Skip to main content
Log in

Influence of isolation degree of spatial patterns on persistence of populations

  • Original Paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

Spatial patterns are ubiquitous in nature, which have been identified as important factors in dynamics of ecosystems. However, how pattern structures have influence on persistence of populations is far from well being understood. Particularly, whether some characters of spatial pattern can be indicators for ecosystems collapse is not well studied. As a result, we presented a predator–prey system with spatial motion and found that isolation degree (average distance between patterns with high density) of spatial patterns plays an important role in the persistence of populations: If isolation degree is much smaller, then the population will persist; if isolation degree is much larger, then the population density will decrease with increasing space size and run a high risk of extinction as space size is large enough. Our results highlight the relationship between pattern structures and ecosystems collapse.

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

Notes

  1. Since this critical value depends on the parameters sets and the lattice size, it can not be expressed mathematically.

References

  1. Murray, J.D.: Mathematical Biology II: Spatial Models and Biomedical Applications. Springer, New York (2003)

    MATH  Google Scholar 

  2. Nowak, M.A., May, R.M.: Evolutionary games and spatial chaos. J. Anim. Ecol. 359, 826–829 (1992)

    Google Scholar 

  3. Segel, L., Jackson, J.: Dissipative structure: an explanation and an ecological example. J. Theor. Biol. 37, 545–559 (1972)

    Article  Google Scholar 

  4. Comins, H.N., Hassell, M.P., May, R.M.: The spatial dynamics of host-parasitoid systems. J. Anim. Ecol. 61, 735–748 (1992)

    Article  Google Scholar 

  5. Turing, A.M.: The chemical basis of morphogenesis. Philos. Trans. R. Soc. Lond. B 237, 37–72 (1952)

    Article  Google Scholar 

  6. Medvinsky, A.B., Petrovskii, S.V., Tikhonova, I.A., Malchow, H., Li, B.-L.: Spatio-temporal complexity of plankton and fish dynamics in simple model ecosystems. SIAM Rev. 44, 311–370 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  7. Sherratt, J.A.: Unstable wavetrains and chaotic wakes in reaction–diffusion systems of lambda-omega type. Phys. D 82, 165–179 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  8. Sherratt, J.A., Lambin, X., Sherratt, T.N.: The effects of the size and shape of landscape features on the formation of traveling waves in cyclic populations. Am. Nat. 162, 503–513 (2003)

    Article  Google Scholar 

  9. Sherratt, J.A., Lambin, X., Thomas, C.J., Sherratt, T.N.: Generation of periodic waves by landscape features in cyclic predator–prey systems. Proc. R. Soc. Lond. B 269, 327–334 (2002)

    Article  Google Scholar 

  10. Chen, J.-X., Peng, L., Zheng, Q., Zhao, Y.-H., Ying, H.-P.: Influences of periodic mechanical deformation on pinned spiral waves. Chaos 24, 033103 (2014)

    Article  MathSciNet  Google Scholar 

  11. Chen, J.-X., Peng, L., Ma, J., Ying, H.-P.: Liberation of a pinned spiral wave by a rotating electric pulse. EPL 107, 38001 (2014)

    Article  Google Scholar 

  12. Lou, Q., Chen, J.-X., Zhao, Y.-H., Shen, F.-R., Fu, Y., Wang, L.-L., Liu, Y.: Control of turbulence in heterogeneous excitable media. Phys. Rev. E 85, 026213 (2012)

    Article  Google Scholar 

  13. Petrovskii, S., Li, B.L., Malchow, H.: Quantification of the spatial aspect of chaotic dynamics in biological and chemical systems. Bull. Math. Biol. 65, 425–446 (2003)

    Article  Google Scholar 

  14. Morozov, A., Petrovskii, S., Li, B.L.: Bifurcations and chaos in a predator–prey system with the Allee effect. Proc. R. Soc. Lond. B 271, 1407–1414 (2004)

    Article  Google Scholar 

  15. Alonso, D., Bartumeus, F., Catalan, J.: Mutual interference between predators can give rise to Turing spatial patterns. Ecology 83, 28–34 (2002)

    Article  Google Scholar 

  16. Baurmann, M., Gross, T., Feudel, U.: Instabilities in spatially extended predator-prey systems: spatiotemporal patterns in the neighborhood of Turing–Hopf bifurcations. J. Theor. Biol. 245, 220–229 (2007)

    Article  MathSciNet  Google Scholar 

  17. Garvie, M.R.: Finite-difference schemes for reaction–diffusion equations modeling predator–prey interactions in Matlab. Bull. Math. Biol. 69, 931–956 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  18. Gurney, W.S.C., Veitch, A.R., Cruickshank, I., Mcgeachin, G.: Circles and spirals: population persistence in a spatially explicit predator–prey model. Ecology 79, 2516–2530 (1998)

    Google Scholar 

  19. Hengeveld, R.: Dynamics of Biological Invasions. Chapman and Hall, London (1989)

    Google Scholar 

  20. Li, Z.-Z., Gao, M., Hui, C., Han, X.-Z., Shi, H.: Impact of predator pursuit and prey evasion on synchrony and spatial patterns in metapopulation. Ecol. Model. 185, 245–254 (2005)

    Article  Google Scholar 

  21. Malchow, H.: Motional instabilities in predator–prey systems. J. Theor. Biol. 204, 639–647 (2000)

    Article  Google Scholar 

  22. Wu, X., Ma, J., Yuan, L., Liu, Y.: Simulating electric activities of neurons by using PSPICE. Nonlinear Dyn. 75, 113–126 (2013)

    Article  MathSciNet  Google Scholar 

  23. Ma, J., Liu, Q., Ying, H., Wu, Y.: Emergence of spiral wave induced by defects block. Commun. Nonlinear Sci. Numer. Simul. 18, 1665–1675 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  24. Liu, T.-B., Ma, J., Zhao, Q., Tang, J.: Force exerted on the spiral tip by the heterogeneity in an excitable medium. EPL 104, 58005 (2013)

    Article  Google Scholar 

  25. Ma, J., Wu, X., Chu, R., Zhang, L.: Selection of multi-scroll attractors in Jerk circuits and their verification using Pspice. Nonlinear Dyn. 76, 1951–1962 (2014)

    Article  Google Scholar 

  26. Ma, J., Wang, C.N., Jin, W.Y., Wu, Y.: Transition from spiral wave to target wave and other coherent structures in the networks of Hodgkin–Huxley neurons. Appl. Math. Comput. 217, 3844–3852 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  27. Panfilov, A.V., Muller, S.C., Zykov, V.S., Keener, J.P.: Elimination of spiral waves in cardiac tissue by multiple electrical shocks. Phys. Rev. E 21, 4644–4647 (2000)

    Article  Google Scholar 

  28. Fenton, F.H., Cherry, E.M., Hastings, H.M., Evans, S.J.: Multiple mechanisms of spiral wave breakup in a model of cardiac electrical activity. Chaos 12, 852 (2002)

    Article  Google Scholar 

  29. Quail, T., Shrier, A., Glass, L.: Spatial symmetry breaking determines spiral wave chirality. Phys. Rev. Lett. 113, 158101 (2014)

    Article  Google Scholar 

  30. Hou, Z., Yang, L., Zuo, X., Xin, H.: Noise induced pattern transition and spatiotemporal stochastic resonance. Phys. Rev. Lett. 81, 2854–2857 (1998)

    Article  Google Scholar 

  31. Lindner, B., Garcia-Ojalvo, J., Neiman, A., Schimansky-Geier, L.: Effects of noise in excitable systems. Phys. Rep. 392, 321–424 (2004)

    Article  Google Scholar 

  32. Ciarletta, P., Balbi, V., Kuhl, E.: Pattern selection in growing tubular tissues. Phys. Rev. Lett. 113, 248101 (2014)

    Article  Google Scholar 

  33. Fernandes, L.D., de Aguiar, M.A.M.: Turing patterns and apparent competition in predator–prey food webs on networks. Phys. Rev. E 86, 056203 (2012)

    Article  Google Scholar 

  34. Rodrigues, L.A.D., Mistro, D.C., Petrovskii, S.: Pattern formation, long-term transients, and the Turing–Hopf bifurcation in a space-and time-discrete predator–prey system. Bull. Math. Biol. 73, 1812–1840 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  35. Sun, G.-Q., Zhang, G., Jin, Z., Li, L.: Predator cannibalism can give rise to regular spatial pattern in a predator–prey system. Nonlinear Dyn. 58, 75–84 (2009)

    Article  MATH  Google Scholar 

  36. Sun, G.-Q., Jin, Z., Li, L., Haque, M., Li, B.-L.: Spatial patterns of a predator–prey model with cross diffusion. Nonlinear Dyn. 69, 1631–1638 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  37. Sun, G.-Q., Jin, Z., Liu, Q.-X., Li, L.: Dynamical complexity of a spatial predator–prey model with migration. Ecol. Model. 219, 248–255 (2009)

    Article  Google Scholar 

  38. Zhang, L., Thygesen, U., Banerjee, M.: Size-dependent diffusion promotes the emergence of spatiotemporal patterns. Phys. Rev. E 90, 012904 (2014)

    Article  Google Scholar 

  39. Kefi, S., Rietkerk, M., Alados, C.L., Pueyo, Y., Papanastasis, V.P., ElAich, A., de Ruiter, P.C.: Spatial vegetation patterns and imminent desertification in Mediterranean arid ecosystems. Nature 449, 213–217 (2007)

    Article  Google Scholar 

  40. Scheffer, M., Bascompte, J., Brock, W.A., Brovkin, V., Carpenter, S.R., Dakos, V., Held, H., van Nes, E.H., Rietkerk, M., Sugihara, G.: Early-warning signals for critical transitions. Nature 461, 53–59 (2009)

    Article  Google Scholar 

  41. Wang, M.X.: Stationary patterns for a prey–predator model with prey-dependent and ratio-dependent functional responses and diffusion. Phys. D 196, 172–192 (2004)

  42. Peng, R., Shi, J.P., Wang, M.X.: Stationary pattern of a ratio-dependent food chain model with diffusion. SIAM J. Appl. Math. 67, 1479–1503 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  43. Yochelis, A., Tintut, Y., Demer, L.L., Garfinkel, A.: The formation of labyrinths, spots and stripe patterns in a biochemical approach to cardiovascular calcification. New J. Phys. 10, 055002 (2008)

    Article  Google Scholar 

  44. Rietkerk, M., van de Koppel, J.: The formation of labyrinths, spots and stripe patterns in a biochemical approach to cardiovascular calcification. Trends Ecol. Evol. 23, 169–175 (2008)

    Article  Google Scholar 

  45. Cross, M.C., Hohenberg, P.C.: Pattern formation outside of equilibrium. Rev. Mod. Phys. 65, 851–1112 (1993)

    Article  Google Scholar 

Download references

Acknowledgments

The project is funded by the National Natural Science Foundation of China under Grants (11171314, 11331009, 31500443, 11501338 and 11301490), and International Exchange Program of Postdoctor in Fudan University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ze-Yan Wu.

Appendices

Appendix 1: Derivation of Laplacian operator

We want to demonstrate that if an individual moves randomly in the space, then Laplacian operator can describe such motion. We firstly show the derivation of Laplacian operator in one-dimensional space. Random walk assumes that an individual takes steps of length \(\triangle x\) to its left or right side along a line, and after each \(\triangle t\) time unit, the individual will take one step. Initially, the individual locates at \(x_0\) at initial time \(t_0\), then after one step, the individual will locate at either \(x_0-\triangle x\) or \(x_0+\triangle x\) at time \(t_0+\triangle t\) with the same probability 1 / 2 (see Fig. 6). Denote P(tx) as the number (or density) of individuals at time t and location x. Then, we have

$$\begin{aligned} P(t+\triangle t, x)=\frac{1}{2}P(t, x-\triangle x)+\frac{1}{2}P(t, x+\triangle x).\nonumber \\ \end{aligned}$$
(5.1)
Fig. 6
figure 6

An individual moves randomly in the one-dimensional space. Space step is \(\triangle x\) and time step is \(\triangle t\)

Taylor expansion gives rise to:

$$\begin{aligned} P(t+\triangle t, x)= & {} P(t,x)+\frac{\partial P}{\partial t}(t,x)\triangle t\nonumber \\&+\,\frac{1}{2}\frac{\partial ^{2} P}{\partial t^{2}}(t,x)(\triangle t)^2+\cdots , \end{aligned}$$
(5.2a)
$$\begin{aligned} P(t, x-\triangle x)= & {} P(t,x)+\frac{\partial P}{\partial x}(t,x)(-\triangle x)\nonumber \\&+\,\frac{1}{2}\frac{\partial ^{2} P}{\partial x^{2}}(t,x)(-\triangle x)^2+\cdots , \nonumber \\\end{aligned}$$
(5.2b)
$$\begin{aligned} P(t, x+\triangle x)= & {} P(t,x)+\frac{\partial P}{\partial x}(t,x)(\triangle x)\nonumber \\&+\,\frac{1}{2}\frac{\partial ^{2} P}{\partial x^{2}}(t,x)(\triangle x)^2+\cdots , \end{aligned}$$
(5.2c)

which induces that:

$$\begin{aligned} \frac{\partial P}{\partial t}(t,x)\triangle t+\cdots =\frac{1}{2}\frac{\partial ^{2} P}{\partial x^{2}}(t,x)(\triangle x)^2+\cdots .\nonumber \\ \end{aligned}$$
(5.3)

Since \(\triangle t\) and \(\triangle x\) are small enough, the higher-order terms are very close to zero. As a result, one can have:

$$\begin{aligned} \frac{\partial P}{\partial t}(t,x)\triangle t=\frac{1}{2}\frac{\partial ^{2} P}{\partial x^{2}}(t,x)(\triangle x)^2. \end{aligned}$$
(5.4)

With the assumption that:

$$\begin{aligned} \triangle t\rightarrow 0,\quad \triangle x\rightarrow 0, \quad \frac{(\triangle x)^{2}}{2 \triangle t}\rightarrow D, \end{aligned}$$
(5.5)

we obtain that:

$$\begin{aligned} \frac{\partial P}{\partial t}(t,x)=D\frac{\partial ^{2} P}{\partial x^{2}}(t,x). \end{aligned}$$
(5.6)

The derivation in two-dimensional space is the same with that in one-dimensional space.

Appendix 2: Conditions of Turing pattern in system (2.3)

The Jacobian matrix corresponding to the equilibrium point \(E^*\) is that

$$\begin{aligned} M=\left( \begin{array}{cc} a_{11} &{} a_{12} \\ a_{21} &{} a_{22} \\ \end{array} \right) . \end{aligned}$$

We address the temporal stability of the uniform states which is associated with non-uniform perturbations

$$\begin{aligned} \left( {\begin{array}{c}u\\ v\end{array}}\right) =\left( {\begin{array}{c}u^{*}\\ v^{*}\end{array}}\right) +\theta \left( {\begin{array}{c}u_\kappa \\ v_\kappa \end{array}}\right) e^{\lambda t +\mathrm{i} \kappa x}+\mathrm{c.c.}+{\mathcal {O}}(\theta ^2),\nonumber \\ \end{aligned}$$
(5.7)

where \(\lambda \) is the perturbation growth rate, \(\kappa \) is the wavenumber, and c.c. stands for complex conjugate. The linear instability \((\theta \ll 1)\) of each one of the uniform states, is deduced from the dispersion relations. Substituting expression (5.7) into system (2.3) and neglecting all nonlinear terms in u and v, one finds the characteristic equation for the growth rate \(\lambda \) as \(\mathrm{det}(J)\), where

$$\begin{aligned} J=\left( \begin{array}{cc} \lambda -a_{11}+d_{1}\kappa ^2 &{} -a_{12}\\ -a_{21} &{} \lambda -a_{22}+d_{2}\kappa ^2\\ \end{array} \right) . \end{aligned}$$
(5.8)

As a result, the corresponding eigenvalues are

$$\begin{aligned} \lambda _{\kappa }=\frac{-\alpha _{\kappa }\pm \sqrt{\alpha _{\kappa }^{2}-4\beta _{\kappa }}}{2}, \end{aligned}$$
(5.9)

where

$$\begin{aligned} \alpha _{\kappa }= & {} (d_{1}+d_{2})\kappa ^{2}-(a_{11}+a_{22}),\\ \beta _{\kappa }= & {} d_{1}d_{2}\kappa ^{4}-(d_{2}a_{11}+d_{1}a_{22})\kappa ^{2} +a_{11}a_{22}-a_{12}a_{21}. \end{aligned}$$

Consequently, the conditions of Turing pattern in system (2.3) are \(\mathrm Re(\lambda _{\kappa =0})<0 \) and \(\mathrm Re(\lambda _{\kappa >0})>0\) for some \(\kappa >0\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sun, GQ., Wu, ZY., Wang, Z. et al. Influence of isolation degree of spatial patterns on persistence of populations. Nonlinear Dyn 83, 811–819 (2016). https://doi.org/10.1007/s11071-015-2369-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-015-2369-6

Keywords

Navigation