Skip to main content

Advertisement

Log in

Coexistence, Extinction, and Optimal Harvesting in Discrete-Time Stochastic Population Models

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

We analyze the long-term behavior of interacting populations which can be controlled through harvesting. The dynamics is assumed to be discrete in time and stochastic due to the effect of environmental fluctuations. We present powerful extinction and coexistence criteria when there are one or two interacting species. We then use these tools in order to see when harvesting leads to extinction or persistence of species, as well as what the optimal harvesting strategies, which maximize the expected long-term yield, look like. For single species systems, we show under certain conditions that the optimal harvesting strategy is of bang-bang type: there is a threshold under which there is no harvesting, while everything above this threshold gets harvested. We are also able to show that stochastic environmental fluctuations will, in most cases, force the expected harvesting yield to be lower than the deterministic maximal sustainable yield. The second part of the paper is concerned with the analysis of ecosystems that have two interacting species which can be harvested. In particular, we carefully study predator–prey and competitive Ricker models. We are able to analytically identify the regions in parameter space where the species coexist, one species persists and the other one goes extinct, as well as when there is bistability. We look at how one can find the optimal proportional harvesting strategy. If the system is of predator–prey type, the optimal proportional harvesting strategy is, depending on the interaction parameters and the price of predators relative to prey, either to harvest the predator to extinction and maximize the asymptotic yield of the prey or to not harvest the prey and to maximize the asymptotic harvesting yield of the predators. If the system is competitive, in certain instances it is optimal to drive one species extinct and to harvest the other one. In other cases, it is best to let the two species coexist and harvest both species while maintaining coexistence. In the setting of the competitive Ricker model, we show that if one competitor is dominant and pushes the other species to extinction, the harvesting of the dominant species can lead to coexistence.

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

Access this article

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

Instant access to the full article PDF.

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

Similar content being viewed by others

References

  • Alvarez, L.H.R., Hening, A.: Optimal sustainable harvesting of populations in random environments. Stoch. Process. Appl. https://doi.org/10.1016/j.spa.2019.02.008

  • Alvarez, L.H.R., Lungu, E., Øksendal, B.: Optimal multi-dimensional stochastic harvesting with density-dependent prices. Afr. Mat. 27(3–4), 427–442 (2016)

    MathSciNet  MATH  Google Scholar 

  • Alvarez, L.H.R., Shepp, L.A.: Optimal harvesting of stochastically fluctuating populations. J. Math. Biol 37, 155–177 (1998)

    MathSciNet  MATH  Google Scholar 

  • Anderson, C.N., Hsieh, C.-H., Sandin, S.A., Hewitt, R., Hollowed, A., Beddington, J., May, R.M., Sugihara, G.: Why fishing magnifies fluctuations in fish abundance. Nature 452(7189), 835–839 (2008)

    Google Scholar 

  • Bayliss, P.: Population dynamics of magpie geese in relation to rainfall and density: implications for harvest models in a fluctuating environment. J. Appl. Ecol. 26, 913–924 (1989)

    Google Scholar 

  • Benaïm, M., Schreiber, S.J.: Persistence and extinction for stochastic ecological models with internal and external variables. J. Math. Biol. 79(1), 393–431 (2019)

    MathSciNet  MATH  Google Scholar 

  • Cameron, T.C., O’Sullivan, D., Reynolds, A., Hicks, J.P., Piertney, S.B., Benton, T.G.: Harvested populations are more variable only in more variable environments. Ecol. Evol. 6(12), 4179–4191 (2016)

    Google Scholar 

  • Caswell, H.: Predator-mediated coexistence: a nonequilibrium model. Am. Nat. 112(983), 127–154 (1978)

    Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Clark, C.W.: Mathematical Bioeconomics: The Mathematics of Conservation, vol. 91. John Wiley & Sons, New Jersey (2010)

    Google Scholar 

  • Crowley, P.H.: Predator-mediated coexistence: an equilibrium interpretation. J. Theor. Biol. 80(1), 129–144 (1979)

    Google Scholar 

  • Dai, G., Tang, M.: Coexistence region and global dynamics of a harvested predator-prey system. SIAM J. Appl. Math. 58(1), 193–210 (1998)

    MathSciNet  MATH  Google Scholar 

  • Ellner, S.: Asymptotic behavior of some stochastic difference equation population models. J. Math. Biol. 19(2), 169–200 (1984)

    MathSciNet  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Fagerholm, H., Högnäs, G.: Stability classification of a ricker model with two random parameters. Adv. Appl. Probab. 34(1), 112–127 (2002)

    MathSciNet  MATH  Google Scholar 

  • Gamelon, M., Sandercock, B.K., Sæther, B.-E.: Does harvesting amplify environmentally induced population fluctuations over time in marine and terrestrial species? J. Appl. Ecol. 56(9), 2186–2194 (2019)

    Google Scholar 

  • Getz, W.M., Haight, R.G.: Population Harvesting: Demographic Models of Fish, Forest, and Animal Resources, vol. 27. Princeton University Press, Princeton (1989)

    Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Hening, A., Nguyen, D., Chesson, P.: A general theory of coexistence and extinction for stochastic ecological communities (2020). arXiv preprint arXiv:2007.09025

  • Hening, A., Nguyen, D.H.: The competitive exclusion principle in stochastic environments. J. Math. Biol. 80(5), 1323–1351 (2020)

    MathSciNet  MATH  Google Scholar 

  • Hening, A., Nguyen, D.H., Ungureanu, S.C., Wong, T.K.: Asymptotic harvesting of populations in random environments. J. Math. Biol. 78(1–2), 293–329 (2019)

    MathSciNet  MATH  Google Scholar 

  • Hening, A., Strickler, E.: On a predator-prey system with random switching that never converges to its equilibrium. SIAM J. Math. Anal. 51(5), 3625–3640 (2019)

    MathSciNet  MATH  Google Scholar 

  • Hening, A., Tran, K.: Harvesting and seeding of stochastic populations: analysis and numerical approximation. J. Math. Biol. 81, 65–112 (2020)

    MathSciNet  MATH  Google Scholar 

  • Hening, A., Tran, K., Phan, T., Yin, G.: Harvesting of interacting stochastic populations. J. Math. Biol. 79(2), 533–570 (2019)

    MathSciNet  MATH  Google Scholar 

  • Hilborn, R., Walters, C.J., et al.: Quantitative fisheries stock assessment: choice, dynamics and uncertainty. Rev. Fish Biol. Fish. 2(2), 177–178 (1992)

    Google Scholar 

  • Hilker, F.M., Liz, E.: Proportional threshold harvesting in discrete-time population models. J. Math. Biol. 79(5), 1927–1951 (2019)

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Hofmann, E.E., Powell, T.: Environmental variability effects on marine fisheries: four case histories. Ecol. Appl. 8(sp1), S23–S32 (1998)

    Google Scholar 

  • Hsu, S.: Predator-mediated coexistence and extinction. Math. Biosci. 54(3–4), 231–248 (1981)

    MathSciNet  MATH  Google Scholar 

  • Hulme, P.E.: Adapting to climate change: is there scope for ecological management in the face of a global threat? J. Appl. Ecol. 42(5), 784–794 (2005)

    Google Scholar 

  • Larkin, P.: Some observations on models of stock and recruitment relationships for fishes. Rapp. et Proces Verbaux des Réun. Cons. Perm. Int. pour l’Explor. de la Mer 164, 316–324 (1973)

    Google Scholar 

  • Lungu, E.M., Øksendal, B.: Optimal harvesting from a population in a stochastic crowded environment. Math. Biosci. 145(1), 47–75 (1997)

    MathSciNet  MATH  Google Scholar 

  • Martin, A., Ruan, S.: Predator-prey models with delay and prey harvesting. J. Math. Biol. 43(3), 247–267 (2001)

    MathSciNet  MATH  Google Scholar 

  • May, R.M., Beddington, J., Horwood, J., Shepherd, J.: Exploiting natural populations in an uncertain world. Math. Biosci. 42(3–4), 219–252 (1978)

    MathSciNet  Google Scholar 

  • Meyn, S.P., Tweedie, R.L.: Stability of Markovian processes. I. criteria for discrete-time chains. Adv. Appl. Probab. 24(3), 542–574 (1992)

    MathSciNet  MATH  Google Scholar 

  • Myerscough, M., Gray, B., Hogarth, W., Norbury, J.: An analysis of an ordinary differential equation model for a two-species predator-prey system with harvesting and stocking. J. Math. Biol. 30(4), 389–411 (1992)

    MathSciNet  MATH  Google Scholar 

  • Reed, W.J.: The steady state of a stochastic harvesting model. Math. Biosci. 41(3–4), 273–307 (1978)

    MathSciNet  MATH  Google Scholar 

  • Reed, W.J.: Optimal escapement levels in stochastic and deterministic harvesting models. J. Environ. Econ. Manag. 6(4), 350–363 (1979)

    MATH  Google Scholar 

  • Rouyer, T., Sadykov, A., Ohlberger, J., Stenseth, N.C.: Does increasing mortality change the response of fish populations to environmental fluctuations? Ecol. Lett. 15(7), 658–665 (2012)

    Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Shelton, A.O., Mangel, M.: Fluctuations of fish populations and the magnifying effects of fishing. Proc. Nat. Acad. Sci. 108(17), 7075–7080 (2011)

    Google Scholar 

  • Sih, A., Crowley, P., McPeek, M., Petranka, J., Strohmeier, K.: Predation, competition, and prey communities: a review of field experiments. Ann. Rev. Ecol. Syst. 16(1), 269–311 (1985)

    Google Scholar 

  • Slobodkin, L.: The Growth and Regulation of Animal Numbers. Holt, Reinhardt and Winston, New York (1961)

    Google Scholar 

  • Tran, K., Yin, G.: Optimal harvesting strategies for stochastic competitive Lotka–Volterra ecosystems. Automatica 55, 236–246 (2015)

    MathSciNet  MATH  Google Scholar 

  • Tran, K., Yin, G.: Numerical methods for optimal harvesting strategies in random environments under partial observations. Automatica 70, 74–85 (2016)

    MathSciNet  MATH  Google Scholar 

  • Turelli, M.: Does environmental variability limit niche overlap? Proc. Nat. Acad. Sci. 75(10), 5085–5089 (1978)

    MATH  Google Scholar 

  • Vellekoop, M., Högnäs, G.: A unifying framework for chaos and stochastic stability in discrete population models. J. Math. Biol. 35(5), 557–588 (1997)

    MathSciNet  MATH  Google Scholar 

  • Xia, J., Liu, Z., Yuan, R., Ruan, S.: The effects of harvesting and time delay on predator-prey systems with holling type ii functional response. SIAM J. Appl. Math. 70(4), 1178–1200 (2009)

    MathSciNet  MATH  Google Scholar 

  • Yodzis, P.: The effects of harvesting on competitive systems. Bull. Math. Biol. 38(2), 97–109 (1976)

    MathSciNet  MATH  Google Scholar 

  • Yodzis, P.: Harvesting and limiting similarity. Am. Nat. 111(981), 833–843 (1977)

    Google Scholar 

Download references

Acknowledgements

The author thanks Dang Nguyen and Sergiu Ungureanu for helpful discussions related to the paper.

Funding

The funding was provided by Division of Mathematical Sciences (Grant No. 1853463).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexandru Hening.

Additional information

Communicated by Charles R. Doering.

Publisher's Note

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

The author is being supported by the NSF through the Grant DMS-1853463.

Appendices

Appendix A. Criteria for Persistence and Extinction

1.1 Single Species System

Suppose we have one species whose dynamics is given by

$$\begin{aligned} X_{t+1} = X_t f(X_t, \xi _{t+1}) \end{aligned}$$
(A.1)

We present a few known results which give the existence of a unique invariant probability measure. These results appear in work by Ellner (1984, 1989), Vellekoop and Högnäs (1997), Fagerholm and Högnäs (2002), Schreiber (2012).

Theorem A.1

Assume that \(F(x,\xi )=xf(x,\xi )\) is continuously differentiable and strictly increasing in x, and \(f(x,\xi )\) is strictly decreasing in x. If \({\mathbb {E}}[\ln f(0,\xi _1)]>0\) and \(\lim _{x\rightarrow \infty } {\mathbb {E}}[\ln f(x,\xi _1)]<0\), then there exists a positive invariant probability measure \(\mu \) and the distribution of \(X_t\) converges weakly to \(\mu \) whenever \(X_0=x>0\).

Sometimes, if monotonicity fails, one can make use of the following result (Vellekoop and Högnäs 1997).

Theorem A.2

Assume that

$$\begin{aligned} f(x,\xi )=\lambda h(x)^{-\xi } \end{aligned}$$

where g is a positive differentiable function such that \(x\mapsto xh'(x)/h(x)\) is strictly increasing on \([0,\infty )\). Assume \({\mathbb {E}}\xi _1, {\mathbb {E}}\xi _1^2 <\infty \) and \(\xi _1\) has a positive density on (0, L) for some \(0<L<\infty \). Then, there is a positive invariant probability measure \(\mu \) and the distribution of \(X_t\) converges to \(\mu \) whenever \(X_0=x>0\).

We note that the above theorem provides a classification of the stochastic Ricker model if the random variable \(\xi _1\) has a density and is supported on (0, L) for some \(L>0\). One can also fully classify (Fagerholm and Högnäs 2002) the stochastic Ricker model if the random coefficients do not have compact support.

Theorem A.3

Consider the stochastic Ricker model \(X_{t+1} = X_t \exp (r_{t+1}-a_{t+1}X_t)\) where

  • \(r_1,\ldots \) is a sequence of i.i.d. random variables such that \({\mathbb {E}}[r_1]<\infty \) and \(r_1\) has positive density on \((-\infty ,+\infty )\),

  • \(a_1,\ldots \) is a sequence of positive i.i.d. random variables independent of \(r_t\) such that \({\mathbb {E}}[a_1]<\infty \) and

  • there exists \(x_c\) such that \({\mathbb {E}}[\exp (r_1x)]<\infty \) for all \(x\in [0, x_c]\).

Then, if \({\mathbb {E}}[r_1]<0\), \(X_t\rightarrow 0\) with probability 1, while if \({\mathbb {E}}[r_1]>0\), there is a positive invariant measure \(\mu \) such that \(X_t\) converges weakly to \(\mu \).

1.2 Two Species Systems

Suppose we have a two species system. The following result appeared in work by Ellner (1989).

Theorem A.4

Assume the following assumptions are satisfied

  • For each \(i=1,2\), there exists a positive invariant measure \(\mu _i\) such that the distribution \(\mu _i\) such that the distribution of \(X_t^i\) converges to \(\mu _i\) weakly whenever \(X_0^i>0\) and \(X_0^j=0\).

  • The mean per capita growth rates \(r_i({\mathbf {x}})\) are continuous functions.

  • The process \({\mathbf {X}}\) is irreducible on \((0,\infty )\times (0,\infty )\),

  • For any Borel measurable \(A\subset {\mathbb {R}}_+^2\) we have \({\mathbb {P}}({\mathbf {X}}_1\in A|{\mathbf {X}}_0={\mathbf {x}}_n) \rightarrow {\mathbb {P}}({\mathbf {X}}_1\in A|{\mathbf {X}}_0={\mathbf {x}})\) whenever \({\mathbf {x}}_n\rightarrow {\mathbf {x}}\).

  • For any \({\mathbf {x}}\in {\mathbb {R}}_+^2\), \(\sup _{t>0}{\mathbb {E}}[\ln ^+ X_t^i~|~{\mathbf {X}}_0={\mathbf {x}}]<\infty \) for \(i=1,2\).

If \(r_1(\mu _2)>0\) and \(r_2(\mu _1)>0\), then there exists a unique positive invariant measure \(\mu \) and the distribution of \({\mathbf {X}}_t\) converges to \(\mu \) weakly whenever \(X_0^1, X_0^2>0\).

1.3 General Criteria for Coexistence

Assume we have a general n species system modeled by

$$\begin{aligned} X^i_{t+1} = X_t^i f_i({\mathbf {X}}_t, \xi _{t+1}), i=1,\ldots ,n. \end{aligned}$$
(A.2)

The subset \({\mathcal {S}}\subset {\mathbb {R}}_+^n\) will denote the state space of the dynamics. It will either be a compact subset of \({\mathbb {R}}_+^n\) or all of \({\mathbb {R}}_+^n\). The coexistence set is the subset \({\mathcal {S}}_+=\{{\mathbf {x}}\in {\mathcal {S}}~|~x_i>0, i=1,\ldots n\}\) of the state space where no species is extinct. We will make the following assumptions:

(A1):

\(\xi _1,\ldots ,\xi _n,\ldots \) is a sequence of i.i.d. random variables taking values in a Polish space E.

(A2):

For each i the fitness function \(f_i({\mathbf {x}},\xi )\) is continuous in \({\mathbf {x}}\), measurable in \(({\mathbf {x}},\xi )\) and strictly positive.

(A3):

If the dynamics is unbounded: There exists a function \(V:{\mathcal {S}}_+\rightarrow {\mathbb {R}}_+\) and constants \(\gamma _1,\gamma _3, C>0\) and \(\rho \in (0,1)\) such that for all \({\mathbf {x}}\in {\mathcal {S}}_+\) we have

(i):

\(V({\mathbf {x}})\ge |{\mathbf {x}}|^{\gamma _1}+1\),

(ii):

\({\mathbb {E}}\left[ V({\mathbf {x}}^\top f({\mathbf {x}},\xi _1))\ell ({\mathbf {x}},\xi _1)\right] \le \rho V({\mathbf {x}})+C\), where

$$\begin{aligned} \ell ({\mathbf {x}},\xi ):=\left( \max _{i=1}^n \left\{ \max \left\{ f_i({\mathbf {x}},\xi ), \frac{1}{f_i({\mathbf {x}},\xi )}\right\} \right\} \right) ^{\gamma _3}. \end{aligned}$$
(A4):

If the dynamics is bounded: There exists a constant \(\gamma _3>0\) such that for all \({\mathbf {x}}\in {\mathcal {S}}_+\) we have

$$\begin{aligned} {\mathbb {E}}\left[ \ell ({\mathbf {x}},\xi _1)\right] <\infty . \end{aligned}$$

Remark A.1

In particular, if one supposes the conditions

  1. 1)

    There is a compact subset \(K\subset {\mathbb {R}}_+^n\times {\mathbb {R}}^{\kappa _0}\) such that all solutions \({\mathbf {X}}_t\) satisfy \({\mathbf {X}}_t\in K\) for \(t\in {\mathbb {Z}}_+\) sufficiently large;

  2. 2)

    For all \(i=1,2,\ldots ,n \),

    $$\begin{aligned} \sup _{{\mathbf {x}},\xi }|\ln f_i({\mathbf {x}},\xi )|<\infty ; \end{aligned}$$

then assumption (A4) is satisfied.

Assumptions (A1) and (A2) ensure that \({\mathbf {X}}_t\) is a Feller process that lives on \({\mathcal {S}}_{+}\), i.e., \({\mathbf {X}}_t\in {\mathcal {S}}_+, t\in {\mathbb {Z}}_+\) whenever \({\mathbf {X}}_0\in {\mathcal {S}}_{+}\). One has to make the extra assumptions (A3) or (A4) in order to ensure the process does not blow up or fluctuate too abruptly between 0 and \(\infty \). We note that most ecological models will satisfy these assumptions. For more details see the work by Benaïm and Schreiber (2019), Hening et al. (2020).

We will follow the notation, methods and results developed by Meyn and Tweedie (1992). A point \({\mathbf {y}}\in {\mathbb {R}}_+^n\) is said to be accessible from \({\mathbf {x}}\in {\mathcal {S}}_+\) if for every neighborhood U of y, there exists \(t\ge 0\) such that \(P_t({\mathbf {x}},U)>0\). Define

$$\begin{aligned} \Gamma _{\mathbf {x}}:=\{{\mathbf {y}}\in {\mathcal {S}}_+ ~|~{\mathbf {y}}~\text {is accessible from}~{\mathbf {x}}\} \end{aligned}$$

and for \(A\subset {\mathbb {R}}_+^n\)

$$\begin{aligned} \Gamma _A=\bigcap _{{\mathbf {x}}\in A}\Gamma _{\mathbf {x}}. \end{aligned}$$

Note that \(\Gamma _A\) is the set of points which are accessible from every point of A. We say a set A is accessible if for all \({\mathbf {x}}\in {\mathbb {R}}_+^{n,\circ }\)

$$\begin{aligned} \Gamma _{{\mathbf {x}}}\cap A\ne \emptyset . \end{aligned}$$

Suppose there exist \({\mathbf {x}}^*\in \Gamma _{{\mathcal {S}}_+}\), a neighborhood U of \({\mathbf {x}}^*\), and a nonzero measure \(\phi \) on \({\mathcal {S}}_+\), such that for any \({\mathbf {x}}\in U\) there is \(m^*\in {\mathbb {Z}}_+\) such that

$$\begin{aligned} {\mathbb {P}}_ {\mathbf {x}}({\mathbf {X}}_{m^*}\in \cdot )\ge \phi (\cdot ). \end{aligned}$$

We will assume that such conditions are satisfied in our models. In many cases, it is not hard to check that these conditions hold—see Ellner (1989), Hening et al. (2020). Suppose the dynamics happens in either a compact subset of \({\mathbb {R}}_+^n\) or in \({\mathbb {R}}_+^n\). We denote the state space of the dynamics by \({\mathcal {S}}\). We define the extinction set, where at least one species is extinct, by

$$\begin{aligned} {\mathcal {S}}_0:=\{{\mathbf {x}}\in {\mathcal {S}}~:~\min _i x_i=0\}. \end{aligned}$$

For any \(\eta >0\) let

$$\begin{aligned} {\mathcal {S}}_\eta :=\{{\mathbf {x}}\in {\mathcal {S}}~:~\min _i x_i\le \eta \} \end{aligned}$$

be the subset of \({\mathcal {S}}\) where at least one species is within \(\eta \) of extinction. Denote by \({\mathcal {M}}\) the set of all ergodic invariant probability measures supported on \({\mathcal {S}}_0\) and by \({{\,\mathrm{Conv}\,}}({\mathcal {M}})\) the set of all invariant probability measures supported on \({\mathcal {S}}_0\). We say (A.2) is stochastically persistent in probability (Chesson 1982) if for all \(\varepsilon >0\) there exists \(\eta (\varepsilon )=\eta >0\) such that for all \({\mathbf {x}}\in {\mathcal {S}}_+\)

$$\begin{aligned} \liminf _{t\rightarrow \infty } {\mathbb {P}}_{{\mathbf {x}}}\{{\mathbf {X}}_t\notin S_\eta \}>1-\varepsilon . \end{aligned}$$

For any \(t\in {\mathbb {N}}\) define the normalized occupation measure

$$\begin{aligned} \Pi _t(B):=\frac{1}{t}\sum _{s=1}^t\delta _{{\mathbf {X}}(s)}(B) \end{aligned}$$

where \(\delta _{{\mathbf {X}}(s)}\) is the Dirac measure at \({\mathbf {X}}(s)\) and B is any Borel subset of \({\mathcal {S}}\). Note that \(\Pi _t\) is a random probability measure and \(\Pi _t(B)\) tells us the proportion of time the system spends in B up to time t. Denote the (random) set of \(\hbox {weak}^*\)-limit points of \((\Pi _t)_{t\in {\mathbb {N}}}\) by \({\mathcal {U}}={\mathcal {U}}(\omega )\). We say (A.2) is almost surely stochastically persistent (Schreiber 2012; Benaïm and Schreiber 2019) if for all \(\varepsilon >0\) there exists \(\eta (\varepsilon )=\eta >0\) such that for all \({\mathbf {x}}\in {\mathcal {S}}_+\)

$$\begin{aligned} \liminf _{t\rightarrow \infty } \Pi _t({\mathcal {S}}\setminus {\mathcal {S}}_\eta )>1-\varepsilon , ~{\mathbf {X}}(0)={\mathbf {x}}. \end{aligned}$$

The following general theorem gives us persistence for a general n species system.

Theorem A.5

Suppose that for all \(\mu \in {{\,\mathrm{Conv}\,}}({\mathcal {M}})\) we have

$$\begin{aligned} \max _{i} r_i(\mu )>0. \end{aligned}$$
(A.3)

Then, the system is almost surely stochastically persistent and stochastically persistent in probability. Under additional irreducibility conditions, there exists a unique invariant probability measure \(\pi \) on \({\mathcal {S}}_+\) and as \(t\rightarrow \infty \) the distribution of \({\mathbf {X}}_t\) converges in total variation to \(\pi \) whenever \({\mathbf {X}}(0)={\mathbf {x}}\in {\mathcal {S}}_+\). Furthermore, if \(w:{\mathcal {S}}_+\rightarrow {\mathbb {R}}\) is continuous and either bounded or satisfies

$$\begin{aligned} w({\mathbf {x}})\le {\mathbb {E}}\left[ V({\mathbf {x}}^\top f({\mathbf {x}},\xi _1))\ell ({\mathbf {x}},\xi _1)\right] , {\mathbf {x}}\in {\mathcal {S}}_+ \end{aligned}$$

then

$$\begin{aligned} {\mathbb {E}}w({\mathbf {X}}_t) \rightarrow \int _{{\mathcal {S}}_+} w({\mathbf {x}})\,\pi (\mathrm{d}{\mathbf {x}}). \end{aligned}$$

Sketch of proof

First, using the Markov property and Assumption A3) one can show that for all \(t\in {\mathbb {Z}}_+\) and \({\mathbf {x}}\in {\mathcal {S}}\)

$$\begin{aligned} {\mathbb {E}}_{\mathbf {x}}(V({\mathbf {X}}_t)\le \rho ^t V({\mathbf {x}})+\dfrac{C}{1-\rho }, \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} {\mathbb {E}}_{\mathbf {x}}\ell ({\mathbf {X}}_t, \xi _{t+1})&\le \rho ^{t+1} V({\mathbf {x}})+\dfrac{C}{1-\rho }. \end{aligned} \end{aligned}$$

As a next step, one can show that if a continuous function w satisfies \(\lim _{{\mathbf {x}}\rightarrow \infty } \frac{w({\mathbf {x}})}{{\mathbb {E}}[V({\mathbf {x}}^T {\mathbf {f}}({\mathbf {x}},\xi _t))\ell ({\mathbf {x}},\xi _t)]}=0\), then w is \(\mu \)-integrable for any invariant probability measure \(\mu \) of \({\mathbf {X}}\). Moreover, the strong law of large numbers for martingales will show that

$$\begin{aligned} \lim _{T\rightarrow \infty } \frac{1}{T} \sum _0^T\left( \log f_i({\mathbf {X}}_{t+1})-P \log f_i({\mathbf {X}}_t)\right) =0, \,\text {when}~ {\mathbf {X}}(0)={\mathbf {x}}\end{aligned}$$
(A.4)

where P is the transition operator of \({\mathbf {X}}_t\). This combined with arguments by Benaïm and Schreiber (2019) implies that if \(\mu ({\mathcal {S}}_+)=1\) then \(r_i(\mu )=0\) for any \(i\in I\).

The next step is to show that there exist \(M, C_2, \gamma _4>0,\rho _2\in (0,1)\) such that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {x}}\left[ V({\mathbf {X}}_1)\prod _{i=1}^n X_i^{p_i}(1)\right] \le \left( {\varvec{1}}_{\{|{\mathbf {x}}|<M\}}(C_2-\rho _2)+\rho _2\right) V({\mathbf {x}})\prod _{i=1}^nx_i^{p_i}, ~{\mathbf {x}}\in {\mathcal {S}}\end{aligned}$$

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.5)

It is shown in Schreiber et al. (2011) by the min-max principle that Assumption (A.3) 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.6)

One can then prove, using arguments by Hening and Nguyen (2018), that there exists an integer \(T^*>0\) such that, for any \(T>T^*\), \({\mathbf {x}}\in {\mathcal {S}}_0, |{\mathbf {x}}|<M\) one has

$$\begin{aligned} \sum _{t=0}^T{\mathbb {E}}_{\mathbf {x}}\left( \ln V({\mathbf {X}}_{t+1})-\ln V({\mathbf {X}}_t)-\sum p_i\ln f_i({\mathbf {X}}_t, \xi _{t+1})\right) \le -r^*(T+1). \end{aligned}$$
(A.7)

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

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

with \({\mathbf {p}}\) and \(r^*\) satisfying (A.6). Let \(n^*\in {\mathbb {N}}\) be such that

$$\begin{aligned} \rho _2^{1-n^*}>C_2. \end{aligned}$$
(A.8)

Using the previous results, as well as the analysis developed by Hening and Nguyen (2018) one can prove the following: 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 \({\mathbf {x}}\in {\mathcal {S}}_+, \Vert {\mathbf {x}}\Vert \le M\),

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

One can show that the process \((\rho ^{-t}_2 U({\mathbf {X}}(t)))_{t\ge 0}\) is a supermartingale and use this in conjunction with the Markov property to show that there exist numbers \(\kappa =\kappa (\theta ,T^*)\in (0,1)\) and \({{\tilde{K}}}={{\tilde{K}}}(\theta ,T^*)>0\) such that

$$\begin{aligned} {\mathbb {E}}_{\mathbf {x}}U^\theta ({\mathbf {X}}_{n^*T^*})\le \kappa U^\theta ({\mathbf {x}})+{{\tilde{K}}}\,\text { for all }\, {\mathbf {x}}\in {\mathcal {S}}_+. \end{aligned}$$
(A.9)

If the Markov chain \({\mathbf {X}}_t\) is irreducible and aperiodic on \({\mathcal {S}}_+\), and a compact set is petite, then one can use the well-known results by Meyn and Tweedie (1992) in conjunction with the Lyapunov condition (A.9) to conclude that there is \(c_4>1\) such that for all \({\mathbf {x}}\in {\mathcal {S}}_+\)

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

where \(\Vert \cdot \Vert _{TV}\) is the total variation distance. In particular, this implies that the distribution of \({\mathbf {X}}_t\) converges weakly to \(\pi \) as \(t\rightarrow \infty \). \(\square \)

Appendix B. Two Species Systems

In general, one needs stronger assumptions for extinction. We will assume for simplicity \(n\le 2\), so that we have one or two species. We need one more condition for extinction. This condition makes sure that the martingale part of \({\mathbf {X}}_t\) is bounded and that the family of occupation measures \((\Pi _t)_{t\in {\mathbb {Z}}_+}\) is tight.

A5) There exists a function \(\phi :{\mathcal {S}}\rightarrow (0,\infty )\) and constants \(C, \delta _\phi >0\) such that for all \({\mathbf {x}}\in {\mathcal {S}}\)

$$\begin{aligned} {\mathbb {E}}_{\mathbf {x}}V({\mathbf {X}}_1)\le V({\mathbf {x}})-\phi ({\mathbf {x}})+C \end{aligned}$$

and

$$\begin{aligned} {\mathbb {E}}_{{\mathbf {x}}}\left( V({\mathbf {X}}_1)-{\mathbb {E}}_{\mathbf {x}}V({\mathbf {X}}_1)\right) ^2+{\mathbb {E}}\left| \log f({\mathbf {x}},\xi _1)-{\mathbb {E}}\log f({\mathbf {x}},\xi _1)\right| ^2\le \delta _\phi \phi ({\mathbf {x}}). \end{aligned}$$

Define \({\mathcal {S}}^j:=\{{\mathbf {x}}\in {\mathcal {S}}~|~x_i=0, i\ne j\}\) to be the subspace supported by the species j. If we restrict the process to \({\mathcal {S}}^j\), then the extinction set is given by \({\mathcal {S}}_0:=\{0\}\) and the persistence set by \({\mathcal {S}}_+^j:={\mathcal {S}}^j\setminus \{0\}\). Let \({\mathcal {M}}^j:=\{\mu \in {\mathcal {M}}~|~\mu ({\mathcal {S}}^j)=1\}, {\mathcal {M}}^{j,+}:=\{\mu \in {\mathcal {M}}~|~\mu ({\mathcal {S}}_+^j)=1\}\) be the sets of ergodic probability measures on \({\mathcal {S}}^j\) and \({\mathcal {S}}^{j}_+\). We also assume that the subspaces \({\mathcal {S}}_0^1, {\mathcal {S}}_0^2, {\mathcal {S}}_+\) are accessible, i.e., we can get close to them from any starting point \({\mathbf {x}}\in {\mathcal {S}}_+\) with positive probability, and each subspace supports at most one ergodic probability measure. Consider two species interacting via the general system

$$\begin{aligned} \begin{aligned} X^1_{t+1}&=X^1_t f_1(X^1_t, X^2_t,\xi _{t+1}),\\ X^2_{t+1}&=X^2_t f_2(X^1_t, X^2_t,\xi _{t+1}), \end{aligned} \end{aligned}$$
(B.1)

The results by Chesson and Ellner (1989), Ellner (1989) assumed some type of monotonicity and only looked at competitive behavior. They can be generalized as follows (see Hening et al. 2020 for proofs). We first look at the Dirac delta measure \(\delta _0\) at the origin (0, 0)

$$\begin{aligned} r_i(\delta _0) = {\mathbb {E}}[\ln f_i(0,\xi _1)], i=1,2. \end{aligned}$$

If \(r_i(\delta _0)>0\), then species i survives on its own and converges to a unique invariant probability measure \(\mu _i\) supported on \({\mathcal {S}}_+^i := \{{\mathbf {x}}\in {\mathcal {S}}~|~x_i\ne 0, x_j=0, i\ne j\}\). Remember that the (random) set of \(\hbox {weak}^*\)-limit points of the family of occupation measures \((\Pi _t)_{t\in {\mathbb {N}}}\) is denoted by \({\mathcal {U}}={\mathcal {U}}(\omega )\). Thus, if we say that \({\mathcal {U}}(\omega )=\{\mu _1\}\), this means that for the realization \(\omega \) we have \(\Pi _t \rightarrow \mu _1\) weakly.

  1. (i)

    Suppose \(r_1(\delta _0)>0, r_2(\delta _0)>0\). The realized per-capita growth rates can be computed via

    $$\begin{aligned} r_i(\mu _j)=\int _{(0,\infty )}{\mathbb {E}}[\ln f_i(x,\xi _1)]\mu _j(\mathrm{d}x). \end{aligned}$$
    • If \(r_1(\mu _2)>0\) and \(r_2(\mu _1)>0\), we have coexistence and convergence of the distribution of \({\mathbf {X}}_t\) to the unique invariant probability measure \(\pi \) on \({\mathcal {S}}_+\).

    • If \(r_1(\mu _2)>0\) and \(r_2(\mu _1)<0\), we have the persistence of \(X^1\) and extinction of \(X^2\). In other words, for any \({\mathbf {x}}\in {\mathcal {S}}_+\)

      $$\begin{aligned} {\mathbb {P}}_{\mathbf {x}}\left\{ {\mathcal {U}}(\omega )=\{\mu _1\} ~\text {and}~\lim _{t\rightarrow \infty }\frac{\ln X^2_t}{t}=r_2(\mu _1)<0, \right\} =1. \end{aligned}$$
    • If \(r_1(\mu _2)<0\) and \(r_2(\mu _1)>0\), we have the persistence of \(X^2\) and extinction of \(X^1\). In other words, for any \({\mathbf {x}}\in {\mathcal {S}}_+\)

      $$\begin{aligned} {\mathbb {P}}_{\mathbf {x}}\left\{ {\mathcal {U}}(\omega )=\{\mu _2\} ~\text {and}~\lim _{t\rightarrow \infty }\frac{\ln X^1_t}{t}=r_1(\mu _2)<0, \right\} =1. \end{aligned}$$
    • If \(r_1(\mu _2)<0\) and \(r_2(\mu _1)<0\), we have that for any \({\mathbf {x}}\in {\mathcal {S}}_+\)

      $$\begin{aligned} p_{{\mathbf {x}},j}:={\mathbb {P}}_{\mathbf {x}}\left\{ {\mathcal {U}}(\omega )=\{\mu _j\} ~\text {and}~\lim _{t\rightarrow \infty }\frac{\ln X^i_t}{t}=r_i(\mu _j)<0, i\ne j \right\} \end{aligned}$$

      and

      $$\begin{aligned} p_{{\mathbf {x}},1}+ p_{{\mathbf {x}},2}=1. \end{aligned}$$
  2. (ii)

    Suppose \(r_1(\delta _0)>0, r_2(\delta _0)<0\). Then, species 1 survives on its own and converges to its unique invariant probability measure \(\mu _1\) on \({\mathcal {S}}^1_+\).

    • If \(r_2(\mu _1)>0\), we have the persistence of both species and convergence of the distribution of \({\mathbf {X}}_t\) to the unique invariant probability measure \(\pi \) on \({\mathcal {S}}_+\).

    • If \(r_2(\mu _1)<0\), we have the persistence of \(X^1\) and the extinction of \(X^2\). In other words, for any \({\mathbf {x}}\in {\mathcal {S}}_+\)

      $$\begin{aligned} {\mathbb {P}}_{\mathbf {x}}\left\{ {\mathcal {U}}(\omega )=\{\mu _1\} ~\text {and}~\lim _{t\rightarrow \infty }\frac{\ln X^2_t}{t}=r_2(\mu _1)<0, \right\} =1. \end{aligned}$$
  3. (iii)

    Suppose \(r_1(\delta _0)<0, r_2(\delta _0)<0\). Then, both species go extinct with probability one. In other words, for any \({\mathbf {x}}\in {\mathcal {S}}_+\)

    $$\begin{aligned} {\mathbb {P}}_{\mathbf {x}}\left\{ \lim _{t\rightarrow \infty }\frac{\ln X^i_t}{t}=r_i(\delta _0)<0 \right\} , i=1,2. \end{aligned}$$

We note that our results are significantly more general than those from Ellner (1989). In Ellner (1989), the author only gives conditions for coexistence and does not treat the possibility of the extinction of one or both species.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hening, A. Coexistence, Extinction, and Optimal Harvesting in Discrete-Time Stochastic Population Models. J Nonlinear Sci 31, 1 (2021). https://doi.org/10.1007/s00332-020-09667-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00332-020-09667-0

Keywords

Mathematics Subject Classification

Navigation