 𝑀 ∑ 𝛽 𝑅 𝛽 , where𝑞 𝑅 = 1 𝑀 ∑ 𝛽 𝑅 2 𝛽 = ⟨ 𝑅 2 ⟩ is the number of surviving resources.𝑀 * -
(ii)
,𝑆 * 𝑆 = 𝜙 𝑁 , and⟨ 𝑁 ⟩ = 1 𝑆 ∑ 𝑗 𝑁 𝑗 , where𝑞 𝑁 = 1 𝑆 ∑ 𝑗 𝑁 2 𝑗 = ⟨ 𝑁 2 ⟩ is the number of surviving species.𝑆 * -
(iii)
assuming𝐶 𝑖 𝛼 ≡ 𝜇 𝑀 + 𝜎 𝑐 𝑑 𝑖 𝛼 ,⟨ 𝑑 𝑖 𝛼 ⟩ = 0 with⟨ 𝑑 𝑖 𝛼 𝑑 𝑗 𝛽 ⟩ = 𝛿 𝑖 𝑗 𝛿 𝛼 𝛽 𝑀 , ,⟨ 𝑐 𝑖 𝛼 ⟩ = 𝜇 𝑀 .⟨ 𝑐 𝑖 𝛼 𝑐 𝑗 𝛽 ⟩ = 𝜎 2 𝑐 𝑀 𝛿 𝑖 𝑗 𝛿 𝛼 𝛽 + 𝜇 2 𝑀 2 ≈ 𝜎 2 𝑐 𝑀 𝛿 𝑖 𝑗 𝛿 𝛼 𝛽 -
(iv)
, with𝐾 𝛼 = 𝐾 + 𝛿 𝐾 𝛼 ,⟨ 𝐾 𝛼 ⟩ = 1 𝑀 ∑ 𝛽 𝐾 𝛽 = 𝐾 ⟨ 𝛿 𝐾 𝛼 𝛿 𝐾 𝛽 ⟩ = 𝛿 𝛼 𝛽 𝜎 2 𝐾 . -
(v)
, with𝑚 𝑖 = 𝑚 + 𝛿 𝑚 𝑖 ,⟨ 𝑚 𝑖 ⟩ = 𝑚 ⟨ 𝛿 𝑚 𝑖 𝛿 𝑚 𝑗 ⟩ = 𝛿 𝑖 𝑗 𝜎 2 𝑚 . -
(vi)
, and for the identity matrix𝛾 = 𝑀 𝑆 .𝛾 = 1
Following similar steps as in Ref. , we perturb the ecosystem with a new species and resource
Denote by
In what follows we assume replica symmetry. In this case, the sums in the equations above can be approximated as Gaussian random variables. For this reason, it is helpful to introduce new auxiliary random variables:
where
Case 1: Both
Case 2: Either
Case 3: Both
Combining the cases above, the steady-state solution is a Gaussian mixture depending on the positivity of
Cavity equations for the susceptibilities can be obtained directly by differentiating these equations:
In this case, the resource never vanishes so that we can fix
The two solutions above are continuous at the transition point:
The comparison between cavity solutions and numerical simulations for
To understand these solutions and behaviors better, it is helpful to consider three regimes: Regime A where
In Regime B, resource extinction has a significant effect on the system's feasibility, shown in Fig. . With resource extinction, Eq. shows there is a sudden change for the linear response function
Without resource extinction, Eq. shows the linear response function
Without resource and species extinction, as shown in Eq. ,
In Regime C, further increasing of
In Regime A (
For Eqs. without resource extinction, the solutions for the steady states become
For ecosystems without resource and species extinction, the solutions for the steady states become
For Regime C (
in agreement with the equations obtained in Ref. for purely random interactions. For Eqs. without resource extinction, the solutions for the steady states become
For ecosystems without resource and species extinction, the solutions for the steady states become
In this section, we show how the generalized Lotka-Volterra model can be related to the CRM, and in particular, the how the steady states of the two models can be made to coincide. Solving for the steady-state values of the nonextinct resources by setting the bottom equation in Eqs. equal to zero gives
Substituting this into the top equation in Eqs. gives
where we have defined an interaction matrix
where
with community matrix
In May's work,
In this section, we prove that the largest eigenvalue
We start by defining the diagonal matrix
where
Since
Now we note that
We conclude that the eigenvalues of
An alternative but much simpler proof can be provided with the properties of the D-stable matrix . A real square matrix
Our numerical simulations show that after the transition, our ecosystems are well described by purely random interactions. This suggests that we should be able to derive our cavity results using random matrix theory (RMT). We now show that this is indeed the case. Our starting point are the average susceptibilities which are defined as
From the cavity calculations, we only care about
We can combine these equations with Eqs. and to obtain
We now show that the cavity solutions are consistent with results from RMT using Eqs. and in Regime A and Regime C described in the main text. In this regime, where The second line of Eq. is obtained by transferring the integral function to a complex analytic function and applying the residue theorem. This result is the same as the cavity solution Eq. when In Regime B, it hard to estimate the minimum eigenvalue. We can use Stieltjes transformation of information-plus-noise-type matrices which are well studied in wireless communications , where Using Theorem 1.1 in Dozier and Silverstein , the Stieltjes transform The asymptotic spectrum of The result is shown in Fig. . The minimum eigenvalue reaches 0 nearly at We emphasize that the phase transition point, derived from Eq. , does not change at different It shows when
All simulations are done with the CVXPY package . The code is available on GitHub .
-
(1) Figures , , , and : the consumer matrix
is sampled from the Gaussian distribution𝐂 .𝒩 ( 𝜇 𝑀 , 𝜎 𝑐 √ 𝑀 ) ,𝑆 = 1 0 0 ,𝑀 = 1 0 0 ,𝜇 = 0 ,𝐾 = 1 ,𝜎 𝐾 = 0 . 1 , and𝑚 = 0 . 1 , and each data point is averaged from 5000 independent realizations. The model is simulated with Eqs. .𝜎 𝑚 = 0 . 0 1 -
(2) Figure : the consumer matrix
is sampled from the uniform distribution𝐂 .𝒰 ( 0 , 𝑏 ) ,𝑆 = 1 0 0 ,𝑀 = 1 0 0 ,𝜇 = 0 ,𝐾 = 1 ,𝜎 𝐾 = 0 . 1 ,𝑚 = 0 . 1 , and each data point is averaged from 5000 independent realizations. The model is described by Eqs. .𝜎 𝑚 = 0 . 0 1 -
(3) Figure : the consumer matrix
is sampled from the Bernoulli distribution𝐂 .𝐵 𝑒 𝑟 𝑛 𝑜 𝑢 𝑙 𝑙 𝑖 ( 𝑝 𝑐 ) ,𝑆 = 1 0 0 ,𝑀 = 1 0 0 ,𝜇 = 0 ,𝐾 = 1 ,𝜎 𝐾 = 0 . 1 ,𝑚 = 0 . 1 , and each data point is averaged from 5000 independent realizations. The model is described by Eqs. .𝜎 𝑚 = 0 . 0 1 -
(4) Figures , : the simulation is the same as Fig. . Each spectrum is drawn from 10 000 independent realizations.
-
(5) Figure : the consumer matrix
is sampled from the Gaussian distribution𝐂 .𝒩 ( 𝜇 𝑀 , 𝜎 𝑐 √ 𝑀 ) ,𝑆 = 1 0 0 ,𝑀 = 1 0 0 ,𝜇 = 0 ,𝐾 = 1 ,𝜎 𝐾 = 0 . 1 , and𝑚 = 0 . 1 . The model without resource extinction simulated with Eqs. , and each data point is averaged from 5000 independent realizations.. The model with resource extinction is simulated with Eqs. , and each data point is averaged from 4000 independent realizations. Each spectrum is drawn from 1 independent realizations for𝜎 𝑚 = 0 . 0 1 .𝑆 = 5 0 0 -
(6) Figure : the simulation is the same as Fig. . Each histogram is drawn from 10 000 independent realizations.
-
(7) Figure : the consumer matrix
is sampled from the Gaussian distribution𝐂 .𝒩 ( 𝜇 𝑀 , 𝜎 𝑐 √ 𝑀 ) ,𝑆 = 1 0 0 ,𝑀 = 1 0 0 ,𝜇 = 1 ,𝐾 = 1 ,𝜎 𝐾 = 0 . 1 , and𝑚 = 0 . 1 . For (C),𝜎 𝑚 = 0 . 0 1 ,𝜔 = 1 , and model details can be found in Ref. ; for (D), the dynamics is described in Eq. (17) in the Supplemental Material of Ref. . The noise is only applied on the consumption matrix and𝜎 𝜔 = 0 is kept the same at different𝐷 . Each data point is averaged from 4000 independent realizations for (A), from 5000 independent realizations for (B), and from 1000 independent realizations for (C, D).𝜎 𝑐 -
(8) Figure : the simulation is the same as Fig. except
,𝑆 = 2 0 0 . For the identity case, the consumer matrix is obtained by concatenating the𝑀 = 1 0 0 identity plus noise matrix and a𝑀 × 𝑀 Gaussian random matrix. The model without resource extinction simulated with Eqs. , and each data point is averaged from 5000 independent realizations.( 𝑆 − 𝑀 ) × 𝑀 -
(9) Figures , , and : the simulation is the same as Figs. and . The model without resource extinction simulated with Eqs. , and each data point is averaged from 5000 independent realizations. The model with resource extinction is simulated with Eqs. , and each data point is averaged from 4000 independent realizations.
In the main text, we show that the value of species packing
In the quasistatic limit, Eqs. become
which can not be reduced to the Lotka–Volterra model. Therefore, we have to rederive the susceptibility matrix for Eqs. .
To have well-defined susceptibilities, we introduce an auxiliary variable
At the end we can set
where the bar
For the extinct species and resources, by definition the susceptibilities are zero. For this reason, we focus only on the surviving resources and species. At steady state, Eqs. give
where
Substituting in for the partial derivatives using the susceptibility matrices defined above, we have
These two equations can be written as a single matrix equation for block matrices:
where
To solve this equation, we define two
We can see that the new susceptibilities:
The figures below are referred to in the main text.
Reproduce Fig : the fraction of surviving species
Community properties for generalized consumer-resource models under Gaussian noise. (a) MacArthur's consumer resource model with resource extinction. (b) Linear resource dynamics: the resource dynamics is changed to
Spectra of
Comparison between numerical simulations (scatter points) and cavity solutions (solid lines) for
Comparison of the minimum eigenvalue
References (48)
- D. V. Spracklen, S. R. Arnold, and C. Taylor, Observations of increased tropical rainfall preceded by air passage over forests, Nature (London) 489, 282 (2012).
- Y. Belkaid and T. W. Hand, Role of the microbiota in immunity and inflammation, Cell 157, 121 (2014).
- J. I. Prosser, B. J. Bohannan, T. P. Curtis, R. J. Ellis, M. K. Firestone, R. P. Freckleton, J. L. Green, L. E. Green, K. Killham, J. J. Lennon et al., The role of ecological theory in microbial ecology, Nat. Rev. Microbiol. 5, 384 (2007).
- J. Friedman, L. M. Higgins, and J. Gore, Community structure follows simple assembly rules in microbial microcosms, Nat. Ecol. Evol. 1, 0109 (2017).
- J. Friedman and J. Gore, Ecological systems biology: The dynamics of interacting populations, Curr. Opin. Syst. Biol. 1, 114 (2017).
- C. Ratzke, J. Denk, and J. Gore, Ecological suicide in microbes, Nat. Ecol. Evol. 2, 867 (2018).
- S.-K. Ma, Modern Theory of Critical Phenomena (Routledge, London, UK, 2018).
- R. M. May, Will a large complex system be stable? Nature (London) 238, 413 (1972).
- J. Ginibre, Statistical ensembles of complex, quaternion, and real matrices, J. Math. Phys. 6, 440 (1965).
- T. Gross, L. Rudolf, S. A. Levin, and U. Dieckmann, Generalized models reveal stabilizing factors in food webs, Science 325, 747 (2009).
- S. Allesina and S. Tang, Stability criteria for complex ecosystems, Nature (London) 483, 205 (2012).
- R. MacArthur and R. Levins, The limiting similarity, convergence, and divergence of coexisting species, Am. Nat. 101, 377 (1967).
- D. Tilman, Resource Competition and Community Structure (Princeton University Press, Princeton, NJ, 1982).
- J. E. Goldford, N. Lu, D. Bajić, S. Estrela, M. Tikhonov, A. Sanchez-Gorostiaga, D. Segrè, P. Mehta, and A. Sanchez, Emergent simplicity in microbial community assembly, Science 361, 469 (2018).
- R. Marsland, W. Cui, J. Goldford, and P. Mehta, The community simulator: A python package for microbial ecology, PLoS One 15, e0230430 (2020).
- R. Marsland, W. Cui, and P. Mehta, A minimal model for microbial biodiversity can reproduce experimentally observed ecological patterns, Sci. Rep. 10, 3308 (2020).
- M. Tikhonov and R. Monasson, Collective Phase in Resource Competition in a Highly Diverse Ecosystem, Phys. Rev. Lett. 118, 048103 (2017).
- A. Posfai, T. Taillefumier, and N. S. Wingreen, Metabolic Trade-Offs Promote Diversity in a Model Ecosystem, Phys. Rev. Lett. 118, 028103 (2017).
- R. Marsland III, W. Cui, J. Goldford, A. Sanchez, K. Korolev, and P. Mehta, Available energy fluxes drive a transition in the diversity, stability, and functional structure of microbial communities, PLoS Comput. Biol. 15, e1006793 (2019).
- C. A. Serván, J. A. Capitán, J. Grilli, K. E. Morrison, and S. Allesina, Coexistence of many species in random ecosystems, Nat. Ecol. Evol. 2, 1237 (2018).
- G. Bunin, Ecological communities with Lotka-Volterra dynamics, Phys. Rev. E 95, 042414 (2017).
- M. Advani, G. Bunin, and P. Mehta, Statistical physics of community ecology: A cavity solution to MacArthur's consumer resource model, J. Stat. Mech.: Theory Exp. (2018) 033406.
- F. J. Dyson, Statistical theory of the energy levels of complex systems, J. Math. Phys. 3, 140 (1962).
- P. Chesson, Macarthur's consumer-resource model, Theor. Popul. Biol. 37, 26 (1990).
- I. Dalmedigos and G. Bunin, Dynamical persistence in high-diversity resource-consumer communities, PLoS Comput. Biol. 16, e1008189 (2020).
- R. P. Rohr, S. Saavedra, and J. Bascompte, On the structural stability of mutualistic systems, Science 345, 1253497 (2014).
- V. A. Marčenko and L. A. Pastur, Distribution of eigenvalues for some sets of random matrices, Math. USSR Sb. 1, 457 (1967).
- T. Rogers, I. P. Castillo, R. Kühn, and K. Takeda, Cavity approach to the spectral density of sparse symmetric random matrices, Phys. Rev. E 78, 031116 (2008).
- A. Altieri and S. Franz, Constraint satisfaction mechanisms for marginal stability and criticality in large ecosystems, Phys. Rev. E 99, 010401(R) (2019).
- E. Agliari, F. Alemanno, A. Barra, and A. Fachechi, On the Marchenko-Pastur law in analog bipartite spin-glasses, J. Phys. A: Math. Theor. 52, 254002 (2019).
- R. Couillet and M. Debbah, Random Matrix Methods for Wireless Communications (Cambridge University Press, Cambridge, UK, 2011).
- P. Loubaton, P. Vallet et al., Almost sure localization of the eigenvalues in a Gaussian information plus noise model: Application to the spiked models, Electron. J. Probab. 16, 1934 (2011).
- G. Biroli, G. Bunin, and C. Cammarota, Marginally stable equilibria in critical ecosystems, New J. Phys. 20, 083051 (2018).
- 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: Math. Theor. 52, 484001 (2019).
- 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).
- M. Barbier, J.-F. Arnoldi, G. Bunin, and M. Loreau, Generic assembly patterns in complex ecological communities, Proc. Natl. Acad. Sci. U.S.A. 115, 2156 (2018).
- S. Butler and J. P. O'Dwyer, Stability criteria for complex microbial communities, Nat. Commun. 9, 2970 (2018).
- https://github.com/Emergent-Behaviors-in-Biology/typical-random-ecosystems.
- W. Cui, R. Marsland III, and P. Mehta, Effect of Resource Dynamics on Species Packing in Diverse Ecosystems, Phys. Rev. Lett. 125, 048101 (2020).
- L. Hogben, Handbook of Linear Algebra (Chapman and Hall/CRC, London, UK, 2013).
- J. Grilli, M. Adorisio, S. Suweis, G. Barabás, J. R. Banavar, S. Allesina, and A. Maritan, Feasibility and coexistence of large ecological communities, Nat. Commun. 8, 14389 (2017).
- R. B. Dozier and J. W. Silverstein, On the empirical distribution of eigenvalues of large dimensional information-plus-noise-type matrices, J. Multivariate Anal. 98, 678 (2007).
- J. Baik, G. B. Arous, S. Péché et al., Phase transition of the largest eigenvalue for nonnull complex sample covariance matrices, Ann. Prob. 33, 1643 (2005).
- F. Benaych-Georges and R. R. Nadakuditi, The singular values and vectors of low rank perturbations of large rectangular random matrices, J. Multivariate Anal. 111, 120 (2012).
- A. Agrawal, R. Verschueren, S. Diamond, and S. Boyd, A rewriting system for convex optimization problems, J. Control Decis. 5, 42 (2018).
- R. Marsland III, W. Cui, and P. Mehta, The minimum environmental perturbation principle: A new perspective on niche theory, Am. Nat. 196, 291 (2020).
- P. Mehta, W. Cui, C.-H. Wang, and R. Marsland III, Constrained optimization as ecological dynamics with applications to random quadratic programming in high dimensions, Phys. Rev. E 99, 052111 (2019).
- W. Cui, J. W. Rocks, and P. Mehta, The perturbative resolvent method: Spectral densities of random matrix ensembles via perturbation theory, arXiv:2012.00663.