Skip to main content
Log in

Stochastic eco-evolutionary model of a prey-predator community

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

Abstract

We are interested in the impact of natural selection in a prey-predator community. We introduce an individual-based model of the community that takes into account both prey and predator phenotypes. Our aim is to understand the phenotypic coevolution of prey and predators. The community evolves as a multi-type birth and death process with mutations. We first consider the infinite particle approximation of the process without mutation. In this limit, the process can be approximated by a system of differential equations. We prove the existence of a unique globally asymptotically stable equilibrium under specific conditions on the interaction among prey individuals. When mutations are rare, the community evolves on the mutational scale according to a Markovian jump process. This process describes the successive equilibria of the prey-predator community and extends the polymorphic evolutionary sequence to a coevolutionary framework. We then assume that mutations have a small impact on phenotypes and consider the evolution of monomorphic prey and predator populations. The limit of small mutation steps leads to a system of two differential equations which is a version of the canonical equation of adaptive dynamics for the prey-predator coevolution. We illustrate these different limits with an example of prey-predator community that takes into account different prey defense mechanisms. We observe through simulations how these various prey strategies impact the community.

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

Similar content being viewed by others

References

  • Abrams P (1983) The theory of limiting similarity. Ann Rev Ecol Systemat, pp 359–376

  • Abrams PA (2000) The evolution of predator-prey interactions: theory and evidence. Ann Rev Ecol Systemat 31(1):79–105

    Article  MathSciNet  Google Scholar 

  • Abrams PA, Matsuda H (1997) Prey adaptation as a cause of predator-prey cycles. Evolution, pp 1742–1750

  • Adler LS, Seifert MG, Wink M, Morse GE (2012) Reliance on pollinators predicts defensive chemistry across tobacco species. Ecol Lett 15(10):1140–1148

    Article  Google Scholar 

  • Agrawal AA, Hastings AP, Johnson MTJ, Maron JL, Salminen JP (2012) Insect herbivores drive real-time ecological and evolutionary change in plant populations. Science 338(6103):113–116

    Article  Google Scholar 

  • Agren J, Schemske DW (1994) Evolution of trichome number in a naturalized population of brassica rapa. Am Nat 143:1–13

    Article  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Athreya K, Ney P (2004) Branching processes. Dover books on mathematics series. Dover Publications, Mineola

    Google Scholar 

  • Bakker ES, Ritchie ME, Olff H, Milchunas DG, Knops JM (2006) Herbivore impact on grassland plant diversity depends on habitat productivity and herbivore size. Ecol Lett 9(7):780–788

    Article  Google Scholar 

  • Baldwin IT (1998) Jasmonate-induced responses are costly but benefit plants under attack in native populations. Proc Natl Acad Sci 95(14):8113–8118

    Article  Google Scholar 

  • Becerra JX, Noge K, Venable DL (2009) Macroevolutionary chemical escalation in an ancient plant-herbivore arms race. Proc Natl Acad Sci 106(43):18062–18066

    Article  Google Scholar 

  • Bennett JA, Lamb EG, Hall JC, Cardinal-McTeague WM, Cahill JF (2013) Increased competition does not lead to increased phylogenetic overdispersion in a native grassland. Ecol Lett 16(9):1168–1176

    Article  Google Scholar 

  • Brännström Å, Johansson J, Loeuille N, Kristensen N, Troost TA, Lambers RHR, Dieckmann U (2012) Modelling the ecology and evolution of communities: a review of past achievements, current efforts, and future promises. Evol Ecol Res 14(5):601–625

    Google Scholar 

  • Brännström Å, Loeuille N, Loreau M, Dieckmann U (2011) Emergence and maintenance of biodiversity in an evolutionary food-web model. Theor Ecol 4(4):467–478

    Article  Google Scholar 

  • Burns JH, Strauss SY (2011) More closely related species are more ecologically similar in an experimental test. Proc Nat Acad Sci 108(13):5302–5307

    Article  Google Scholar 

  • Caldarelli G, Higgs PG, McKane AJ (1998) Modelling coevolution in multispecies communities. J Theor Biol 193(2):345–358

    Article  Google Scholar 

  • Champagnat N (2006) A microscopic interpretation for adaptive dynamics trait substitution sequence models. Stoch Process Appl 116(8):1127–1160

    Article  MathSciNet  MATH  Google Scholar 

  • Champagnat N, Ferrière R, Méléard S (2006) Unifying evolutionary dynamics: from individual stochastic processes to macroscopic models. Theor Popul Biol 69(3):297–321

    Article  MATH  Google Scholar 

  • Champagnat N, Jabin PE, Méléard S (2014) Adaptation in a stochastic multi-resources chemostat model. J Math Pures Appl 101(6):755–788

    Article  MathSciNet  MATH  Google Scholar 

  • Champagnat N, Méléard S (2011) Polymorphic evolution sequence and evolutionary branching. Probab Theory Relat Fields 151(1–2):45–94

    Article  MATH  Google Scholar 

  • Cottle R, Pang J, Stone R (1992) The linear complementarity problem. Classics in applied mathematics. Society for industrial and applied mathematics (SIAM, 3600 Market Street, Floor 6, Philadelphia, PA 19104)

  • Courtois EA, Baraloto C, Timothy Paine C, Petronelli P, Blandinieres PA, Stien D, Höuel E, Bessière JM, Chave J (2012) Differences in volatile terpene composition between the bark and leaves of tropical tree species. Phytochemistry 82:81–88

    Article  Google Scholar 

  • Denison RF, Kiers ET, West SA (2003) Darwinian agriculture: when can humans find solutions beyond the reach of natural selection? Quart Rev Biol 78(2):145–168

    Article  Google Scholar 

  • Dercole F, Ferriere R, Gragnani A, Rinaldi S (2006) Coevolution of slow-fast populations: evolutionary sliding, evolutionary pseudo-equilibria and complex red queen dynamics. Proc R Soc B Biol Sci 273(1589):983–990

    Article  Google Scholar 

  • Dieckmann U, Law R (1996) The dynamical theory of coevolution: a derivation from stochastic ecological processes. J Math Biol 34(5–6):579–612

    Article  MathSciNet  MATH  Google Scholar 

  • Dieckmann U, Marrow P, Law R (1995) Evolutionary cycling in predator-prey interactions: population dynamics and the red queen. J Theor Biol 176(1):91–102

    Article  Google Scholar 

  • Doebeli M, Koella JC (1995) Evolution of simple population dynamics. Proc R Soc Lon Ser B Biol Sci 260(1358):119–125

    Article  Google Scholar 

  • Drossel B, Higgs PG, McKane AJ (2001) The influence of predator-prey population dynamics on the long-term evolution of food web structure. J Theor Biol 208(1):91–107

    Article  Google Scholar 

  • Durrett R, Mayberry J (2010) Evolution in predator-prey systems. Stoch Process Appl 120(7):1364–1392

    Article  MathSciNet  MATH  Google Scholar 

  • Ehrlich PR, Raven PH (1964) Butterflies and plants: a study in coevolution. Evolution 18:586–608

    Article  Google Scholar 

  • Ethier N, Kurtz T (1986) Markov processes characterization and convergence. Wiley

  • Ferrière R, Dieckmann U, Couvet D (2004) Introduction. In: Evolutionary conservation biology, pp 1–16. Cambridge University Press, Cambridge

  • Ferriere R, Gatto M (1993) Chaotic population dynamics can result from natural selection. Proc R Soc Lond Ser B Biol Sci 251(1330):33–38

    Article  Google Scholar 

  • Fournier N, Méléard S (2004) A microscopic probabilistic description of a locally regulated population and macroscopic approximations. Ann Appl Prob 14(4):1880–1919

    Article  MATH  Google Scholar 

  • Frankham R (1996) Relationship of genetic variation to population size in wildlife. Conserv Biol 10(6):1500–1508

    Article  Google Scholar 

  • Georgelin E, Kylafis G, Loeuille N (2015) Eco-evolutionary dynamics influence the maintenance of antagonistic-mutualistic communities facing disturbances. Adv Ecol Res. doi:10.1016/bs.aecr.2015.01.005

  • Goh B (1978) Sector stability of a complex ecosystem model. Math Biosci 40(1):157–166

    Article  MathSciNet  MATH  Google Scholar 

  • Graham C, Méléard S (1997) An upper bound of large deviations for a generalized star-shaped loss network. Markov Process Relat Fields 3(2):199–223

    MATH  Google Scholar 

  • Herms DA, Mattson WJ (1992) The dilemma of plants: to grow or defend. Quart Rev Biol, pp 283–335

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

    Book  MATH  Google Scholar 

  • Illius A, Fitzgibbon C (1994) Costs of vigilance in foraging ungulates. Anim Behav 47(2):481–484

    Article  Google Scholar 

  • Ives AR, Carpenter SR (2007) Stability and diversity of ecosystems. Science 317(5834):58–62

    Article  Google Scholar 

  • Kessler A, Baldwin IT (2001) Defensive function of herbivore-induced plant volatile emissions in nature. Science 291(5511):2141–2144

    Article  Google Scholar 

  • Leimu R, Mutikainen P, Koricheva J, Fischer M (2006) How general are positive relationships between plant population size, fitness and genetic variation? J Ecol 94(5):942–952

    Article  Google Scholar 

  • Lind E, Borer E, Seabloom E, Adler P, Bakker J, Blumenthal D, Crawley M, Davies K, Firn J, Gruner D (2013) Life-history constraints in grassland plant species: a growth-defence trade-off is the norm. Ecol Lett 16(4):513–521

    Article  Google Scholar 

  • Loeuille N (2010) Influence of evolution on the stability of ecological communities. Ecol Lett 13(12):1536–1545

    Article  Google Scholar 

  • Loeuille N, Barot S, Georgelin E, Kylafis G, Lavigne C (2013) Eco-evolutionary dynamics of agricultural networks: implications for sustainable management. Ecol Netw Agric World 49:339–435

    Article  Google Scholar 

  • Loeuille N, Leibold M (2008) Ecological consequences of evolution in plant defenses in a metacommunity. Theor Populat Biol 74(1):34–45

    Article  MATH  Google Scholar 

  • Loeuille N, Loreau M (2004) Nutrient enrichment and food chains: can evolution buffer top-down control? Theor Populat Biol 65(3):285–298

    Article  MATH  Google Scholar 

  • Loeuille N, Loreau M (2005) Evolutionary emergence of size-structured food webs. Proc Natl Acad Sci USA 102(16):5761–5766

    Article  Google Scholar 

  • Loeuille N, Loreau M (2006) Evolution of body size in food webs: does the energetic equivalence rule hold? Ecol Lett 9(2):171–178

    Article  Google Scholar 

  • Loeuille N, Loreau M, Ferrière R (2002) Consequences of plant-herbivore coevolution on the dynamics and functioning of ecosystems. J Theor Biol 217(3):369–381

    Article  Google Scholar 

  • Lotka AJ (1926) Elements of physical biology. Am Math Mon 33(8):426–428

    Article  MathSciNet  Google Scholar 

  • Marrow P, Dieckmann U, Law R (1996) Evolutionary dynamics of predator-prey systems: an ecological perspective. J Math Biol 34(5–6):556–578

    Article  MATH  Google Scholar 

  • Marrow P, Law R, Cannings C (1992) The coevolution of predator-prey interactions: ESSS and red queen dynamics. Proc R Soc Lon Ser B Biol Sci 250(1328):133–141

    Article  Google Scholar 

  • Mauricio R, Rausher MD (1997) Experimental manipulation of putative selective agents provides evidence for the role of natural enemies in the evolution of plant defense. Evolution 51:1435–1444

    Article  Google Scholar 

  • May RM (2001) Stability and complexity in model ecosystems, vol 6. Princeton University Press, Princeton

    MATH  Google Scholar 

  • McCann K, Hastings A, Huxel GR (1998) Weak trophic interactions and the balance of nature. Nature 395(6704):794–798

    Article  Google Scholar 

  • Metz J, Nisbet R, Geritz S (1992) How should we define ‘fitness’ for general ecological scenarios? Trends Ecol Evolut 7(6):198–202

    Article  Google Scholar 

  • Metz JA, Geritz SA, Meszéna G, Jacobs FJ, Van Heerwaarden J (1996) Adaptive dynamics, a geometrical study of the consequences of nearly faithful reproduction. Stoch Spat Struct Dyn Syst 45:183–231

    Google Scholar 

  • Meyer JR, Ellner SP, Hairston NG, Jones LE, Yoshida T (2006) Prey evolution on the time scale of predator-prey dynamics revealed by allele-specific quantitative PCR. Proc Natl Acad Sci 103(28):10690–10695

    Article  Google Scholar 

  • Müller-Schärer H, Schaffner U, Steinger T (2004) Evolution in invasive plants: implications for biological control. Trends Ecol Evolut 19(8):417–422

    Article  Google Scholar 

  • Murray J (2002) Mathematical Biology: I. An introduction. Interdisciplinary applied mathematics. Springer, Berlin

    Google Scholar 

  • Poelman EH, van Loon JJ, Dicke M (2008) Consequences of variation in plant defense for biodiversity at higher trophic levels. Trends Plant Sci 13(10):534–541

    Article  Google Scholar 

  • Poorter H, De Jong R (1999) A comparison of specific leaf area, chemical composition and leaf construction costs of field plants from 15 habitats differing in productivity. New Phytol 143(1):163–176

    Article  Google Scholar 

  • Robinson KM, Ingvarsson PK, Jansson S, Albrectsen BR (2012) Genetic variation in functional traits influences arthropod community composition in aspen (Populus tremula l.). PLoS ONE 7(5):e37679

    Article  Google Scholar 

  • Rossberg A, Matsuda H, Amemiya T, Itoh K (2006) Food webs: experts consuming families of experts. J Theor Biol 241(3):552–563

    Article  MathSciNet  Google Scholar 

  • Soulé M (1976) Allozyme variation: its determinants in space and time. In: Molecular Evolution. Sinauer, Sunderland, pp 60–77

  • Strauss SY (1997) Floral characters link herbivores, pollinators, and plant fitness. Ecology 78(6):1640–1645

    Article  Google Scholar 

  • Strauss SY, Rudgers JA, Lau JA, Irwin RE (2002) Direct and ecological costs of resistance to herbivory. Trends Ecol Evolut 17(6):278–285

    Article  Google Scholar 

  • Takeuchi Y, Adachi N (1982) Stable equilibrium of systems of generalized volterra type. J Math Anal Appl 88(1):157–169

    Article  MathSciNet  MATH  Google Scholar 

  • Takeuchi Y, Adachi N (1983) Existence and bifurcaction of stable equilibrium in two-prey one-predator communities. Bull Math Biol 45(6):877–900

    Article  MathSciNet  MATH  Google Scholar 

  • Thébault E, Fontaine C (2010) Stability of ecological communities and the architecture of mutualistic and trophic networks. Science 329(5993):853–856

    Article  Google Scholar 

  • Thrall P, Oakeshott J, Fitt G, Southerton S, Burdon J, Sheppard A, Russell R, Zalucki M, Heino M, Ford Denison R (2011) Evolution in agriculture: the application of evolutionary approaches to the management of biotic interactions in agro-ecosystems. Evol Appl 4(2):200–215

    Article  Google Scholar 

  • Tobias JA, Cornwallis CK, Derryberry EP, Claramunt S, Brumfield RT, Seddon N (2013) Species coexistence and the dynamics of phenotypic evolution in adaptive radiation. Nature 506:359–363

    Article  Google Scholar 

  • Trussell GC, Ewanchuk PJ, Matassa CM (2006) The fear of being eaten reduces energy transfer in a simple food chain. Ecology 87(12):2979–2984

    Article  Google Scholar 

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

    Article  MATH  Google Scholar 

  • Watanabe S, Ikeda N (1981) Stochastic differential equations and diffusion processes. Elsevier, Amsterdam

    MATH  Google Scholar 

  • Yoder JB, Nuismer SL (2010) When does coevolution promote diversification? Am Nat 176(6):802–817

    Article  Google Scholar 

  • Yoshida T, Jones LE, Ellner SP, Fussmann GF, Hairston NG (2003) Rapid evolution drives ecological dynamics in a predator-prey system. Nature 424(6946):303–306

    Article  Google Scholar 

  • Zhang R, Leshak A, Shea K (2012) Decreased structural defence of an invasive thistle under warming. Plant Biol 14(1):249–252

    Google Scholar 

Download references

Acknowledgments

The authors are grateful to Frédéric Bonnans who provided insight and expertise on Linear Complementarity Problems. This article benefited from the support of the ANR MANEGE (ANR-09-BLAN-0215) and from the Chair “Modélisation Mathématique et Biodiversité” of Veolia Environnement - École Polytechnique - Museum National d’Histoire Naturelle - Fondation X.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Manon Costa.

Appendices

Appendix A: Construction of a trajectory of the prey-predator community process

We construct a trajectory of the prey-predator community process as solution of a system stochastic differential equations driven by Poisson point measures (see Fournier and Méléard 2004; Champagnat et al. 2006). We introduce two families of independent Poisson point measures on \((\mathbb {R}_+)^2\) with intensity \(dsd\theta \): \((R_j)_{1\le j\le d+m}\) for the prey and predators reproduction events and \((M_j)_{1\le j\le d+m}\) for the death events. Then, \(\forall 1\le i\le d\) and \(\forall 1\le l\le m\)

$$\begin{aligned} \begin{aligned} N^K_i(t)&=N^K_i(0) +\int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le b(x_i)N^K_i(s-)} R_i(ds,d\theta ) \\&\quad -\int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le \lambda (x_i,\mathbf {Z}(s-))N^K_i(s-) } M_i(ds, d\theta ) ,\\ H^K_l(t)&=H^K_l(0) + \int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le rH^K_l(s-)\bigl (\sum _{i=1}^d \frac{B(x_i,y_l)}{K}N^K_i(s-) \bigr )} R_{d+l}(ds,d\theta ) \\&\quad -\int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le D(y_l)H^K_l(s-) } M_{d+l}(ds, d\theta ). \end{aligned} \end{aligned}$$
(36)

Let us explain briefly these equations. We focus on the prey population \(N^K_i\) with trait \(x_i\). A trajectory is constructed using two Poisson point measures \(R_i\) and \(M_i\). The measure \(R_i\) handles the reproduction events and \(M_i\) the death events. A Poisson point measure \(R\) on \((\mathbb {R}_+)^2\) with intensity \(dsd\theta \) charges a countable set of points \(\varOmega =\{(s_u,\theta _u),u\in \mathbb {N}\}\) (with mass \(1\) on each point) (e.g. Watanabe and Ikeda 1981 Chapter I.8 for a complete definition). Then \( \int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le b(x_i)N^K_i(s-)} R_i(ds,d\theta ) \) only counts the points \((s^i_u,\theta ^i_u)_{u\in \mathbb {N}}\) such that \(s_u^i\le t\) and \(\theta ^i_u \le b(x_i)N^K_i(s^i_u-)\). Thus, we select the points of \(R_i\) which correspond to birth events of the prey population. The other integrals have similar interpretations.

The existence of solutions of (36) is justified by Proposition 2.1(i). From this construction, we deduce the expression of the prey and the predator population sizes:

$$\begin{aligned} N^K(t)= & {} N^K(0) +\sum _{i=1}^d \left[ \int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le b(x_i)N^K_i(s-)} R_i(ds,d\theta )\right. \\&\quad -\left. \int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le \lambda (x_i,\mathbf {Z}(s-))N^K_i(s-) } M_i(ds, d\theta )\right] ,\\ H^K(t)= & {} H^K(0) +\sum _{l=1}^m \Bigl [ \int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le rH^K_l(s-)\left( \sum _{i=1}^d \frac{B(x_i,y_l)}{K}N^K_i(s-) \right) } R_{d+l}(ds,d\theta ) \\&\quad -\int _0^t\int _{\mathbb {R}_+} \mathbf {1}_{\theta \le D(y_l)H^K_l(s-) } M_{d+l}(ds, d\theta )\Bigr ]. \end{aligned}$$

Appendix B: Proof of Proposition 2.1

(i) For the first part, we compare the prey population with a population evolving in the absence of predators. Let us denote by \((\widetilde{N}^K_1,\dots ,\widetilde{N}^K_d)\) the sizes of the prey sub-populations evolving without predators and set \(\widetilde{N}^K=\sum _{i=1}^d\widetilde{N}^K_i\). Using the description given in Appendix A, we construct the processes \(N^K\) and \(\widetilde{N}^K\) on the same probability space in such a way that \(\forall t\ge 0\), \(\widetilde{N}^K(t)\ge N^K(t)\) almost surely. Fournier and Méléard (2004, Theorem 5.3) and Champagnat et al. (2006, Lemma 1) established that

$$\begin{aligned} \sup _K\mathbb {E}\left( \sup _{t\in [0,T]} \left( \frac{\widetilde{N}^K(t)}{K} \right) ^3 \right) <\infty \quad \text {and }\quad \mathbb {E}\left( \sup _K \sup _{t\in [0,T]} \left( \frac{\widetilde{N}^K(t)}{K} \right) ^3 \right) <\infty . \end{aligned}$$
(37)

The process \(N^K\) then satisfies the same moment properties. To study the number of predators, we define \(\tau _n=\inf \{t\ge 0,H^K(t)\ge n\}\). By neglecting the death events, we obtain that

$$\begin{aligned}&\mathbb {E}\left( \sup _{t\in [0,T\wedge \tau _n]}\left( \frac{H^{K}(t)}{K}\right) ^3 \right) \le \mathbb {E}\left( \left( \frac{H^{K}(0)}{K}\right) ^3\right) \\&\quad +\,4\mathbb {E}\left( \int _0^{T\wedge \tau _n} \left( 1+\left( \frac{H^{K}(s)}{K}\right) ^{2}\right) \frac{H^{K}(s)}{K} \frac{N^K(s)}{K} r\bar{B}ds \right) , \end{aligned}$$

where we used that \((1+x)^3-x^3\le 4(1+x^2)\), \(\forall x\ge 0\). Since the process \(\widetilde{N}^K\) is independent of the number \(H^K\) of predators we get that

$$\begin{aligned} \begin{aligned}&\mathbb {E}\left( \sup _{t\in [0,T\wedge \tau _n]}\left( \frac{H^{K}(t)}{K}\right) ^3\right) \\&\quad \le \phi (T) +2r\bar{B}\int _0^{T} \mathbb {E}\left( \sup _{t\in [0,s\wedge \tau _n]}\left( \frac{H^{K}(t)}{K}\right) ^{3}\right) \mathbb {E}\left( \sup _{t\in [0,s]} \frac{\widetilde{N}^K(t)}{K}\right) ds,\\ \end{aligned} \end{aligned}$$

where \(\phi (T)=\mathbb {E}\left( \left( \frac{H^{K}(0)}{K}\right) ^3\right) +2r\bar{B}T\mathbb {E}\left( \sup _{t\in [0,T]}\frac{\widetilde{N}^K(t)}{K}\right) \). By Gronwall’s Lemma and (37), we obtain that

$$\begin{aligned} \mathbb {E}\left( \sup _{t\in [0,T\wedge \tau _n]}\left( \frac{H^{K}(t)}{K}\right) ^3\right) \le C(T) \end{aligned}$$
(38)

which concludes point (i) and proves the existence of \(\mathbf {Z}^K\) for all times.

(ii) The second part is much more difficult since using such a coupling is not possible: the constant \(C(T)\) obtained in (38) goes to \(\infty \) as \(T\rightarrow \infty \). In the sequel we study the behavior of the time derivative of \(\mathbb {E}\Bigl (\bigl ( \frac{N^K(t)+H^K(t)}{K} \bigr )^2\Bigr )\). We gather together the terms related to predation and bound the other terms using Assumption A to obtain

$$\begin{aligned} \begin{aligned} \frac{d}{dt}\mathbb {E}&\left( \left( \frac{N^K(t)}{K}+\frac{H^K(t)}{K} \right) ^2\right) \le \mathbb {E}\bigl (K\Psi (\mathbf {Z}^K(t))\bigr ), \end{aligned} \end{aligned}$$
(39)

where

$$\begin{aligned} \begin{aligned} \Psi (\mathbf {Z}^K)&\,= \sum _{i=1}^d\sum _{l=1}^m \frac{H^K_l}{K}\frac{N^K_i}{K}B(x_i,y_l)\\&\quad \times \left[ \left( \frac{N^K+H^K-1}{K}\right) ^2-\left( \frac{N^K+H^K}{K}\right) ^2\right. \\&\left. \qquad +\,r\left( \frac{N^K+H^K+1}{K}\right) ^2-r\left( \frac{N^K+H^K}{K}\right) ^2 \right] \\&\quad +\,\frac{N^K}{K}\bar{b} \left[ \left( \frac{N^K+H^K+1}{K}\right) ^2-\left( \frac{N^K+H^K}{K}\right) ^2\right] \\&\quad +\underline{c}\frac{ (N^K)^2 }{K^2} \left[ \left( \frac{N^K+H^K-1}{K}\right) ^2-\left( \frac{N^K+H^K}{K}\right) ^2\right] \\&\quad +\frac{H^K}{K} \underline{D}\left[ \left( \frac{N^K+H^K-1}{K} \right) ^2-\left( \frac{N^K+H^K}{K}\right) ^2\right] . \end{aligned} \end{aligned}$$
(40)

The function \(\Psi \) is the sum of three terms that we handle separately. The first term gathers together all the predation effects. The second term (sum of the second and third terms) only depends on the prey population. The last term is related to the death of predators. We start with the first term. To remove the dependence on the traits, we search for conditions on the term between square brackets to be non positive. This is equivalent to consider the sign of \((1-\frac{1}{n+h})^2-1+r(1+\frac{1}{n+h})^2-r,\) for \((n,h)\in \mathbb {N}^2{\setminus }\{(0,0)\}\). It is non positive as soon as \(n+h\ge \frac{(1+r)}{2(1-r)}=n_1\). Thus if \(N^K>n_1\),

$$\begin{aligned} \begin{aligned}&\sum _{i=1}^d\sum _{l=1}^m \frac{H^K_l}{K}\frac{N^K_i}{K}B(x_i,y_l)\left( \frac{N^K+H^K}{K}\right) ^2\\&\quad \times \left[ \left( 1-\frac{1}{N^K+H^K }\right) ^2-1+r(1+\frac{1}{N^K+H^K})^2 -r \right] \\&\le \frac{N^KH^K}{K^2}\left( \frac{N^K+H^K}{K}\right) ^2\underline{B}\left[ \left( 1-\frac{1}{N^K+H^K }\right) ^2\right. \\&\left. \qquad -1+r\left( 1+\frac{1}{N^K+H^K}\right) ^2 -r \right] , \end{aligned} \end{aligned}$$
(41)

which is non positive.

For the second term, let us remark that if \(N^K>K \frac{2\bar{b}}{\underline{c}}=Kn_2\), then

$$\begin{aligned} \begin{aligned}&N^K\bar{b}\left( \left( 1+\frac{1}{N^K+H^K}\right) ^2-1\right) +\frac{\underline{c}}{K}(N^K)^2\left( \left( 1-\frac{1}{N^K+H^K}\right) ^2-1\right) \\&\quad \le N^K\bar{b}\left[ \left( \left( 1+\frac{1}{N^K+H^K}\right) ^2-1\right) +2\left( \left( 1-\frac{1}{N^K+H^K}\right) ^2-1\right) \right] , \end{aligned} \end{aligned}$$
(42)

We set \(n_0=\max (n_1,n_2)\). If \(N^K\ge Kn_0\), we obtain by combining (41) and (42) that:

$$\begin{aligned} \begin{aligned} \Psi (\mathbf {Z}^K)&\le \frac{1}{K}\left( \frac{N^K+H^K}{K}\right) ^2 \left[ N^K\bar{b}\left[ \left( \left( 1+\frac{1}{N^K+H^K}\right) ^2-1\right) \right. \right. \\&\left. \left. \qquad +\,2\left( \left( 1-\frac{1}{N^K+H^K}\right) ^2-1\right) \right] \right. \\&\quad +\left. H^K\underline{D}\left[ \left( 1 -\frac{1}{N^K+H^K}\right) ^2-1\right] \right] . \end{aligned} \end{aligned}$$
(43)

Finally the term between square brackets in (43) is smaller than \(-\min (\bar{b},\underline{D})\), as soon as \(N^K\ge Kn_0\) for \(n_0\) large enough. Thus \(\forall t\ge 0\),

$$\begin{aligned} \begin{aligned} \frac{d}{dt}\mathbb {E}&\left( \left( \frac{N^K(t)+H^K(t)}{K}\right) ^2\right) \\&\le \mathbb {E}\left( -\min (\bar{b},\underline{D})\left( \frac{N^K(t)+H^K(t)}{K}\right) ^2\mathbf {1}_{\{N^K(t)>Kn_0\}} +K\Psi (\mathbf {Z}^K(t))\mathbf {1}_{\{N^K\le Kn_0\}} \right) . \end{aligned} \end{aligned}$$

We now consider the event \(\{N^K\le Kn_0\}\). On this event we aim at bounding from above the function \(\Psi \) with

$$\begin{aligned} \Psi (\mathbf {Z}^K)\le \frac{1}{K}\left( \frac{N^K+H^K}{K}\right) ^2\varPhi ^K\left( \frac{N^K}{K},\frac{H^K}{K}\right) . \end{aligned}$$
(44)

Since for \((n,h)\in \mathbb {N}^2{\setminus }\{(0,0)\}\),

$$\begin{aligned} \left( 1-\frac{1}{n+h}\right) ^2-1 +r\left( 1+\frac{1}{n+h}\right) ^2-r = -2(1-r)\frac{1}{n+h}+(1+r)\frac{1}{(n+h)^2}, \end{aligned}$$

and Assumption 7, we set for every \((u,v)\in (\mathbb {R}_+)^2{\setminus }\{(0,0)\}\),

$$\begin{aligned} \varPhi ^K(u,v)= & {} \frac{2u}{u+v}(\bar{b}-\underline{c}u)-\frac{2v}{u+v}((1-r)\underline{B}u+\underline{D})\nonumber \\&\quad +\,\frac{u}{K(u+v)^2}(\bar{b}+\underline{c}u+(1+r)\bar{B}v)+\underline{D}\frac{v}{K(u+v)^2}. \end{aligned}$$
(45)

We seek a condition on \(v\) to obtain that \(\varPhi ^K(u,v)\le -D\), \(\forall K\ge 0\), \(\forall 0\le u\le n_0\). This inequality can be written as a polynomial

$$\begin{aligned} v^2 \alpha (u)+ v\beta (u,K)+\gamma (u,K)\le 0, \end{aligned}$$
(46)

where the coefficients are given by

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \alpha (u)=-2(1-r)\underline{B}u-\underline{D},\\ \beta (u,K)=2u(\bar{b}-\underline{c}u)-2u^2(1-r)\underline{B}+\frac{u}{K}(1+r)\bar{B}+\frac{\underline{D}}{K},\\ \displaystyle \gamma (u,K)=\frac{u}{K}(\bar{b}+\underline{c}u +\underline{D}\displaystyle u^2+2u^2(\bar{b}-\underline{c}u)).\\ \end{array}\right. \end{aligned}$$

As \(\alpha (u)<0\), this polynomial remains negative for every \(v\) greater than its largest real root. If the polynomial (46) has real roots, then we can bound from above the largest one with

$$\begin{aligned} \frac{|\beta (u,K)|+\sqrt{\beta (u,K)^2-4\alpha (u)\gamma (u,K)}}{-2\alpha (u)}. \end{aligned}$$

The coefficient \(\beta (u,K)\) decreases with \(K\), thus for every \(K\ge 1\) and \(0\le u\le n_0\),

$$\begin{aligned} \begin{aligned}&2u(\bar{b}-\underline{c}u)-2u^2(1-r)\underline{B}\le \beta (u,K)\le 2u(\bar{b}-\underline{c}u)\\&\qquad -2u^2(1-r)\underline{B}+u(1+r)\bar{B}+\underline{D}\\&\qquad -2(1+\underline{c}) n_0^2\underline{B}\le \beta (u,K)\le 2n_0 \bar{b}+n_0(1+r)\underline{B}+\underline{D}. \end{aligned} \end{aligned}$$

In the case where \(\gamma (u,K)<0\), the discriminant \(\Delta (u,K)=\beta (u,K)^2-4\alpha (u)\gamma (u,K)\) is bounded by \(|\beta (u,K)|\). Otherwise \(\Delta \le \beta (u,K)^2+8((1-r)\underline{B}u+\underline{D})u (\bar{b}+\underline{c}u +\underline{D}u^2+2u^2(u-\underline{c}u))\) which can be bounded uniformly for \(u\in [0,n_0]\). Thus there exists \(h_0\) independent on \(K\) such that

$$\begin{aligned} \forall n\le Kn_0,\quad \forall \, h>Kh_0,\quad \varPhi ^K\left( \frac{n}{K},\frac{h}{K}\right) \le -D. \end{aligned}$$

Finally

$$\begin{aligned} \begin{aligned}&\frac{d}{dt}\mathbb {E}\left( \left( \frac{N^K(t)+H^K(t)}{K} \right) ^2\right) \le \mathbb {E}\left( K\Psi (\mathbf {Z}^K(t))\mathbf {1}_{\{N^K(t)\le Kn_0,H^K(t)\le Kh_0\}} \right) \\&\quad +\mathbb {E}\left( -C\left( \frac{N^K(t)+H^K(t)}{K}\right) ^2\left( \mathbf {1}_{\{N^K(t)>Kn_0\}}+\mathbf {1}_{\{N^K(t)\le Kn_0, H^K(t)>Kh_0\}}\right) \right) , \end{aligned} \end{aligned}$$

with \(C>0\). To conclude it remains to bound the expectation of \(\Psi \) on the event \(\{N^K\le Kn_0 \text { and }H^K\le Kh_0\}\). Keeping only the positive terms we obtain that

$$\begin{aligned} \begin{aligned}&\mathbb {E}\left( K\Psi (\mathbf {Z}^K(t))=\mathbf {1}_{\{N^K(t)\le Kn_0,H^K(t)\le Kh_0\}} \right) \,\le \mathbb {E}\left( \mathbf {1}_{\{N^K(t)\le Kn_0,H^K(t)\le Kh_0\}}\right. \\&\quad \left. \times \left( \bar{b}N^K(t)+r\bar{B} N^K(t)\frac{H^K(t)}{K}\right) \left( \left( \frac{N^K(t)+H^K(t)}{K}+\frac{1}{K} \right) ^2-\left( \frac{N^K(t)+H^K(t)}{K} \right) ^2\right) \right) \\&\quad \,\le \sum _{n=0}^{Kn_0}\sum _{h=0}^{Kh_0}\left( \frac{n+h}{K}\right) ^2\left( \bar{b}n+r\bar{B} n\frac{h}{K}\right) \left( \left( 1+\frac{1}{n+h}\right) ^2-1\right) \\&\quad \,\le \sum _{n=0}^{Kn_0}\sum _{h=0}^{Kh_0}\left( \frac{n+h}{K}\right) ^2 3\left( \bar{b}+r\bar{B} \frac{h}{K}\right) , \end{aligned} \end{aligned}$$

where the last inequality derives from \((1+u)^2-1\le 3u\), for all \(u\in [0,1]\).

Combining all these results

$$\begin{aligned} \begin{aligned} \frac{d}{dt}\mathbb {E}\left( \left( \frac{N^K(t)+H^K(t)}{K}\right) ^2\right)&\le \mathbb {E}\left( -C\left( \frac{N^K(t)+H^K(t)}{K})^2\right) \right) \\&+ \int _0^{n_0}\int _{0}^{h_0}\left( 3( n+h)^2(C+\bar{b})+ (n+h)^3r\underline{B}\right) dh dn \\&\le C^{\prime }-C\mathbb {E}\left( \left( \frac{N^K(t)+H^K(t)}{K}\right) ^2\right) , \end{aligned} \end{aligned}$$

with \(C^{\prime }>0\). We solve this inequality to get that

$$\begin{aligned} \begin{aligned} \mathbb {E}\left( \left( \frac{N^K(t)+H^K(t)}{K}\right) ^2\right)&\le C^{\prime } +\left( \mathbb {E}\left( \left( \frac{N^K(0)+H^K(0)}{K}\right) ^2\right) -C^{\prime }\right) e^{-Ct}. \end{aligned} \end{aligned}$$

which gives the uniform bound.

Appendix C: Proof of Theorem 3.2

The proof relies on the expression of Linear Complementarity Problems as variational inequality problems.

Definition 2

The variational inequality problem associated with a function \(f:\mathbb {R}^u\rightarrow \mathbb {R}^u\) and a subset \(E\subset \mathbb {R}^u\) seeks a vector \(z\in E\) such that

$$\begin{aligned} \forall a\in E,\quad (a-z)^T f(z)\ge 0. \end{aligned}$$
(47)

The existence of solutions is not true in a general setting but we are interested in a specific framework where the subset \(E\) is compact and convex.

Theorem C1

Let \(E\) be a non empty compact convex of \(\mathbb {R}^u\) and \(f\) continuous function, then the variational inequality problem associated to \((f,E)\) admits a solution.

The proof of Theorem C1 is rather classical and requires to express a solution as a fix point of a projection of the subset \(E\) (see Cottle et al. 1992, Theorem 3.7.1). With this result we can prove the Theorem 3.2.

Proof

(Proof of Theorem 3.2 ) Let us recall that a solution to the Linear complementarity problem associated to the couple \((\tilde{M},\tilde{q})\) defined in (16) is a vector \(\mathbf {z}=(\mathbf {n},\mathbf {h})\in \mathbb {R}^d\times \mathbb {R}^m\) such that: for every \(1\le i\le d\) and \(1\le l\le m\),

$$\begin{aligned} n_i\ge 0,\quad (q+M\mathbf {n}+B\mathbf {h})_i\ge 0,\quad (\mathbf {n})^T(q+M\mathbf {n}+B\mathbf {h})=0 \end{aligned}$$
(48)

and

$$\begin{aligned} h_l\ge 0,\quad (D-B^T\mathbf {n})_l\ge 0, \quad (\mathbf {h})^T(D -B^T\mathbf {n})=0 \end{aligned}$$
(49)

These conditions (48) entail that the vector \(\mathbf {n}\) is a solution to \(LCP(M,q+B\mathbf {h})\).

Note that if \(\mathbf {n}\in \mathbb {R}^d\) is solution to the restricted problem \(LCP(M,q)\) satisfying moreover \((-B^T\mathbf {n}+D)_l\ge 0\) for all \(1\le l\le m\), then the vector \((\mathbf {n},0)\) is solution to \(LCP(\tilde{M},\tilde{q})\). Similarly we seek a suitable vector \(\mathbf {n}\) and adjust it thanks to the vector \(\mathbf {h}\).

We consider the variational inequality problem associated to the set

$$\begin{aligned} E=\{\mathbf {n}\in (\mathbb {R}_+)^d,\quad \forall 1\le l\le m \quad (D-B^T\mathbf {n})_l\ge 0 \}, \end{aligned}$$

and the continuous function \(f(\mathbf {n})=q+M\mathbf {n}\).

Since \(D\) is non negative, the set \(E\) is not empty. Moreover \(E\) is convex, closed and bounded thus compact. Theorem C1 ensures the existence of a solution \(\mathbf {n}^*\) to this problem. Note that (47) can be written as

$$\begin{aligned} \forall a\in E, \quad a^Tf(\mathbf {n}^*)\ge (\mathbf {n}^*)^Tf(\mathbf {n}^*). \end{aligned}$$

Thus \(\mathbf {n}^*\) minimizes the function \(a\rightarrow a^Tf(\mathbf {n}^*)\) on \(E\). Therefore

  • either \(\mathbf {n}^*\) is in the interior of \(E\) and is therefore a global minimizer of the function \(a\rightarrow a^Tf(\mathbf {n}^*)\) on \(\mathbb {R}^d\) and \((\mathbf {n}^*,0)\) is a solution to \(LCP(\widetilde{M},\widetilde{q})\).

  • otherwise we can define the Lagrange multipliers for this problem. There exist \(d+m\) non negative real \(h_1,\ldots ,h_{d+m}\) such that \(\forall 1\le i\le d \), \(\forall 1\le l\le m \),

    $$\begin{aligned} (q+M\mathbf {n}^*)_i=h_i-\sum _{l=1}^mB_{il}h_{d+l},\quad h_in_i^*=0,\text { and }\quad h_{d+l}(-B^T \mathbf {n}^*+D)_k=0. \end{aligned}$$

    The first condition entails that \(h_i=(q+M\mathbf {n}^*)_i+\sum _{l=1}^mB_{il}h_{d+l}\) and therefore the vector \((\mathbf {n}^*,h_{d+1},\dots ,h_{d+m})\) is a solution to \(LCP(\widetilde{M},\widetilde{q})\).

Appendix D: Proof of Theorem 4.2

A perturbation \(\fancyscript{Z}^K=(\fancyscript{N}^K_1,\ldots ,\fancyscript{N}^K_d,\fancyscript{H}^K_1,\ldots ,\fancyscript{H}^K_m)\) of the prey-predator community process is defined by \(2\) families of \(d+m\) real-valued random processes \((u^K_i)_{1\le i \le d+m}\) and \((v^K_i)_{1\le i \le d+m}\) which are predictable with respect to the filtration \(\fancyscript{F}_t\) generated by the processes \(\mathbf {Z}^K\). Both families are uniformly bounded by a parameter \(\kappa >0\).

The perturbation \(\fancyscript{Z}^K\) is solution of the following system of stochastic differential equations driven by the Poisson point measures \(R_i\) and \(M_i\) introduced in Appendix A.

$$\begin{aligned} \begin{aligned} \fancyscript{Z}^K(t)&\,=\fancyscript{Z}^K(0) +\sum _{i=1}^d \left[ \int _0^t\int _{\mathbb {R}_+} \frac{e_i}{K} \mathbf {1}_{\theta \le b(x_i)\fancyscript{N}^K_i(s-)\, +\,u^K_i(s)} R_i(ds,d\theta ) \right. \\&\quad \left. -\int _0^t\int _{\mathbb {R}_+}\frac{e_i}{K} \mathbf {1}_{\theta \le \fancyscript{N}^K_i(s-))\lambda (x_,\fancyscript{Z}^K(s-)) \,+\,v^K_i(s)} M_i(ds, d\theta )\right] \\&\quad +\sum _{l=1}^m \left[ \int _0^t\int _{\mathbb {R}_+} \frac{e_{d+l}}{K} \mathbf {1}_{\theta \le r\fancyscript{H}^K_l(s-)\left( \sum _{i=1}^d \frac{B(x_i,y_l)}{K}\fancyscript{N}^K_i(s-) \right) \,+\,u^K_{d+l}(s)} R_{d+l}(ds,d\theta ) \right. \\&\quad \left. -\int _0^t\int _{\mathbb {R}_+}\frac{e_{d+l}}{K} \mathbf {1}_{\theta \le D(y_l)\fancyscript{H}^K_l(s-) \,+\,v^K_{d+l}(s)} M_{d+l}(ds, d\theta )\right] . \end{aligned} \end{aligned}$$
(50)

where \((e_1,\dots ,e_d,e_{d+1},\dots ,e_{d+m})\) is the canonical basis of \(\mathbb {R}^{d+m}\).

The proof relies on the study of the stochastic process \(L(\fancyscript{Z}^K)\) where \(L\) is the Lyapunov function for the system \(LVP(\mathbf {x},\mathbf {y})\) introduced in (11) with an appropriate choice of \(\gamma \). The function \(L\) is the sum of two functions \(V\) and \(W\). \(V\) defined in (8) is linear in the coordinate \(n_i\), \(i\in P\) and \(h_l\), \(l\in Q\) and strictly convex in the other coordinates. Moreover, its Hessian matrix at \(\mathbf {z}^*\) is diagonal. \(W\) defined (12) is a quadratic form in \((\mathbf {z}-\mathbf {z}^*)\). This justifies the inequality (20):

$$\begin{aligned} \begin{aligned} ||\mathbf {z}-\mathbf {z}^*||^2&\le \sum _{i\notin P}|n_i-n_i^*|^2 +\sum _{i\in P}|n_i|+\sum _{l\notin Q}|h_l-h_l^*| +\sum _{l\in Q}|h_l|\\&\le C\left( L(\mathbf {z})-L(\mathbf {z}^*)\right) \le CC^{\prime }\nonumber \\&\quad \,\times \left( \sum _{i\notin P}|n_i-n_i^*|^2 +\sum _{i\in P}|n_i|+\sum _{l\notin Q}|h_l-h_l^*| +\sum _{l\in Q}|h_l|\right) , \end{aligned} \end{aligned}$$

where \(P\) and \(Q\) have been defined in (6). We set in the following

$$\begin{aligned} ||\mathbf {z}-\mathbf {z}^*||_{PQ}=\sum _{i\notin P}|n_i-n_i^*|^2 +\sum _{i\in P}|n_i|+\sum _{l\notin Q}|h_l-h_l^*| +\sum _{l\in Q}|h_l| \end{aligned}$$

The derivative of \(L(\mathbf {z}(t))\) given in (13) can be bounded from above in the neighbourhood of \(\mathbf {z}^*\) by

$$\begin{aligned} \frac{d}{dt}L(\mathbf {z}(t))\le & {} -C_1||\mathbf {n}(t)-\mathbf {n}^*||^2 -C_1\left( \sum _{i\in P} n_i(t) +\sum _{l\in Q} h_l(t) \right) \\&-\,C_1 \sum _{i\notin P}\left( \sum _{l\notin Q} B_{il} (h_l(t)-h^*_l)\right) ^2, \end{aligned}$$

for a positive real number \(C_1\). If we set

$$\begin{aligned} C_2=\inf \left\{ \sum _{i\notin P}\left( \sum _{l\notin Q} B_{il} (h_l-h^*_l)\right) ^2, \mathbf {h}\in (\mathbb {R}_+)^m, ||\mathbf {h}-\mathbf {h}^*||=1\right\} >0, \end{aligned}$$

then

$$\begin{aligned} \begin{aligned} \frac{d}{dt}L(\mathbf {z}(t))&\le -C_1||\mathbf {n}(t)-\mathbf {n}^*||^2 -C_1\left( \sum _{i\in P} n_i(t) +\sum _{l\in Q} h_l(t) \right) \\&\quad -C_1C_2 \sum _{l\notin Q}(h_l(t)-h^*_l)^2. \end{aligned} \end{aligned}$$

We then obtain (21):

$$\begin{aligned} \frac{d}{dt}L(\mathbf {z}(t)) \le -C^{\prime \prime } ||\mathbf {z}-\mathbf {z}^*||^2. \end{aligned}$$

We introduce \(\tau _{\varepsilon }^K=\inf \{t\ge 0, \fancyscript{Z}^K(t)\notin B_{\varepsilon }\}\). In the sequel we prove that there exist \(\varepsilon ^{\prime \prime }<\varepsilon \) and \(V>0\) such that if \(\fancyscript{Z}^K(0)\in \fancyscript{B}{_{\varepsilon ^{\prime \prime }}}\), then

$$\begin{aligned} \lim _{K\rightarrow \infty } \mathbb {P}\Bigl ( \tau ^K_{\varepsilon }>e^{KV}\Bigr ) =1. \end{aligned}$$
(51)

For every \(t\le \tau _{\varepsilon }^K\),

$$\begin{aligned} \begin{aligned} L(\fancyscript{Z}^K(t))&=L(\fancyscript{Z}^K(0)) +M^K(t)\\&\quad +\int _0^t\sum _{i=1}^d \left( L\left( \fancyscript{Z}^K(s)+\frac{e_i}{K}\right) -L\left( \fancyscript{Z}^K(s)\right) \right) \left( \fancyscript{N}^K_i(s)b(x_i)+u^K_i(s)\right) ds \\&\quad +\int _0^t\sum _{i=1}^d\left( L\left( \fancyscript{Z}^K(s)-\frac{e_i}{K}\right) -L\left( \fancyscript{Z}^K(s)\right) \right) \left( \fancyscript{N}^K_i(s)\lambda (x_i,\fancyscript{Z}^K(s))+v^K_i(s)\right) ds\\&\quad + \int _0^t\sum _{l=1}^m\left( L\left( \fancyscript{Z}^K(s)+\frac{e_{d+l}}{K}\right) -L\left( \fancyscript{Z}^K(s)\right) \right) \\&\qquad \times \left( \fancyscript{H}^K_l(s)\left( r\sum _{i=1}^dB(x_i,y_m)\frac{\fancyscript{N}^K_i(s)}{K}\right) +u^K_{d+l}\right) ds \\&\quad + \int _0^t\sum _{l=1}^m\left( L\left( \fancyscript{Z}^K(s)-\frac{e_{d+l}}{K}\right) -L\left( \fancyscript{Z}^K(s)\right) \right) \left( \fancyscript{H}^K_l(s)D(y_l)+v^K_{d+l}\right) ds . \end{aligned} \end{aligned}$$

where \(M^K_t\) is a local martingale which can be expressed with respect to the compensated Poisson point measures \((\widetilde{R}_i)_{1\le i\le d+m}\) and \((\widetilde{M}_{i})_{1\le i\le d+ m}\):

(52)

For every \(t\le \tau _{\varepsilon }^K\) and \(1\le i\le d\) we give the second order expansion of the terms

$$\begin{aligned}&L\left( \fancyscript{Z}^K(t)+\frac{e_i}{K}\right) -L(\fancyscript{Z}^K(t))=\frac{1}{K}\frac{\partial }{\partial e_i} L(\fancyscript{Z}^K(t))\\&\quad +\frac{1}{2} \int _{0}^{\frac{1}{K}}\left( \frac{\fancyscript{N}^K_i(t)}{K}+\frac{1}{K}-u\right) \frac{\partial ^2}{\partial e_i^2} L\left( \fancyscript{Z}^K(t)-\left( \frac{\fancyscript{N}^K_i(t)}{K}-u\right) e_i\right) du. \end{aligned}$$

We obtain a similar equality for the derivative with respect to \(e_{d+l}\) for \(1\le l\le m\).

Let us remark that \(\sup \{\frac{\partial ^2}{\partial e_j^2} L(u,v) ,(u,v)\in \fancyscript{B}_{\varepsilon }\}<\infty \) for \(\varepsilon \) small enough, for all \(1\le j \le d+m\). Therefore the integrated term is of order \(1/K^2\) for large \(K\). The impact of the perturbed terms can be bounded similarly using the first derivative. Thus

$$\begin{aligned} \begin{aligned} L(\fancyscript{Z}^K(t))&= L(\fancyscript{Z}^K(0)) +M^K(t)\\&\quad +\int _0^t \sum _{i=1}^d \frac{\partial L(\fancyscript{Z}^K(s))}{\partial e_i} \frac{\fancyscript{N}^K_i(s)}{K}\\&\quad \times \left[ b(x_i)-d(x_i)-\sum _{j=1}^dc(x_i,x_j)\frac{\fancyscript{N}^K_j(s)}{K}-\sum _{l=1}^m\frac{\fancyscript{H}^K_l(s)}{K}B(x_i,y_l) \right] ds\\&\quad + \int _0^t \sum _{l=1}^m \frac{\partial L(\fancyscript{Z}^K(s))}{\partial e_{d+l}} \frac{\fancyscript{H}^K_l(s)}{K} \left[ r\sum _{i=1}^dB(x_i)\frac{\fancyscript{N}^K_i(s)}{K}-D(y_l)\right] ds \\&\quad + \fancyscript{O}\Bigl (\frac{t}{K}\Bigr )+\fancyscript{O}\bigl (\kappa t\bigr ). \end{aligned} \end{aligned}$$

Note that if \(\mathbf {z}(t)\) is a solution of \(LVP(\mathbf {x},\mathbf {y})\) then:

$$\begin{aligned} \begin{aligned} \frac{\partial L(\mathbf {z}(t))}{\partial t}&=\sum _{i=1}^d \frac{\partial }{\partial e_i} L(\mathbf {z}(t))n_i(t)\\&\quad \times \, \left[ b(x_i)-d(x_i)-\sum _{j=1}^dc(x_i,x_j)n_j(t)-\sum _{k=1}^mB(x_i,y_k)h_k(t) \right] \\&\quad + \sum _{l=1}^m \frac{\partial }{\partial e_{d+l}}L(\mathbf {z}(t)) h_l(t)\left[ r\sum _{i=1}^dB(x_i,y_l)n_i(t)-D(y_l)\right] . \end{aligned} \end{aligned}$$

We denote by \(\frac{\partial L(\fancyscript{Z}^K(t))}{\partial t}\) the derivative along the solution \(\mathbf {z}\) such that \(\mathbf {z}(t)=\fancyscript{Z}^K(t)\). Then for \(\kappa \ge 1{/}K\):

$$\begin{aligned} L(\fancyscript{Z}^K(t))=&L(\fancyscript{Z}^K(0)) +M^K(t)+ \int _0^t \frac{\partial L(\fancyscript{Z}^K(s))}{\partial t}ds +\fancyscript{O}\bigl (\kappa t\bigr ). \end{aligned}$$

Using inequalities (20) and (21) we obtain that there exists \(C^{\prime \prime \prime }>0\), such that if \( t\le T\wedge \tau _{\varepsilon }^K\) then

$$\begin{aligned} \begin{aligned} ||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2&\le C\left[ C^{\prime }\left( ||\fancyscript{Z}^K(0)-\mathbf {z}^*||_{PQ}\right) + \sup _{t\in [0,T]}|M^K(t)|\right. \\&\quad \left. -\,C^{\prime \prime }\int _0^t\left( ||\fancyscript{Z}^K(s)-\mathbf {z}^*||^2-C^{\prime \prime \prime }\kappa \right) ds\right] .\\ \end{aligned} \end{aligned}$$
(53)

This inequality is the main tool of the proof. It connects the time spent by the process above a given threshold with the values it takes during this time interval.

We define \(S_{\kappa }=\inf \{t\ge 0,||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2 \le 2C^{\prime \prime \prime }\kappa \}\). Then for every \(t\le S_{\kappa }\wedge T \wedge \tau _{\varepsilon }^K\):

$$\begin{aligned} ||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2\le C\left[ C^{\prime }\left( ||\fancyscript{Z}^K(0)-\mathbf {z}^*||_{PQ}\right) + \sup _{[0,T]}|M^K(t)|-C^{\prime \prime }C^{\prime \prime \prime }\kappa t \bigr ) \right] . \end{aligned}$$

As the l.h.s. is nonnegative we define

$$\begin{aligned} T_{\kappa }=\frac{C^{\prime }(||\fancyscript{Z}^K(0)-\mathbf {z}^*||_{PQ}+ \sup _{[0,T]}|M^K(t)|}{C^{\prime \prime }C^{\prime \prime \prime }\kappa }\ge 0, \end{aligned}$$
(54)

which can be seen as the maximal time spent by the process \(||\mathbf {Z}^K(t)-\mathbf {z}^*||^2\) above \(2C^{\prime \prime \prime }\kappa \) before the time \(T\wedge \tau _{\varepsilon }^K\). Therefore for every \(t\le S_{\kappa }\wedge T \wedge \tau _{\varepsilon }^K\):

$$\begin{aligned} ||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2\le CC^{\prime \prime }C^{\prime \prime \prime }\kappa T_{\kappa }. \end{aligned}$$

To control the norm \(||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2\) it remains to control \(T_{\kappa }\) and thus the martingale \(M^K\). To obtain the uniform bound, we use the exponential bound given by Lemma 1. On the event

$$\begin{aligned} \Bigl \{ T_{\kappa }\le T\wedge \frac{\varepsilon ^2}{2CC^{\prime \prime }C^{\prime \prime \prime }\kappa } \Bigr \}, \end{aligned}$$
(55)

then \(\sup _{[0,S_{\kappa }]}(||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2) \le \frac{\varepsilon ^2}{2},\) and in particular \(S_{\kappa }\le \tau _{\varepsilon }^K\wedge T_{\kappa }.\)

Moreover applying (53) on the same event we get

$$\begin{aligned} \sup _{[0,T\wedge \tau _{\varepsilon }^K]}(||\mathbf {Z}^K(t)-\mathbf {z}^*||^2)\le CC^{\prime \prime }C^{\prime \prime \prime }\kappa (T+T_{\kappa })\le \frac{\varepsilon ^2}{2}+CC^{\prime \prime }C^{\prime \prime \prime }\kappa T. \end{aligned}$$
(56)

Thus if furthermore \(\kappa <\varepsilon ^2{/}(2CC^{\prime \prime }C^{\prime \prime \prime }T)\) then \(\tau _{\varepsilon }^K> T\).

These results lead to the Theorem. Let \(\varepsilon ^{\prime }>0\) such that \(\varepsilon ^{\prime \prime }<\varepsilon ^{\prime }{/}2<\varepsilon ^{\prime }<\varepsilon \).

We introduce a sequence of stopping times that describes the back and forth of the process \(\fancyscript{Z}^K\) between the balls \(\fancyscript{B}_{\varepsilon ^{\prime \prime }}\) and \(\fancyscript{B}_{\varepsilon ^{\prime }/2}\) (see Fig. 6). Set \(\tau _0=0\) and for every \(k\ge 1\) such that \(\tau _k<\tau _{\varepsilon }^K\):

$$\begin{aligned} \begin{aligned} \tau ^{\prime }_k=&\inf \bigl \{t\ge \tau _{k-1}: \fancyscript{Z}^K(t)\notin B_{\varepsilon ^{\prime }/2} \bigr \},\\ \tau _k=&\inf \bigl \{t\ge \tau ^{\prime }_{k}: \fancyscript{Z}^K(t)\in B_{\varepsilon ^{\prime \prime }} \text { ou } \fancyscript{Z}^K(t)\notin B_{\varepsilon } \bigr \}. \end{aligned} \end{aligned}$$
(57)

We denote by \(k_{\varepsilon }\) the number of back and forths before the exit:

$$\begin{aligned} k_{\varepsilon }=\inf \{k\in \mathbb {N}, \tau _k=\tau _{\varepsilon }^K\}. \end{aligned}$$

In the sequel we bound \(k_{\varepsilon }\) from below.

Fig. 6
figure 6

A trajectory of \(\fancyscript{Z}^K\) in the neighbourhood of \(\mathbf {z}^*\) for \(d=m=1\)

We consider an initial condition \(\fancyscript{Z}^K(0)\in \fancyscript{B}_{\varepsilon ^{\prime }}\). We set \(\kappa =(\varepsilon ^{\prime \prime })^2/2C^{\prime \prime \prime }\) and apply the previous results. The time \(\tau _1\) corresponds to the first return in \(\fancyscript{B}_{\varepsilon ^{\prime \prime }}\) therefore it is equal to the time \(S_{\kappa }\) introduced before. We deduce from the previous computations that on the event (55)

$$\begin{aligned} \mathbb {P}\bigl ( \tau _{1}<\tau _{\varepsilon }^K\bigr )&= \mathbb {P}\bigl ( \sup _{[0,\tau _{1}]} ||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2 <\varepsilon ^2 \bigr )\ge \mathbb {P}\bigl ( T_{\kappa }\le T\wedge \frac{\varepsilon ^2}{2CC^{\prime \prime }C^{\prime \prime \prime }\kappa } \bigr ). \end{aligned}$$

We replace \(T_{\kappa }\) by its value (54) to get that

$$\begin{aligned} \begin{aligned}&\mathbb {P}\bigl ( T_{\kappa }>T\wedge \frac{\varepsilon ^2}{2CC^{\prime \prime }C^{\prime \prime \prime }\kappa } \bigr ) \\&\quad =\mathbb {P}\Bigl ( \sup _{[0,T]}|M^K(t)| >\bigl (C^{\prime \prime }C^{\prime \prime \prime }\kappa T\wedge \frac{\varepsilon ^2}{2C} \bigr )-C^{\prime }(||\fancyscript{Z}^K(0)-\mathbf {z}^*||_{PQ})\Bigr )\\&\quad \le \mathbb {P}\Bigl ( \sup _{[0,T]}|M^K(t)| >\bigl (C^{\prime \prime }C^{\prime \prime \prime }\kappa T\wedge \frac{\varepsilon ^2}{2C} \bigr )-C^{\prime }\varepsilon ^{\prime }\Bigr ), \end{aligned} \end{aligned}$$

where we used that \(\fancyscript{Z}^K(0)\in \fancyscript{B}_{\varepsilon ^{\prime }}\) to obtain the last inequality.

If we choose \(T=2C^{\prime }\varepsilon ^{\prime }/C^{\prime \prime }C^{\prime \prime \prime }\kappa \) and \(\varepsilon ^{\prime }\) such that \(2C^{\prime }\varepsilon ^{\prime }<\frac{\epsilon ^2}{2C}\) then the inequality becomes

$$\begin{aligned} \mathbb {P}\bigl ( T_{\kappa }>T\wedge \frac{\varepsilon ^2}{2CC^{\prime \prime }C^{\prime \prime \prime }\kappa } \bigr )&\le \mathbb {P}\bigl ( \sup _{[0,T]}|M^K(t)| >C^{\prime }\varepsilon ^{\prime }\bigr ). \end{aligned}$$

We finally use Lemma 1 to obtain

$$\begin{aligned} \mathbb {P}\bigl ( T_{\kappa }>T\wedge \frac{\varepsilon ^2}{2CC^{\prime \prime }C^{\prime \prime \prime }\kappa } \bigr )&\le \exp (-KV), \end{aligned}$$

where \(V>0\) only depends on \(\varepsilon ^{\prime }\) and \(\varepsilon ^{\prime \prime }\).

Since this inequality remains true as long as the initial condition is in \(B_{\varepsilon ^{\prime }}\) we deduce that

$$\begin{aligned} \sup _{\fancyscript{Z}^K(0)\in B_{\varepsilon ^{\prime }}}\mathbb {P}\Bigl ( \tau _1<\tau _{\varepsilon }^K\Bigr )\ge 1-\exp (-KV). \end{aligned}$$
(58)

Applying the strong Markov property at the stopping time \(\tau _k\) for \(k\ge 1\)

$$\begin{aligned} \sup _{\fancyscript{Z}^K(0)\in B_{\varepsilon ^{\prime }}}\mathbb {P}\Bigl ( \tau _k<\tau _{\varepsilon }^K|\tau _{k-1}<\tau _{\varepsilon }^K\Bigr )\ge 1-\exp (-KV). \end{aligned}$$

therefore we can bound \(k_{\varepsilon }\) from below by a random variable distributed according to a geometric law of parameter \(\exp (-KV)\). Then

$$\begin{aligned} \lim _{K\rightarrow \infty } \mathbb {P}(k_{\varepsilon }>\exp (\textit{KV}{/}2))=1. \end{aligned}$$
(59)

It remains to prove that these back and forths do not happen too fast. We establish that the time intervals \(\tau _k-\tau _{k-1}\) are of order \(1\) for \(k\ge 2\). To this aim we search for \(T^{\prime }\) such that for every \(k\ge 2\), \(\mathbb {P}(\tau ^{\prime }_k-\tau _{k-1}>T^{\prime })>0\). Using the strong Markov property again, it is sufficient to prove that \(\inf _{\fancyscript{Z}^K(0)\in B_{\varepsilon ^{\prime \prime }}}\mathbb {P}( \tau ^{\prime }_1>T^{\prime })>0\):

$$\begin{aligned} \inf _{\fancyscript{Z}^K(0)\in B_{\varepsilon ^{\prime \prime }}}\mathbb {P}( \tau ^{\prime }_1>T^{\prime })=\inf _{\fancyscript{Z}^K(0)\in B_{\varepsilon ^{\prime \prime }}}\mathbb {P}\left( \sup _{[0,T^{\prime }\wedge \tau ^{\prime }_1]}||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2<\frac{\varepsilon ^{\prime 2}}{4}\right) . \end{aligned}$$
(60)

We deduce from (56) with \(\varepsilon =\varepsilon ^{\prime }/2\) that on the event \(\{T_{\kappa }\le T^{\prime }\wedge \frac{\varepsilon ^{\prime 2}}{8CC^{\prime \prime }C^{\prime \prime \prime }\kappa }\}\):

$$\begin{aligned} \sup _{[0,T^{\prime }\wedge \tau ^{\prime }_1]}\left( ||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2\right) ^2\le CC^{\prime \prime }C^{\prime \prime \prime }\kappa (T^{\prime }+T_{\kappa })\le \frac{\varepsilon ^{\prime 2}}{8}+CC^{\prime \prime }C^{\prime \prime \prime }\kappa T^{\prime }. \end{aligned}$$

Setting \(T^{\prime }=2C^{\prime }\varepsilon ^{\prime \prime }/C^{\prime \prime }C^{\prime \prime \prime }\kappa \) and \(\varepsilon ^{\prime \prime }\) such that \(2C^{\prime }\varepsilon ^{\prime \prime }<\varepsilon ^{\prime 2}/4C\), we get that

$$\begin{aligned} \sup _{[0,T^{\prime }\wedge \tau ^{\prime }_1]}||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2< \frac{\varepsilon ^{\prime 2}}{4}, \end{aligned}$$

and thus \(\tau ^{\prime }_1>T^{\prime }\).

Lemma 1 ensures again that for any initial condition in \( \fancyscript{B}_{\varepsilon ^{\prime \prime }}\):

$$\begin{aligned} \begin{aligned} \mathbb {P}\left( T_{\kappa }> T^{\prime }\wedge \frac{\varepsilon ^{\prime 2}}{8CC^{\prime \prime }C^{\prime \prime \prime }\kappa } \right) \le \mathbb {P}\left( \sup _{[0,T^{\prime }]}|M^K(t)| > C^{\prime }\varepsilon ^{\prime \prime } \right) \underset{{K\rightarrow \infty }}{\longrightarrow }0. \end{aligned} \end{aligned}$$

and thus

$$\begin{aligned} \inf _{\fancyscript{Z}^K(0)\in B_{\varepsilon ^{\prime \prime }}}\mathbb {P}\left( \sup _{[0,T^{\prime }\wedge \tau ^{\prime }_1]}||\fancyscript{Z}^K(t)-\mathbf {z}^*||^2<\frac{\varepsilon ^{\prime 2}}{4}\right) \underset{{K\rightarrow \infty }}{\longrightarrow } 1. \end{aligned}$$

Finally (51) is deduced from (59).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Costa, M., Hauzy, C., Loeuille, N. et al. Stochastic eco-evolutionary model of a prey-predator community. J. Math. Biol. 72, 573–622 (2016). https://doi.org/10.1007/s00285-015-0895-y

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-015-0895-y

Keywords

Mathematics Subject Classification

Navigation