Skip to main content
Log in

A nonlocal kinetic model for predator–prey interactions

  • Published:
Swarm Intelligence Aims and scope Submit manuscript

Abstract

We extend the aggregation model from Fetecau (2011) by adding a field of vision to individuals and by including a second species. The two species, assumed to have a predator–prey relationship, have dynamics governed by nonlocal kinetic equations that include advection and turning. The latter is the main mechanism for aggregation and orientation, which results from interactions among individuals of the same species as well as predator–prey relationships. We illustrate numerically a diverse set of predator–prey behaviors that can be captured by this model. We show that a prey’s escape outcome depends on the social interactions between its group members, the prey’s field of vision and the sophistication of the predator’s hunting strategies.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Notes

  1. Superscripts d and o in the interaction kernels refer to distance and orientation (angle), respectively. Subscripts r,al,a stand for repulsion, alignment and attraction, respectively.

References

  • Albi, G., & Pareschi, L. (2013). Binary interaction algorithms for the simulation of flocking and swarming dynamics. Multiscale Modeling & Simulation. SIAM Interdisciplinary Journal, 11(1), 1–29.

    Article  MathSciNet  MATH  Google Scholar 

  • Balch, T., & Arkin, R. C. (1998). Behavior-based formation control for multi-robot teams. IEEE Transactions on Robotics and Automation, 14(6), 926–939.

    Article  Google Scholar 

  • Barbaro, A. B. T., Taylor, K., Trethewey, P. F., Youseff, L., & Birnir, B. (2009). Discrete and continuous models of the dynamics of pelagic fish: application to the capelin. Mathematics and Computers in Simulation, 79(12), 3397–3414.

    Article  MathSciNet  MATH  Google Scholar 

  • Barbera, E., Curró, C., & Valentia, G. (2010). Wave features of a hyperbolic prey–predator model. Mathematical Methods in the Applied Sciences, 33, 1504–1515.

    MathSciNet  MATH  Google Scholar 

  • Bertin, E., Droz, M., & Grégoire, G. (2009). Hydrodynamic equations for self-propelled particles: microscopic derivation and stability analysis. Journal of Physics A: Mathematical and Theoretical, 42, 445001.

    Article  Google Scholar 

  • Bertozzi, A. L., Carrillo, J. A., & Laurent, T. (2009). Blow-up in multidimensional aggregation equations with mildly singular interaction kernels. Nonlinearity, 22(3), 683–710.

    Article  MathSciNet  MATH  Google Scholar 

  • Brauer, F., & Castillo-Chavez, C. (2001). Texts in applied mathematics: Vol. 40. Mathematical models in population biology and epidemiology. New York: Springer.

    Book  MATH  Google Scholar 

  • Camazine, S., Deneubourg, J.-L., Franks, N. R., Sneyd, J., Theraulaz, G., & Bonabeau, E. (2001). Princeton studies in complexity. Self-organization in biological systems. Princeton: Princeton University Press.

    Google Scholar 

  • Cantrell, R. S., & Cosner, C. (1996). Models for predator–prey systems at multiple scales. SIAM Review, 38(2), 256–286.

    Article  MathSciNet  MATH  Google Scholar 

  • Canuto, C., Hussaini, M. Y., Quarteroni, A., & Zang, T. A. (1988). Springer series in computational physics. Spectral methods in fluid dynamics. New York: Springer.

    Book  MATH  Google Scholar 

  • Carey, G., Teal, J. M., Kanwisher, J. W., Lawson, K. D., & Beckett, J. S. (1971). Warm-bodied fish. American Zoologist, 11, 135–143.

    Google Scholar 

  • Carrillo, J. A., Fornasier, M., Toscani, G., & Vecil, F. (2010). Particle, kinetic, and hydrodynamic models of swarming. In G. Naldi, L. Pareschi, & G. Toscani (Eds.), Mathematical modeling of collective behavior in socio-economic and life sciences (pp. 297–336). New York: Birkhäuser.

    Chapter  Google Scholar 

  • Cook, D. J., Gmytrasiewicz, P., & Holder, L. B. (1996). Decision-theoretic cooperative sensor planning. IEEE Transactions on Pattern Analysis and Machine Intelligence, 18(10), 1013–1023.

    Article  Google Scholar 

  • Couzin, I. D., Krause, J., James, R., Ruxton, G., & Franks, N. R. (2002). Collective memory and spatial sorting in animal groups. Journal of Theoretical Biology, 218, 1–11.

    Article  MathSciNet  Google Scholar 

  • Cucker, F., & Smale, S. (2007). Emergent behavior in flocks. IEEE Transactions on Automatic Control, 52(5), 852–862.

    Article  MathSciNet  Google Scholar 

  • Degond, P., & Motsch, S. (2008). Continuum limit of self-driven particles with orientation interaction. Mathematical Models and Methods in Applied Sciences, 18 (suppl.), 1193–1215.

    Article  MathSciNet  Google Scholar 

  • D’Orsogna, M. R., Chuang, Y.-L., Bertozzi, A. L., & Chayes, L. S. (2006). Self-propelled particles with soft–core interactions: patterns, stability and collapse. Physical Review Letters, 96(10), 104302.

    Article  Google Scholar 

  • Eftimie, R., de Vries, G., & Lewis, M. A. (2007b). Complex spatial group patterns result from different animal communication mechanisms. Proceedings of the National Academy of Sciences of the United States of America, 104(17), 6974–6979.

    Article  MathSciNet  MATH  Google Scholar 

  • Eftimie, R., de Vries, G., Lewis, M. A., & Lutscher, F. (2007a). Modeling group formation and activity patterns in self-organizing collectives of individuals. Bulletin of Mathematical Biology, 69(5), 1537–1566.

    Article  MathSciNet  MATH  Google Scholar 

  • Eftimie, R., de Vries, G., & Lewis, M. (2009). Weakly nonlinear analysis of a hyperbolic model for animal group formation. Journal of Mathematical Biology, 59, 37–74.

    Article  MathSciNet  MATH  Google Scholar 

  • Feder, T. (2007). Statistical physics is for the birds. Physics Today, 60(10), 28–30.

    Article  Google Scholar 

  • Fetecau, R. (2011). Collective behavior of biological aggregations in two dimensions: a nonlocal kinetic model. Mathematical Models and Methods in Applied Sciences, 21(7), 1539–1569.

    Article  MathSciNet  MATH  Google Scholar 

  • Grobecker, D. B., & Pietsch, T. W. (1979). High-speed cinematographic evidence for ultrafast feeding in antennariid anglerfishes. Am. Ass. Adv. Sci., 205(4411), 1161–1162.

    Google Scholar 

  • Ha, S.-Y., & Tadmor, E. (2008). From particle to kinetic and hydrodynamic descriptions of flocking. Kinetic and Related Models, 1(3), 415–435.

    Article  MathSciNet  MATH  Google Scholar 

  • Hadeler, K. (1999). Reaction transport systems in biological modeling. In Lecture notes in mathematics. Mathematics inspired by biology (pp. 95–150). New York: Springer.

    Chapter  Google Scholar 

  • Hamilton, W. D. (1971). Geometry for the selfish herd. Journal of Theoretical Biology, 31, 295–311.

    Article  Google Scholar 

  • Hayward, M. W., O’Brien, J., Hofmeyr, M., & Kerley, G. I. H. (2006). Prey preferences of the African wild dog Lycaon pictus (Canidae: Carnivora): ecological requirements for conservation. Journal of Mammalogy, 87, 1122–1131.

    Article  Google Scholar 

  • Hedenström, A., & Rosén, M. (2001). Predator versus prey: on aerial hunting and escape strategies in birds. Int. Soc. Behav. Ecology, 12(2), 150–156.

    Article  Google Scholar 

  • Kamimura, A., & Ohira, T. (2010). Group chase and escape. New Journal of Physics, 12, 053013.

    Article  Google Scholar 

  • Kunz, H., & Hemelrijk, C. K. (2012). Simulations of the social organization of large schools of fish whose perception is obstructed. Applied Animal Behaviour Science, 138, 142–151.

    Article  Google Scholar 

  • Kuwamura, M., Nakazawa, T., & Ogawa, T. (2009). A minimum model of prey–predator system with dormancy of predators and the paradox of enrichment. Journal of Mathematical Biology, 58(3), 459–479.

    Article  MathSciNet  Google Scholar 

  • Liu, S., & Beretta, E. (2006). A stage-structured predator–prey model of Beddington–DeAngelis type. SIAM Journal on Applied Mathematics, 66(4), 1101–1129.

    Article  MathSciNet  MATH  Google Scholar 

  • Magurran, A. E. (1990). The adaptive significance of schooling as an anti-predator defence in fish. Annales Zoologici Fennici, 27, 51–66.

    Google Scholar 

  • Mogilner, A., & Edelstein-Keshet, L. (1999). A non-local model for a swarm. Journal of Mathematical Biology, 38, 534–570.

    Article  MathSciNet  MATH  Google Scholar 

  • Murdoch, W. W. (1969). Switching in general predators: experiments on predator specificity and stability of prey. Ecological Monographs, 39(4), 335–354.

    Article  Google Scholar 

  • Myerscough, M. R., Gray, B. F., Hogarth, W. L., & Norbury, J. (1992). An analysis of an ordinary differential equation model for a two-species predator–prey system with harvesting and stocking. Journal of Mathematical Biology, 30(6), 389–411.

    MathSciNet  MATH  Google Scholar 

  • Naldi, G., Pareschi, L., & Toscani, G. (Eds.) (2010). Mathematical modeling of collective behavior in socio-economic and life sciences, modeling and simulation in science, engineering and technology. New York: Birkhäuser.

    Google Scholar 

  • Newman, J. P., & Sayama, H. (2008). Effect of sensory blind zones on milling behavior in a dynamic self-propelled particle model. Phys. Rev. E, 78, 011913.

    Article  Google Scholar 

  • Niizeki, S. (1989). Existence and stability of traveling wave solutions of competition models with migration effects. Japan Journal of Applied Mathematics, 6(3), 421–478.

    Article  MathSciNet  MATH  Google Scholar 

  • Okubo, A. (1980). Diffusion and ecological problems, mathematical models. New York: Springer.

    MATH  Google Scholar 

  • Othmer, H. G., Dunbar, S. R., & Alt, W. (1988). Models of dispersal in biological systems. Journal of Mathematical Biology, 26, 263–298.

    Article  MathSciNet  MATH  Google Scholar 

  • Parrish, J. K., & Keshet, L. E. (1999). Complexity, pattern, and evolutionary trade-offs in animal aggregation. Science, 284, 99–101.

    Article  Google Scholar 

  • Partridge, B. L. (1982). Structure and function of fish schools. Scientific American, 246(6), 114–123.

    Article  Google Scholar 

  • Pitcher, T. J., & Parrish, J. K. (1993). Functions of shoaling behavior in teleosts. London: Chapman & Hall.

    Book  Google Scholar 

  • Robinson, K. P., & Tetley, M. J. (2007). Behavioural observations of foraging minke whales (Balaenoptera acutorostrata) in the outer moray firth, North-East Scotland. Journal of the Marine Biological Association of the UK, 87, 85–86.

    Article  Google Scholar 

  • Romanczuk, P., & Schimansky-Geier, L. (2012). Swarming and pattern formation due to selective attraction and repulsion. Interface Focus, 2(6), 746–756.

    Article  Google Scholar 

  • Stander, P. (1992). Cooperative hunting in lions: the role of the individual. Behavioral Ecology and Sociobiology, 29(6), 445–454.

    Article  Google Scholar 

  • Topaz, C. M., Bertozzi, A. L., & Lewis, M. A. (2006). A nonlocal continuum model for biological aggregation. Bulletin of Mathematical Biology, 68, 1601–1623.

    Article  MathSciNet  Google Scholar 

  • Uvarov, B. (1966). Grasshoppers and locusts. London: Centre for Overseas Pest Research.

    Google Scholar 

  • Vaughn, R. L., Muzi, E., Richardson, J., & Würsig, B. (2011). Dolphin bait-balling behaviors in relation to prey ball escape behaviors. Journal of Ethology, 117, 859–871.

    Article  Google Scholar 

  • White, K. A. J., Murray, J. D., & Lewis, M. A. (1996). Wolf–deer interactions: a mathematical model. Proceedings of the Royal Society of London. Series B, Biological Sciences, 263, 299–305.

    Article  Google Scholar 

Download references

Acknowledgements

R.F. was supported by NSERC Discovery Grant PIN-341834. Both R.F. and J.M. want to thank John Stockie for various suggestions and stimulating discussions during this research.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to R. C. Fetecau.

Appendix: Modelling the probability functions in equations (13)–(15)

Appendix: Modelling the probability functions in equations (13)–(15)

We first describe the modeling of w a . Since w a is a probability function, we have

$$ \int_{-\pi}^{\pi} w_a \bigl( \phi' -\phi, \phi'-\psi \bigr) \,d\phi= 1. $$
(35)

The probability function w a is modeled by

$$ w_a \bigl( \phi' -\phi, \phi'-\psi \bigr) = g_{\sigma_a} \bigl( \phi' -\phi- v_a \bigl( \phi'-\psi \bigr) \bigr), $$
(36)

where \(g_{\sigma_{a}}\) is an approximation of the delta function with width σ a , and v a is a turning function. The decision making individual can turn to any direction within a specific range. This range is centered around the direction

$$\phi= \phi' - v_a \bigl( \phi'-\psi \bigr), $$

which happens when the argument of the function \(g_{\sigma_{a}}\) is zero. The parameter σ a >0 measures the width of the turning range the decision making individual will move into—see Fig. 3(b). The smaller the σ a , the more accurate the turning. If σ a is large, then the range is wide and the decision making individual can move anywhere within the range.

The function to describe \(g_{\sigma_{a}}\) is taken to be

$$ g_{\sigma_a} ( \eta ) = \frac{1}{\sqrt{\pi}\sigma_a} \sum _{z \in\mathbb{Z}}e^{- (\frac{\eta+ 2\pi z}{\sigma_a} )^2}, $$
(37)

which is a periodic Gaussian with extra contributions from full rotations.

The turning function v a is modeled as

$$ v_a ( \eta ) = k_a \sin\eta, $$
(38)

where k a is a constant between 0 and 1 that describes how much the decision making individual will turn due to attraction (see Fig. 3(b)).

An alternative choice is v a (η)=k a η, which may be more biologically realistic. However the choice (38) for v a is made because it is periodic and works well with the fast Fourier transform, which the numerics in this paper is based on, as discussed in Sect. 3.

The probability functions w r and w al are defined through the same steps as equations (36)–(38), using approximations \(g_{\sigma_{r}}\), \(g_{\sigma_{\mathit{al}}}\) to the delta function, with widths σ r , σ al , turning functions v r , v al , and turning strengths k r , k al , respectively.

There is a major difference between w a and w r which has to be pointed out. Namely, in the definition of w r , k r must be between −1 and 0 instead of between 0 and 1, as is k a . This is the only place where the negative factor comes in to enforce the negative behavior of the repulsion interaction. The attraction and alignment interactions are positive because they respond positively to the surrounding individuals.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Fetecau, R.C., Meskas, J. A nonlocal kinetic model for predator–prey interactions. Swarm Intell 7, 279–305 (2013). https://doi.org/10.1007/s11721-013-0084-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11721-013-0084-9

Keywords

Navigation