Skip to main content
Log in

The competitive exclusion principle in stochastic environments

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

Abstract

In its simplest form, the competitive exclusion principle states that a number of species competing for a smaller number of resources cannot coexist. However, it has been observed empirically that in some settings it is possible to have coexistence. One example is Hutchinson’s ‘paradox of the plankton’. This is an instance where a large number of phytoplankton species coexist while competing for a very limited number of resources. Both experimental and theoretical studies have shown that temporal fluctuations of the environment can facilitate coexistence for competing species. Hutchinson conjectured that one can get coexistence because nonequilibrium conditions would make it possible for different species to be favored by the environment at different times. In this paper we show in various settings how a variable (stochastic) environment enables a set of competing species limited by a smaller number of resources or other density dependent factors to coexist. If the environmental fluctuations are modeled by white noise, and the per-capita growth rates of the competitors depend linearly on the resources, we prove that there is competitive exclusion. However, if either the dependence between the growth rates and the resources is not linear or the white noise term is nonlinear we show that coexistence on fewer resources than species is possible. Even more surprisingly, if the temporal environmental variation comes from switching the environment at random times between a finite number of possible states, it is possible for all species to coexist even if the growth rates depend linearly on the resources. We show in an example (a variant of which first appeared in Benaim and Lobry ’16) that, contrary to Hutchinson’s explanation, one can switch between two environments in which the same species is favored and still get coexistence.

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

Similar content being viewed by others

References

  • Abrams PA, Holt RD, Roth JD (1998) Apparent competition or apparent mutualism? Shared predation when populations cycle. Ecology 79(1):201–212

    Google Scholar 

  • Armstrong RA, McGehee R (1976a) Coexistence of species competing for shared resources. Theor Popul Biol 9(3):317–328

    MathSciNet  MATH  Google Scholar 

  • Armstrong RA, McGehee R (1976b) Coexistence of two competitors on one resource. J Theor Biol 56(2):499–502

    Google Scholar 

  • Armstrong RA, McGehee R (1980) Competitive exclusion. Am Nat 115(2):151–170

    MathSciNet  Google Scholar 

  • Bansaye V, Delmas J-F, Marsalle L, Tran VC (2011) Limit theorems for markov processes indexed by continuous time Galton–Watson trees. Ann Appl Probab 21(6):2263–2314

    MathSciNet  MATH  Google Scholar 

  • Benaim M (2018) Stochastic persistence. arXiv preprint arXiv:1806.08450

  • Benaïm M, Lobry C (2016) Lotka Volterra in fluctuating environment or “how switching between beneficial environments can make survival harder”. Ann Appl Probab 26(6):3754–3785

    MathSciNet  MATH  Google Scholar 

  • Benaïm M, Strickler E (2019) Random switching between vector fields having a common zero. Ann Appl Probab 29(1):326–375

    MathSciNet  MATH  Google Scholar 

  • Borodin AN, Salminen P (2016) Handbook of Brownian motion-facts and formulae, corrected, 2nd edn. Birkhäuser, Basel, Boston and Berlin

    MATH  Google Scholar 

  • Braumann CA (2002) Variable effort harvesting models in random environments: generalization to density-dependent noise intensities. Math Biosci 177/178:229–245 Deterministic and stochastic modeling of biointeraction (West Lafayette, IN, 2000)

    MathSciNet  MATH  Google Scholar 

  • Chesson P (1978) Predator–prey theory and variability. Annu Rev Ecol Syst 9(1):323–347

    Google Scholar 

  • Chesson PL (1982) The stabilizing effect of a random environment. J Math Biol 15(1):1–36

    MathSciNet  MATH  Google Scholar 

  • Chesson P (1994) Multispecies competition in variable environments. Theor Popul Biol 45(3):227–276

    MathSciNet  MATH  Google Scholar 

  • Chesson P (2000) General theory of competitive coexistence in spatially-varying environments. Theor Popul Biol 58(3):211–237

    MathSciNet  MATH  Google Scholar 

  • Chesson P (2009) Scale transition theory with special reference to species coexistence in a variable environment. J Biol Dyn 3(2–3):149–163

    MathSciNet  MATH  Google Scholar 

  • Chesson P (2018) Updates on mechanisms of maintenance of species diversity. J Ecol 106(5):1773–1794

    Google Scholar 

  • Chesson PL, Ellner S (1989) Invasibility and stochastic boundedness in monotonic competition models. J Math Biol 27(2):117–138

    MathSciNet  MATH  Google Scholar 

  • Chesson P, Huntly N (1997) The roles of harsh and fluctuating conditions in the dynamics of ecological communities. Am Nat 150(5):519–553

    Google Scholar 

  • Chesson PL, Warner RR (1981) Environmental variability promotes coexistence in lottery competitive systems. Am Nat 117(6):923–943

    MathSciNet  Google Scholar 

  • Cloez B, Dessalles R, Genadot A, Malrieu F, Marguet A, Yvinec R (2017) Probabilistic and piecewise deterministic models in biology. ESAIM Proc Surv 60:225–245

    MathSciNet  MATH  Google Scholar 

  • Cushing JM (1980) Two species competition in a periodic environment. J Math Biol 10(4):385–400

    MathSciNet  MATH  Google Scholar 

  • Davis MHA (1984) Piecewise-deterministic Markov processes: a general class of non-diffusion stochastic models. J R Stat Soc Ser B (Methodol) 46:353–388

    MATH  Google Scholar 

  • De Mottoni P, Schiaffino A (1981) Competition systems with periodic coefficients: a geometric approach. J Math Biol 11(3):319–335

    MathSciNet  MATH  Google Scholar 

  • Ditlevsen S, Löcherbach E (2017) Multi-class oscillating systems of interacting neurons. Stoch Process Appl 127(6):1840–1869

    MathSciNet  MATH  Google Scholar 

  • Du NH, Dang NH (2011) Dynamics of kolmogorov systems of competitive type under the telegraph noise. J Differ Equ 250(1):386–409

    MathSciNet  MATH  Google Scholar 

  • Du NH, Dang NH (2014) Asymptotic behavior of Kolmogorov systems with predator–prey type in random environment. Commun Pure Appl Anal 13(6):32

    MathSciNet  MATH  Google Scholar 

  • Ellner S (1989) Convergence to stationary distributions in two-species stochastic competition models. J Math Biol 27(4):451–462

    MathSciNet  MATH  Google Scholar 

  • Evans SN, Ralph PL, Schreiber SJ, Sen A (2013) Stochastic population growth in spatially heterogeneous environments. J Math Biol 66(3):423–476

    MathSciNet  MATH  Google Scholar 

  • Evans SN, Hening A, Schreiber SJ (2015) Protected polymorphisms and evolutionary stability of patch-selection strategies in stochastic environments. J Math Biol 71(2):325–359

    MathSciNet  MATH  Google Scholar 

  • Gard TC (1984) Persistence in stochastic food web models. Bull Math Biol 46(3):357–370

    MathSciNet  MATH  Google Scholar 

  • Gard TC (1988) Introduction to stochastic differential equations. M. Dekker, New York

    MATH  Google Scholar 

  • Gause GF (1932) Experimental studies on the struggle for existence: I. Mixed population of two species of yeast. J. Exp. Biol. 9(4):389–402

    Google Scholar 

  • Haigh J, Smith JM (1972) Can there be more predators than prey? Theor Popul Biol 3(3):290–299

    MATH  Google Scholar 

  • Hardin G (1960) The competitive exclusion principle. Science 131(3409):1292–1297

    Google Scholar 

  • Hening A, Nguyen DH (2018a) Coexistence and extinction for stochastic Kolmogorov systems. Ann Appl Probab 28(3):1893–1942

    MathSciNet  MATH  Google Scholar 

  • Hening A, Nguyen DH (2018) Persistence in stochastic Lotka–Volterra food chains with intraspecific competition. Bull Math Biol 80(10):2527–2560

    MathSciNet  MATH  Google Scholar 

  • Hening A, Nguyen DH (2018) Stochastic Lotka–Volterra food chains. J Math Biol 77(1):135–163

    MathSciNet  MATH  Google Scholar 

  • Hening A, Strickler E (2019) On a predator–prey system with random switching that never converges to its equilibrium. SIAM J Math Anal 51:3625–3640

    MathSciNet  MATH  Google Scholar 

  • Hening A, Nguyen DH, Yin G (2018) Stochastic population growth in spatially heterogeneous environments: the density-dependent case. J Math Biol 76(3):697–754

    MathSciNet  MATH  Google Scholar 

  • Hepp B, Gupta A, Khammash M (2015) Adaptive hybrid simulations for multiscale stochastic reaction networks. J Chem Phys 142(3):034118

    Google Scholar 

  • Hofbauer J (1981) A general cooperation theorem for hypercycles. Monatshefte für Mathematik 91(3):233–240

    MathSciNet  MATH  Google Scholar 

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

    MATH  Google Scholar 

  • Hofbauer J, So JW-H (1989) Uniform persistence and repellors for maps. Proc Am Math Soc 107(4):1137–1142

    MathSciNet  MATH  Google Scholar 

  • Holt RD (1977) Predation, apparent competition, and the structure of prey communities. Theor Popul Biol 12(2):197–229

    MathSciNet  Google Scholar 

  • Hsu SB (1980) A competition model for a seasonally fluctuating nutrient. J Math Biol 9(2):115–132

    MathSciNet  MATH  Google Scholar 

  • Hutchinson GE (1961) The paradox of the plankton. Am Nat 95(882):137–145

    Google Scholar 

  • Hutson V (1984) A theorem on average Liapunov functions. Monatshefte für Mathematik 98(4):267–275

    MathSciNet  MATH  Google Scholar 

  • Kaplan JL, Yorke JA (1977) Competitive exclusion and nonequilibrium coexistence. Am Nat 111(981):1030–1036

    Google Scholar 

  • Kesten H, Ogura Y (1981) Recurrence properties of Lotka–Volterra models with random fluctuations. J Math Soc Jpn 33(2):335–366

    MathSciNet  MATH  Google Scholar 

  • King AA, Schaffer WM (1999) The rainbow bridge: Hamiltonian limits and resonance in predator–prey dynamics. J Math Biol 39(5):439–469

    MathSciNet  MATH  Google Scholar 

  • Klausmeier CA (2010) Successional state dynamics: a novel approach to modeling nonequilibrium foodweb dynamics. J Theor Biol 262(4):584–595

    MathSciNet  MATH  Google Scholar 

  • Koch AL (1974a) Coexistence resulting from an alternation of density dependent and density independent growth. J Theor Biol 44(2):373–386

    Google Scholar 

  • Koch AL (1974b) Competitive coexistence of two predators utilizing the same prey under constant environmental conditions. J Theor Biol 44(2):387–395

    Google Scholar 

  • Lasota A, Mackey MC (1999) Cell division and the stability of cellular populations. J Math Biol 38(3):241–261

    MathSciNet  MATH  Google Scholar 

  • Levin SA (1970) Community equilibria and stability, and an extension of the competitive exclusion principle. Am Nat 104(939):413–423

    Google Scholar 

  • Li L, Chesson P (2016) The effects of dynamical rates on species coexistence in a variable environment: the paradox of the plankton revisited. Am Nat 188(2):E46–E58

    Google Scholar 

  • Litchman E, Klausmeier CA (2001) Competition of phytoplankton under fluctuating light. Am Nat 157(2):170–187

    Google Scholar 

  • Malrieu F, Zitt P-A (2017) On the persistence regime for Lotka–Volterra in randomly fluctuating environments. ALEA 14(2):733–749

    MathSciNet  MATH  Google Scholar 

  • Malrieu F, Phu TH (2016) Lotka–Volterra with randomly fluctuating environments: a full description. arXiv preprint arXiv:1607.04395

  • Mao X (1997) Stochastic differential equations and their applications. Horwood Publishing series in mathematics & applications. Horwood Publishing Limited, Chichester

    Google Scholar 

  • McGehee R, Armstrong RA (1977) Some mathematical problems concerning the ecological principle of competitive exclusion. J Differ Equ 23(1):30–52

    MathSciNet  MATH  Google Scholar 

  • Rinaldi S, Muratori S, Kuznetsov Y (1993) Multiple attractors, catastrophes and chaos in seasonally perturbed predator–prey communities. Bull Math Biol 55(1):15–35

    MATH  Google Scholar 

  • Schreiber SJ, Benaïm M, Atchadé KAS (2011) Persistence in fluctuating environments. J Math Biol 62(5):655–683

    MathSciNet  MATH  Google Scholar 

  • Smith HL, Thieme HR (2011) Dynamical systems and population persistence, vol 118. American Mathematical Society, Providence

    MATH  Google Scholar 

  • Stewart FM, Levin BR (1973) Partitioning of resources and the outcome of interspecific competition: a model and some general considerations. Am Nat 107(954):171–198

    Google Scholar 

  • Turelli M (1977) Random environments and stochastic calculus. Theor Popul Biol 12(2):140–178

    MathSciNet  MATH  Google Scholar 

  • Turelli M, Gillespie JH (1980) Conditions for the existence of stationary densities for some two-dimensional diffusion processes with applications in population biology. Theor Popul Biol 17(2):167–189

    MathSciNet  MATH  Google Scholar 

  • Tyson R, Lutscher F (2016) Seasonally varying predation behavior and climate shifts are predicted to affect predator–prey cycles. Am Nat 188(5):539–553

    Google Scholar 

  • Volterra V (1928) Variations and fluctuations of the number of individuals in animal species living together. J Cons Int Explor Mer 3(1):3–51

    MathSciNet  Google Scholar 

  • Yuan C, Chesson P (2015) The relative importance of relative nonlinearity and the storage effect in the lottery model. Theor Popul Biol 105:39–52

    MATH  Google Scholar 

  • Yvinec R, Zhuge C, Lei J, Mackey MC (2014) Adiabatic reduction of a model of stochastic gene expression with jump Markov process. J Math Biol 68(5):1051–1070

    MathSciNet  MATH  Google Scholar 

  • Zicarelli JD (1975) Mathematical analysis of a population model with several predators on a single prey. University of Minnesota, Saint Paul

    Google Scholar 

Download references

Acknowledgements

We thank Jim Cushing and Simon Levin for their helpful suggestions. The manuscript has improved significantly due to the comments of Peter Chesson and one anonymous referee. The authors have been in part supported by the NSF through the grants DMS 1853463 (A. Hening) and DMS 1853467 (D. Nguyen). Part of this work has been done while AH was visiting the University of Sydney through an international visitor program fellowship.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexandru Hening.

Additional information

Publisher's Note

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

Appendices

Appendix A: Proof of Theorem 4.1

If the number of species is strictly greater than the number of resources, \(n>m\), the system

$$\begin{aligned} \sum _{i=1}^nc_{i} b_{ij}=0, ~j=1,\dots ,m \end{aligned}$$
(A.1)

admits a nontrivial solution \((c_1, \dots , c_n)\).

Theorem A.1

Assume that \(\lim _{\Vert \mathbf {x}\Vert \rightarrow \infty } R_j(\mathbf {x})=-\infty , j=1,\dots ,m\). Suppose further that n species interact according to (4.3), the number of species is greater than the number of resources \(n>m\) and the resources depend on the species densities according to (2.2) so that they eventually get exhausted. Suppose further that \(g_i(\mathbf {x})=1\) and

$$\begin{aligned} 0<r_m\le \liminf _{\Vert \mathbf {x}\Vert \rightarrow \infty } \frac{|R_{j}(\mathbf {x})|}{|R_1(\mathbf {x})|}\le r^M<\infty , \end{aligned}$$

for \(j=1,\dots ,m\). Let \((c_1,\dots ,c_n)\) be a non-trivial solution to (A.1) and assume that \(\sum _{i=1}^n c_i\left( \alpha _i+\frac{\sigma _{ii}}{2}\right) \ne 0\). Then, for any starting densities \(\mathbf {x}(0)\in (0,\infty )^n\) with probability 1

$$\begin{aligned} \limsup _{t\rightarrow 0}\dfrac{\ln \min \{x_1(t),\dots , x_n(t)\}}{t}<0. \end{aligned}$$

Proof of Theorem 4.1

Suppose \(g_i(\mathbf {x})=1\), \(i=1,\dots , n\) and \(\sum _jb_{ij}>b_m>0\) for any i and some \(b_m>0\). Note that if \(\sum _j b_{ij}=0\) then we can remove \(R_j\) from the equation. Assume that

$$\begin{aligned} 0<r_m\le \liminf _{\Vert \mathbf {x}\Vert \rightarrow \infty } \frac{|R_{j}(\mathbf {x})|}{|R_1(\mathbf {x})|}\le r^M<\infty . \end{aligned}$$

Then, since \(\lim _{\Vert \mathbf {x}\Vert \rightarrow \infty } R_j(\mathbf {x})=-\infty \), we have when \(|\mathbf {x}|\) large that:

$$\begin{aligned} \sum _{i}\left( -x_i\alpha _i+x_i\sum _j b_{ij} R_j(\mathbf {x})\right)\le & {} -\sum _{i}\left( x_i\alpha _i +x_i\sum _j b_{ij} |R_{j}(\mathbf {x})|\right) \\\le & {} -\sum _{i}\left( x_i\alpha _i +x_i r_m \sum _j b_{ij} |R_{1}(\mathbf {x})|\right) \\\le & {} - r_mb_m\left( \sum _i x_i\right) |R_1(\mathbf {x})|\\\le & {} -\frac{r_m b_m}{mR_m}\left( \sum _i x_i\right) \sum _j |R_j(\mathbf {x})| \end{aligned}$$

which together with the linearity of the diffusion part implies that Assumption 1.1 from the work by HN16 holds with \(\mathbf {c}=(1,\dots , 1)\). As a result, for any starting point \(\mathbf {x}(0)\in (0,\infty )^n\) the SDE (4.3) has a unique positive solution and by Hening and Nguyen (2018a) (equation (5.22)) with probability 1

$$\begin{aligned} \limsup _{\Vert \mathbf {x}\Vert \rightarrow \infty }\frac{\ln \Vert \mathbf {x}\Vert }{t}\le 0. \end{aligned}$$
(A.2)

By possibly replacing all \(c_i\) by \(-c_i\) we can assume that \(\sum _{i=1}^n c_i\left( \alpha _i+\frac{\sigma _{ii}}{2}\right) >0\). Using this in conjunction with (4.3), (A.1) and Itô’s Lemma we see that

$$\begin{aligned} \dfrac{\sum _{i=1}^n c_i\ln x_i(t)}{t}=-\sum _{i=1}^nc_i\left( \alpha _i+\frac{\sigma _{ii}}{2}\right) +\frac{1}{t}\sum _{i=1}^n c_i\int _0^tE_i(s)\,ds \end{aligned}$$

Letting \(t\rightarrow \infty \) and using that \(\lim _{t\rightarrow \infty }\dfrac{\sum _{i=1}^n c_iE_i(t)}{t}=0\) with probability 1, we obtain that with probability 1

$$\begin{aligned} \lim _{t\rightarrow \infty }\dfrac{\sum _{i=1}^n c_i\ln x_i(t)}{t}=-\sum _{i=1}^nc_i\left( \alpha _i+\frac{\sigma _{ii}}{2}\right) < 0. \end{aligned}$$

In view of (A.2) this implies that with probability 1

$$\begin{aligned} \limsup _{t\rightarrow \infty } \dfrac{\ln \left( \min \{x_1(t),\dots , x_n(t)\}\right) }{t}<0. \end{aligned}$$

\(\square \)

Appendix B: Proof of Theorem 4.2

Theorem B.14

Assume two species interact according to

$$\begin{aligned} dx_i(t) = x_i(t)\left( -\alpha _i+b_iR(\mathbf {x}(t))\right) \,dt + x_i(t)\sqrt{\beta _i x_i(t)} \,dB_i(t), ~i=1,2, \end{aligned}$$

the resource R depends linearly on the species densities

$$\begin{aligned} R(\mathbf {x})={\overline{R}} - a_1 x_1(t) - a_2x_2(t) \end{aligned}$$

and \(b_i{\overline{R}}>\alpha _i, i=1,2\). Then there exist \(\beta _1, \beta _2>0\) such that the two species coexist.

Proof

Consider

$$\begin{aligned} dx_i(t) = x_i(t)\left( -\alpha _i+b_iR(\mathbf {x}(t))\right) \,dt + x_i(t)\sqrt{\beta _i x_i(t)} \,dB_i(t), ~i=1,2. \end{aligned}$$

If the species \(x_2\) is absent species \(x_1\) has the one-dimensional dynamics

$$\begin{aligned} dx(t) = x(t)\left( -\alpha _1 + b_1 ({\overline{R}}-a_1x(t))\right) dt+ x(t)\sqrt{\beta _1 x(t)} \,dB_1(t). \end{aligned}$$

Since \(b_1{\overline{R}}>\alpha _1\), we can use Hening and Nguyen (2018a) to show that the process x(t) has a unique invariant measure on \((0,\infty )\), say \(\mu _1\). Moreover, (Hening and Nguyen 2018a, Lemma 2.1 ) shows that

$$\begin{aligned} \int _0^\infty \left( -\alpha _1+b_1({\overline{R}}-a_1 x)-\beta _1x\right) \mu _1(dx)=0 \end{aligned}$$

or

$$\begin{aligned} \int _0^\infty x\mu _1(dx)=\dfrac{b_1{\overline{R}}-\alpha _1}{b_1a_1+\beta _1}. \end{aligned}$$

The invasion rate of \(x_2\) with respect to \(x_1\) can be computed by (4.2) as

$$\begin{aligned} \Lambda _{x_2}=\int _0^\infty \left( -\alpha _2+b_2({\overline{R}}-a_1 x)\right) \mu _1(dx) =(b_2{\overline{R}}-\alpha _2)- b_2a_1 \dfrac{b_1\overline{R}-\alpha _1}{b_1a_1+\beta _1}. \end{aligned}$$

Similarly, one can compute the invasion rate of \(x_1\) with respect to \(x_2\) as

$$\begin{aligned} \Lambda _{x_1}=(b_1{\overline{R}}-\alpha _1)- b_1a_2 \dfrac{b_2\overline{R}-\alpha _2}{b_2a_2+\beta _2}. \end{aligned}$$

Since \(b_i{\overline{R}}-\alpha _i>0\), one can easily see that \( \Lambda _{x_2}>0\), and \(\Lambda _{x_1}>0\) if \( \beta _1>\dfrac{b_2a_1(b_1{\overline{R}}-\alpha _1)}{b_2\overline{R}-\alpha _2}-b_1a_1. \) and \( \beta _2>\dfrac{b_1a_2(b_2\overline{R}-\alpha _2)}{b_1{\overline{R}}-\alpha _1}-b_2a_2. \) If both invasion rates are positive we get by Hening and Nguyen (2018a) that the species coexist. \(\square \)

Appendix C: Proof of Theorem 4.3

We construct an SDE example of two species competing for one abiotic resource and coexisting. We remark that this happens solely because of the random temporal environmental variation term.

Theorem C.1

Suppose the dynamics of the two species is given by

$$\begin{aligned} \begin{aligned} dx_1(t)&=x_1(t)(-\alpha _1+f({\overline{R}}-a_1x_1(t)-a_2x_2(t)))\,dt +\sigma _1x_1dB_1(t)\\ dx_2(t)&=x_2(t)(-\alpha _2+({\overline{R}}-a_1x_1(t)-a_2x_2(t)))\,dt \end{aligned} \end{aligned}$$
(C.1)

where f is a continuously differentiable Lipschitz function satisfying \(\lim _{x\rightarrow -\infty }f(x)=-\infty \), \(\dfrac{df(x)}{dx}>0, \dfrac{d^2f(x)}{dx^2}\le 0\) for all \(x\in \mathbb {R}\) and \(\dfrac{d^2f(x)}{dx^2}<0\) for x in some subinterval of \(\left( -\infty ,\frac{{\overline{R}}}{a_1}\right) \). Let \(a_1, a_2, \sigma _1, \alpha _1,\sigma _1, {\overline{R}}\) be any fixed positive constants satisfying \(f({\overline{R}})>\alpha _1+\dfrac{\sigma _1^2}{2}\). Then there exists an interval \((c_0,c_1)\subset (0,\infty )\) such that the two species coexist for all \(\alpha _2\in (c_0,c_1)\).

Proof

The dynamics of species \(x_1\) in the absence of species \(x_2\) is given by the one-dimensional SDE

$$\begin{aligned} dx(t)=x(t)(-\alpha _1+f({\overline{R}}-a_1x(t)))\,dt +\sigma _1xdB_1(t).\\ \end{aligned}$$

Since \(\lim _{x\rightarrow \infty }f({\overline{R}}-a_1x)=-\infty \), and \(f(\overline{R})>\alpha _1+\dfrac{\sigma _1^2}{2}\), this diffusion has a unique invariant probability measure \(\mu \) on \((0,\infty )\) whose density is strictly positive on \((0,\infty )\) (see Borodin and Salminen (2016) or Mao (1997)). Moreover, by noting that \(\lim _{t\rightarrow \infty }\frac{\ln x(t)}{t}=0\) with probability 1 (using Lemma 5.1 of Hening and Nguyen (2018a)) and using Itô’s formula one sees that

$$\begin{aligned} \int _0^\infty f({\overline{R}}-a_1x)\mu (dx)=\alpha _1+\dfrac{\sigma _1^2}{2}. \end{aligned}$$
(C.2)

Since f is a concave function and \(\dfrac{d^2f(x)}{dx^2}>0\) for all x in some subinterval of \(\left( -\infty ,\frac{\overline{R}}{a_1}\right) \) we must have by Jensen’s inequality that

$$\begin{aligned} \int _0^\infty f({\overline{R}}-a_1x)\mu (dx)< f\left( \int _0^\infty (\overline{R}-a_1x)\mu (dx)\right) . \end{aligned}$$
(C.3)

The fact that the function f is strictly increasing together with (C.2) and (C.3) forces

$$\begin{aligned} \varepsilon _0:=\int _0^\infty ({\overline{R}}-a_1x)\mu (dx)- f^{-1}\left( \alpha _1+\dfrac{\sigma _1^2}{2}\right) >0, \end{aligned}$$
(C.4)

where \(f^{-1}\) is the inverse of f—it exists because f is strictly increasing. As a result, the invasion rate of species \(x_2\) with respect to \(x_1\), of the invariant probability measure \(\mu \), can be computed using (4.2) as

$$\begin{aligned} \Lambda _{x_2}=-\alpha _2+\int _0^\infty ({\overline{R}}-a_1x)\mu (dx)=\varepsilon _0+ f^{-1}\left( \alpha _1+\dfrac{\sigma _1^2}{2}\right) - \alpha _2. \end{aligned}$$

This implies that \(\Lambda _{x_2}>0\) if and only if

$$\begin{aligned} \alpha _2<\varepsilon _0+ f^{-1}\left( \alpha _1+\dfrac{\sigma _1^2}{2}\right) . \end{aligned}$$
(C.5)

The dynamics of species \(x_2\) in the absence of species \(x_1\) is

$$\begin{aligned} dy(t)=y(t)({\overline{R}}-\alpha _2-a_2y(t))dt. \end{aligned}$$

The positive solutions of this equation converge to the point \(y^*=\dfrac{{\overline{R}}-\alpha _2}{a_2}\) if and only if

$$\begin{aligned} {\overline{R}}>\alpha _2. \end{aligned}$$
(C.6)

The invasion rate of \(x_1\) with respect to \(x_2\) will be

$$\begin{aligned} \Lambda _{x_1}=-\alpha _1-\frac{\sigma _1^2}{2}+f(\alpha _2). \end{aligned}$$

Note that since the function f is increasing we get \(\Lambda _{x_1}>0\) if and only if

$$\begin{aligned} \alpha _2>f^{-1}\left( \alpha _1+\frac{\sigma _1^2}{2}\right) . \end{aligned}$$
(C.7)

Note that \(f^{-1}\left( \alpha _1+\dfrac{\sigma _1^2}{2}\right) <{\overline{R}}\) since by assumption \(f({\overline{R}})>\alpha _1+\dfrac{\sigma _1^2}{2}\). As a result, making use of the inequalities (C.5), (C.6) and (C.7) we get that \(\Lambda _{x_2}>0, \Lambda _{x_1}>0\) if any only if

$$\begin{aligned} \alpha _2\in \left( f^{-1}\left( \alpha _1+\dfrac{\sigma _1^2}{2}\right) , \overline{R}\wedge \left( f^{-1}\left( \alpha _1+\dfrac{\sigma _1^2}{2}\right) +\varepsilon _0\right) \right) . \end{aligned}$$

This implies by Theorem 3.1 or by Benaim (2018) [Theorem 4.4 and Definition 4.3] that the two species coexist. \(\square \)

Appendix D: Proof of Theorem 5.1

Theorem D.2

Assume that

$$\begin{aligned} \lim _{\Vert \mathbf {x}\Vert \rightarrow \infty } \left( -\alpha _i(u)+\sum _{j=1}^mb_{ij}(u)R_j(\mathbf {x},u) \right) <0, i=1,\dots , n, u=1,\dots ,N.\quad \quad \quad \end{aligned}$$
(D.1)

Suppose further that there exists a vector \((c_1,\dots ,c_n)\) that is simultaneously a solution to the systems (5.5) for all \(u\in \{1,\dots ,N\}\). Then, with probability 1,

$$\begin{aligned} \limsup _{t\rightarrow 0}\dfrac{\ln \min \{x_1(t),\dots , x_n(t)\}}{t}<0 \end{aligned}$$

except possibly for the critical case when

$$\begin{aligned} \sum _{i=1}^n c_i\sum _{k=1}^N\alpha _i(k)\nu _k=0, \end{aligned}$$
(D.2)

where \((\nu _k)_{k\in \mathcal {N}}\) is the invariant probability measure of the Markov chain (r(t)).

Proof

Under the condition (D.1), there exists an \(M>0\) such that the set \(K_M:=\{\mathbf {x}\in \mathbb {R}^n: \Vert \mathbf {x}\Vert \le M\}\) is a global attractor of (5.4). As a result, the solution to (5.4) eventually enters and never leaves the compact set \(K_M\). In particular, this shows that the process \(\mathbf {x}(t)\) is bounded. Next, note that we can assume that

$$\begin{aligned} \sum _{i=1}^n c_i\sum _{k=1}^N\alpha _i(k)\pi _k>0. \end{aligned}$$
(D.3)

Otherwise, if \(\sum _{i=1}^n c_i\sum _{k=1}^N\alpha _i(k)\pi _k<0\), we can replace \(c_i\) by \(-c_i\), \(i=1,\dots , n\) and then get (D.3). Using (5.1) and the fact that that \(c_i\)’s solve (5.5) simultaneously we get

$$\begin{aligned} \dfrac{\sum _{i=1}^n c_i\ln x_i(t)}{t}=-\dfrac{1}{t}\int _0^t\sum _{i=1}^n c_i\alpha _i(r(s))ds \end{aligned}$$

Letting \(t\rightarrow \infty \) and using the ergodicity of the Markov chain (r(t)) we obtain that with probability 1

$$\begin{aligned} \limsup _{t\rightarrow \infty }\dfrac{\sum _{i=1}^n c_i\ln x_i(t)}{t}= & {} -\liminf _{t\rightarrow \infty }\dfrac{1}{t}\int _0^t\sum _{i=1}^n c_i\alpha _i(r(s))ds\\= & {} -\sum _{i=1}^n c_i\sum _{k=1}^N\alpha _i(k)\pi _k< 0 \text { a.s.} \end{aligned}$$

Since \(\mathbf {x}(t)\) is bounded, this implies that with probability 1

$$\begin{aligned} \limsup _{t\rightarrow \infty } \dfrac{\ln \min \{x_1(t),\dots , x_n(t)\}}{t}<0. \end{aligned}$$

\(\square \)

Appendix E: Proof of Theorem 5.2

According to Benaïm and Lobry (2016), Malrieu and Zitt (2017), and Malrieu and Phu (2016) it is enough to find an example for which the invasion rates \(\Lambda _{x_1}, \Lambda _{x_2}\) are positive. We will follow Benaïm and Lobry (2016) in order to compute the invasion rates of the two species. Set for \(u=1,2\)\(\mu _u=-\alpha _1(u)+b_1(u){\overline{R}}, \nu _u=-\alpha _2(u)+b_2(u){\overline{R}}\)\({\overline{a}}_u=\dfrac{b_1(u)a_1(u)}{\mu _u}, \overline{b}_u=\dfrac{b_1(u)a_2(u)}{\mu _u}, \overline{c}_u=\dfrac{b_2(u)a_1(u)}{\nu _u}, \overline{d}(u)=\dfrac{b_2(u)a_2(u)}{\nu _u}\), \(p_u=\dfrac{1}{{\overline{a}}_u}, q_u=\dfrac{1}{{\overline{d}}_u}, \gamma _1=\dfrac{q_{12}}{\mu _u}, \gamma _2=\dfrac{q_{21}}{\nu _u}\). If \(p_1\ne p_2\), suppose without loss of generality that \(p_1<p_2\). Define the functions

$$\begin{aligned} \theta (x)=\dfrac{|x-p_1|^{\gamma _1-1}|p_2-x|^{\gamma _2-1}}{x^{1+\gamma _1+\gamma _2}} \end{aligned}$$

and

$$\begin{aligned} P(x)=\dfrac{{\overline{a}}_2-{\overline{a}}_1}{|{\overline{a}}_2-\overline{a}_1|}\left[ \dfrac{\nu _2}{\mu _2}(1-{\overline{c}}_2 x)(1-\overline{a}_1x)-\dfrac{\nu _1}{\mu _1}(1-{\overline{c}}_1 x)(1-{\overline{a}}_2x)\right] . \end{aligned}$$

By Benaïm and Lobry (2016) we have

$$\begin{aligned} \Lambda _{x_2}={\left\{ \begin{array}{ll} \dfrac{1}{q_{12}+q_{21}}\left( q_{21}\nu _1(1-{\overline{c}}_1 p)+q_{12}\nu _2(1-{\overline{c}}_2p)\right) &{}\text { if } p_1=p_2=p\\ p_1p_2\dfrac{\int _{p_1}^{p_2}\theta (x)P(x)dx}{\int _{p_1}^{p_2}\theta (x)dx}&{}\text { if } p_1<p_2 \end{array}\right. } \end{aligned}$$
(E.1)

The expression for \(\Lambda _{x_1}\) can be obtained by swapping \(\mu _i\) and \(\nu _i\), \(({\overline{a}}_i,{\overline{c}}_i)\) with \(({\overline{d}}_i,\overline{b}_i)\), and \(p_i\) with \(q_i\).

For the example from Figs. 3 and 4 we have used the integral equation from (E.1) together with the numerical integration package of Mathematica in order to find \(\Lambda _{x_1}>0\) and \(\Lambda _{x_2}>0\). This implies by Theorem 3.1, Benaïm and Lobry (2016) or by Benaim (2018) [Theorem 4.4 and Definition 4.3] that the two species coexist.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hening, A., Nguyen, D.H. The competitive exclusion principle in stochastic environments. J. Math. Biol. 80, 1323–1351 (2020). https://doi.org/10.1007/s00285-019-01464-y

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-019-01464-y

Keywords

Mathematics Subject Classification

Navigation