Skip to main content

Advertisement

Log in

A general theory of coexistence and extinction for stochastic ecological communities

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

Abstract

We analyze a general theory for coexistence and extinction of ecological communities that are influenced by stochastic temporal environmental fluctuations. The results apply to discrete time (stochastic difference equations), continuous time (stochastic differential equations), compact and non-compact state spaces and degenerate or non-degenerate noise. In addition, we can also include in the dynamics auxiliary variables that model environmental fluctuations, population structure, eco-environmental feedbacks or other internal or external factors. We are able to significantly generalize the recent discrete time results by Benaim and Schreiber (J Math Biol 79:393–431, 2019) to non-compact state spaces, and we provide stronger persistence and extinction results. The continuous time results by Hening and Nguyen (Ann Appl Probab 28(3):1893–1942, 2018a) are strengthened to include degenerate noise and auxiliary variables. Using the general theory, we work out several examples. In discrete time, we classify the dynamics when there are one or two species, and look at the Ricker model, Log-normally distributed offspring models, lottery models, discrete Lotka–Volterra models as well as models of perennial and annual organisms. For the continuous time setting we explore models with a resource variable, stochastic replicator models, and three dimensional Lotka–Volterra models.

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

Abbreviations

\({E_1}\) :

The set of subsets of \(\{1,\dots ,n\}\) such that if\(I\in E_1\) then the invariant measures living on \({\mathcal {S}}_+^I\) are attractors. See Definition 2.1

\(E_2\) :

The set \(E_2:=\mathcal {P}(n)\setminus E_1\)

\(F_i\) :

The fitness functions of species i

\(G\) :

Function describing the dynamics of the auxiliary variable

\(P_t\) :

The t-transition operator or semigroup operator that acts on functions \(H\in \mathcal {B}_b\) as \(P_t(H)({\mathbf {z}}) = {\mathbb {E}}_{\mathbf {z}}(H({\mathbf {Z}}(t))\)

\(P\) :

The transition operator of the discrete-time process \({\mathbf {Z}}(t)\)

\(V\) :

Lyapunov function coming up in Assumption 2.1 and in Assumption 3.1

\(X_i(t)\) :

The density of species i at time t

\({\mathbf {E}}, {\mathbf {W}}, {\mathbf {B}}\) :

Brownian motions on \({\mathbb {R}}^n, {\mathbb {R}}^{\kappa _0}, {\mathbb {R}}^{n+\kappa _0}\)

\({\mathbf {Z}}(t)\) :

The process \(({\mathbf {X}}(t),{\mathbf {Y}}(t))\) at time t

\({{\,\mathrm{Conv}\,}}({\mathcal {M}})\) :

The set of invariant probability measures supported on \({\mathcal {S}}_0\)

\(\Delta \) :

The set \(\Delta :=\{{\mathbf {x}}\in {\mathbb {R}}^n_+~|~\sum _i x_i=1\}\)

\({\mathbb {E}}_{\mathbf {z}}\) :

The expectation under the probability measure \({\mathbb {P}}_{\mathbf {z}}\)

\({\mathbb {E}}\) :

The expectation under the probability measure \({\mathbb {P}}\)

\(\Gamma , \Sigma , (\sigma _{ij})\) :

\(\Gamma \) is a \((n+{\kappa _0})\times (n+{\kappa _0})\) matrix such that \(\Gamma ^\top \Gamma =\Sigma =(\sigma _{ij})_{(n+{\kappa _0})\times (n+{\kappa _0})}\). The matrix \(\Sigma \) encodes the covariance structure of the Brownian motions from the continuous time setting (3.1)

\(\Gamma _{\mathbf {z}}\) :

The set \(\Gamma _{\mathbf {z}}:=\{\tilde{\mathbf {z}}\in {\mathcal {S}}~|~\tilde{\mathbf {z}}~\text {is accessible from}~{\mathbf {z}}\}\) of points \(\tilde{\mathbf {z}}\) such that for every neighborhood U of \(\tilde{\mathbf {z}}\) there is \(t\ge 0\) for which \(P_t({\mathbf {z}}, U)>0\)

\({\mathcal {M}}^1\) :

The set of ergodic invariant probability measures supported on \(\partial {\mathbb {R}}_+^{n}\) and which are attractors: \({\mathcal {M}}^1:=\{\mu \in {\mathcal {M}}: \mu ~~\text {satisfies Assumption}\) 3.4\(\}\). This is in the continuous time setting without an auxiliary variable

\({\mathcal {M}}^2\) :

The set of ergodic invariant probability measures supported on \(\partial {\mathbb {R}}_+^{n}\) and which are not attractors: \({\mathcal {M}}^2:={\mathcal {M}}\setminus {\mathcal {M}}^1\). This is in the continuous time setting without an auxiliary variable

\({\mathcal {M}}^I\) :

\({\mathcal {M}}^I:=\{\mu \in {\mathcal {M}}~|~\mu ({\mathcal {S}}^I)=1\}\) is the set of ergodic measures supported on the subspace \({\mathcal {S}}^I\)

\({\mathcal {M}}^{I,+}\) :

\({\mathcal {M}}^{I,+}:=\{\mu \in {\mathcal {M}}~|~\mu ({\mathcal {S}}_+^I)=1\}\) be the set of ergodic measures supported on the subspace \({\mathcal {S}}^{I}_+\)

\({\mathcal {M}}^{I,\partial }\) :

\({\mathcal {M}}^{I,\partial }:=\{\mu \in {\mathcal {M}}~|~\mu ({\mathcal {S}}_0^I)=1\}\) is the set of ergodic probability measures supported on \({\mathcal {S}}^I_0\)

\({\mathcal {M}}\) :

The set of ergodic invariant probability measures supported on \({\mathcal {S}}_0\)

\({\mathbb {P}}_{\mathbf {z}}\) :

\({\mathbb {P}}_{\mathbf {z}}(\cdot )={\mathbb {P}}(\cdot ~|~{\mathbf {Z}}(0)={\mathbf {z}})\)

\({\mathbb {P}}\) :

The probability measure

\(\Pi _{t,{\mathbf {z}}}(B)\) :

The occupation measure \(\Pi _{t,{\mathbf {z}}}(B):={\mathbb {E}}_{{\mathbf {z}}}\widetilde{\Pi }_t(B)\)

\({\mathbb {R}}_+^\mu \) :

The set \({\mathbb {R}}_+^\mu :=\{(x_1,\dots ,x_n)\in {\mathbb {R}}^n_+: x_i=0\text { if } i\in {\mathcal {S}}(\mu )^c\}\). This is in the continuous time setting without an auxiliary variable

\({\mathbb {R}}_+^{\mu ,\circ }\) :

The set \({\mathbb {R}}_+^{\mu ,\circ }:=\{(x_1,\dots ,x_n)\in {\mathbb {R}}^n_+: x_i=0\text { if } i\in {\mathcal {S}}(\mu )^c\text { and }x_i>0\text { if }x_i\in {\mathcal {S}}(\mu )\}\). This is in the continuous time setting without an auxiliary variable

\({\mathcal {S}}(\mu )\) :

The species supported by the ergodic measure \(\mu \), i.e. \({\mathcal {S}}(\mu )=\{1\le i\le n~:~\mu ({\mathcal {S}}^i)=1\}\)

\({\mathcal {S}}^I\) :

For a nonempty subset \(I\subset \{1,\dots ,n\}\) we define \({\mathcal {S}}^I:=\{({\mathbf {x}},{\mathbf {y}})\in {\mathcal {S}}~|~x_i=0, i\notin I\}\). \({\mathcal {S}}^I\) is the subspace in which all species not in I are absent, and some or all species from I are present. The set \({\mathcal {S}}^I\) represents a subcommunity where we can define persistence and extinction sets relative to that subcommunity

\({\mathcal {S}}^\emptyset \) :

Let \({\mathcal {S}}^\emptyset = \{(0,{\mathbf {y}})\in {\mathcal {S}}\}\) be the set where all species are extinct

\({\mathcal {S}}^i\) :

The set \({\mathcal {S}}^i:=\{{\mathbf {z}}=({\mathbf {x}},{\mathbf {y}})\in {\mathcal {S}}~:~x_i>0\}\) is the subset of the state space where species i persists

\({\mathcal {S}}_+^I\) :

If we restrict the process to \({\mathcal {S}}^I\) then the persistence set, where all species from I persist, is given by \({\mathcal {S}}_+^I:={\mathcal {S}}^I\setminus {\mathcal {S}}_0^I\)

\({\mathcal {S}}_+\) :

The persistence set \({\mathcal {S}}_+:={\mathcal {S}}\setminus {\mathcal {S}}_0\)

\({\mathcal {S}}_0^I\) :

If we restrict the process to \({\mathcal {S}}^I\) then the extinction set, where at least one species from I is extinct, is given by \({\mathcal {S}}_0^I:=\{{\mathbf {z}}\in {\mathcal {S}}^I~|~\prod _{j\in I}x_j=0\}\)

\({\mathcal {S}}_0\) :

The extinction set \({\mathcal {S}}_0:=\{({\mathbf {x}},{\mathbf {y}})\in {\mathcal {S}}~:~\min _i x_i=0\}\), where at leaste one of the n species is extinct

\({\mathcal {S}}_\eta \) :

\({\mathcal {S}}_\eta :=\{({\mathbf {x}},{\mathbf {y}})\in {\mathcal {S}}~:~\min _i x_i\le \eta \}\)

\({\mathcal {S}}\) :

The state space \({\mathbb {R}}_+^n\times {\mathbb {R}}^{\kappa _0}\) of the process \({\mathbf {Z}}\)

\({\mathcal {U}}={\mathcal {U}}(\omega )\) :

The (random) set of weak\(^*\)-limit points of \((\widetilde{\Pi }_t)_{t\in {\mathbb {N}}}\)

\({\mathbf {c}}^\top {\mathbf {x}}\) :

The scalar product \({\mathbf {c}}^\top {\mathbf {x}}= \sum _{i} c_ix_i\)

\(\delta _{\mathbf {z}}\) :

The Dirac mass at \({\mathbf {z}}\)

\(\gamma _1,\gamma _3, C, \rho \) :

Strictly positive constants coming up in Assumption 2.1

\(\gamma _4, \gamma _5, C_4\) :

Strictly positive constants coming up in Assumption 3.1

\(\mathbb {R}_+^n\) :

The positive orthant \([0,\infty )^n\)

\(\mathbb {Z_+}\) :

The set of positive integers

\(\mathbf {X}(t)\) :

The species densities \((X_1(t),\dots ,X_n(t))\) at time t

\(\mathbf {Y}(t)\) :

The value of the auxiliary variable at time t

\(\mathcal {B}_b\) :

The set \(\mathcal {B}_b:=\{H:{\mathcal {S}}\rightarrow {\mathbb {R}}~|~H~\text {Borel, bounded}\}\) of bounded Borel functions

\(\mathcal {C}_b\) :

The set \(\mathcal {C}_b:=\{H:{\mathcal {S}}\rightarrow {\mathbb {R}}~|~H~\text {continuous, bounded}\}\) of bounded continuous functions

\(\mathcal {L}\) :

The infinitesimal generator of the process \({\mathbf {Z}}(t)\) in the continuous time setting

\(\mathcal {P}(n)\) :

The set of all subsets of \(\{1,\dots ,n\}\)

\(\partial {\mathbb {R}}_+^{\mu }\) :

The set \(\partial {\mathbb {R}}_+^{\mu }:={\mathbb {R}}_+^\mu \setminus {\mathbb {R}}_+^{\mu ,\circ }\). This is in the continuous time setting without an auxiliary variable

\(\phi \) :

A function that comes up in Assumption 2.2 which is needed for our extinction results

\(\widetilde{\Pi }_t(B)\) :

The random occupation measure of a Borel set B: in discrete time \(\widetilde{\Pi }_t(B):=\frac{1}{t}\sum _{s=1}^t\delta _{{\mathbf {Z}}(s)}(B)\) and in continuous time \(\tilde{\Pi }_t(B) = \frac{1}{t}\int _0^t {\varvec{1}}_{\{{\mathbf {Z}}(s)\in B\}}\,ds\)

\(\xi (t)\) :

Random variable describing the environment at time t (continuous time case) or on the interval \([t,t+1)\) (discrete time case)

\(a \circ b\) :

\(a \circ b := (a_1b_1, a_2b_2,\dots , a_nb_n)\)

\(f_i\) :

The drift of the dynamics of species i is \(X_if_i({\mathbf {X}})\,dt\) in the continuous time setting

\(g_i\) :

The diffusion term of the dynamics of species i is \(X_ig_i({\mathbf {X}})\,dE^i\) in the continuous time setting

\(h\) :

The function \(h({\mathbf {z}},\xi )=\left( \max _{i=1}^n \left\{ \max \left\{ F_i({\mathbf {z}},\xi ), \frac{1}{F_i({\mathbf {z}},\xi )}\right\} \right\} \right) ^{\gamma _3}\) coming up in Assumption 2.1

\(r_i(\mu )\) :

The expected per-capita growth rate of species i when introduced in the community described by \(\mu \). In the discrete-time case \(r_i(\mu ) =\int _{{\mathcal {S}}}{\mathbb {E}}[\log F_i({\mathbf {z}},\xi (1))]\,\mu (d{\mathbf {z}}),\) while in the continuous case \(r_i(\mu ) = \int _{{\mathcal {S}}}\left( f_i({\mathbf {z}})-\frac{\sigma _{ii}g_i^2({\mathbf {z}})}{2}\right) \mu (d{\mathbf {z}})\)

References

  • Abrams P (1984) Variability in resource consumption rates and the coexistence of competing species. Theor Popul Biol 25:106–124

    MATH  Google Scholar 

  • Abrams P, Holt R (2002) The impact of consumer-resource cycles on the coexistence of competing consumers. Theor Popul Biol 62:281–295

    MATH  Google Scholar 

  • Andrewartha H, Birch L (1954) The distribution and abundance of animals. University of Chicago Press, Chicago

    Google Scholar 

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

    MathSciNet  Google Scholar 

  • Bartlett MS (1960) Stochastic population models in ecology and epidemiology. Methuen, London

    MATH  Google Scholar 

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

  • Benaïm M, Hofbauer J, Sandholm WH (2008) Robust permanence and impermanence for stochastic replicator dynamics. J Biol Dyn 2(2):180–195

    MathSciNet  MATH  Google Scholar 

  • Benaïm M, Schreiber SJ (2019) Persistence and extinction for stochastic ecological models with internal and external variables. J Math Biol 79:393–431

    MathSciNet  MATH  Google Scholar 

  • Cantrell RS, Cosner C, Lou Y (2012) Evolutionary stability of ideal free dispersal strategies in patchy environments. J Math Biol 65(5):943–965

    MathSciNet  MATH  Google Scholar 

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

    Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Chesson P (2000b) Mechanisms of maintenance of species diversity. Annu Rev Ecol Syst 31:343–66

    Google Scholar 

  • Chesson P (2003) Quantifying and testing coexistence mechanisms arising from recruitment fluctuations. Theor Popul Biol 64(3):345–357

    MathSciNet  MATH  Google Scholar 

  • Chesson P (2008) Quantifying and testing species coexistence mechanisms. Fundacion BBVA, Bilbao, pp 119–164

    Google Scholar 

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

    Google Scholar 

  • Chesson P (2019) Chesson’s coexistence theory: comment. Ecology. https://doi.org/10.1002/ecy.2851

    Article  Google Scholar 

  • Chesson P (2020) Species coexistence, book section 2. Oxford University Press, Oxford, pp 5–27

    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 P, Huntly NJ, Roxburgh SH, Pantastico-Caldas M, Facelli JM (2013) The storage effect: definition and tests in two plant communities, book section 2. Cambridge University Press, New York, pp 11–40

    Google Scholar 

  • Chesson P, Kuang JJ (2010) The storage effect due to frequency-dependent predation in multispecies plant communities. Theor Popul Biol 78(2):148–164

    MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Chesson PL (1983) Coexistence of competitors in a stochastic environment: the storage effect. Lect Notes Biomath 52:188–198

    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

    MathSciNet  MATH  Google Scholar 

  • Chesson PL (1989) A general model of the role of environmental variability in communities of competing species. Am Math Soc Lect Math Life Sci 20:97–123

    MathSciNet  MATH  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 PL, Huntly N (1988) Community consequences of life-history traits in a variable environment. Ann Zool Fenn 25:5–16

    Google Scholar 

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

    MathSciNet  Google Scholar 

  • Connell J (1978) Diversity in tropical rain forests and coral reefs. Science 199:1302–1310

    Google Scholar 

  • Diaconis P, Freedman D (1999) Iterated random functions. SIAM review 41(1):45–76

    MathSciNet  MATH  Google Scholar 

  • Durrett R, Remenik D (2012) Evolution of dispersal distance. J Math Biol 64(4):657–666

    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, 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 

  • 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 

  • Fudenberg D, Harris C (1992) Evolutionary dynamics with aggregate shocks. J Econ Theory 57(2):420–441

    MathSciNet  MATH  Google Scholar 

  • Garay BM, Hofbauer J (2003) Robust permanence for ecological differential equations, minimax, and discretizations. SIAM J Math Anal 34(5):1007–1039

    MathSciNet  MATH  Google Scholar 

  • Gonzalez A, Holt RD (2002) The inflationary effects of environmental fluctuations in source–sink systems. Proc Natl Acad Sci 99(23):14872–14877

    Google Scholar 

  • Gyllenberg M, Jiang J, Niu L, Yan P (2019) On the dynamics of multi-species ricker models admitting a carrying simplex. J Differ Equ Appl 25:1489–1530

    MathSciNet  MATH  Google Scholar 

  • Hardin DP, Takáč P, Webb GF (1988a) Asymptotic properties of a continuous-space discrete-time population model in a random environment. J Math Biol 26(4):361–374

    MathSciNet  MATH  Google Scholar 

  • Hardin DP, Takáč P, Webb GF (1988b) A comparison of dispersal strategies for survival of spatially heterogeneous populations. SIAM J Appl Math 48(6):1396–1423

    MathSciNet  MATH  Google Scholar 

  • Hardin DP, Takáč P, Webb GF (1990) Dispersion population models discrete in time and continuous in space. J Math Biol 28(1):1–20

    MathSciNet  MATH  Google Scholar 

  • Hastings A (1983) Can spatial variation alone lead to selection for dispersal? Theor Popul Biol 24(3):244–251

    MATH  Google Scholar 

  • Hening A (2021) Coexistence, extinction, and optimal harvesting in discrete-time stochastic population models. J Nonlinear Sci 31:1–50

    MathSciNet  MATH  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 (2018b) 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 (2018c) Stochastic Lotka–Volterra food chains. J Math Biol 77(1):135–163

    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 

  • Hening A, Nguyen D, Schreiber SJ (2021) A classification of the dynamics of three-dimensional stochastic ecological systems. Ann Appl Probab

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

    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

    MathSciNet  MATH  Google Scholar 

  • Hofbauer J, Imhof LA (2009) Time averages, recurrence and transience in the stochastic replicator dynamics. Ann Appl Probab 19(4):1347–1368

    MathSciNet  MATH  Google Scholar 

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

    Google Scholar 

  • Imhof LA (2005) The long-run behavior of the stochastic replicator dynamics. Ann Appl Probab 15(1B):1019–1045

    MathSciNet  MATH  Google Scholar 

  • Kendall D (1948) On the generalized birth and death process. Ann Math Stat 19:1–15

    MathSciNet  MATH  Google Scholar 

  • Kuang JJ, Chesson P (2008) Predation-competition interactions for seasonally recruiting species. Am Nat 171(3):E119–E133

    Google Scholar 

  • Kuang JJ, Chesson P (2009) Coexistence of annual plants: generalist seed predation weakens the storage effect. Ecology 90(1):170–182

    Google Scholar 

  • Kuang JJ, Chesson P (2010) Interacting coexistence mechanisms in annual plant communities: frequency-dependent predation and the storage effect. Theor Popul Biol 77(1):56–70

    MATH  Google Scholar 

  • Kurtz TG (1970) Solutions of ordinary differential equations as limits of pure jump Markov processes. J Appl Probab 7(1):49–58

    MathSciNet  MATH  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 

  • May RM (1974) Stability and complexity in model ecosystems, 2nd edn. Princeton University Press, Princeton, NJ

    Google Scholar 

  • Meyn SP, Tweedie RL (1992) Stability of Markovian processes. I. Criteria for discrete-time chains. Adv Appl Probab 24(3):542–574

    MathSciNet  MATH  Google Scholar 

  • Ripa J, Ives AR (2003) Food web dynamics in correlated and autocorrelated environments. Theor Popul Biol 64(3):369–384

    MATH  Google Scholar 

  • Roth G, Schreiber SJ (2014) Persistence in fluctuating environments for interacting structured populations. J Math Biol 69(5):1267–1317

    MathSciNet  MATH  Google Scholar 

  • Roy M, Holt RD, Barfield M (2005) Temporal autocorrelation can enhance the persistence and abundance of metapopulations comprised of coupled sinks. Am Nat 166(2):246–261

    Google Scholar 

  • Ryals B, Sacker RJ (2015) Global stability in the 2d Ricker equation. J Differ Equ Appl 21(11):1068–1081

    MathSciNet  MATH  Google Scholar 

  • Ryals B, Sacker RJ (2016) Global stability in the 2d ricker equation revisited. Discrete Contin Dyn Syst B 22(2):585

    MathSciNet  MATH  Google Scholar 

  • Schmidt KA (2004) Site fidelity in temporally correlated environments enhances population persistence. Ecol Lett 7(3):176–184

    Google Scholar 

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

    Google Scholar 

  • Schreiber SJ (2012) Persistence for stochastic difference equations: a mini-review. J Differ Equ Appl 18(8):1381–1403

    MathSciNet  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 

  • Strong DR (1986) Density vagueness: abiding the variance in the demography of real populations. Harper & Row Publishing, New York

    Google Scholar 

  • Turelli M (1978) Does environmental variability limit niche overlap? Proc Natl Acad Sci USA 75:5085–5089

    MATH  Google Scholar 

  • Turelli M (1981) Niche overlap and invasion of competitors in random environments i: models without demographic stochasticity. Theor Popul Biol 20:1–56

    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 

  • 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 

Download references

Acknowledgements

The authors acknowledge generous support from the NSF through the Grants DMS-1853463 for Alexandru Hening, DMS-1853467 for Dang Nguyen, and DEB-1353715 for Peter Chesson.

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. Persistence proofs

Lemma A.1

For any \({\mathbf {z}}\in {\mathcal {S}}, t\in {\mathbb {N}}\) we have the bounds

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}(V({\mathbf {Z}}(t))\le \rho ^t V({\mathbf {z}})+C\sum _{s=0}^t\rho ^s\le \rho ^t V({\mathbf {z}})+\dfrac{C}{1-\rho }, \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}h({\mathbf {Z}}(t), \xi (t)) \le&\, {\mathbb {E}}_{\mathbf {z}}\left[ V({\mathbf {Z}}(t)) h({\mathbf {Z}}(t), \xi (t))\right] \le&\, \rho {\mathbb {E}}_{\mathbf {z}}V({\mathbf {Z}}(t))+C\\ \le&\,\rho ^{t+1} V({\mathbf {z}})+\dfrac{C}{1-\rho }. \end{aligned} \end{aligned}$$

If a function \(\psi \) satisfies

$$\begin{aligned} \lim _{{\mathbf {z}}\rightarrow \infty } \frac{\psi ({\mathbf {z}})}{{\mathbb {E}}V({\mathbf {z}},\xi (t))h({\mathbf {z}},\xi (t))}=0 \end{aligned}$$

then \(\psi \) is \(\mu \)-integrable for any invariant probability measure \(\mu \). Moreover, if \(\mu ({\mathcal {S}}_+)=1\) then \(r_i(\mu )=0\) for any \(i\in I\).

Proof

By Assumption A3) we have \(PV({\mathbf {z}})\le \rho V({\mathbf {z}})+C\). Using this and the Markov property yields \({\mathbb {E}}_{\mathbf {z}}(V({\mathbf {Z}}(t+1))\le \rho {\mathbb {E}}_{\mathbf {z}}V({\mathbf {Z}}(t))+C\). As a result

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}(V({\mathbf {Z}}(t))\le \rho ^t V({\mathbf {z}})+C\sum _{s=0}^t\rho ^s\le \rho ^t V({\mathbf {z}})+\dfrac{C}{1-\rho }. \end{aligned}$$

On the other hand Assumption A3) and the above imply that

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}\left[ V({\mathbf {Z}}(t+1)) h({\mathbf {Z}}(t), \xi (t))\right] \le&\, \rho {\mathbb {E}}_{\mathbf {z}}V({\mathbf {Z}}(t))+C\\ \le&\,\rho ^{t} V({\mathbf {z}})+\dfrac{C}{1-\rho }. \end{aligned} \end{aligned}$$
(A.1)

Since by A3) part i) \(V({\mathbf {z}})\ge 1\) we also get that

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}h({\mathbf {Z}}(t), \xi (t)) \le&\rho ^{t+1} V({\mathbf {z}})+\dfrac{C}{1-\rho }. \end{aligned} \end{aligned}$$

This implies that if \(\mu \) is an ergodic invariant probability measure then

$$\begin{aligned} {\mathbb {E}}_\mu h({\mathbf {Z}}(t), \xi (t)) = \int {\mathbb {E}}h({\mathbf {z}}, \xi (1)) \mu (d{\mathbf {z}})\le \frac{C}{1-\rho }. \end{aligned}$$

Since

$$\begin{aligned} |\log F_i({\mathbf {z}},\xi )|^k\le c_k \max \left\{ F_i({\mathbf {z}},\xi ),\frac{1}{F_i({\mathbf {z}},\xi )}\right\} ^{\gamma _3}\le c_k h({\mathbf {z}},\xi ) \end{aligned}$$

for some \(c_k>0\), we see that for any \(k\in {\mathbb {N}}\) there exists \(C_k>0\) such that

$$\begin{aligned} \int {\mathbb {E}}|\log F_i({\mathbf {z}},\xi (1))|^k\mu (d{\mathbf {z}})\le C_k. \end{aligned}$$

The strong law of large numbers for martingales implies that for \(\mu \) almost every \({\mathbf {z}}\) we have

$$\begin{aligned} \lim _{T\rightarrow \infty } \frac{1}{T} \sum _0^T\left( \log F_i({\mathbf {Z}}(t+1))-P \log F_i({\mathbf {Z}}(t))\right) =0, \,\text {when}~ {\mathbf {Z}}(0)={\mathbf {z}}. \end{aligned}$$
(A.2)

Having (A.2), we can follow the arguments by Benaïm and Schreiber (2019, Lemma 3 and Proposition 1) to obtain that if \(\mu ({\mathcal {S}}^{I}_+)=1\) then \(r_i(\mu )=0\) for any \(i\in I\).

Finally, because of the boundedness (A.1), it is standard to show that if a function \(\psi \) satisfies

$$\begin{aligned} \lim _{{\mathbf {z}}\rightarrow \infty } \frac{\psi ({\mathbf {z}})}{{\mathbb {E}}V({\mathbf {z}},\xi (t))h({\mathbf {z}},\xi (t))}=0 \end{aligned}$$

then \(\psi \) is \(\mu \)-integrable for any invariant probability measure \(\mu \). See Lemma 3.3 in Hening and Nguyen (2018a) for a similar proof. \(\square \)

Lemma A.2

There exist \(M, C_2, \gamma _4>0,\rho _2\in (0,1)\) such that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}\left[ V({\mathbf {Z}}(1))\prod _{i=1}^n X_i^{p_i}(1)\right] \le \left( {\varvec{1}}_{\{|{\mathbf {z}}|<M\}}(C_2-\rho _2)+\rho _2\right) V({\mathbf {z}})\prod _{i=1}^nx_i^{p_i}, ~{\mathbf {z}}\in {\mathcal {S}}\end{aligned}$$

for any \({\mathbf {p}}=(p_1,\ldots ,p_n)\in {\mathbb {R}}^n\) satisfying

$$\begin{aligned} |{\mathbf {p}}|_1:=\sum |p_i|\le \gamma _4. \end{aligned}$$
(A.3)

Proof

We note that for \(w_1,\ldots ,w_n>0\)

$$\begin{aligned} \begin{aligned} \prod _{i=1}^n w_i^{p_i}\le&\, \prod _{i=1}^n \left( w_i^{-1}\vee {w_i}\right) ^{|p_i|}\\ \le&\, \left( \max _{i=1,\ldots ,n}\left\{ w_i^{-1}\vee {w_i}\right\} \right) ^{\sum _i|p_i|}\\ \le&\, 1+\frac{\sum _i|p_i|}{\gamma _3}\left( \max _{i=1,\ldots ,n}\left\{ w_i^{-1}\vee {w_i}\right\} \right) ^{\gamma _3} \end{aligned} \end{aligned}$$

if \(\sum _i |p_i|<\gamma _3.\) The last inequality follows from the inequality \(x^p\le 1+px\) for \(x\ge 0, p\in (0,1)\). Thus,

$$\begin{aligned} \prod _{i=1}^n F_i^{p_i}({\mathbf {z}},\xi ) \le 1+ \frac{|{\mathbf {p}}|_1}{\gamma _3} h({\mathbf {z}},\xi ). \end{aligned}$$

Since \(\lim _{{\mathbf {z}}\rightarrow \infty } V({\mathbf {z}})=\infty \), we can select \(\rho _1\in (\rho ,1)\), \(M>0\) and \(C_1>0\) such that

$$\begin{aligned} \rho V({\mathbf {z}})+C\le (C_1{\varvec{1}}_{\{|z|<M\}}+\rho _1)V({\mathbf {z}}). \end{aligned}$$

Let \(\gamma _4\in (0,\gamma _3)\) be such that \(\left( 1+\frac{\gamma _4}{\gamma _3}\right) \rho _1=:\rho _2<1.\) There exists \(C_2>0\) satisfying

$$\begin{aligned} (C_1{\varvec{1}}_{\{|z|<M\}}+\rho _1)\left( 1+\frac{\gamma _4}{\gamma _3}\right) \le (C_2-\rho _2){\varvec{1}}_{\{|z|<M\}}+\rho _2. \end{aligned}$$

The above estimates together with (2.1), Lemma A.1, and (A.3) yield

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}\left[ V\left( {\mathbf {Z}}(1)\right) \prod _{i=1}^n X_i^{p_i}(1)\right] =&\,\prod _{i=1}^nx_i^{p_i} {\mathbb {E}}\left[ V({\mathbf {x}}\circ F({\mathbf {z}},\xi ), G({\mathbf {z}},\xi ))\prod _{i=1}^n F_i^{p_i}({\mathbf {z}},\xi )\right] \\ \le&\,\prod _{i=1}^nx_i^{p_i} {\mathbb {E}}\left[ V({\mathbf {x}}\circ F({\mathbf {z}},\xi ), G({\mathbf {z}},\xi ))\left( 1+\frac{|{\mathbf {p}}|_1}{\gamma _3}h({\mathbf {z}},\xi )\right) \right] \\ \le&\, \prod _{i=1}^nx_i^{p_i}(\rho V({\mathbf {z}})+C)\left( 1+\frac{|{\mathbf {p}}|_1}{\gamma _3}\right) \\ \le&\, \prod _{i=1}^nx_i^{p_i}(C_1{\varvec{1}}_{\{|z|<M\}}+\rho _1) V({\mathbf {z}})\left( 1+\frac{|{\mathbf {p}}|_1}{\gamma _3}\right) \\ \le&\,((C_2-\rho ){\varvec{1}}_{\{|z|<M\}}+\rho _2) \prod _{i=1}^nx_i^{p_i}V({\mathbf {z}}). \end{aligned}$$

\(\square \)

We will denote by \(B_M :=\{{\mathbf {u}}\in {\mathcal {S}}:\Vert {\mathbf {u}}\Vert \le M\}\) the closed ball of radius \(M>0\) around the origin. Let \({\mathcal {M}}\) be the set of ergodic invariant probability measures of \({\mathbf {X}}\) supported on the boundary \({\mathcal {S}}_0:=\partial {\mathbb {R}}^n_+\times {\mathbb {R}}^{\kappa _0}\). Remember that for a subset \(\widetilde{\mathcal {M}}\subset {\mathcal {M}}\) we denote by \({{\,\mathrm{Conv}\,}}(\widetilde{\mathcal {M}})\) the convex hull of \(\widetilde{\mathcal {M}}\), that is the set of probability measures \(\pi \) of the form \(\pi (\cdot )=\sum _{\mu \in \widetilde{\mathcal {M}}}p_\mu \mu (\cdot )\) with \(p_\mu \ge 0,\sum _{\mu \in \widetilde{\mathcal {M}}}p_\mu =1\).

Consider \(\mu \in {\mathcal {M}}\). Assume \(\mu (\{0\}\times {\mathbb {R}}^{\kappa _0})=0\). Since \(\mu \) is ergodic there exist \(0<n_1<\cdots < n_k\le n\) such that \({{\,\mathrm{supp}\,}}(\mu )\subset S^\mu :={\mathbb {R}}^\mu _+\times {\mathbb {R}}^{\kappa _0}\) where

$$\begin{aligned} {\mathbb {R}}_+^\mu :=\{(x_1,\ldots ,x_n)\in {\mathbb {R}}^n_+: x_i=0\text { if } i\in I_\mu ^c\} \end{aligned}$$

for \(I_\mu :={\mathcal {S}}(\mu )=\{n_1,\ldots , n_k\}\) and \(I_\mu ^c:=\{1,\ldots ,n\}{\setminus }\{n_1,\ldots , n_k\}\).

$$\begin{aligned} {\mathbb {R}}_+^{\mu ,\circ }:=\{(x_1,\ldots ,x_n)\in {\mathbb {R}}^n_+: x_i=0\text { if } i\in I_\mu ^c\text { and }x_i>0\text { if }x_i\in I_\mu \} \end{aligned}$$

and \(\partial {\mathbb {R}}_+^{\mu }:={\mathbb {R}}_+^\mu {\setminus }{\mathbb {R}}_+^{\mu ,\circ }.\)

The following condition ensures persistence.

Assumption A.1

For any \(\mu \in {{\,\mathrm{Conv}\,}}({\mathcal {M}})\) one has

$$\begin{aligned} \max _i r_i(\mu )>0, \end{aligned}$$

where

$$\begin{aligned} r_i(\mu ):=\int _{{\mathcal {S}}_0}\left[ {\mathbb {E}}\ln F_i({\mathbf {z}},\xi )\right] \mu (d{\mathbf {z}}). \end{aligned}$$

Remember that the occupation measures are defined as

$$\begin{aligned} \Pi _{t,{\mathbf {z}}}(\cdot )=\frac{1}{t}\sum _{s=0}^t{\mathbb {P}}_{\mathbf {z}}({\mathbf {Z}}(s)\in \cdot )\,ds,~{\mathbf {z}}\in {\mathcal {S}}, t\in {\mathbb {Z}}_+ \end{aligned}$$

Lemma A.3

Suppose the following

  • The sequences \(({\mathbf {z}}_k)_{k\in N}\subset {\mathbb {R}}_+^n\times {\mathbb {R}}^{\kappa _0}, (T_k)_{k\in {\mathbb {N}}}\subset {\mathbb {N}}\) are such that \(\Vert {\mathbf {z}}_k\Vert \le M\), \(T_k>1\) for all \(k\in {\mathbb {N}}\) and \(\lim _{k\rightarrow \infty }T_k=\infty \).

  • The sequence \((\Pi _{T_k,{\mathbf {z}}_k})_{k\in {\mathbb {N}}}\) converges weakly to an invariant probability measure \(\pi \).

  • The function \(h:{\mathbb {R}}^n_+\times {\mathbb {R}}^{\kappa _0}\rightarrow {\mathbb {R}}\) is any continuous function satisfying

    $$\begin{aligned} \lim _{{\mathbf {z}}\rightarrow \infty }\frac{|h({\mathbf {z}})|}{V({\mathbf {z}})\max _{i=1}^n\{(x_i^{\gamma _3}\wedge x^{-\gamma _3}_i)\}} =0 \end{aligned}$$

Then one has

$$\begin{aligned} \lim _{k\rightarrow \infty }\int _{{\mathbb {R}}^n_+}h({\mathbf {x}})\Pi _{T_k,{\mathbf {z}}_k}(d{\mathbf {x}})= \int _{{\mathbb {R}}^n_+}h({\mathbf {x}})\pi (d{\mathbf {x}}). \end{aligned}$$

Proof

The proof is almost identical to that of Lemma 3.4 by Hening and Nguyen (2018a) and is therefore omitted. \(\square \)

It is shown in Schreiber (2011, Lemma 4) by the min–max principle that Assumption A.1 is equivalent to the existence of \({\mathbf {p}}>0\) such that

$$\begin{aligned} \min \limits _{\mu \in {\mathcal {M}}}\left\{ \sum _{i}p_i r_i(\mu )\right\} :=2r^*>0. \end{aligned}$$
(A.4)

By rescaling if necessary, we can assume that \(|{\mathbf {p}}|_1=\gamma _4\).

Lemma A.4

Suppose that Assumption A.1 holds. Let \({\mathbf {p}}\) and \(r^*\) be as in (A.3). There exists an integer \(T^*>0\) such that, for any \(T>T^*\), \({\mathbf {x}}\in \partial {\mathbb {R}}^n_+, {\mathbf {z}}= ({\mathbf {x}},{\mathbf {y}}) \in B_M\) one has

$$\begin{aligned} \sum _{t=0}^T{\mathbb {E}}_{\mathbf {z}}\left( \ln V({\mathbf {Z}}(t+1))-\ln V({\mathbf {Z}}(t))-\sum p_i\ln F_i({\mathbf {Z}}(t), \xi (t))\right) \le -r^*(T+1). \end{aligned}$$
(A.5)

Proof

In view of Lemma A.1,

$$\begin{aligned} \sup _{t\in {\mathbb {N}}, \Vert {\mathbf {z}}\Vert \le M} {\mathbb {E}}_{\mathbf {z}}V({\mathbf {Z}}(t))<\infty \end{aligned}$$

which implies, since

$$\begin{aligned} \frac{{\mathbb {E}}_{\mathbf {z}}\ln V({\mathbf {Z}}(T))}{T+1}\le \frac{{\mathbb {E}}_{\mathbf {z}}V({\mathbf {Z}}(T))}{T+1} \end{aligned}$$

that

$$\begin{aligned} \lim _{T\rightarrow \infty } \sup _{\Vert {\mathbf {z}}\Vert \le M} \frac{1}{T+1}\sum _{t=0}^T{\mathbb {E}}_{\mathbf {z}}\left( \ln V({\mathbf {Z}}(t+1))-\ln V({\mathbf {Z}}(t))\right) =0. \end{aligned}$$

With (A.4) and Lemma A.2 and Lemma A.3, we can argue by contradiction in the same manner as in Lemma 4.1 by Hening and Nguyen (2018a) to show that

$$\begin{aligned} \limsup _{T\rightarrow \infty } \frac{1}{T+1}\sup _{\Vert {\mathbf {z}}\Vert \le M} \sum _{t=0}^T{\mathbb {E}}_{\mathbf {z}}\left( -\sum p_i\ln F_i({\mathbf {Z}}(t), \xi (t))\right) \le -2r^*. \end{aligned}$$

Combining the two above limits finishes the proof. \(\square \)

Let \(n^*\in {\mathbb {N}}\) be such that

$$\begin{aligned} \rho _2^{1-n^*}>C_2. \end{aligned}$$
(A.6)

Proposition A.1

Define \(U:{\mathcal {S}}_+\rightarrow {\mathbb {R}}_+\) by

$$\begin{aligned} U({\mathbf {z}})=V({\mathbf {z}})\prod _{i=1}^nx_{i}^{-p_i} \end{aligned}$$

with \({\mathbf {p}}\) and \(r^*\) satisfying (A.3) and \(T^*>0\) satisfying the assumptions of Lemma A.4. There exist numbers \(\theta \in \left( 0,\frac{\gamma _4}{2}\right) \), \(K_\theta >0\), such that for any \(T\in [T^*,n^*T^*]\cap {\mathbb {Z}}\) and \({\mathbf {z}}\in {\mathcal {S}}_+, \Vert {\mathbf {z}}\Vert \le M\),

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}U^\theta ({\mathbf {Z}}(T))\le U^\theta ({\mathbf {x}})\exp \left( -\frac{1}{2}\theta r^*T\right) +K_\theta . \end{aligned}$$

Proof

$$\begin{aligned} \begin{aligned} \ln U({\mathbf {Z}}(T))=&\ln U({\mathbf {Z}}(0)) + \sum _{t=0}^{T-1} \left( \ln U({\mathbf {Z}}(t+1))-\ln U({\mathbf {Z}}(t))\right) \\ =&\ln U({\mathbf {Z}}(0)) + G(T) \end{aligned} \end{aligned}$$
(A.7)

where

$$\begin{aligned} \begin{aligned} G(T)=\sum _{t=0}^{T-1} \left( \ln V({\mathbf {Z}}(t+1))-\ln V({\mathbf {Z}}(t))-\sum p_i\ln F_i({\mathbf {Z}}(t), \xi (t))\right) . \end{aligned} \end{aligned}$$
(A.8)

In view of (A.7) and Lemma A.2

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}\exp ( G(T))=\dfrac{{\mathbb {E}}_{\mathbf {z}}U({\mathbf {Z}}(T))}{U({\mathbf {z}})}\le (C_2)^T. \end{aligned}$$
(A.9)

Let \({\widehat{U}}(\cdot ):{\mathbb {R}}^{n,\circ }_+\times {\mathbb {R}}^{\kappa _0}\mapsto {\mathbb {R}}_+\) be defined by \({\widehat{U}}({\mathbf {z}})=V({\mathbf {z}})\prod _{i=1}^n x_i^{p_i}\). We also have

$$\begin{aligned} \dfrac{{\mathbb {E}}_{\mathbf {z}}{\widehat{U}}({\mathbf {Z}}(T))}{{\widehat{U}}({\mathbf {z}})}\le (C_2)^T. \end{aligned}$$
(A.10)

Note that

$$\begin{aligned} U^{-1}({\mathbf {z}})={\widehat{U}}({\mathbf {z}})\frac{1}{V^2({\mathbf {z}})}\le {\widehat{U}}({\mathbf {z}}). \end{aligned}$$
(A.11)

Using (A.11) and (A.10) yields

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}\exp (-G(T))=&\,\dfrac{{\mathbb {E}}_{\mathbf {z}}U^{-1}({\mathbf {Z}}(T))}{U^{-1}({\mathbf {z}})}\\ \le&\,\dfrac{{\mathbb {E}}_{\mathbf {z}}{\widehat{U}}({\mathbf {Z}}(T))}{V^2({\mathbf {z}})U^{-1}({\mathbf {x}})}\\ \le&\, \dfrac{{\mathbb {E}}_{\mathbf {z}}{\widehat{U}}({\mathbf {Z}}(T))}{{\widehat{U}}({\mathbf {z}})}\\ \le&\, (C_2)^T. \end{aligned} \end{aligned}$$
(A.12)

By (A.9) and (A.12) the assumptions of Hening and Nguyen (2018a, Lemma 3.5) hold for the random variable G(T). Therefore, there exists \({\tilde{K}}_2\ge 0\) such that

$$\begin{aligned} 0\le \dfrac{d^2{\tilde{\phi }}_{{\mathbf {z}},T}}{d\theta ^2}(\theta )\le {\tilde{K}}_2\,\text { for all }\,\theta \in \left[ 0,\frac{1}{2}\right) ,\, {\mathbf {z}}\in {\mathbb {R}}^{n,\circ }_+\times {\mathbb {R}}^{\kappa _0}, \Vert {\mathbf {z}}\Vert \le M, T\in [T^*,n^*T^*]\cap {\mathbb {Z}}\end{aligned}$$

where

$$\begin{aligned} {\tilde{\phi }}_{{\mathbf {z}},T}(\theta )=\ln {\mathbb {E}}_{\mathbf {z}}\exp (\theta G(T)). \end{aligned}$$

In view of Lemma A.4 and the Feller property of \(({\mathbf {Z}}(t))\), there exists a \({\tilde{\delta }}>0\) such that if \(\Vert {\mathbf {z}}\Vert \le M\), \(\mathrm {dist}({\mathbf {z}},\partial {\mathbb {R}}^n_+)<{\tilde{\delta }}\) and \(T\in [T^*,n^*T^*]\cap {\mathbb {Z}}\) then

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}G(T)\le -\dfrac{3}{4}r^*T. \end{aligned} \end{aligned}$$
(A.13)

Another application of Hening and Nguyen (2018a, Lemma 3.5) yields

$$\begin{aligned} \dfrac{d{\tilde{\phi }}_{{\mathbf {z}},T}}{d\theta }(0)= & {} {\mathbb {E}}_{\mathbf {z}}G(T)\le -\dfrac{3}{4} r^*T\,\text { for }\, {\mathbf {z}}\in {\mathbb {R}}^{n,\circ }_+\times {\mathbb {R}}^{\kappa _0}, \Vert {\mathbf {z}}\Vert \le M, \mathrm {dist}({\mathbf {z}},\partial {\mathbb {R}}^n_+)\\< & {} {\tilde{\delta }}, T\in [T^*,n^*T^*]\cap {\mathbb {Z}}. \end{aligned}$$

By a Taylor expansion around \(\theta =0\), for \(\Vert {\mathbf {z}}\Vert \le M, \mathrm {dist}({\mathbf {z}},\partial {\mathbb {R}}^n_+)<{\tilde{\delta }}, T\in [T^*,n^*T^*]\cap {\mathbb {Z}}\) and \(\theta \in \left[ 0,\frac{1}{2}\right) \) we have

$$\begin{aligned} {\tilde{\phi }}_{{\mathbf {x}},T}(\theta )\le -\dfrac{3}{4} r^*T\theta +\theta ^2{\tilde{K}}_2. \end{aligned}$$

If we choose any \(\theta \in \left( 0,\frac{1}{2}\right) \) satisfying \(\theta <\frac{r^*T^*}{4{\tilde{K}}_2}\), we obtain that

$$\begin{aligned} {\tilde{\phi }}_{{\mathbf {z}},T}(\theta )\le -\dfrac{1}{2} r^*T\theta \,\,\text { for all }\,{\mathbf {z}}\in {\mathbb {R}}^{n,\circ }\times {\mathbb {R}}^{\kappa _0},\Vert {\mathbf {z}}\Vert \le M, \mathrm {dist}({\mathbf {z}},\partial {\mathbb {R}}^n_+)<{\tilde{\delta }}, T\in [T^*,n^*T^*]\cap {\mathbb {Z}}.\nonumber \\ \end{aligned}$$
(A.14)

In light of (A.14), we have for all \(\theta <\frac{r^*T^*}{4{\tilde{K}}_2}\), \(\Vert {\mathbf {z}}\Vert \le M, 0<\mathrm {dist}({\mathbf {z}},\partial {\mathbb {R}}^n_+)<{\tilde{\delta }}, T\in [T^*,n^*T^*]\) that

$$\begin{aligned} \dfrac{{\mathbb {E}}_{\mathbf {z}}U^\theta ({\mathbf {Z}}(T))}{U^\theta ({\mathbf {z}})}=\exp {\tilde{\phi }}_{{\mathbf {z}},T}(\theta )\le \exp \left( -\frac{1}{2}r^*T\theta \right) . \end{aligned}$$
(A.15)

In view of Lemma A.2, we have for \({\mathbf {z}}\) satisfying \(\Vert {\mathbf {z}}\Vert \le M, \mathrm {dist}({\mathbf {z}},\partial {\mathbb {R}}^n_+)\ge {\tilde{\delta }}\) and \(T\in [T^*,n^*T^*]\) that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}U^\theta ({\mathbf {Z}}(T))\le (C_2)^{\theta n^*T^*}\sup \limits _{\Vert {\mathbf {z}}\Vert \le M, \mathrm {dist}({\mathbf {z}},\partial {\mathbb {R}}^n_+)\ge {\tilde{\delta }}}\{U^\theta ({\mathbf {z}})\}=:K_\theta <\infty . \end{aligned}$$
(A.16)

Combining (A.15) and (A.16) we are done. \(\square \)

Theorem A.1

Suppose that Assumption A.1 holds. Let \(\theta \) be as in Proposition A.1, \(T^*\) as in Lemma A.4 and \(n^*\) as in (A.6). There exist numbers \(\kappa =\kappa (\theta ,T^*)\in (0,1)\) and \({\tilde{K}}={\tilde{K}}(\theta ,T^*)>0\) such that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}U^\theta ({\mathbf {Z}}(n^*T^*))\le \kappa U^\theta ({\mathbf {z}})+\tilde{K}\,\text { for all }\, {\mathbf {z}}\in {\mathbb {R}}^{n,\circ }_+\times {\mathbb {R}}^{\kappa _0}. \end{aligned}$$
(A.17)

We have

$$\begin{aligned} \limsup _{t\rightarrow \infty }{\mathbb {P}}_{\mathbf {z}}\{|X_i(t)|\vee |X^{-1}_i(t)|>m \text { for some } i=1,\ldots , n\}\le c_2 m^{-c_3} \end{aligned}$$
(A.18)

for some positive \(c_2, c_3>0.\) Moreover, for any compact set \(K\subset {\mathbb {R}}^{n,\circ }\times {\mathbb {R}}^{\kappa _0}\),

$$\begin{aligned} {\mathbb {P}}_{\mathbf {z}}(\tau _K>k)\le c_KU^\theta ({\mathbf {z}})\kappa ^k \end{aligned}$$
(A.19)

If the Markov chain \({\mathbf {Z}}(t)\) is irreducible and aperiodic on \({\mathbb {R}}^{n,\circ }_+\times {\mathbb {R}}^{\kappa _0}\), and a compact set is petite, then there is \(c_4>1\) such that

$$\begin{aligned} c_4^t\Vert P_t({\mathbf {z}},\ldots )-\pi \Vert _{TV}\rightarrow 0\text { as } t\rightarrow \infty . \end{aligned}$$

Proof

Define

$$\begin{aligned} \tau =\inf \{t\ge 0: \Vert {\mathbf {Z}}(t)\Vert \le M\}. \end{aligned}$$
(A.20)

By Lemma A.2 for all \({\mathbf {z}}\in {\mathcal {S}}\)

$$\begin{aligned} PU({\mathbf {z}})\le \rho _2 U({\mathbf {z}}), ~\Vert {\mathbf {z}}\Vert \ge M. \end{aligned}$$

This implies that the process \(\rho ^{-t}_2 U({\mathbf {Z}}(t))\) is a supermartingale and therefore

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}&\left[ \rho _2^{-\theta (\tau \wedge n^*T^*)}U^\theta ({\mathbf {Z}}(\tau \wedge n^*T^*))\right] \le U^\theta ({\mathbf {z}}), {\mathbf {z}}\in {\mathcal {S}}. \end{aligned} \end{aligned}$$

Thus,

$$\begin{aligned} \begin{aligned} U^\theta ({\mathbf {z}})\ge&{\mathbb {E}}_{\mathbf {z}}\left[ \rho _2^{-\theta (\tau \wedge n^*T^*)}U^\theta ({\mathbf {Z}}(\tau \wedge n^*T^*))\right] \\ =&{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \le (n^*-1)T^*\}}\rho _2^{-\theta (\tau \wedge n^*T^*)}U^\theta ({\mathbf {Z}}(\tau \wedge n^*T^*))\right] \\&+{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{ (n^*-1)T^*<\tau<n^*T^*\}}\rho _2^{-\theta (\tau \wedge n^*T^*)}U^\theta ({\mathbf {Z}}(\tau \wedge n^*T^*))\right] \\&+ {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \ge n^*T^*\}}\rho _2^{-\theta (\tau \wedge n^*T^*)}U^\theta ({\mathbf {Z}}(\tau \wedge n^*T^*))\right] \\ \ge&{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \le (n^*-1)T^*\}}U^\theta ({\mathbf {Z}}(\tau ))\right] \\&+\rho _2^{-\theta (n^*-1)T^*}{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{ (n^*-1)T^*<\tau <n^*T^*\}}U^\theta ({\mathbf {Z}}(\tau ))\right] ~~\\&+\rho _2^{-\theta n^*T^*}{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \ge n^*T^*\}}U^\theta ({\mathbf {Z}}(n^*T^*))\right] .\\ \end{aligned} \end{aligned}$$
(A.21)

By the strong Markov property of \({\mathbf {Z}}(t)\) and Proposition A.1, we obtain

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}&\left[ {\varvec{1}}_{\{\tau \le (n^*-1)T^*\}}U^\theta ({\mathbf {Z}}(n^*T^*))\right] \\&\le {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \le (n^*-1)T^*\}}\big [K_\theta +e^{-\frac{1}{2}\theta r^*(n^*T^*-\tau )}U^\theta ({\mathbf {Z}}(\tau ))\big ]\right] \\&\le K_\theta + \exp \left( -\frac{1}{2}\theta r^*T^*\right) {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \le (n^*-1)T^*\}}U^\theta ({\mathbf {Z}}(\tau ))\right] \end{aligned} \end{aligned}$$
(A.22)

Similarly, the strong Markov property of \({\mathbf {Z}}(t)\), Jensen’s inequality and Lemma A.2 imply

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}&\left[ {\varvec{1}}_{\{(n^*-1)T^*<\tau<n^*T^*\}}U^\theta ({\mathbf {Z}}(n^*T^*))\right] \\&\le {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{(n^*-1)T^*<\tau<n^*T^*\}}C_2^{\theta (n^*T^*-\tau )}U^\theta ({\mathbf {Z}}(\tau ))\right] \\&\le C_2^{\theta T^*}{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{(n^*-1)T^*<\tau <n^*T^*\}}U^\theta ({\mathbf {Z}}(\tau ))\right] . \end{aligned} \end{aligned}$$
(A.23)

Applying (A.22) and (A.23) to (A.21) yields

$$\begin{aligned} \begin{aligned} U^\theta (x) \ge&{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \le (n^*-1)T^*\}}U^\theta ({\mathbf {Z}}(\tau ))\right] \\&+\rho _2^{-\theta (n^*-1)T^*}{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{ (n^*-1)T^*<\tau<n^*T^*\}}U^\theta ({\mathbf {Z}}(\tau ))\right] \\&+\rho _2^{-\theta n^*T^*} {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \ge n^*T^*\}}U^\theta ({\mathbf {Z}}(n^*T^*))\right] \\ \ge&\exp \left( \frac{1}{2}\theta r^*T^*\right) {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \le (n^*-1)T^*\}}U^\theta ({\mathbf {Z}}(n^*T^*))\right] -\exp \left( \frac{1}{2}\theta r^*T^*\right) K_\theta \\&+C_2^{-\theta T^*}\rho _2^{-\theta (n^*-1)T^*}{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{ (n^*-1)T^*<\tau <n^*T^*\}}U^\theta ({\mathbf {Z}}(n^*T^*))\right] \\&+\rho _2^{-\theta n^*T^*} {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \ge n^*T^*\}}U^\theta ({\mathbf {Z}}(n^*T^*))\right] \\ \ge&\kappa ^{-1}{\mathbb {E}}_{\mathbf {z}}U^\theta ({\mathbf {Z}}(n^*T^*))-K_\theta \exp \left( \frac{1}{2}\theta \rho ^*T^*\right) \end{aligned} \end{aligned}$$
(A.24)

where \(\kappa =\max \left\{ \exp \left( -\frac{1}{2}\theta r^*T^*\right) , C^{\theta T^*}\rho _2^{\theta (n^*-1)T^*}, \rho _2^{\theta n^*T^*}\right\} <1\) by (A.6). The proof of (A.17) is complete by taking

$$\begin{aligned} {\tilde{K}}=K_\theta \exp \left( \frac{1}{2}\theta \rho ^*T^*\right) \kappa . \end{aligned}$$

Having (A.17), the claims (A.19) and (A.18) follow by Benaïm and Schreiber (2019, Proposition 3.3 and Theorem 3.1). \(\square \)

Appendix B. Extinction Proofs

For \(I\subset \{1,\ldots ,n\}\), denote by \({\mathcal {M}}^I, {\mathcal {M}}^{I,+}, {\mathcal {M}}^{I,\partial }\) the set of ergodic probability measures on \({\mathcal {S}}^I, {\mathcal {S}}^{I}_+, {\mathcal {S}}_0^I\) respectively.

Lemma B.1

Assume that there exists a function \(\phi :{\mathcal {S}}\rightarrow {\mathbb {R}}_+\) and constants \(C, \delta _\phi >0\) such that for all \({\mathbf {z}}\in {\mathcal {S}}\)

$$\begin{aligned} P V({\mathbf {z}})\le V({\mathbf {z}})-\phi ({\mathbf {z}})+C \end{aligned}$$
(B.1)

and

$$\begin{aligned} {\mathbb {E}}_{{\mathbf {z}}}\left( V({\mathbf {Z}}(1))-PV({\mathbf {z}})\right) ^2+{\mathbb {E}}\left| \log F({\mathbf {z}},\xi (1))-{\mathbb {E}}\log F({\mathbf {z}},\xi (1))\right| ^2\le \delta _\phi \phi ({\mathbf {z}}).\nonumber \\ \end{aligned}$$
(B.2)

Then, the family of random occupation measures \(({\widetilde{\Pi }}_t)_{t\in {\mathbb {N}}}\) is tight. and with probability one

$$\begin{aligned} \lim _{T\rightarrow \infty }\frac{1}{T}\sum _{t=0}^T\left( \log F({\mathbf {Z}}(t),\xi (t))-{\mathbb {E}}\left[ \log F({\mathbf {Z}}(t),\xi (t))\big |{\mathcal {F}}_t\right] \right) =0. \end{aligned}$$
(B.3)

where \(({\mathcal {F}}_t)_{t\in {\mathbb {N}}}\) is the filtration generated by the process \({\mathbf {Z}}\).

Proof

Suppose \({\mathbf {Z}}(0)={\mathbf {z}}\in {\mathcal {S}}\). We have

$$\begin{aligned} \begin{aligned} V({\mathbf {Z}}(t+1))\le&V({\mathbf {Z}}(t))-\phi ({\mathbf {Z}}(t))+C +(V({\mathbf {Z}}(t+1))-PV({\mathbf {Z}}(t))\\ \le&V({\mathbf {Z}}(t))-\frac{1}{2}\phi ({\mathbf {Z}}(t))+2C -\left( \frac{1}{2}\phi ({\mathbf {Z}}(t))+C -(V({\mathbf {Z}}(t+1))-PV({\mathbf {Z}}(t)))\right) \end{aligned}\nonumber \\ \end{aligned}$$
(B.4)

We see from (B.2) that the quadratic variation of the martingale \(V({\mathbf {Z}}(t+1))-PV({\mathbf {Z}}(t))\) is bounded by \(\delta \phi ({\mathbf {Z}}(t))\). As a result we can use the strong law of large numbers for martingales and the bound (B.2), to get

$$\begin{aligned} \lim _{T\rightarrow \infty } \frac{\frac{1}{T}\sum _{t=0}^T(V({\mathbf {Z}}(t+1))-PV({\mathbf {Z}}(t)))}{\frac{1}{T}\sum _{t=0}^T \left( \phi ({\mathbf {Z}}(t))+C\right) }=\lim _{T\rightarrow \infty } \frac{\sum _{t=0}^T(V({\mathbf {Z}}(t+1))-PV({\mathbf {Z}}(t)))}{\sum _{t=0}^T \left( \phi ({\mathbf {Z}}(t))+C\right) }=0\,\text { a.s.} \end{aligned}$$

which implies

$$\begin{aligned} \limsup _{T\rightarrow \infty } \frac{1}{T}\sum _{t=0}^T \left( \frac{1}{2}\phi ({\mathbf {Z}}(t))+C -(V({\mathbf {Z}}(t+1))-PV({\mathbf {Z}}(t)))\right) \le 0 \text { a.s.} \end{aligned}$$
(B.5)

Taking sums in (B.4), noting that

$$\begin{aligned} \liminf _{T\rightarrow \infty }\frac{1}{T} \sum _0^T (V({\mathbf {Z}}(t+1)) - V({\mathbf {Z}}(t)))= \liminf _{T\rightarrow 0}\frac{1}{T} (V({\mathbf {Z}}(T+1)) - V({\mathbf {z}})) \ge 0\text { a.s.} \end{aligned}$$

(because V is nonnegative) and using (B.5) yields that with probability one

$$\begin{aligned} \liminf _{T\rightarrow \infty } \frac{1}{T}\sum _0^T \left( -\frac{1}{2}\phi ({\mathbf {Z}}(t))+2C\right) \ge 0. \end{aligned}$$

As a result

$$\begin{aligned} \limsup _{T\rightarrow \infty } \frac{1}{T}\sum _0^T \phi ({\mathbf {Z}}(t))\le 4C, \end{aligned}$$
(B.6)

almost surely. Since \(\lim _{|{\mathbf {z}}|\rightarrow \infty }\phi ({\mathbf {z}})=\infty \), the boundedness of \(\widetilde{\Pi }_T\phi \) in (B.6) implies that the family of randomized occupation measures \(\{\widetilde{\Pi }_t, t\in {\mathbb {N}}\}\) is tight. Moreover, the strong law of large numbers together with (B.6) and (B.2) implies (B.3). \(\square \)

Theorem B.1

If \(E_1\) is nonempty, then for any \(I\in E_1\), there exists \(\alpha _I>0\) such that, for any a compact set \({\mathcal {K}}^I\subset {\mathcal {S}}^{I}_+\), we have

$$\begin{aligned} \lim _{\mathrm {dist}({\mathbf {z}},{\mathcal {K}}^I)\rightarrow 0, {\mathbf {z}}\in S^\circ }{\mathbb {P}}_{\mathbf {z}}\left\{ \lim _{t\rightarrow \infty }\dfrac{\ln X_i(t)}{t}\le -\alpha _I, i\in I^c\right\} =1. \end{aligned}$$

Theorem B.2

If \(E_2\) is empty or \(\max _{i}\{ r_i(\nu )\}>0\) for any \(\nu \) with \(\mu ({\mathcal {S}}^{J}_+)=1\) for some \(J\in E_2\) and \(\cup _{I\in E_1}{\mathcal {S}}^{I}_+\) is accessible then

$$\begin{aligned} \sum _{I\in E_1} p_{{\mathbf {z}},I}=1 \end{aligned}$$

where

$$\begin{aligned} p_{{\mathbf {z}},I}= & {} {\mathbb {P}}_{\mathbf {z}}\left\{ \emptyset \ne {\mathcal {U}}(\omega )\subset {{\,\mathrm{Conv}\,}}\{{\mathcal {M}}^{I,+}\} ~\text {and}~\lim _{t\rightarrow \infty }\right. \\&\left. \frac{\ln X_j(t)}{t}\in \{r_j(\mu ):\mu \in {{\,\mathrm{Conv}\,}}({\mathcal {M}}^{I,+})\},\right. \\&\left. j\in I^c\right\} . \end{aligned}$$

Proof

Once Theorem B.1 is proved, Theorem B.2 can be obtained using the fact that any weak limit of a family of random occupation measures is an invariant probability measure supported on \(S_0\) (Benaim 2018; Hening and Nguyen 2018a) and by using the arguments from Lemma 5.8, Lemma 5.9 and Theorem 5.2 by Hening and Nguyen (2018a). \(\square \)

Fix \(I\in E_1\). Since by Lemma B.1 the family \(({\widetilde{\Pi }}_t)_{t\in {\mathbb {N}}}\) of random occupation measures is tight, condition (2.9) is equivalent to the existence of \(0<{\widehat{p}}_i<\gamma _3/n, i\in I\)

$$\begin{aligned} \inf _{\nu \in {{\,\mathrm{Conv}\,}}({\mathcal {M}}^{I,\partial })}\sum _{i\in I}{\widehat{p}}_i r_i(\nu )>0. \end{aligned}$$

As a result, there exists a small \({\check{p}}\in (0,\gamma _3/n)\) such that

$$\begin{aligned} \begin{aligned} \sum _{i\in I}{\widehat{p}}_i r_i(\nu )-{\check{p}}\max _{i\notin I}\{ r_i(\nu )\}>0 \text { for any }\nu \in {{\,\mathrm{Conv}\,}}({\mathcal {M}}^{I,\partial }). \end{aligned} \end{aligned}$$
(B.7)

Define \({\tilde{p}}_i\) by \({\tilde{p}}_i={\widehat{p}}_i\) if \(i\in I_\mu \) and \({\tilde{p}}_i=-{\check{p}}\) if \(i\in I_\mu ^c\). In view of (B.7), (2.8) and Lemma B.1, there is \( r_e>0\) such that for any \(\nu \in {{\,\mathrm{Conv}\,}}({\mathcal {M}}^I)\),

$$\begin{aligned} \sum _{i\in I}{\widehat{p}}_i r_i(\nu )-{\check{p}}\max _{i\in I^c}\left\{ r_i(\nu )\right\} >3r_e. \end{aligned}$$
(B.8)

Lemma B.2

Let \(I\in E_1\) and suppose that Assumption 2.1 holds. Suppose \({\widehat{p}}_i, {\check{p}}, r_e\) are the quantities from (B.8) and \(n^*\) is defined by (A.6). There exist constants \(T_e\ge 0\), \(\delta _e>0\) such that, for any \(T\in [T_e,n^*T_e]\cap {\mathbb {Z}}\), \(\Vert {\mathbf {z}}\Vert \le M, x_i<\delta _e, i\in I^c\), we have

$$\begin{aligned} \begin{aligned}&\sum _{t=0}^T{\mathbb {E}}_{\mathbf {z}}\left( \ln V({\mathbf {Z}}(t+1))-\ln V({\mathbf {Z}}(t))-\sum _{i\in I} \ln F_i({\mathbf {Z}}(t), \xi (t))+{\check{p}} \max _{i\in I^c} \ln F_i({\mathbf {Z}}(t), \xi (t))\right) dt\\&\quad \le -r_e(T+1). \end{aligned} \end{aligned}$$
(B.9)

Proof

This is very similar to the proof of Lemma A.4 and is therefore omitted. \(\square \)

Proposition B.1

Let \(I\in E_1\) and suppose that Assumption 2.1 holds. There exists \(\theta \in (0,1)\) such that for any \(T\in [T_e,n^*T_e]\cap {\mathbb {Z}}\) and \({\mathbf {z}}\in {\mathcal {S}}_+\) satisfying \( \Vert {\mathbf {z}}\Vert \le M,\) \(x_i<\delta _e,\) \(i\in I^c\) one has

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}W_\theta ({\mathbf {Z}}(T))\le \exp \left( -\frac{1}{2}\theta r_eT\right) W_\theta ({\mathbf {z}}) \end{aligned}$$

where \(M, T_e, {\widehat{p}}_i,{\check{p}}, \delta _e, n^*\) are as in Lemma B.2 and

$$\begin{aligned} W_\theta ({\mathbf {z}}):=\sum _{i\in I^c}\left[ V({\mathbf {z}})\dfrac{x_i^{\check{p}}}{\prod _{j\in I} x_j^{{\widehat{p}}_j}}\right] ^\theta , {\mathbf {z}}\in {\mathcal {S}}_+. \end{aligned}$$

Proof

For \(i\in I^c\), let \(W({\mathbf {z}},i):=V({\mathbf {z}})\dfrac{x_i^{{\check{p}}}}{\prod _{j\in I} x_j^{{\widehat{p}}_j}}.\) Similarly to Proposition A.1, by making use of Lemma B.2, one can find a \(\theta >0\) such that for \(T\in [T_e,n^*T_e]\cap {\mathbb {Z}}\), \({\mathbf {z}}\in {\mathbb {R}}^{n,\circ }_+\times {\mathbb {R}}^{\kappa _0}\) with \(\Vert {\mathbf {z}}\Vert \le M,\) and \(x_i<\delta _e\) we have

$$\begin{aligned} {\mathbb {E}}W^\theta ({\mathbf {Z}}(T),i)\le \exp \left( -\frac{1}{2}\theta r_eT\right) W^\theta ({\mathbf {z}}, i). \end{aligned}$$

The proof is complete by noting that

$$\begin{aligned} W_\theta ({\mathbf {z}})=\sum _{i\in I^c}W^\theta ({\mathbf {z}},i). \end{aligned}$$

\(\square \)

Proof of Theorem B.1

Let \(I\in E_1\). By Lemma A.2 for any \(i\in I^c\)

$$\begin{aligned} P W({\mathbf {z}},i)\le \rho _2 W({\mathbf {z}},i), |{\mathbf {z}}|\ge M. \end{aligned}$$

Then using Jensen’s inequality, we have

$$\begin{aligned} P W_\theta ({\mathbf {z}})\le \rho _2^\theta W_\theta ({\mathbf {z}}) \text { if } |{\mathbf {z}}|\ge M. \end{aligned}$$
(B.10)

Define the constants

$$\begin{aligned} C_U:=\sup \left\{ \dfrac{\prod _{i\in I} x_i^{{\widehat{p}}_i}}{V({\mathbf {z}})}: {\mathbf {z}}\in {\mathbb {R}}^{n,\circ }_+\times {\mathbb {R}}^{\kappa _0}\right\} <\infty , \\ \varsigma :=\dfrac{\delta _e^{{\check{p}}\theta }}{C_U^\theta } \end{aligned}$$

and the stopping time

$$\begin{aligned} \eta :=\inf \left\{ t\ge 0: W_\theta ({\mathbf {Z}}(t))\ge \varsigma \right\} . \end{aligned}$$

Clearly, if \(W_\theta ({\mathbf {z}})<\varsigma \), then \(\eta >0\) and for any \(i\in I^c\), we get

$$\begin{aligned} X_i(t)\le \delta _e\,, t\in [0,\eta ). \end{aligned}$$
(B.11)

Let

$$\begin{aligned} \widetilde{W}_\theta ({\mathbf {z}}):=\varsigma \wedge W_\theta ({\mathbf {z}}). \end{aligned}$$

We have from the concavity of \(x\mapsto x\wedge \varsigma \) that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}\widetilde{W}_\theta ({\mathbf {Z}}(T))\le \varsigma \wedge {\mathbb {E}}_{\mathbf {z}}W_\theta ({\mathbf {Z}}(T)). \end{aligned}$$

Let \(\tau \) be defined as in (A.20). By (B.10) we have that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}\left[ \rho _2^{-\theta (\tau \wedge \eta \wedge n^*T_e)}W_\theta ({\mathbf {Z}}(\theta \gamma _b(\tau \wedge \xi \wedge n^*T_e))\right] \le W_\theta ({\mathbf {z}}), {\mathbf {z}}\in {\mathcal {S}}_+. \end{aligned}$$

As a result for all \({\mathbf {z}}\in {\mathcal {S}}_+\)

$$\begin{aligned} \begin{aligned} W_\theta ({\mathbf {z}})\ge&\, {\mathbb {E}}_{\mathbf {z}}\left[ \rho _2^{-\theta (\tau \wedge \eta \wedge n^*T_e)}W_\theta ({\mathbf {Z}}(\tau \wedge \eta \wedge n^*T_e))\right] \\ \ge&\, {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\tau \}}W_\theta ({\mathbf {Z}}(\tau ))\right] \\&+\, {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\eta \}}W_\theta ({\mathbf {Z}}(\eta ))\right] \\&+\,\rho _2^{-\theta (n^*-1)T_e} {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{(n^*-1)T_e<\tau \wedge \eta \le n^*T_e\}}W_\theta ({\mathbf {Z}}(\tau \wedge \eta ))\right] \\&+\,\rho _2^{-\theta n^*T_e} {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta > n^*T_e\}}W_\theta ({\mathbf {Z}}(n^*T_e))\right] .\\ \end{aligned} \end{aligned}$$
(B.12)

By the strong Markov property of \(({\mathbf {Z}}(t))\) and Proposition B.1 (which we can use because of (B.11)) we see that for all \({\mathbf {z}}\in {\mathcal {S}}_+\)

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}&\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\tau \}}W_\theta ({\mathbf {Z}}(n^*T_e))\right] \\&\le {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\tau \}}\exp \left( -\frac{1}{2}\theta r_e(n^*T_e-\tau )\right) W_\theta ({\mathbf {Z}}(\tau ))\right] \\&\le \exp \left( -\frac{1}{2}\theta r_eT_e)\right) {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\tau \}}W_\theta ({\mathbf {Z}}(\tau ))\right] . \ \end{aligned} \end{aligned}$$
(B.13)

Similarly, by the strong Markov property of \(({\mathbf {Z}}(t))\) and Lemma A.2, we obtain for any \({\mathbf {z}}\in {\mathcal {S}}_+\) that

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {z}}&\left[ {\varvec{1}}_{\{(n^*-1)T_e<\tau \wedge \eta \le n^*T_e\}}W_\theta ({\mathbf {Z}}(n^*T_e))\right] \\&\le {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{(n^*-1)T_e<\tau \wedge \eta \le n^*T_e\}}C_2^{\theta (n^*T_e-\tau \wedge \eta )}W_\theta ({\mathbf {Z}}(\tau \wedge \eta ))\right] \\&\le C_2^{\theta T_e}{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{(n^*-1)T_e<\tau \wedge \eta \le n^*T_e\}}W_\theta ({\mathbf {Z}}( \tau \wedge \eta ))\right] . \end{aligned} \end{aligned}$$
(B.14)

If \(W_\theta ({\mathbf {z}})<\varsigma \) then applying (B.13), (B.14) and the inequality \(\widetilde{W}_\theta ({\mathbf {Z}}(n^*T_e))\le W_\theta ({\mathbf {Z}}(n^*T_e\wedge \eta ))\) to (B.12) yields

$$\begin{aligned} \begin{aligned} \widetilde{W}_\theta ({\mathbf {z}})=W_\theta ({\mathbf {z}}) \ge&{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\tau \}}W_\theta ({\mathbf {Z}}(\tau ))\right] \\&+ {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\eta \}}W_\theta ({\mathbf {Z}}(\eta ))\right] \\&+\rho _2^{-\theta (n^*-1)T_e} {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{(n^*-1)T_e<\tau \wedge \eta \le n^*T\}}W_\theta ({\mathbf {Z}}(\tau \wedge \eta ))\right] \\&+\rho _2^{-\theta n^*T_e} {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta> n^*T_e\}}W_\theta ({\mathbf {Z}}(n^*T_e))\right] \\ \ge&\exp \left( \frac{1}{2}\theta r_eT_e)\right) {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\tau \}}W_\theta ({\mathbf {Z}}(n^*T_e))\right] \\&+ {\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta \wedge (n^*-1)T_e=\eta \}}\widetilde{W}_\theta ({\mathbf {Z}}(n^*T_e))\right] \\&+\rho _2^{-\theta (n^*-1)T_e}C_2^{-\theta T_e}{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{(n^*-1)T_e<\tau \wedge \eta \le n^*T_e\}}W_\theta ({\mathbf {Z}}(n^*T_e))\right] \\&+\rho _2^{-\theta n^*T_e}{\mathbb {E}}_{\mathbf {z}}\left[ {\varvec{1}}_{\{\tau \wedge \eta > n^*T_e\}}W_\theta ({\mathbf {Z}}(n^*T_e))\right] \\ \ge&{\mathbb {E}}_{\mathbf {z}}\widetilde{W}_\theta ({\mathbf {Z}}(n^*T_e)) \,\quad \text { (since } \widetilde{W}_\theta (\cdot )\le W_\theta (\cdot )). \end{aligned} \end{aligned}$$
(B.15)

Clearly, if \(W_\theta ({\mathbf {z}})\ge \varsigma \) then

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}\widetilde{W}_\theta ({\mathbf {Z}}(n^*T_e)) \le \varsigma =\widetilde{W}_\theta ({\mathbf {z}}). \end{aligned}$$
(B.16)

As a result of (B.15), (B.16) and the Markov property of \(({\mathbf {Z}}(t))\), the sequence \(\{w_2(k): k\ge 0\}\) where \(w_2(k):=\widetilde{W}_\theta ({\mathbf {Z}}(kn^*T_e))\) is a supermartingale. Define the discrete stopping time

$$\begin{aligned} \eta ^*:=\inf \{k\in {\mathbb {N}}: W_\theta ({\mathbf {Z}}(kn^*T_e))\ge \varsigma \}. \end{aligned}$$

Moreover, we also deduce from (B.15) that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}{\varvec{1}}_{\{\eta ^*>1\}}W_\theta ({\mathbf {Z}}(n^*T_e))\le \kappa _e W_\theta ({\mathbf {z}}), {\mathbf {z}}\in {\mathcal {S}}_+ \end{aligned}$$

where \(\kappa _e^{-1}=\min \left\{ \rho _2^{-\theta (n^*-1)T_e}C_2^{-\theta T_e},\exp \left( \frac{1}{2}\theta r_eT_e)\right) ,\rho _2^{-\theta n^*T_e}\right\} >1.\) As a result, \(\{w_3(k): k\ge 0\}\) with

$$\begin{aligned} w_3(k):=\kappa _e^{-k}{\varvec{1}}_{\{\eta ^*>k\}}W_\theta ({\mathbf {Z}}(kn^*T_e)) \end{aligned}$$

is also a supermartingle. For any \(\varepsilon \in (0,1)\), if \(W_\theta ({\mathbf {z}})\le \varsigma \varepsilon \) we have

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}w_2(k\wedge \eta ^*)\le {\mathbb {E}}_{\mathbf {z}}w_2(0)=W_\theta ({\mathbf {z}})\le \varsigma \varepsilon \,, k\ge 0. \end{aligned}$$
(B.17)

Subsequently, (B.17) combined with the Markov inequality and the fact that \(w_2(\eta )=W_\theta ({\mathbf {Z}}(\eta n^*T_e))\ge \varsigma \) yields

$$\begin{aligned} {\mathbb {P}}_{\mathbf {z}}\{\eta ^*<k\}\le \varsigma ^{-1}{\mathbb {E}}_{\mathbf {z}}w_2(k\wedge \eta ^*)\le \varepsilon , ~\text {if}~k\in {\mathbb {N}}, W_\theta ({\mathbf {z}})\le \varsigma \varepsilon . \end{aligned}$$

Next, let \(k\rightarrow \infty \) to get

$$\begin{aligned} {\mathbb {P}}_{\mathbf {z}}\{\eta ^*<\infty \}\le \varepsilon \,\text { if }\,W_\theta ({\mathbf {z}})\le \varsigma \varepsilon . \end{aligned}$$
(B.18)

Let \(\kappa _3\in (\kappa _e,1)\). Since \(P_t \widetilde{W}_\theta ({\mathbf {z}})\le C_2^{\theta t} \widetilde{W}_\theta ({\mathbf {z}})\) for any \(t>0, {\mathbf {z}}\in {\mathcal {S}}_+\), we have that

$$\begin{aligned} {\mathbb {P}}_{\mathbf {z}}\left\{ \widetilde{W}_\theta ({\mathbf {Z}}(t))\le c, t=1,\ldots , n^*T_e\right\} \le \dfrac{n^*T_e C_2^{\theta n^*T_e}\widetilde{W}_\theta ({\mathbf {z}})}{c}. \end{aligned}$$

The last inequality, the Markov property of \({\mathbf {Z}}(t)\), and the fact that \((w_3(k))_{k\ge 0}\) is a supermartingale imply that for any \({\mathbf {z}}\in {\mathcal {S}}_+\)

$$\begin{aligned} \begin{aligned} {\mathbb {P}}_{\mathbf {z}}&\left\{ {\varvec{1}}_{\{\eta ^*>k\}}\widetilde{W}_\theta ({\mathbf {Z}}(t))\le \left( \frac{\kappa _e}{\kappa _3}\right) ^{-k}, t\in [kn^*T_e+1, (k+1)n^*T_e]\right\} \\&\le n^*T_e C_2^{\theta n^*T_e}{\mathbb {E}}_{\mathbf {z}}w_3(k)\dfrac{\kappa _e^{-k}}{\kappa _3^{-k}}\le n^*T_e C_2^{\theta n^*T_e}\dfrac{{\mathbb {E}}_{\mathbf {z}}w_3(0)}{\kappa _3^{-k}}\le n^*T_e C_2^{\theta n^*T_e}\dfrac{W_\theta ({\mathbf {z}})}{\kappa _3^{-k}} \end{aligned} \end{aligned}$$

Since \(\sum _k\left( n^*T_e C_2^{\theta n^*T_e}\dfrac{W_\theta ({\mathbf {z}})}{\kappa _3^{-k}}\right) <\infty \), we deduce from the Borel–Cantelli lemma that

$$\begin{aligned} {\mathbb {P}}_{\mathbf {z}}\left\{ \limsup _{k\rightarrow \infty } \sup _{t\in [kn^*T_e+1, (k+1)n^*T_e]}{\varvec{1}}_{\{\eta ^*>k\}}\widetilde{W}_\theta ({\mathbf {Z}}(t))\left( \frac{\kappa _3}{\kappa _e}\right) ^{k}<1\right\} =1. \end{aligned}$$

This and (B.18) imply that if \(W_\theta ({\mathbf {z}})\le \varsigma \varepsilon \) then

$$\begin{aligned} {\mathbb {P}}_{\mathbf {z}}\left\{ \limsup _{t\rightarrow \infty } \left( \left( \frac{\kappa _3}{\kappa _e}\right) ^{\frac{t}{n^*T_e+1}} W_\theta ({\mathbf {Z}}(t))\right) <1\right\} \ge 1-\varepsilon . \end{aligned}$$

Since \(\sup _{{\mathbf {z}}\in {\mathbb {R}}^n\times {\mathbb {R}}^{\kappa _0}}\frac{x_i^{\theta \check{p}}}{W_\theta ({\mathbf {z}})}<\infty , i\in I^c\), we can easily obtain the extinction result from Theorem 2.5.

\(\square \)

Appendix C. Robustness proofs

Let \(\circ \) denote the element-wise product and \({\varvec{1}}_n\) be the vector in \({\mathbb {R}}^n\) whose components are all 1. Assume that the function V from Assumption A3) satisfies the robust estimate

$$\begin{aligned} {\mathbb {E}}\left[ V({\mathbf {x}}^\top (F({\mathbf {z}},\xi )\circ ({\varvec{1}}_n+\widetilde{\varepsilon }_1)), G({\mathbf {z}},\xi )+\widetilde{\varepsilon }_2))h({\mathbf {z}},\xi )\right] \le \rho V({\mathbf {z}})+C \end{aligned}$$

for any vectors \(\widetilde{\varepsilon }_1,\widetilde{\varepsilon }_2\in {\mathbb {R}}^n\) such that \(|\widetilde{\varepsilon }_1|\vee |\widetilde{\varepsilon }_2|<\delta \). Note that

$$\begin{aligned} {\tilde{h}}({\mathbf {z}},\xi )=\max _{i=1}^n \left\{ \widetilde{F}_i({\mathbf {z}},\xi ),\frac{1}{\widetilde{F}_i({\mathbf {z}},\xi )}\right\} ^{\gamma _3}. \end{aligned}$$

Consequently \({\tilde{h}}({\mathbf {z}},\xi )\le e^\delta h({\mathbf {z}},\xi )\) and

$$\begin{aligned} {\mathbb {E}}\left[ V({\mathbf {x}}^\top (\widetilde{F}({\mathbf {z}},\xi )), \widetilde{G}({\mathbf {z}},\xi ))\widetilde{h}({\mathbf {z}},\xi )\right] \le \rho V({\mathbf {z}})+C \end{aligned}$$

if \(\delta >0\) is sufficiently small. This shows that there exist \(C_2>0,\rho _2\in (0,1),\gamma _4\in (0,\gamma _3)\) such that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}V\left( \widetilde{\mathbf {Z}}(1)\prod _{i=1}^n\widetilde{X}_i^{p_i}(1)\right) \le \left( {\varvec{1}}_{\{|{\mathbf {z}}|<M\}}(C_2-\rho _2)+\rho _2\right) V^{\frac{1}{2}}({\mathbf {z}})\prod _{i=1}^nx_i^{p_i} \end{aligned}$$

for any \({\mathbf {p}}=(p_1,\ldots ,p_n)\in {\mathbb {R}}^n\) satisfying

$$\begin{aligned} |{\mathbf {p}}|_1:=\sum |p_i|\le \gamma _4. \end{aligned}$$

Analogously to Lemma A.1 one can show the following uniform bound

$$\begin{aligned} \sup _{|{\mathbf {z}}|\le M, t\in {\mathbb {N}}} \left( {\mathbb {E}}_{\mathbf {z}}V(\widetilde{\mathbf {Z}}(t))+{\mathbb {E}}_{{\mathbf {z}}}\tilde{h}(\widetilde{\mathbf {Z}}(t),\xi (t))\right) =K_{M}<\infty \end{aligned}$$
(C.1)

when \(\delta \) is sufficiently small.

Slight changing in the factor on the right hand side of A.5, we have

Lemma C.1

Suppose that Assumption A.1 holds. Let \({\mathbf {p}}\) and \(r^*\) be as in (A.4). There exists a \(\widetilde{T}^*>0\) such that, for any \(T>\widetilde{T}^*\), \({\mathbf {z}}\in {\mathcal {S}}_0, \Vert {\mathbf {z}}\Vert \le M\) one has

$$\begin{aligned} \sum _{t=0}^T{\mathbb {E}}_{\mathbf {z}}\left( \ln V({\mathbf {Z}}(t+1))-\ln V({\mathbf {Z}}(t))-\sum p_i\ln F_i({\mathbf {Z}}(t), \xi (t))\right) \le -1.5r^*(T+1). \end{aligned}$$
(C.2)

On the other hand, it obviously follows from (2.11) that for any \(\varepsilon>0, M>0, T>0, n_0>0\), we have

$$\begin{aligned} {\mathbb {E}}_{\mathbf {z}}\left\{ {\varvec{1}}_{\{|\widetilde{\mathbf {Z}}(t))|\vee |{\mathbf {Z}}(t)|<n_0\}}|\widetilde{\mathbf {Z}}(t))-{\mathbf {Z}}(t)|\right\} \le \varepsilon ,\, {\mathbf {z}}\le M, t\in \{0,\ldots , T\} \end{aligned}$$

when \(\delta \) is sufficiently small. This and the uniform integrability (C.1) imply that, for any \(\varepsilon , T>0\), there exists \(\delta >0\) such that for any \(\delta \)-pertubation of (3.1), we have

$$\begin{aligned} \begin{aligned}&{\mathbb {E}}_{{\mathbf {z}}}|\ln \widetilde{F}_i(\widetilde{\mathbf {Z}}(t))-\ln F_i({\mathbf {Z}}(t))| \le {\mathbb {E}}_{{\mathbf {z}}}|\ln \widetilde{F}_i(\widetilde{\mathbf {Z}}(t))-\ln \widetilde{F}_i({\mathbf {Z}}(t))|\\&\quad +{\mathbb {E}}_{{\mathbf {z}}}|\ln \widetilde{F}_i({\mathbf {Z}}(t))-\ln F_i({\mathbf {Z}}(t))| \varepsilon ,\, {\mathbf {z}}\in {\mathcal {S}}_0,|{\mathbf {z}}|\le M, t\in \{0,\ldots , T\}. \end{aligned} \end{aligned}$$
(C.3)

By virtue of (C.2) and (C.3), for sufficiently small \(\delta \),

$$\begin{aligned} \sum _{t=0}^T{\mathbb {E}}_{\mathbf {z}}\left( \ln V(\widetilde{\mathbf {Z}}(t+1))-\ln V(\widetilde{\mathbf {Z}}(t))-\sum p_i\ln F_i(\widetilde{\mathbf {Z}}(t)), \xi (t))\right) \le -r^*(T+1) \end{aligned}$$
(C.4)

for any \({\mathbf {z}}\in {\mathcal {S}}_0, |{\mathbf {z}}|\le M\) and \(T\ge \widetilde{T}^*.\) With this estimate, it follows from the arguments in Appendix A that Theorem A.1 holds for \(\widetilde{\mathbf {Z}}(t)\). Similarly, Theorem 2.5 and Theorem 2.6 hold.

Appendix D. Proof of Lemma 4.1

Lemma D.1

Suppose \(\{(E_i(t), S(t))_{i=1,\ldots , n}\}, t\in {\mathbb {N}}\) is a sequence of \(n+1\)-dimensional random variables, i.i.d. over t such that \({\mathbb {E}}\left[ S(t)e^{E_j(t)}\right] ^2<\infty .\) Then the model given by (2.4) and (4.11) satisfies Assumption 2.1 by taking a small enough \(\gamma _3>0\), and

$$\begin{aligned} V({\mathbf {z}}) = \sum _j z_j + 1. \end{aligned}$$

Assumption 2.2 holds with

$$\begin{aligned} \phi ({\mathbf {z}}) = \delta V({\mathbf {z}}) \end{aligned}$$

for some \(\delta >0\). Moreover, if the support of \(\ln S(t) + max_j (E_j(t)\ln \delta _j)\) contains values less than 0 then the boundary is accessible.

Proof

We denote by K below a positive generic constant. Using (4.11)

$$\begin{aligned} \sum _j N_j(t)e^{E_j(t)-D_j(t)}\le \sum _j N_j(t) e^{E_j(t)}\exp \{-\alpha _{j} N_j(t) e^{E_j(t)}\}< K<\infty .\nonumber \\ \end{aligned}$$
(D.1)

On the other hand

$$\begin{aligned} \left( 1-\delta _j+ S(t)e^{E_j(t)-D_j(t)}\right) \vee \left( 1-\delta _j+ S(t)e^{E_j(t)-D_j(t)}\right) ^{-1} \le K+\sum _j S(t)e^{E_j(t)-D_j(t))} \end{aligned}$$

Using this and the inequality

$$\begin{aligned} (a+b)^\gamma \le 2^\gamma (a^\gamma +b^\gamma ) \end{aligned}$$

we see that for any \(\varepsilon >0\) there is \(\gamma _3>0\) sufficiently small such that

$$\begin{aligned} \begin{aligned} h(t):=&\max _j\left( \left( 1-\delta _j)+ S(t)e^{E_j(t)-D_j(t)}\right) ^{\gamma _3}\vee \left( 1-\delta _j+S(t) e^{E_j(t)-D_j(t)}\right) ^{-\gamma _3}\right) \\ \le&\left( 1+\varepsilon +\left( S(t)\sum e^{E_j(t)-D_j(t)}\right) ^{\gamma _3}\right) \end{aligned} \end{aligned}$$

Using (D.1) and the fact that \({\mathbb {E}}S(t)<\infty \) we get

$$\begin{aligned} {\mathbb {E}}^t\sum _j \left( (1-\delta _j)N_j(t)(1+\varepsilon +\left( S(t)e^{E_j(t)-D_j(t)}\right) ^{\gamma _3})\right) \le (1-0.5{\check{d}}) \sum _j N_j(t) +K \end{aligned}$$

where \({\mathbb {E}}^t[\cdot ]:={\mathbb {E}}[\cdot ~|~N_1(t),\ldots , N_n(t)]\) and \({\check{d}}=\min \{\delta _i\}\). Since \(N_j(t) e^{E_j(t)}\exp \{-a_{j}(1+\gamma _3) N_j(t) e^{E_j(t)}\) is bounded above by a nonrandom constant, then

$$\begin{aligned} N_j(t)e^{(1+\gamma _3)(E_j(t)-D_j(t))}\le e^{\gamma _3 E_j(t)} N_j(t) e^{E_j(t)}\exp \{-a_{j}(1+\gamma _3) N_j(t) e^{E_j(t)}\}< Ke^{\gamma _3 E_j(t)} \end{aligned}$$

Therefore, since \({\mathbb {E}}\left[ S^2(t)e^{\gamma _3 E_j(t)}\right] <\infty \) we have

$$\begin{aligned} {\mathbb {E}}^t S^{1+\gamma _3}(t)N_j(t)e^{(1+\gamma _3)(E_j(t)-D_j(t))} \le K{\mathbb {E}}^t \left( S^{1+\gamma _3}(t)e^{\gamma _3 E_j(t)}\right) \le K_2. \end{aligned}$$

As a result,

$$\begin{aligned}&{\mathbb {E}}^t \sum _j \left( 1-\delta _j+S(t)e^{E_j(t)-D_j(t)}\right) N_j(t)h(t) \le \sum _j (1-0.5\delta _j) N_j(t) \\&\quad + K_2 \le (1-0.25{\check{d}}) \sum _j N_j(t)+K_2. \end{aligned}$$

This implies that Assumption A3) is satisfied with \(V({\mathbf {z}})=\sum _j N_j+1=\sum _j z_j+1\) and small \(\gamma _3>0\). Note that

$$\begin{aligned} \sum _j N_j(t+1)-P_1 \sum _j N_j(t)=S(t)\sum _j e^{E_j(t)-D_j(t)} N_j(t)- {\mathbb {E}}^t S(t) \sum _j e^{E_j(t)-D_j(t)} N_j(t) \end{aligned}$$

This shows that

$$\begin{aligned} \begin{aligned} {\mathbb {E}}^t \left( \sum _j N_j(t+1)-P_1 \sum _j N_j(t)\right) ^2\le&\, {\mathbb {E}}^t \left( S(t)\sum _j e^{E_j(t)-D_j(t)} N_j(t)\right) ^2\\ \le&\, K_3^2{\mathbb {E}}^t \left[ S(t)e^{E_j(t)}\right] ^2 = K_3^2 {\mathbb {E}}^t \left[ S(t)e^{E_j(t)}\right] ^2 \end{aligned} \end{aligned}$$

since \(\sum _j e^{-D_j(t)} N_j(t)\) is bounded by a constant \(K_3\). If \({\mathbb {E}}\left[ S(t)e^{E_j(t)}\right] ^2<\infty \) then Assumption 2.2 is satisfied with \(\phi ({\mathbf {z}})=\delta (|{\mathbf {z}}|+1)\) for some \(\delta >0\).

Moreover, if support of \(\ln S(t) + \max _j \{E_j(t)\ln \delta _j\}\) contains values less than 0 then it is clear that the boundary is accessible since when \(\ln S(t) + \max _j \{E_j(t) \ln \delta _j\}\) is less than a negative constant, \(N_j(t+1)\le \rho N_j(t)\) for \(\rho <1\). \(\square \)

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. & Chesson, P. A general theory of coexistence and extinction for stochastic ecological communities. J. Math. Biol. 82, 56 (2021). https://doi.org/10.1007/s00285-021-01606-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-021-01606-1

Keywords

Mathematics Subject Classification

Navigation