Skip to main content
Log in

Equilibrium and surviving species in a large Lotka–Volterra system of differential equations

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Lotka–Volterra (LV) equations play a key role in the mathematical modeling of various ecological, biological and chemical systems. When the number of species (or, depending on the viewpoint, chemical components) becomes large, basic but fundamental questions such as computing the number of surviving species still lack theoretical answers. In this paper, we consider a large system of LV equations where the interactions between the various species are a realization of a random matrix. We provide conditions to have a unique equilibrium and present a heuristics to compute the number of surviving species. This heuristics combines arguments from Random Matrix Theory, mathematical optimization (LCP), and standard extreme value theory. Numerical simulations, together with an empirical study where the strength of interactions evolves with time, illustrate the accuracy and scope of the results.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  • Akjouj I, Najim J (2022) Feasibility of sparse large Lotka–Volterra ecosystems. J Math Biol 85(6–7):66

    Article  MathSciNet  MATH  Google Scholar 

  • Akjouj I, Hachem W, Maïda M, Najim J (2023) Equilibria of large random Lotka–Volterra systems with vanishing species: a mathematical approach. arXiv preprint arXiv:2302.07820

  • Allesina S, Tang S (2012) Stability criteria for complex ecosystems. Nature 483(7388):205

    Article  Google Scholar 

  • Arnoldi J-F, Bideault A, Loreau M, Haegeman B (2018) How ecosystems recover from pulse perturbations: a theory of short- to long-term responses. J Theor Biol 436:79–92

    Article  MathSciNet  MATH  Google Scholar 

  • Bai ZD, Silverstein JW (2010) Spectral analysis of large dimensional random matrices, 2nd edn. Springer series in statistics. Springer, New York

    Book  MATH  Google Scholar 

  • Balkema A, De Haan L (1978) Limit distributions for order statistics. I. Theory Prob Appl 23(1):77–92

    Article  MathSciNet  MATH  Google Scholar 

  • Baskerville EB, Dobson AP, Bedford T, Allesina S, Anderson TM, Pascual M (2011) Spatial guilds in the Serengeti food web revealed by a Bayesian group model. PLoS Comput Biol 7(12):e1002321

    Article  MathSciNet  Google Scholar 

  • Bizeul P, Najim J (2021) Positive solutions for large random linear systems. Proc Am Math Soc 149(6):2333–2348

    Article  MathSciNet  MATH  Google Scholar 

  • Bordenave C, Chafaï D (2012) Around the circular law. Probab Surv 9:1–89

    Article  MathSciNet  MATH  Google Scholar 

  • Brose U, Archambault P, Barnes A, Bersier L-F, Boy T, Canning-Clode J, Conti E, Dias M, Digel C, Dissanayake A, Flores A, Fussmann K, Gauzens B, Gray C, Häussler J, Hirt M, Jacob U, Jochum M, Kéfi S, McLaughlin O, MacPherson M, Latz E, Layer-Dobra K, Legagneux P, Li Y, Madeira C, Martinez N, Mendonça V, Mulder C, Navarrete S, O’Gorman E, Ott D, Paula J, Perkins D, Piechnik De, Pokrovsky I, Raffaelli D, Rall B, Rosenbaum B, Ryser R, Silva A, Sohlström Es, N. Sokolova, M. Thompson, R. Thompson, F. Vermandele, C. Vinagre, S. Wang, J. Wefer, R. Williams, E. Wieters, G. Woodward, and A. Iles. (2019) Predator traits determine food-web architecture across ecosystems. Nat Ecol Evol 3(6):919–927

  • Bunin G (2017) Ecological communities with Lotka–Volterra dynamics. Phys Rev E 95(4):042414

    Article  MathSciNet  Google Scholar 

  • Busiello DM, Suweis S, Hidalgo J, Maritan A (2017) Explorability and the origin of network sparsity in living systems. Sci Rep 7(1):12323

    Article  Google Scholar 

  • Capitaine M, Donati-Martin C, Féral D (2009) The largest eigenvalues of finite rank deformation of large Wigner matrices: convergence and nonuniversality of the fluctuations. Ann Probab 37(1)

  • Clenet M (2022) Equilibrium and surviving species in a large Lotka–Volterra system of differential equations. https://github.com/maxime-clenet/Equilibrium-and-surviving-species-in-a-large-Lotka-Volterra-system

  • Cottle RW, Pang J-S, Stone RE (2009) The linear complementarity problem. SIAM

  • Dougoud M, Vinckenbosch L, Rohr RP, Bersier L-F, Mazza C (2018) The feasibility of equilibria in large ecosystems: a primary but neglected concept in the complexity-stability debate. PLoS Comput Biol 14(2):e1005988

    Article  Google Scholar 

  • Fukami T, Nakajima M (2011) Community assembly: Alternative stable states or alternative transient states? Ecol Lett 14(10):973–984

    Article  Google Scholar 

  • Galla T (2018) Dynamically evolved community size and stability of random Lotka–Volterra ecosystems (a). Europhys Lett 123(4):48004

    Article  Google Scholar 

  • Geman S, Hwang C-R (1982) A chaos hypothesis for some large systems of random equations. Z Wahrsch Verw Gebiete 60(3):291–314

    Article  MathSciNet  MATH  Google Scholar 

  • Gibbs T, Grilli J, Rogers T, Allesina S (2018) Effect of population abundances on the stability of large random ecosystems. Phys Rev E 98(2):022410

    Article  Google Scholar 

  • Grilli J, Adorisio M, Suweis S, Barabás G, Banavar JR, Allesina S, Maritan A (2017) Feasibility and coexistence of large ecological communities. Nat Commun 8(1):14389

    Article  Google Scholar 

  • Hastings A (2001) Transient dynamics and persistence of ecological systems. Ecol Lett 4(3):215–220

    Article  Google Scholar 

  • Hofbauer J, Sigmund K (1998) Evolutionary games and population dynamics. Cambridge University Press

  • Jost L (2006) Entropy and diversity. Oikos 113(2):363–375

    Article  Google Scholar 

  • Jost L (2007) Partitioning diversity into independent alpha and beta components. Ecology 88(10):2427–2439

    Article  Google Scholar 

  • Kéfi S, Dakos V, Scheffer M, Van Nes EH, Rietkerk M (2013) Early warning signals also precede non-catastrophic transitions. Oikos 122(5):641–648

  • Lamperski A (2019) Lemke’s algorithm for linear complementarity problems. https://github.com/AndyLamperski/lemkelcp

  • Law R, Morton RD (1996) Permanence and the assembly of ecological communities. Ecology 77(3):762–775

    Article  Google Scholar 

  • Logofet DO (2018) Matrices and graphs: stability problems in mathematical ecology, 1st edn. CRC Press

  • Lotka AJ (1925) Elements of physical biology. Williams & Wilkins

  • May RM (1972) Will a large complex system be stable? Nature 238(5364):413

    Article  Google Scholar 

  • Murty KG (1988) Linear complementarity, linear and nonlinear programming. Sigma series in applied mathematics, vol 3. Heldermann

  • Murty KG (1972) On the number of solutions to the complementarity problem and spanning properties of complementary cones. Linear Algebra Appl 5(1):65–108

    Article  MathSciNet  MATH  Google Scholar 

  • Neubert MG, Caswell H (1997) Alternatives to resilience for measuring the responses of ecological systems to perturbations. Ecology 78(3):653–665

    Article  Google Scholar 

  • Nolting BC, Abbott KC (2016) Balls, cups, and quasi-potentials: quantifying stability in stochastic systems. Ecology 97(4):850–864

    Article  Google Scholar 

  • Opper M, Diederich S (1992) Phase transition and 1/f noise in a game dynamical model. Phys Rev Lett 69(10):1616–1619

    Article  Google Scholar 

  • Pettersson S, Savage VM, Nilsson Jacobi M (2020) Predicting collapse of complex ecological systems: quantifying the stability-complexity continuum. J R Soc Interface 17(166):20190391

    Article  Google Scholar 

  • Serván CA, Capitán JA, Grilli J, Morrison KE, Allesina S (2018) Coexistence of many species in random ecosystems. Nat Ecol Evol 2(8):1237–1242

    Article  Google Scholar 

  • Smirnov NV (1949) Limit distributions for the terms of a variational series. Trudy Matematicheskogo Instituta imeni VA Steklova 25:3–60

    MathSciNet  Google Scholar 

  • Stone L (2018) The feasibility and stability of large complex biological networks: a random matrix approach. Sci Rep 8(1):8246

    Article  Google Scholar 

  • Takeuchi Y (1996) Global dynamical properties of Lotka–Volterra systems. World Scientific

  • Takeuchi Y, Adachi N (1980) The existence of globally stable equilibria of ecosystems of the generalized Volterra type. J Math Biol 10(4):401–415

    Article  MathSciNet  MATH  Google Scholar 

  • Tao T (2013) Outliers in the spectrum of IID matrices with bounded rank perturbations. Probab Theory Relat Fields 155(1):231–263

    Article  MathSciNet  MATH  Google Scholar 

  • Volterra V (1931) Théorie mathématique de la lutte pour la vie. Gauthiers-Villars

  • Volterra V (1926) Fluctuations in the abundance of a species considered mathematically. Nature 118(2972):558–560

    Article  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Maxime Clenet.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supported by CNRS Project 80 Prime - KARATE.

Appendices

A Simulation details

Simulations were performed in Python. All the figures and the code are available in Clenet (2022).

Simulations on the properties of surviving species are performed in two different ways. The theoretical solutions are obtained resolving numerically the system of equations of heuristics 1. We use a solver (cf. scipy.optimize) to find a local minimum of the function defined by the system of equations (a modification of the Powell hybrid method). The empirical solutions are computed using a Monte Carlo experiment. We simulate a large number of matrix matrix B, we resolve the associated LCP problem using the Lemke’s algorithm. Then, we use the LCP solution to calculate the properties of the surviving species: proportion of survivors, etc. Finally, we make an average on the ensemble of experiments. The Lemke algorithm is implemented in the lemkelcp package and can be found on Lamperski (2019). The dynamics of the Lotka–Volterra are achieved by a Runge–Kutta of order 4 (RK4) implemented in the code.

B Proof of Theorem 2

We have

$$\begin{aligned} I- B +I -B^T = 2I - (B+B^T) = 2I - \left( \frac{A+A^T}{\alpha \sqrt{n}}+\frac{2\mu }{n} \varvec{1}\varvec{1}^*\right) . \end{aligned}$$

Notice that \(2I- (B+B^T)\) is positive definite iff the top eigenvalue of \(B+B^T\) is lower than 2:

$$\begin{aligned} \lambda _{\max }(B+B^T) < 2 \,. \end{aligned}$$
(23)

We first focus on the random part \((A+A^T)/\alpha \) which is a symmetric matrix with independent \({{\mathcal {N}}}(0,2/\alpha ^2)\) entries above the diagonal (note that the distribution of the diagonal entries is different from the off-diagonal entries, with no asymptotic effect). In this case, it is well known that the largest eigenvalue of the normalized matrix (or equivalently its spectral norm since the matrix is symmetric) a.s. converges to the right edge of the support of the semi-circle law (see (Bai and Silverstein 2010, Th. 5.2)):

$$\begin{aligned} \lambda _{\max }\left( \frac{A+A^T}{\alpha \sqrt{n}} \right) \xrightarrow [n\rightarrow \infty ]{a.s.} \frac{2 \sqrt{2}}{\alpha }\,. \end{aligned}$$
(24)

In the centered case (\(\mu = 0\)), condition (23) occurs if \(\alpha > \sqrt{2}\).

We now consider the general case where \(\mu \ne 0\). Notice that the rank-one perturbation matrix \(P = \frac{2\mu }{n} \varvec{1}\varvec{1}^*\) admits a unique non zero eigenvalue \(2\mu \). Denote by \(\check{A}=\frac{A+A^T}{\alpha \sqrt{n}}\). We are interested in the top eigenvalue of the symmetric matrix \(\check{A}+P\). Based on a result by Capitaine et al. (Capitaine et al. 2009, Th. 2.1), we have:

$$\begin{aligned} \lambda _{\max }(\check{A}+P) \xrightarrow [n \rightarrow \infty ]{a.s}\left\{ \begin{array}{ll} 2\mu +\frac{1}{\alpha ^2 \mu } &{}\text {if } \mu > \frac{1}{\sqrt{2} \alpha }\,, \\ \frac{2 \sqrt{2}}{\alpha } &{}\text {else.} \end{array}\right. \end{aligned}$$

This result is illustrated in Fig. 11.

Fig. 11
figure 11

Spectrum (histogram) of the Hermitian random matrix \(B+B^T\) (\(n=1000\), \(\alpha =\sqrt{2}\)). The solid line represents the semi-circular law. In Fig. 11a, \(\mu =0\) and there is no oulier. In Fig. 11b, \(\mu =1.5\) and one can notice the presence of an eigenvalue outside the bulk of the semi-circular law. The dashed line indicates its theoretical value

Assume first that \(\mu \le \frac{1}{\alpha \sqrt{2}}\) (corresponding to zone \({{\mathcal {C}}}\) in Fig. 2), then \(\lambda _{\max }( \check{A}+P) \xrightarrow [n\rightarrow \infty ]{a.s.} \frac{2\sqrt{2}}{\alpha }\), which is strictly lower than 2 (cf. condition (23)) if \(\alpha >\sqrt{2}\). Hence \(\lambda _{\max }(\check{A} +P)\) is eventually strictly lower than 2 under this condition.

Assume now that \(\mu >\frac{1}{\alpha \sqrt{2}}\) (corresponding to zone \({{\mathcal {B}}}\) in Fig. 2), then

$$\begin{aligned} \lambda _{\max }( \check{A}+P) \xrightarrow [n\rightarrow \infty ]{a.s.} 2\mu +\frac{1}{\alpha ^2 \mu }\,. \end{aligned}$$

We are interested in the conditions for which \(2\mu +\frac{1}{\alpha ^2 \mu }<2\) or equivalently

$$\begin{aligned} 2\alpha ^2 \mu ^2 - 2\alpha ^2 \mu +1<0\,. \end{aligned}$$
(25)

An elementary study of the polynomial \(\xi (X)= 2\alpha ^2 X^2 - 2\alpha ^2 X +1\) yields that \(\xi \)’s discriminant is positive if \(\alpha >\sqrt{2}\),

$$\begin{aligned} \xi (\mu ^{\pm })=0\quad \Leftrightarrow \quad \mu ^{\pm }= \frac{1}{2} \pm \frac{1}{2} \sqrt{1- \frac{2}{\alpha ^2}}\, \end{aligned}$$

and \(\xi \left( \frac{1}{\alpha \sqrt{2}}\right) <0\), so that \(\frac{1}{\alpha \sqrt{2}}\in (\mu ^-, \mu ^+)\). In particular condition (25) is fulfilled if

$$\begin{aligned} \mu \in \left( \frac{1}{\alpha \sqrt{2}}\,\ \frac{1}{2}+\frac{1}{2} \sqrt{1-\frac{2}{\alpha ^2}}\right) . \end{aligned}$$

Under this condition, (25) is fulfilled and a.s. \(\limsup _{n\rightarrow \infty }\lambda _{\max }(\check{A}+P) < 2\), which completes the proof: we can then rely on Theorem 1 to conclude.

C Construction of the heuristics

We first discuss Heuristics 1 and establish Equations (12), (13) and (14).

1.1 C.1 Equation (12)

We first recall a result on order statistics of a Gaussian sample. Consider a family \((Z_k)_{k\in [n]}\) of i.i.d. random variables \({{\mathcal {N}}}(0,1)\) and the associated order statistics

$$\begin{aligned} Z_1^*\ \le \ Z_2^*\ \le \cdots \le \ Z_n^*\,. \end{aligned}$$

Consider an index \(\lfloor n\alpha \rfloor \in [n]\) where \(\alpha \in (0,1)\) is fixed, then the typical location of \(Z^*_{ \lfloor n\alpha \rfloor }\) is \(\Phi ^{-1}(\alpha )\):

$$\begin{aligned} Z^*_{ \lfloor n\alpha \rfloor }\simeq \Phi ^{-1}(\alpha )\quad \text {as}\quad n\rightarrow \infty \,, \end{aligned}$$
(26)

see for instance (Smirnov 1949; Balkema and De Haan 1978).

Let \(\varvec{x}^*\) be the equilibrium of (1) and consider the random variable

$$\begin{aligned} \check{Z}_k= \sum _{i\in {{\mathcal {S}}}} B_{ki} x_i^* = (B\varvec{x}^*)_k. \end{aligned}$$

We assume that asymptotically the \(x_i^*\)’s are independent from the \(B_{ki}\)’s, an assumption supported by the chaos hypothesis, see for instance Geman and Hwang (1982). Denote by \(\mathbb {E}_{\varvec{x}^*}=\mathbb {E}(\,\cdot \mid \varvec{x}^*)\) the conditional expectation with respect to \(\varvec{x}^*\). Notice that conditionally to \(\varvec{x}^*\), the \(\check{Z}_k\)’s are independent Gaussian random variables, whose two first moments can easily be computed, see Sect. 1 below for the details:

$$\begin{aligned} \mathbb {E}_{\varvec{x}^*} \check{Z}_k= \mu \,\hat{p}\, \hat{m}\quad \text {and}\quad \text {var}_{\varvec{x}^*}(\check{Z}_k)= \frac{\hat{p}\hat{\sigma }^2}{\alpha ^2} \,. \end{aligned}$$

Notice that the fact that \(\mathbb {E}_{\varvec{x}^*}\) and \(\text {var}_{\varvec{x}^*}(\check{Z}_k)\) only depend on \(\hat{p}, \hat{\sigma }\) and \(\hat{m}\) which are (supposedly) converging quantities supports the idea that \(\check{Z}_k\) is unconditionally a Gaussian random variable with moments:

$$\begin{aligned} \mathbb {E} \check{Z}_k= \mu \,p^*\, m^*\quad \text {and}\quad \text {var}(\check{Z}_k)= \frac{p^* (\sigma ^*)^2}{\alpha ^2} \,, \end{aligned}$$

where \(p^*,m^*,\sigma ^*\) are resp. the limits of \(\hat{p}, \hat{m}, \hat{\sigma }\). We now introduce the standard Gaussian random variables \((Z_k)_{k\in [n]}\) where

$$\begin{aligned} Z_k= \frac{\check{Z}_k - \mathbb {E} \check{Z}_k}{\sqrt{\text {var}(\check{Z}_k)}} = \alpha \frac{\check{Z}_k - \mu \,p^*\, m^*}{\sigma ^* \sqrt{p^*}}\,. \end{aligned}$$

Consider the equilibrium \(\varvec{x}^*=(x_k^*)_{k\in [n]}\). If \(k\in {{\mathcal {S}}}\), that is \(x_k^*>0\), we have

$$\begin{aligned} 1- x_k^* + (B\varvec{x}^*)_k=0 \quad \Rightarrow \quad 1+ (B\varvec{x}^*)_k>0 \,. \end{aligned}$$

This identity has two implications:

$$\begin{aligned} x_k^*= 1 + (B\varvec{x}^*)_k\quad \text {and}\quad 1+ (B\varvec{x}^*)_k>0 \qquad \text {if}\ k\in {{\mathcal {S}}}\,. \end{aligned}$$

Relying on the representation \((B\varvec{x}^*)_k=\check{Z}_k\), we obtain the representation

$$\begin{aligned} x_k^* = 1+(B\varvec{x}^*)_k \ =\ 1+\mu \,p^* \,m^*+\frac{\sigma ^*\sqrt{p^*}}{\alpha } Z_k \qquad \text {if}\quad k\in {{\mathcal {S}}}. \end{aligned}$$
(27)

and the condition:

$$\begin{aligned} 1+(B\varvec{x}^*)_k \ =\ 1+\mu \,p^* \,m^*+\frac{\sigma ^*\sqrt{p^*}}{\alpha } Z_k\, >\, 0. \end{aligned}$$

If \(k\notin {{\mathcal {S}}}\) then

$$\begin{aligned} 1+ (B\varvec{x}^*)_k\ =\ 1+\mu \,p^*\,m^* +\frac{\sigma ^*\sqrt{p^*}}{\alpha } Z_k\ \le \ 0 \end{aligned}$$

by the non invadability condition. Otherwise stated,

$$\begin{aligned} \left\{ \begin{array}{lcll} Z_k&{}\le &{} -\frac{\alpha (1+\mu \, p^* m^*)}{\sigma ^*\sqrt{p^*}} &{}\text {if}\ k\notin {{\mathcal {S}}},\\ Z_k&{}>&{} -\frac{\alpha (1+\mu \, p^* m^*)}{\sigma ^*\sqrt{p^*}} &{}\text {if}\ k\in {{\mathcal {S}}}.\\ \end{array} \right. \end{aligned}$$

Considering the order statistics of the \(Z_k\)’s we obtain:

$$\begin{aligned} Z_1^*\le \cdots \le Z_{\varvec{i}}^*\le -\frac{\alpha (1+\mu \, p^* m^*)}{\sigma ^*\sqrt{p^*}} \le Z_{\varvec{i}+1}^*\le \cdots \le Z_n^*\,. \end{aligned}$$

Now, there are exactly \(n-|{{\mathcal {S}}}|=n(1-\hat{p})\) indices before the threshold corresponding to the components of \(\varvec{x}^*\) equal to zero. In particular, index \(\varvec{i}=n(1-\hat{p})\) corresponds to the value

$$\begin{aligned} Z^*_{\varvec{i}}\simeq -\frac{\alpha (1+\mu \,p^*\,m^*)}{\sigma ^*\sqrt{p^*}}. \end{aligned}$$

Relying on (26), we finally obtain

$$\begin{aligned} \Phi ^{-1}(1-\hat{p}) = -\frac{\alpha (1+\mu \,p^* \, m^*)}{\sigma ^*\sqrt{p^*}}. \end{aligned}$$

It remains to replace \(\hat{p}\) by its limit \(p^*\) to obtain (12).

1.2 C.2 Details on Eq. (12): moments of \(\check{Z}_k\)

We compute hereafter the conditional mean and variance of \(\check{Z}_k=(B\varvec{x}^*)_k\) with respect to \(\varvec{x}^*\). We rely on the following identities:

$$\begin{aligned} \mathbb {E} B_{ki} =\frac{\mu }{n},\quad \mathbb {E} (B_{ki})^2 = \frac{1}{\alpha ^2 n } + \frac{\mu ^2}{n^2} \simeq \frac{1}{\alpha ^2 n},\quad \mathbb {E} B_{ki}B_{kj} = \frac{\mu ^2}{n^2}\quad (i\ne j)\,. \end{aligned}$$

We first compute the conditional mean:

$$\begin{aligned} \mathbb {E}_{\varvec{x}^*}(\check{Z}_k)= & {} \sum _{i\in [n]} \mathbb {E}(B_{ki}) x_i^* = \sum _{i\in {{\mathcal {S}}}} \mathbb {E}(B_{ki}) x_i^* = \frac{\mu }{n}\sum _{i\in {{\mathcal {S}}}} x_i^*, = \mu \frac{|{{\mathcal {S}}}|}{n} \frac{1}{|{{\mathcal {S}}}|}\sum _{i\in {{\mathcal {S}}}} x_i^*, \\= & {} \mu \, \hat{p}\, \hat{m}\,. \end{aligned}$$

We now compute the second moment:

where the approximation in (a) follows from the fact that

$$\begin{aligned} \frac{1}{|{{\mathcal {S}}}|^2} \sum _{i,j\in {{\mathcal {S}}}}x_i^*x_j^* = \frac{1}{|{{\mathcal {S}}}|^2} \sum _{i \ne j}x_i^*x_j^* + {{\mathcal {O}}} \left( \frac{1}{|{{\mathcal {S}}}|} \right) \,. \end{aligned}$$

We can now compute the variance:

$$\begin{aligned} \text {var}_{\varvec{x}^*} \left( \check{Z}_k\right) = \mathbb {E}_{\varvec{x}^*}\left( \check{Z}_k^2\right) - \left( \mathbb {E}_{\varvec{x}^*}\check{Z}_k\right) ^2 \ =\ \frac{\hat{p}\, \hat{\sigma }^2}{\alpha ^2}\,. \end{aligned}$$

1.3 C.3 Equation (13)

Our starting point is the following generic representation of an abundance at equilibrium (either of a surviving or vanishing species):

$$\begin{aligned}&x_k^* = \left( 1+\mu \,p^* m^* +\frac{\sigma ^*\sqrt{p^*}}{\alpha } Z_k \right) \varvec{1}_{\{Z_k> - \delta ^*\}},\\&\quad = \left( 1+\mu \,p^* m^* \right) \varvec{1}_{\{Z_k> - \delta ^*\}} +\left( \frac{\sigma ^*\sqrt{p^*}}{\alpha } Z_k \right) \varvec{1}_{\{Z_k> - \delta ^*\}}. \end{aligned}$$

Summing over \({{\mathcal {S}}}\) and normalizing,

$$\begin{aligned} \begin{aligned} \frac{1}{|{{\mathcal {S}}}|}\sum _{k \in \mathcal {S}}x_k^{*}&=(1+\mu \, p^* m^*)\frac{1}{|{{\mathcal {S}}}|}\sum _{k \in \mathcal {S}}\varvec{1}_{\{Z_k> -\delta ^*\}}+\frac{\sigma ^*\sqrt{p^*}}{\alpha }\frac{1}{|{{\mathcal {S}}}|}\sum _{k \in \mathcal {S}}Z_k\varvec{1}_{\{Z_k> -\delta ^*\}}, \\ \hat{m}&{\mathop {=}\limits ^{(a)}} (1+\mu \, p^* m^*)+\frac{\sigma ^*\sqrt{p^*}}{\alpha }\frac{n}{|{{\mathcal {S}}}|}\frac{1}{n}\sum _{k \in [n]}Z_k\varvec{1}_{\{Z_k> -\delta ^*\}}, \\ \hat{m}&{\mathop {\simeq }\limits ^{(b)}}(1+\mu \,p^* m^*)+\frac{\sigma ^*\sqrt{p^*}}{\alpha } \frac{1}{\mathbb {P}(Z> -\delta ^*)}\mathbb {E}(Z\varvec{1}_{\{Z> -\delta ^*\}}), \\ \hat{m}&\simeq (1+\mu \, p^* m^*)+\frac{\sigma ^*\sqrt{p^*}}{\alpha }\mathbb {E}(Z \mid Z > -\delta ^*). \end{aligned} \end{aligned}$$

where (a) follows from the fact that \(|{{\mathcal {S}}}| = \sum _{k \in \mathcal {S}}\varvec{1}_{\{Z_k > -\delta ^*\}}\) (by definition of \({{\mathcal {S}}}\)), (b) from the law of large numbers \(\frac{1}{n} \sum _{k\in [n]} Z_k \varvec{1}_{\{Z_k>-\delta \}} \xrightarrow [n\rightarrow \infty ]{} \mathbb {E}Z \varvec{1}_{\{Z>-\delta \}}\) and \(\frac{|{{\mathcal {S}}}|}{n} \xrightarrow [n\rightarrow \infty ]{} \mathbb {P}(Z>-\delta ^*)\) with \(Z\sim {{\mathcal {N}}}(0,1)\). It remains to replace \(\hat{m}\) by its limit \(m^*\) to obtain (13).

Equation (14) can be obtained similarly.

1.4 C.4 Equation (14)

As for the proof of (13), we start from the generic representation of \(x_k^*\):

$$\begin{aligned} x_k^*\ {}&= \left( 1+\mu \, p^* m^* +\frac{\sigma ^*\sqrt{p^*}}{\alpha } Z_k\right) \varvec{1}_{\{Z_k> -\delta ^*\}},\\&= \left( 1+\mu \,p^* m^*\right) \varvec{1}_{\{Z_k> -\delta ^*\}}+\frac{\sigma ^*\sqrt{p^*}}{\alpha }Z_k\varvec{1}_{\{Z_k > -\delta ^*\}}\, . \end{aligned}$$

Taking the square, we get

$$\begin{aligned} x_k^{*2}&= \left( 1+\mu \,p^* m^*\right) ^2\varvec{1}_{\{Z_k> -\delta ^*\}}\\&\quad +2(1+\mu \,p^* m^*)\frac{\sigma ^*\sqrt{p^*}}{\alpha }Z_k\varvec{1}_{\{Z_k> -\delta ^*\}} +\frac{(\sigma ^*)^2p^*}{\alpha ^2}Z^2_k\varvec{1}_{\{Z_k > -\delta ^*\}}\,. \end{aligned}$$

Summing over \({{\mathcal {S}}}\) and normalizing, we get

$$\begin{aligned}&\frac{1}{|{{\mathcal {S}}}|}\sum _{k \in \mathcal {S}}(x_k^*)^{2} = (1+\mu \, p^* m^*)^2\frac{1}{|{{\mathcal {S}}}|}\sum _{k \in \mathcal {S}}\varvec{1}_{\{Z_k> -\delta ^*\}} \\&\quad +2(1+\mu \,p^* m^*)\frac{\sigma ^*\sqrt{p^*}}{\alpha } \frac{1}{|{{\mathcal {S}}}|}\sum _{k \in \mathcal {S}}Z_k\varvec{1}_{\{Z_k> -\delta ^*\}}\\&\quad +\frac{(\sigma ^*)^2 p^*}{\alpha ^2}\frac{1}{|{{\mathcal {S}}}|}\sum _{k \in \mathcal {S}}Z^2_k\varvec{1}_{\{Z_k > -\delta ^*\}}\,. \end{aligned}$$

Finally, we conclude by replacing the empirical means by their limits

$$\begin{aligned} \frac{1}{|{{\mathcal {S}}}|}\sum _{k \in \mathcal {S}}Z^i_k\varvec{1}_{\{Z_k> -\delta ^*\}}= & {} \mathbb {E}(Z^i\mid Z>-\delta ^*)\,\quad i=1,2\,. \end{aligned}$$

and get

$$\begin{aligned} \hat{\sigma }^2&= (1+\mu \,p^* m^*)^2+2(1+\mu \,p^* m^*)\frac{\sigma ^*\sqrt{p^*}}{\alpha }\mathbb {E}(Z \mid Z> -\delta ^*)\\&\quad +\frac{(\sigma ^*)^2p^*}{\alpha ^2}\mathbb {E}(Z^2 \mid Z > -\delta ^*)\,. \end{aligned}$$

It remains to replace \(\hat{\sigma }\) by its limit \(\sigma ^*\) to obtain (14).

D Density of the distribution of the persistent species

Assume that \(x^*>0\), and let \(f=\mathbb {R}\rightarrow \mathbb {R}\) be a bounded continuous test function. We have

$$\begin{aligned} \mathbb {E}f(x_k^*)= & {} \textbf{E}\left[ f\left( 1+ \frac{\sigma ^*\sqrt{p^*}}{\alpha } Z_k+\mu \,p^* m^*\right) \ \bigg |\ Z_k> -\delta ^* \right] \,, \\= & {} \int _{-\infty }^{\infty }f\left( 1+\mu \,p^* m^*+\frac{\sigma ^*\sqrt{p^*}}{\alpha }u\right) \frac{\varvec{1}_{\{u>-\delta ^*\}}}{1-\Phi (-\delta ^*)}\frac{e^{-\frac{u^2}{2}}}{\sqrt{2\pi }}du\,, \\= & {} \int _{0}^{\infty }f(y) e^{-\frac{1}{2}\left( \frac{\alpha }{\sigma ^* \sqrt{p^*}}y-\delta ^*\right) ^2} \frac{\alpha }{\sqrt{2\pi } \Phi (\delta ^*)\,p^*\, \sigma ^*}\,dy\,, \end{aligned}$$

hence the density of \(x^*_k\).

1.1 D.1 Theoretical estimation of the diversity index

Recall that \(|\mathcal {S}| = n \hat{p}\) is the number of surviving species and that

$$\begin{aligned} p_i = \frac{x_i}{\sum _{j\in {{\mathcal {S}}}} x_j} \end{aligned}$$

is the frequency of (surviving) species i.

To find a theoretical estimate of Hill number of order 1, we proceed by expansion and set

$$\begin{aligned} p_i = \frac{1}{|\mathcal {S}|}+\delta _i,\quad |\delta _i|\ll \frac{1}{|S|} \end{aligned}$$

where \(\delta _i\) represents the deviation of species i from the standard frequency if all surviving species have the same frequency. Notice that \(\sum _{i\in \mathcal {S}}\delta _i=0\).

$$\begin{aligned} H' \ =\ - \sum _{i\in \mathcal {S}} p_i \log (p_i) = - \sum _{i\in \mathcal {S}} \left( \frac{1}{|\mathcal {S}|}+ \delta _i \right) \log \left( \frac{1}{|\mathcal {S}|}+\delta _i \right) . \end{aligned}$$

We use the Taylor-Young formula of order 2 to decompose the log:

$$\begin{aligned} \log \left( \frac{1}{|\mathcal {S}|}+\delta _i \right)= & {} \log \left( \frac{1}{|\mathcal {S}|}\right) +|\mathcal {S}|\delta _i-\frac{\delta _i^2|\mathcal {S}|^2}{2}+\delta _i^3\varepsilon (\delta _i) \,, \\\approx & {} \log \left( \frac{1}{|\mathcal {S}|}\right) +|\mathcal {S}|\delta _i-\frac{\delta _i^2|\mathcal {S}|^2}{2} .\\ H'\approx & {} - \sum _{i\in \mathcal {S}} \left( \frac{1}{|\mathcal {S}|}+\delta _i \right) \left( \log \left( \frac{1}{|\mathcal {S}|}\right) +|\mathcal {S}|\delta _i-\frac{\delta _i^2|\mathcal {S}|^2}{2}\right) \,, \\= & {} - \sum _{i\in \mathcal {S}}\left[ \frac{1}{|\mathcal {S}|}\log \left( \frac{1}{|\mathcal {S}|}\right) +\delta _i-\frac{\delta _i^2|\mathcal {S}|}{2}+\delta _i \log \left( \frac{1}{|\mathcal {S}|} \right) + |\mathcal {S}|\delta _i^2-\frac{\delta _i^3 |\mathcal {S}|^2}{2} \right] \,, \\= & {} \log (|\mathcal {S}|) - \sum _{i\in \mathcal {S}} \frac{\delta _i^2|\mathcal {S}|}{2} +\sum _{i\in \mathcal {S}} \frac{\delta _i^3|\mathcal {S}|^2}{2}. \end{aligned}$$

Notice that \(\sum _{i=1}^{|\mathcal {S}|} \frac{\delta _i^3|\mathcal {S}|^2}{2}\) is negligible since \(|\delta _i|\ll |S|^{-1}\). The term 1 corresponds to the maximum value that the Shannon diversity index can take if \(|\mathcal {S}|\) are present in the system. It remains to develop the second term of the r.h.s.

$$\begin{aligned} -\frac{1}{2}\sum _{i\in \mathcal {S}}\delta _i^2|\mathcal {S}|= & {} -\frac{|\mathcal {S}|}{2}\sum _{i\in \mathcal {S}} \left( \frac{x_i}{\sum _{j\in \mathcal {S}} x_j}-\frac{1}{|\mathcal {S}|} \right) ^2\, \\= & {} -\frac{|\mathcal {S}|}{2} \sum _{i\in \mathcal {S}} \left( \frac{x_i^2}{(\sum _{j\in \mathcal {S}}x_j)^2}-\frac{2}{|\mathcal {S}|}\frac{x_i}{\sum _{j\in \mathcal {S}} x_j}+\frac{1}{|\mathcal {S}|^2} \right) \, \\= & {} -\frac{|\mathcal {S}|}{2} \sum _{i\in \mathcal {S}} \left( \frac{x_i^2}{(\sum _{j\in \mathcal {S}}x_j)^2}\right) +\frac{1}{2}\, \\= & {} -\frac{|\mathcal {S}|}{2}\frac{\sum _{i\in \mathcal {S}}x_i^2}{|\mathcal {S}|^2 (\frac{1}{|\mathcal {S}|}\sum _{j\in \mathcal {S}}x_j)^2}+\frac{1}{2}\,\\= & {} -\frac{1}{2}\frac{\frac{1}{|\mathcal {S}|}\sum _{i\in \mathcal {S}}x_i^2}{ (\frac{1}{|\mathcal {S}|}\sum _{j\in \mathcal {S}}x_j)^2}+\frac{1}{2}\, \\= & {} -\frac{1}{2} \frac{\hat{\sigma }^2}{(\hat{m})^2}+\frac{1}{2}\, \\= & {} -\frac{1}{2}\left( \frac{\hat{\sigma }^2}{\hat{m}^2}-1 \right) . \end{aligned}$$

Finally the Hill number of order 1 can be computed as:

$$\begin{aligned} e^{H'}\approx & {} e^{\log (|\mathcal {S}|)-\frac{|\mathcal {S}|}{2}\sum _{i=1}^{|\mathcal {S}|}\delta _i^2}\,, \\\approx & {} |\mathcal {S}|\left( 1-\frac{|\mathcal {S}|}{2}\sum _{i=1}^{|\mathcal {S}|}\delta _i^2\right) \ =\ |\mathcal {S}|\left( 1-\frac{1}{2} \frac{\hat{\sigma }^2}{(\hat{m})^2}+\frac{1}{2}\right) \ =\ \frac{|\mathcal {S}|}{2}\left( 3- \frac{\hat{\sigma }^2}{(\hat{m})^2}\right) . \end{aligned}$$

Replacing \(|{{\mathcal {S}}}|\) by \(np^*\) and \(\hat{\sigma }\) and \(\hat{m}\) by their limits, we get the desired result:

$$\begin{aligned} e^{H'}\approx & {} \frac{np^*}{2}\left( 3- \frac{(\sigma ^*)^2}{(m^*)^2}\right) . \end{aligned}$$

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Clenet, M., Massol, F. & Najim, J. Equilibrium and surviving species in a large Lotka–Volterra system of differential equations. J. Math. Biol. 87, 13 (2023). https://doi.org/10.1007/s00285-023-01939-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-023-01939-z

Keywords

Mathematics Subject Classification

Navigation