Skip to main content

Advertisement

Log in

Persistence and extinction for stochastic ecological models with internal and external variables

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

Abstract

The dynamics of species’ densities depend both on internal and external variables. Internal variables include frequencies of individuals exhibiting different phenotypes or living in different spatial locations. External variables include abiotic factors or non-focal species. These internal or external variables may fluctuate due to stochastic fluctuations in environmental conditions. The interplay between these variables and species densities can determine whether a particular population persists or goes extinct. To understand this interplay, we prove theorems for stochastic persistence and exclusion for stochastic ecological difference equations accounting for internal and external variables. Specifically, we use a stochastic analog of average Lyapunov functions to develop sufficient and necessary conditions for (i) all population densities spending little time at low densities i.e. stochastic persistence, and (ii) population trajectories asymptotically approaching the extinction set with positive probability. For (i) and (ii), respectively, we provide quantitative estimates on the fraction of time that the system is near the extinction set, and the probability of asymptotic extinction as a function of the initial state of the system. Furthermore, in the case of persistence, we provide lower bounds for the expected time to escape neighborhoods of the extinction set. To illustrate the applicability of our results, we analyze stochastic models of evolutionary games, Lotka–Volterra dynamics, trait evolution, and spatially structured disease dynamics. Our analysis of these models demonstrates environmental stochasticity facilitates coexistence of strategies in the hawk–dove game, but inhibits coexistence in the rock–paper–scissors game and a Lotka–Volterra predator–prey model. Furthermore, environmental fluctuations with positive auto-correlations can promote persistence of evolving populations and persistence of diseases in patchy landscapes. While our results help close the gap between the persistence theories for deterministic and stochastic systems, we highlight several challenges for future research.

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

Similar content being viewed by others

References

  • Altizer S, Dobson P, Hosseini A, Hudson P, Pascual M, Rohani P (2006) Seasonality and the dynamics of infectious diseases. Ecol Lett 9:467–484

    Article  Google Scholar 

  • Axelrod R (1984) The evolution of cooperation. Basic Books, New York

    MATH  Google Scholar 

  • Axelrod R, Hamilton W (1981) The evolution of cooperation. Science 211:1390–1396

    Article  MathSciNet  MATH  Google Scholar 

  • Benaïm M (2018) Stochastic persistence, part I. arXiv preprint arXiv:1806.08450

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

    Article  MathSciNet  MATH  Google Scholar 

  • Benaïm M, Schreiber SJ (2009) Persistence of structured populations in random environments. Theor Popul Biol 76:19–34

    Article  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

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

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Chesson PL (1985) Coexistence of competitors in spatially and temporally varying environments: a look at the combined effects of different sorts of variability. Theor Popul Biol 28:263–287

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Cuddington K, Wilson WG, Hastings A (2009) Ecosystem engineers: feedback and population dynamics. Am Nat 173:488–498

    Article  Google Scholar 

  • Diaconis P, Freedman D (1999) Iterated random function. SIAM Rev 41:45–76

    Article  MathSciNet  MATH  Google Scholar 

  • Eltahir EAB (1998) A soil moisture-rainfall feedback mechanism: 1. Theory and observations. Water Resour Res 34:765–776

    Article  Google Scholar 

  • Evans SN, Ralph P, Schreiber SJ, Sen A (2013) Stochastic growth rates in spatio-temporal heterogeneous environments. J Math Biol 66:423–476

    Article  MathSciNet  MATH  Google Scholar 

  • Ferriere R, Gatto M (1995) Lyapunov exponents and the mathematics of invasion in oscillatory or chaotic populations. Theor Popul Biol 48:126–171

    Article  MATH  Google Scholar 

  • Garay BM, Hofbauer J (2003) Robust permanence for ecological differential equations, minimax, and discretizations. SIAM J Math Anal 34:1007–1039

    Article  MathSciNet  MATH  Google Scholar 

  • Gillespie J (1973) Polymorphism in random environments. Theor Popul Biol 4:193–195

    Article  Google Scholar 

  • Gillespie JH (1978) A general model to account for enzyme variation in natural populationsV. The SAS-CFF model. Theor Popul Biol 14:1–45

    Article  MATH  Google Scholar 

  • Gillespie JH, Turelli M (1989) Genotype-environment interactions and the maintenance of polygenic variation. Genetics 121:129–138

    Google Scholar 

  • Gyllenberg M, Hognas G, Koski T (1994) Population models with environmental stochasticity. J Math Biol 32:93–108

    Article  MathSciNet  MATH  Google Scholar 

  • Hairer M (2006) Ergodic properties of Markov processes. Lectures given at The University of Warwick, Spring. http://www.hairer.org/notes/Markov.pdf

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

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  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 76:697–754

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Hofbauer J, Schreiber SJ (2010) Robust permanence for interacting structured populations. J Differ Eq 248:1955–1971

    Article  MathSciNet  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

  • Hofbauer J, Sigmund K (2003) Evolutionary game dynamics. Bull Am Math Soc 40:479–519

    Article  MathSciNet  MATH  Google Scholar 

  • Hofbauer J, Hutson V, Jansen W (1987) Coexistence for systems governed by difference equations of Lotka–Volterra type. J Math Biol 25(5):553–570

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  • Jones CG, Lawton JH, Shachak M (1994) Organisms as ecosystem engineers. Oikos 69:373–386

    Article  Google Scholar 

  • Jordan-Cooley WC, Lipcius RN, Shaw LB, Shen J, Shi J (2011) Bistability in a differential equation model of oyster reef height and sediment accumulation. J Theor Biol 289:1–11

    Article  MathSciNet  MATH  Google Scholar 

  • Kerr B, Riley M, Feldman M, Bohannan J (2002) Local dispersal promotes biodiversity in a real-life game of rock-paper-scissors. Nature 418:171–174

    Article  Google Scholar 

  • Kifer Y (1986) Ergodic theory of random transformations. Birkhauser, New York

    Book  MATH  Google Scholar 

  • Kifer Y (1988) Random perturbations of dynamical systems, volume 16 of progress in probability and statistics. Birkhäuser Boston Inc., Boston

    Book  Google Scholar 

  • Kirkup B, Riley M (2004) Antibiotic-mediated antagonism leads to a bacterial game of rock–paper–scissors in vivo. Nature 428:412–414

    Article  Google Scholar 

  • Kuang JJ, Chesson PL (2009) Coexistence of annual plants: generalist seed predation weakens the storage effect. Ecology 90:170–182

    Article  Google Scholar 

  • Kucharski F, Zeng N, Kalnay E (2013) A further assessment of vegetation feedback on decadal sahel rainfall variability. Clim Dyn 40:1453–1466

    Article  Google Scholar 

  • Lande R (1976) Natural selection and random genetic drift in phenotypic evolution. Evolution 30:314–334

    Article  Google Scholar 

  • Lande R, Shannon S (1996) The role of genetic variation in adaptation and population persistence in a changing environment. Evolution 50:434–437

    Article  Google Scholar 

  • Lewontin RC, Cohen D (1969) On population growth in a randomly varying environment. Proc Natl Acad Sci USA 62:1056–1060

    Article  MathSciNet  Google Scholar 

  • Maynard Smith J (1982) Evolution and the theory of games. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Maynard Smith J, Price GR (1973) The logic of animal conflict. Nature 246:15–18

    Article  MATH  Google Scholar 

  • McLaughlin JF, Hellmann JJ, Boggs CL, Ehrlich PR (2002) Climate change hastens population extinctions. Proc Nat Acad Sci USA 99:6070–6074

    Article  Google Scholar 

  • Metz JAJ, de Jong TJ, Klinkhamer PGL (1983) What are the advantages of dispersing; a paper by Kuno extended. Oecologia 57:166–169

    Article  Google Scholar 

  • Metz JAJ, Geritz SAH, Meszéna G, Jacobs EJA, van Heerwaarden JS (1996) Adaptive dynamics, a geometrical study of the consequences of nearly faithful reproduction. Dynamical systems and their applications. Elsevier, North Holland, pp 147–194

    Google Scholar 

  • Meyn SP, Tweedie RL (2009) Markov Chains and stochastic stability. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Michod RE (2000) Darwinian dynamics: evolutionary transitions in fitness and individuality. Princeton University Press, Princeton

    Google Scholar 

  • Moore J, Puckett B, Schreiber SJ (2018) Restoration of eastern oyster populations with positive density dependence. Ecol Appl 28:897–909

    Article  Google Scholar 

  • Moore JL, Lipcius RN, Puckett B, Schreiber SJ (2016) The demographic consequences of growing older and bigger in oyster populations. Ecol Appl 26:2206–2217

    Article  Google Scholar 

  • Nahum JR, Harding BN, Kerr B (2011) Evolution of restraint in a structured rock–paper–scissors community. Proc Nat Acad Sci 108:10831–10838

    Article  Google Scholar 

  • Patel S, Schreiber SJ (2018) Robust permanence for ecological equations with internal and external feedbacks. J Math Biol 77:79–105

    Article  MathSciNet  MATH  Google Scholar 

  • Roth G, Schreiber SJ (2014a) Pushed to brink: Allee effects, environmental stochasticity, and extinction. J Biol Dyn 8:187–205

    Article  MathSciNet  Google Scholar 

  • Roth G, Schreiber SJ (2014b) Persistence in fluctuating environments for interacting structured populations. J Math Biol 68:1267–1317

    Article  MathSciNet  MATH  Google Scholar 

  • Schoener TW (2011) The newest synthesis: understanding the interplay of evolutionary and ecological dynamics. Science 331:426–429

    Article  Google Scholar 

  • Schreiber SJ (2000) Criteria for \({C}^r\) robust permanence. J Differ Eq 162:400–426

    Article  MATH  MathSciNet  Google Scholar 

  • Schreiber SJ (2006) Persistence despite perturbations for interacting populations. J Theor Biol 242:844–52

    Article  MathSciNet  Google Scholar 

  • Schreiber SJ (2012) Persistence for stochastic difference equations: a mini-review. J Differ Eq Appl 18:1381–1403

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Schreiber SJ (2010) Interactive effects of temporal correlations, spatial heterogeneity, and dispersal on population persistence. Proc R Soc Biol Sci 277:1907–1914

    Article  Google Scholar 

  • Schreiber SJ (2019) When do factors promoting balanced selection also promote population persistence? a demographic perspective on Gillespie’s SAS-CFF model. arXiv preprint arXiv:1902.03507

  • Schreiber SJ, Moore J (2018) The structured demography of open populations in fluctuating environments. Methods Ecol Evol 9:1569–1580

    Article  Google Scholar 

  • Schreiber SJ, Patel S, TerHorst C (2018) Evolution as a coexistence mechanism: does genetic architecture matter? Am Nat 191:407–420

    Article  Google Scholar 

  • Sinervo B, Lively C (1996) The rock-paper-scissors game and the evolution of alternative male strategies. Nature 380:240–243

    Article  Google Scholar 

  • Staver AC, Levin SA (2012) Integrating theoretical climate and fire effects on savanna and forest systems. Am Nat 180:211–224

    Article  Google Scholar 

  • Turelli M (1981) Niche overlap and invasion of competitors in random environments I. Models without demographic stochasticity. Theoretical Population Biology 20:1–56

    Article  MathSciNet  MATH  Google Scholar 

  • Vincent TL, Brown JS (2005) Evolutionary game theory, natural selection, and Darwinian dynamics. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Whitlock MC, Davis B (2011) Genetic load. eLS. https://doi.org/10.1002/9780470015902.a0001787.pub2

  • Williams D (1991) Probability with martingales. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Zeng N, Neelin JD, Lau KM, Tucker CJ (1999) Enhancement of interdecadal climate variability in the sahel by vegetation interaction. Science 286:1537–1540

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported in part by United States National Science Foundation Grant DMS-1716803 to SJS and Swiss National Science Foundation Grant 2000021-175728 to MB. We thank three anonymous reviewers for comments on an earlier draft of their paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sebastian J. Schreiber.

Additional information

Publisher's Note

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

Proofs

Proofs

1.1 Proof of Proposition 1

We begin with the following fundamental lemma. Recall \(\xi _t\) take values in a polish space \(\varXi \) and have a common law \(m(d\xi )\).

Lemma 3

Let \(g : \mathcal {S}\times \varXi \mapsto \mathbb {R}\) be a measurable map such that \(\sup _{z\in \mathcal {S}} \int g(z,\xi )^2 m(d\xi )<\infty . \) Define \(\overline{g}(z) = \int g(z,\xi ) m(d\xi ).\) Then

  1. (i)

    For all \(z \in \mathcal {S}\) and \(Z_0 = z\)

    $$\begin{aligned} \lim _{t \rightarrow \infty } \frac{\sum _{s = 0}^{t-1} g(Z_s,\xi _{s+1}) - \sum _{s = 0}^{t-1} \overline{g}(Z_s)}{t} = 0 \text{ with } \text{ probability } \text{ one. } \end{aligned}$$
  2. (ii)

    Let \(\mu \) be an invariant (respectively ergodic) probability measure for \((Z_t)\), then there exists a bounded measurable map \(\widehat{g}\) such that with probability one and for \(\mu \)-almost every z

    $$\begin{aligned} \lim _{t \rightarrow \infty } \frac{\sum _{s = 0}^{t-1} g(Z_s,\xi _{s+1})}{t} = \lim _{t \rightarrow \infty } \frac{\sum _{s = 0}^{t-1} \overline{g}(Z_s)}{t} = \widehat{g}(z) \text{ when } Z_0=z. \end{aligned}$$

    Furthermore

    $$\begin{aligned}&\int \overline{g}(z) \mu (dz) = \int \widehat{g}(z) \mu (dz) \text{(respectively } \widehat{g}(z)\\&\quad = \int \overline{g}(z) \mu (dz) \quad \mu -\text{ almost } \text{ surely } ). \end{aligned}$$

Proof

To prove the first assertion, let \(C=\sup _{z\in \mathcal {S}}\int g(z,\xi )^2m(d\xi )\) and define the martingale

$$\begin{aligned} M_t=\sum _{s=0}^{t-1}\left( g(Z_s,\xi _{s+1})-\overline{g}(Z_s)\right) . \end{aligned}$$

with respect to the sigma-algebra \({\mathcal {F}}_t\) generated by \((Z_0,\xi _0),\dots ,(Z_t,\xi _t)\). \(M_t\) has square integrable martingale differences as \(\mathbb {E}[(M_t-M_{t-1})^2|{\mathcal {F}}_{t-1}]\le 2C^2.\) Define the previsible increasing process \(\langle M\rangle _t\) by \(\langle M\rangle _0=0\) and \(\langle M\rangle _t=\langle M\rangle _{t-1}+\mathbb {E}[(M_t-M_{t-1})^2|{{\mathcal {F}}}_{t-1}]\). \(\langle M\rangle _t\) is known as the angle-brackets process (Williams 1991, Sect. 12.12). By construction, we have \(\langle M\rangle _t\le 2C^2 t\). Define \(\langle M\rangle _\infty =\lim _{t\rightarrow \infty } \langle M\rangle _t\), which exists (possibly infinite) as \(\langle M\rangle _t\) is increasing. On the event \(\langle M\rangle _\infty <+\infty \), Williams (1991, Theorem 12.13a) implies that \(\lim _{t\rightarrow \infty }M_t\) exists and is finite. In particular, \(\lim _{t\rightarrow \infty }M_t/t=0\) on the event \(\langle M\rangle _\infty <+\infty .\) On the event \(\langle M\rangle _\infty =+\infty \), Williams (1991, Theorem 12.14a) implies that \(\lim _{t\rightarrow \infty }M_t/\langle M\rangle _t=0\). In particular, as \(\langle M\rangle _t\le 2C^2t\), \(\lim _{t\rightarrow \infty }M_t/t=0\) on the event \(\langle M\rangle _\infty =+\infty \). Thus, \(\lim _{t\rightarrow \infty }M_t/t=0\) with probability one which completes the proof of the first assertion.

The second assertion follows from Birkhoff’s ergodic theorem applied to stationary Markov Chains (see Meyn and Tweedie (2009), Theorem 17.1.2) \(\square \)

Equation (3.2) and the second assertions of Proposition 1 follow directly from the preceding lemma applied to \(g(z,\xi ) = \log f^i(z,\xi )\). Assumption A4 implies that \(\log f^i(z,\xi _t)\) are uniformly integrable (UI). Therefore, equation (3.3) follows from (3.2) and the UI convergence theorem. The first half of the third assertion follows from ergodicity. To prove the second half of the third assertion, let \(i\in S(\mu )\). By assertion (i) of Proposition 1

$$\begin{aligned} \lim _{t \rightarrow \infty } \frac{\log X^i_t}{t} = {\widehat{r}}^{i}(z)\quad \text{ where } Z_0=z \end{aligned}$$

for \(\mu \)-almost \(z \in \{(x,y)\in \mathcal {S}: x^i>0\}.\) Let \(\mathcal {S}^{i,\eta } = \{(x,y) \in \mathcal {S}\, : x^i \ge \eta \}\) and \(\eta ^*>0\) be such that \(\mu (\mathcal {S}^{i,\eta })>0\) for all \(\eta \le \eta ^*\). By Poincaré Recurrence Theorem, for \(\mu \) almost all \(z \in \mathcal {S}^{i,\eta }\)

$$\begin{aligned} \mathbb {P}_x[ Z_t \in \mathcal {S}^{i,\eta } \text{ infinitely } \text{ often }] = 1 \end{aligned}$$

for \(\eta \le \eta ^*\). Thus \(\widehat{r}_{i}(z) = 0\) for \(\mu \)-almost all \(z \in \mathcal {S}^{i,\eta }\) with \(\eta \le \eta ^*\). Hence \(\widehat{r}^{i}(z) = 0\) for \(\mu \)-almost all \(z \in \bigcup _{n \in \mathbf {N}} \mathcal {S}^{i,1/n} = \{(x,y) \in \mathcal {S}\, : x^i > 0\}.\) This proves assertion (iii).

1.2 Proof of Proposition 2

We begin by showing there is \(T\ge 1\) and \(\alpha \in (0,1)\) such that inequality (3.5) holds i.e. \(P_TV(z)-V(z)\le -\alpha \) for all \(z\in \mathcal {S}_0\). Suppose to the contrary. Then there exists an increasing sequence of times \(t_k\uparrow \infty \), a decreasing sequence of positive reals \(\alpha _k\downarrow 0\) and a sequence of points \(z_k\) in \(\mathcal {S}_0\) such that \(P_{t_k}V(z_k)-V(z_k)\ge -\alpha _k\) for all k. Define a sequence of Borel probability measures \(\mu _k\) on \(\mathcal {S}_0\) by

$$\begin{aligned} \int h(z) \mu _k(dz)= \frac{1}{t_k}\mathbb {E}_{z_k}\left[ \sum _{\tau =0}^{t_k-1} h(Z_\tau )\right] \text{ for } \text{ any } \text{ continuous } h:\mathcal {S}_0\rightarrow \mathbb {R}\text{. } \end{aligned}$$

Let \(\mu \) be a weak* limit point of the \(\mu _i\), which exists as \(\mathcal {S}_0\) is compact. By a standard argument due to Khasminskii (see, e.g., Kifer (1988, Theorem 1.1)), \(\mu \) is an invariant measure for the Markov chain and by weak* compactness, \(\sum _i p^i r^i(\mu )\le 0\). By the ergodic decomposition theorem, there exists an ergodic measure \(\nu \) supported on \(\mathcal {S}_0\) such that inequality (3.4) is violated; a contradiction.

Let \(\phi (x)=e^x-1-x\). For any real C,

$$\begin{aligned} |\phi (-\theta C)|\le \sum _{k\ge 2} \frac{|\theta C|^k}{k!} \le \theta ^2 e^C\quad \text{ whenever } |\theta |\le 1. \end{aligned}$$

Our assumption that \(\sup _{z,\xi }|\log f^i(z,\xi )|<\infty \) implies that there exists \(C>0\) such that

$$\begin{aligned} |\sum _{\tau =0}^{T-1}\sum _i p^i \log f^i (z_\tau ,\xi _{\tau })|\le C \end{aligned}$$

for any \(z_0,\dots ,z_{T-1}\in \mathcal {S}_0\) and \(\xi _1,\dots ,\xi _T\in \varXi .\) Thus, for \(z\in \mathcal {S}{\setminus } \mathcal {S}_0\)

$$\begin{aligned} P_T V_\theta (z)= & {} \mathbb {E}_z\left[ \exp (-\theta \sum _i p^i \log X_T^i) \right] \\= & {} \mathbb {E}_z\left[ \exp \left( -\theta \sum _i p^i \left( \sum _{\tau =0}^{T-1}\log f^i(Z_\tau ,\xi _{\tau +1})+\log x^i\right) \right) \right] \\= & {} \mathbb {E}_z\left[ \exp \left( -\theta \sum _i p^i \sum _{\tau =0}^{T-1}\log f^i(Z_\tau ,\xi _{\tau +1}\right) \right] V_\theta (z)\\= & {} \mathbb {E}_z\left[ 1-\theta \sum _i p^i \sum _{\tau =0}^{T-1}\log f^i(Z_\tau ,\xi _{\tau +1})\right. \\&\left. +\phi \left( -\theta \sum _i p^i \sum _{\tau =0}^{T-1}\log f^i(Z_\tau ,\xi _{\tau +1})\right) \right] V_\theta (z)\\\le & {} (1-\theta \alpha +\theta ^2e^C )V_\theta (z). \end{aligned}$$

Choosing \(\theta =\alpha \,e^{-C}/2\) and \(\rho =1-\theta \alpha /2\) completes the proof of the proposition.

1.3 Proof of Theorem 1

In this section, we provide the details of the proof of Theorem 1 that are not presented in the main text. Through out this section, we assume that (3.4) holds. Let \(\theta >0\), \(T\ge 1\), \(\rho \in (0,1)\), and \(\beta >0\) be as given in Proposition 2.

We begin proving persistence in probability for the general case of \(T\ge 1\). Given any \(z_0\in \mathcal {S}{\setminus } \mathcal {S}_0\), any integer \(t=\ell T+s\ge 1\) with \(\ell \ge 0\) and \(0\le s\le T-1\), and \(\eta >0\), Proposition 2 implies

$$\begin{aligned} \mathbb {P}_{z_0}\left[ Z_{t}\in \mathcal {S}_\eta \right] \min _{z\in \mathcal {S}_\eta {\setminus }\mathcal {S}_0}V_\theta (z)\le & {} \int _\mathcal {S}V_\theta (z)(\delta _{z_0}P_{\ell T+s})(dz)\\\le & {} \rho ^\ell \int _\mathcal {S}V_\theta (z) (\delta _{z_0}P_s)(dz)+\frac{\beta }{1-\rho }\\\le & {} \rho ^\ell \max _{0\le \tau \le T-1}\int _\mathcal {S}V_\theta (z) (\delta _{z_0}P_\tau )(dz)+\frac{\beta }{1-\rho } \end{aligned}$$

As \(\rho ^\ell \rightarrow 0\) as \(t\rightarrow \infty \), inequality (3.7) implies

$$\begin{aligned} \limsup _{t\rightarrow \infty }\mathbb {P}_{z_0}[Z_t\in \eta ]\le a \eta ^b \text{ where } a=\frac{\beta }{a_0(1-\rho )} \end{aligned}$$

which completes the proof of persistence in probability.

To complete the proof of almost-sure persistence presented in the main text, we need the following lemma which follows the strategy of proof of Schreiber et al. (2011, Lemma 6).

Lemma 4

For all \(z\in \mathcal {S}{\setminus } \mathcal {S}_0\), the weak* limit points \(\mu \) of \(\varPi _t\) almost surely satisfy \(\mu (\mathcal {S}_0)=0\).

Proof

The process \(\{Z_t\}_{t=0}^\infty \) being a (weak) Feller Markov chain over a compact set \(\mathcal {S}\) implies that the set of weak* limit points of \(\{\varPi _t\}_{t=0}^\infty \) is almost surely a non-empty compact subset of probability measures supported by \(\mathcal {S}.\) Almost sure invariance of these weak* limit points follows from Lemma 3 (i).

Assertion (i) of Lemma 3 applied to \(g(z,\xi ) = \sum _i p^i\log f^i(z,\xi )\) gives we have

$$\begin{aligned} \lim _{t \rightarrow \infty } \frac{\sum _i p^i\log X^i_t-\sum _ip^i\log x^i - \sum _{s=0}^{t-1}\sum _ip^i \overline{\log f^i}(Z_s )}{t} = 0 \end{aligned}$$

where \(\overline{\log f^i}(z)=\int \log f^i (z,\xi )m(d\xi ).\) Since \(\limsup _{t\rightarrow \infty }\frac{1}{t} \sum _ip^i\left( \log X^i_t-\log x^i\right) \le 0\) almost surely, we get that

$$\begin{aligned} \sum _ip^ir^i(\mu ) \le 0 \end{aligned}$$
(7.1)

almost surely for any weak* limit point \(\mu \) of \(\{\varPi _t\}_{t=0}^\infty \).

Since \(\mathcal {S}_0\) and \(\mathcal {S}{\setminus }\mathcal {S}_0\) are invariant, there exists \(\alpha \in (0,1]\) such that \(\mu =(1-\alpha )\nu _0+\alpha \nu _1\) where \(\nu _0\) is an invariant probability measure with \(\nu _0( \mathcal {S}_0)=1\) and \(\nu _1\) is an invariant probability measure with \(\nu _1(\mathcal {S}_0)=0\). By Proposition 1, \(\sum _ip^ir^i(\nu _1) = 0\). Thus, \((1-\alpha ) \sum _ip^ir^i(\nu _0) \le 0\). Since by assumption \(\sum _ip^ir^i(\nu _0) > 0\), \(\alpha \) must be 1.

\(\square \)

It follows that with probability one, the weak* limit points \(\mu \) of \(\varPi _t\) (given \(Z_0=z\in \mathcal {S}{\setminus }\mathcal {S}_0\)) are invariant probability measures satisfying \(\mu (\mathcal {S}_0)=0\). To complete, the proof of almost-sure persistence, we need to provide uniform upperbounds to the amount of weight that invariant probability measures \(\mu \) with \(\mu (\mathcal {S}_0)=0\) place near \(\mathcal {S}_0\). To this end, let \(\mu \) be an invariant probability measure with \(\mu (\mathcal {S}_0)=0\). As, in general, we can not assume that \(\int V_\theta (z)\mu (dz)\) is well-defined and finite, we present a slightly longer argument than shown in the main text, to deal with this issue. This argument follows Hairer (2006, Proposition 4.24). Let \(M>0\) be any positive real. For two real numbers ab, let \(a\wedge b\) denote \(\min \{a,b\}.\) Then invariance, Jensen’s inequality and Proposition 2 imply

$$\begin{aligned} \begin{aligned} \int (V_\theta \wedge M)(z) \mu (dz)=&\int (P_{T}(V_\theta \wedge M))(z) \mu (dz) \\ =&\int \mathbb {E}_z[V_\theta (Z_T)\wedge M]\mu (dz) \le \int \mathbb {E}_z[V_\theta (Z_T)]\wedge M \mu (dz)\\ \le&\int (\rho V_\theta (z)+\beta )\wedge M \mu (dz). \end{aligned} \end{aligned}$$

Iterating this inequality \(k\ge 1\) times yields

$$\begin{aligned} \int (V_\theta \wedge M)(z) \mu (dz)\le \int ((\rho )^k V_\theta (z)+\beta /(1-\rho ))\wedge M \mu (dz). \end{aligned}$$
(7.2)

By the dominated convergence theorem, taking the limit \(k\rightarrow \infty \) yields

$$\begin{aligned} \int (V_\theta \wedge M)(z) \mu (dz)\le \frac{\beta }{1-\rho }. \end{aligned}$$
(7.3)

By the dominated convergence theorem, taking the limit \(M\rightarrow \infty \) yields

$$\begin{aligned} \int V_\theta (z) \mu (dz)\le \frac{\beta }{1-\rho }. \end{aligned}$$
(7.4)

Thus, as in the main text, for any \(\eta >0\),

$$\begin{aligned} \mu (\mathcal {S}_\eta )\min _{z\in \mathcal {S}_\eta }V_\theta (z)\le \int V_\theta (z) \mu (dz) \le \frac{\beta }{1-\rho }. \end{aligned}$$

Inequality (3.7) implies that

$$\begin{aligned} \mu (\mathcal {S}_\eta )\le a (\eta )^b \text{ for } \text{ all } \eta \le 1 \text{ where } a=\frac{\beta }{a_0(1-\rho )} \end{aligned}$$

which completes the proof of almost-sure persistence.

1.4 Proof of Theorem 2 and Corollary 1

The strategy of this proof is based on the proof of Theorem 3.3 by Benaïm and Lobry (2016). Define \(V(x,y)=\sum _i p^i \log x^i \). As in the proof of Theorem 1, assumption (4.1) and Proposition 1 imply there exists \(T\ge 1\), \(\eta >0\) and \(\alpha >0\) such that

$$\begin{aligned} P_T V(z)-V(z)\le -\alpha \quad \text{ for } \text{ all } z\in \mathcal {S}_\eta . \end{aligned}$$
(7.5)

Moreover, using the same argument as found in Proposition 2, there exists \(\rho \in (0,1)\) and \(\theta >0\) such that (choosing \(\eta >0\) to be smaller if necessary)

$$\begin{aligned} P_T V_\theta (z)\le \rho V_\theta (z)\quad \text{ for } \text{ all } z\in \mathcal {S}_\eta \text{ where } V_\theta (z)=\exp (\theta V(z)). \end{aligned}$$
(7.6)

Define the stopping time \(\tau =\inf \{k: Z_{kT}\notin \mathcal {S}_\eta \}\) and the event \(\mathcal {A}=\{\limsup _{t\rightarrow \infty } V(Z_t)/t\le -\alpha \}\). We will show that there exists a function \(q(\varepsilon )\in (0,\eta )\) such that \(\mathbb {P}_z(\mathcal {A})\ge q(\varepsilon )\) for \(z\in \mathcal {S}_\varepsilon \) and \(q(\varepsilon )\uparrow 1\) as \(\varepsilon \downarrow 0.\) To this end, define \(W_k=V_\theta (Z_{kT})\). Equation(7.6) implies that \(W_{k\wedge \tau }\) is a super martingale. Hence, if we define \(C(\varepsilon )=\sup _{z\in \mathcal {S}_\varepsilon } \exp (V(z))\) for any \(\varepsilon >0\) (note that \(C(\varepsilon )\downarrow 0\) as \(\varepsilon \downarrow 0\)), then

$$\begin{aligned} \mathbb {E}\left[ W_{k\wedge \tau }\mathbf {1}_{\tau <\infty }\right] \le W_0 \le C(\varepsilon )^\theta \quad \text{ whenever } Z_0=z\in \mathcal {S}_\varepsilon . \end{aligned}$$

Taking the limit as \(k\rightarrow \infty \) and defining \(D=\min _{z\in \mathcal {S}{\setminus } \mathcal {S}_\eta } \exp (V(z))>0\), the dominated convergence theorem implies that

$$\begin{aligned} \mathbb {P}_z[\tau<\infty ]D^\theta \le \mathbb {E}_z[W_\tau \mathbf {1}_{\tau <\infty }]\le C(\varepsilon )^\theta \quad \text{ whenever } z\in \mathcal {S}_\varepsilon . \end{aligned}$$

Thus,

$$\begin{aligned} \mathbb {P}_z[\tau =\infty ]\ge 1-\left( \frac{C(\varepsilon )}{D} \right) ^\theta =:q(\varepsilon )\quad \text{ whenever } z\in \mathcal {S}_\varepsilon \end{aligned}$$

and where \(q(\varepsilon )\uparrow 1\) as \(\varepsilon \downarrow 0.\)

Next, consider the martingale, \(M_n=\sum _{\ell =1}^n V(\ell T)-P_TV((\ell -1)T)\). By the strong law for martingales and inequality (7.5), \(\limsup _{n\rightarrow \infty } V(nT)/n\le -\alpha \) on the event \(\tau =\infty .\) As we have assumed that \(|\log f^i|\) is uniformly bounded on \(\mathcal {S}\times \varXi \), \(\limsup _{t\rightarrow \infty }V(t)/t \le -\alpha \) on the event \(\tau =\infty \). Thus, as claimed, we have shown that

$$\begin{aligned} \mathbb {P}_z[\mathcal {A}]\ge q(\varepsilon )\quad \text{ for } z\in \mathcal {S}_\varepsilon . \end{aligned}$$

As \(V(Z_t)\downarrow -\infty \) implies that \(\mathrm {dist}(Z_t,\mathcal {S}_0)\downarrow 0\), the proof of Theorem 2 is complete.

To prove Corollary 1, define the event

$$\begin{aligned} \mathcal {E}=\left\{ \lim _{t\rightarrow \infty } \mathrm {dist}(Z_t,\mathcal {S}_0)=0\right\} . \end{aligned}$$

Choose \(\varepsilon >0\) such that \(q(\varepsilon )>1/2.\) Define the stopping time

$$\begin{aligned} {{\widetilde{\tau }}} = \inf \{t\ge 0 \ : Z_t \in \mathcal {S}_\epsilon \}. \end{aligned}$$

Since \(\mathcal {S}_0\) is accessible, there exists \(\gamma >0\) such that \(\mathbb {P}_{z}[{{\widetilde{\tau }}}<\infty ]>\gamma \) for all \(z\in \mathcal {S}\). The strong Markov property implies that for all \(z\in \mathcal {S}\)

$$\begin{aligned} \mathbb {P}_z\left[ \mathcal {E}\right]= & {} \mathbb {E}_{z}\left[ \mathbb {P}_{Z_{{\widetilde{\tau }}}} \left[ \mathcal {E} \right] \mathbf {1}_{\{ {{\widetilde{\tau }}}<\infty \}} \right] +\mathbb {E}_{z}\left[ \mathbb {P}_{Z_{{\widetilde{\tau }}}} \left[ \mathcal {E} \right] \mathbf {1}_{\{ {{\widetilde{\tau }}}=\infty \}} \right] \\= & {} \mathbb {E}_{z}\left[ \mathbb {P}_{Z_{{\widetilde{\tau }}}} \left[ \mathcal {E} \right] \mathbf {1}_{\{ {{\widetilde{\tau }}}<\infty \}} \right] \ge \gamma /2. \end{aligned}$$

Let \(\mathcal {F}_t\) be the \(\sigma \)-algebra generated by \(\{Z_1,\dots ,Z_t\}\). The Lévy zero-one law implies that for all \(z\in \mathcal {S}\), \(\lim _{t\rightarrow \infty } \mathbb {E}_{z}\left[ \mathbf {1}_{\mathcal {E}} | \mathcal {F}_t\right] = \mathbf {1}_{\mathcal {E}}\) almost surely. On the other hand, the Markov property implies that \(\mathbb {E}_{z}\left[ \mathbf {1}_{\mathcal {E}} | \mathcal {F}_t\right] =\mathbb {E}_z[\mathbb {P}_{Z_t}[{{\mathcal {E}}}]]\ge \gamma /2\) for all \(z\in \mathcal {S}\). Hence \(\mathbb {P}_{z}[\mathcal {E}]=1\) for all \(z\in \mathcal {S}\).

1.5 Proof of Theorem 3

This proof follows the strategy of Theorem 2, but using a V function introduced by Hening and Nguyen (2018a).

Let \(I\subset \{1,2,\dots ,n\}\) be the subset of species and \(\{p^i\}_{i\in I}\) the set of positive reals such that \(\sum _{i\in I} p^i r^i(\mu )>0\) for every ergodic \(\mu \) supported on \(\mathcal {S}^I_0=\{z=(x,y)\in \mathcal {S}^I: \prod _{i\in I}x^i=0\}\) where \(\mathcal {S}^{I}=\{z=(x,y)\in \mathcal {S}: x^i=0\) for all \(i\notin I\}.\) Assume that \(r^i(\mu )<0\) for all \(i\notin I\) and ergodic \(\mu \) such that \(\mu (\mathcal {S}^I_+)=1\) where \(\mathcal {S}^I_+=\mathcal {S}^I{\setminus } \mathcal {S}^I_0.\) Choose \(\delta >0\) and \(\alpha >0\) such that \(-\sum _{i\in I} p^i r^i(\mu )+\delta \max _{i\notin I} r^i(\mu )\le -2\alpha \) for all ergodic probability measures \(\mu \) such that \(\mu (\mathcal {S}^I)=1.\)

Define \(V(x,y)=-\sum _{i\in I} p^i \log x^i+\delta \max _{i\notin I}\log x^i \) and

$$\begin{aligned} V_\theta (x,y)=e^{\theta V(x,y)}=\left( \prod _{i\in I}(x^i)^{-\theta p^i}\right) \max _{i\notin I}(x^i)^{\theta \delta }. \end{aligned}$$

As in the proof of Theorem 1 and Proposition 1, there exists \(T\ge 1\), \(\eta \in (0,1]\) such that

$$\begin{aligned} P_T V(z)-V(z)\le -\alpha \text{ for } \text{ all } z\in K_\eta \end{aligned}$$
(7.7)

where \(K_\eta =\{(x,y)\in \mathcal {S}: \max _{i\notin I} x^i \le \eta \}.\) Moreover, using the same argument as found in Proposition 2, there exists \(\rho \in (0,1)\) and \(\theta >0\) such that (choosing \(\eta >0\) to be smaller if necessary)

$$\begin{aligned} P_T V_\theta (z)\le \rho V_\theta (z) \text{ for } \text{ all } z\in K_\eta {\setminus }\mathcal {S}^I_0. \end{aligned}$$
(7.8)

Choose \({\widetilde{\eta }}>0\) such that \(\{z: V_\theta (z)\le {\widetilde{\eta }}\}\subset K_\eta .\) Define the stopping time \(\tau =\{k: V_\theta (Z_{kT})\ge {\widetilde{\eta }}\}\) and the event \(\mathcal {A}=\{\limsup _{t\rightarrow \infty } V(Z_t)/t\le -\alpha \}\). Define \(W_k=V_\theta (Z_{kT})\) and \({\widetilde{W}}_k = {\widetilde{\eta }} \wedge W_k\). Equation (7.8) implies that \(W_{k\wedge \tau }\) is a super martingale. Thus, for any k, concavity of \(t\mapsto \delta \wedge t\) and Jensen’s inequality implies that

$$\begin{aligned} \begin{aligned} \mathbb {E}\left[ {\widetilde{W}}_{k\wedge \tau }\mathbf {1}_{\tau<\infty }\right] \le&{\widetilde{\eta }}\wedge \mathbb {E}\left[ W_{k\wedge \tau }\mathbf {1}_{\tau <\infty }\right] \\ \le&{\widetilde{\eta }} \wedge W_0 = {\widetilde{\eta }}\wedge \frac{\max _{i\notin I}(x^i)^{\theta \delta }}{\prod _{i\in I}(x^i)^{\theta p^i}} =:C(z)\\&\text{ whenever } Z_0=z=(x,y)\in K_\eta {\setminus }\mathcal {S}^I_0. \end{aligned} \end{aligned}$$

Taking the limit as \(k\rightarrow \infty \), the dominated convergence theorem implies that

$$\begin{aligned} \mathbb {P}_z[\tau<\infty ]{\widetilde{\eta }} \le \mathbb {E}_z[{\widetilde{W}}_\tau \mathbf {1}_{\tau <\infty }]\le C(z)\quad \text{ for } \text{ all } z\in K_\eta {\setminus }\mathcal {S}^I_0. \end{aligned}$$

Thus,

$$\begin{aligned} \mathbb {P}_z[\tau =\infty ]\ge 1-\frac{C(z)}{{\widetilde{\eta }}}\quad \text{ for } \text{ all } z\in K_\eta {\setminus }\mathcal {S}^I_0. \end{aligned}$$

Next, consider the martingale, \(M_n=\sum _{\ell =1}^n V(\ell T)-P_TV((\ell -1)T)\). By the strong law for martingales and (7.7), \(\limsup _{n\rightarrow \infty } V(nT)/n\le -\alpha \) on the event \(\tau =\infty .\) As we have assumed that \(|\log f^i|\) is uniformly bounded on \(\mathcal {S}\times \varXi \), \(\limsup _{t\rightarrow \infty }V(t)/t \le -\alpha \) on the event \(\tau =\infty \). Thus, we have shown that

$$\begin{aligned} \mathbb {P}_z[\mathcal {A}]\ge 1-\frac{C(z)}{{\widetilde{\eta }}}\quad \text{ for } \text{ all } z\in K_\eta {\setminus }\mathcal {S}^I_0. \end{aligned}$$

Let \(M=\sup _{(x,y)\in \mathcal {S}} \max _i x^i>0.\) As \(\limsup _{t\rightarrow \infty } -\frac{\log X^i_t}{t}\ge \limsup _{t\rightarrow \infty } -\frac{\log M}{t} =0\) with probability one whenever \(X_0^i>0\),

$$\begin{aligned} \begin{aligned} -\alpha \ge \limsup _{t\rightarrow \infty }\frac{V(Z_t)}{t}=&\limsup _{t\rightarrow \infty }\frac{1}{t}\left( -\sum _{i\in I} p^i \log X_t^i + \delta \max _{i\notin I}\log X_t^i \right) \\ \ge&\limsup _{t\rightarrow \infty }\frac{\delta }{t} \max _{i\notin I}\log X_t^i \end{aligned} \end{aligned}$$

almost surely on the event \(\mathcal {A}\) whenever \(Z_0=z\in K_\eta {\setminus } \mathcal {S}_0^I\). Hence, on this event, \(\lim \)\(\mathrm {dist}(Z_t,\mathcal {S}_0^I)\downarrow 0\) and the proof of Theorem 3 is complete.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Benaïm, M., Schreiber, S.J. Persistence and extinction for stochastic ecological models with internal and external variables. J. Math. Biol. 79, 393–431 (2019). https://doi.org/10.1007/s00285-019-01361-4

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-019-01361-4

Keywords

Mathematics Subject Classification

Navigation