Research article Special Issues

Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response

  • In this paper, we have investigated a spatial predator-prey system with hunting cooperation in predators and type-I functional response. Using linear stability analysis, we obtain the stipulations for diffusive instability and identify the corresponding domain in the space of control parameters. Using qualitative and quantitative analysis, we obtain complex patterns, namely, spotted pattern, stripe pattern and mixed pattern in the Turing domain, by varying the rate of hunting cooperation in predators and diffusion coefficients of prey and predators. The results focus on the effect of hunting cooperation in pattern dynamics of a diffusive predator-prey model and help us in better understanding of the dynamics of the predator-prey interaction in real environment.

    Citation: Teekam Singh, Ramu Dubey, Vishnu Narayan Mishra. Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response[J]. AIMS Mathematics, 2020, 5(1): 673-684. doi: 10.3934/math.2020045

    Related Papers:

    [1] Nazmul Sk, Bapin Mondal, Abhijit Sarkar, Shyam Sundar Santra, Dumitru Baleanu, Mohamed Altanji . Chaos emergence and dissipation in a three-species food web model with intraguild predation and cooperative hunting. AIMS Mathematics, 2024, 9(1): 1023-1045. doi: 10.3934/math.2024051
    [2] Na Min, Hongyang Zhang, Xiaobin Gao, Pengyu Zeng . Impacts of hunting cooperation and prey harvesting in a Leslie-Gower prey-predator system with strong Allee effect. AIMS Mathematics, 2024, 9(12): 34618-34646. doi: 10.3934/math.20241649
    [3] Weili Kong, Yuanfu Shao . Bifurcations of a Leslie-Gower predator-prey model with fear, strong Allee effect and hunting cooperation. AIMS Mathematics, 2024, 9(11): 31607-31635. doi: 10.3934/math.20241520
    [4] Naveed Iqbal, Ranchao Wu, Yeliz Karaca, Rasool Shah, Wajaree Weera . Pattern dynamics and Turing instability induced by self-super-cross-diffusive predator-prey model via amplitude equations. AIMS Mathematics, 2023, 8(2): 2940-2960. doi: 10.3934/math.2023153
    [5] Wei Li, Qingkai Xu, Xingjian Wang, Chunrui Zhang . Dynamics analysis of spatiotemporal discrete predator-prey model based on coupled map lattices. AIMS Mathematics, 2025, 10(1): 1248-1299. doi: 10.3934/math.2025059
    [6] Rongjie Yu, Hengguo Yu, Min Zhao . Steady states and spatiotemporal dynamics of a diffusive predator-prey system with predator harvesting. AIMS Mathematics, 2024, 9(9): 24058-24088. doi: 10.3934/math.20241170
    [7] Heping Jiang . Complex dynamics induced by harvesting rate and delay in a diffusive Leslie-Gower predator-prey model. AIMS Mathematics, 2023, 8(9): 20718-20730. doi: 10.3934/math.20231056
    [8] Jing Zhang, Shengmao Fu . Hopf bifurcation and Turing pattern of a diffusive Rosenzweig-MacArthur model with fear factor. AIMS Mathematics, 2024, 9(11): 32514-32551. doi: 10.3934/math.20241558
    [9] Ali Yousef, Ashraf Adnan Thirthar, Abdesslem Larmani Alaoui, Prabir Panja, Thabet Abdeljawad . The hunting cooperation of a predator under two prey's competition and fear-effect in the prey-predator fractional-order model. AIMS Mathematics, 2022, 7(4): 5463-5479. doi: 10.3934/math.2022303
    [10] Hairong Li, Yanling Tian, Ting Huang, Pinghua Yang . Hopf bifurcation and hybrid control of a delayed diffusive semi-ratio-dependent predator-prey model. AIMS Mathematics, 2024, 9(10): 29608-29632. doi: 10.3934/math.20241434
  • In this paper, we have investigated a spatial predator-prey system with hunting cooperation in predators and type-I functional response. Using linear stability analysis, we obtain the stipulations for diffusive instability and identify the corresponding domain in the space of control parameters. Using qualitative and quantitative analysis, we obtain complex patterns, namely, spotted pattern, stripe pattern and mixed pattern in the Turing domain, by varying the rate of hunting cooperation in predators and diffusion coefficients of prey and predators. The results focus on the effect of hunting cooperation in pattern dynamics of a diffusive predator-prey model and help us in better understanding of the dynamics of the predator-prey interaction in real environment.


    Spatial patterns formation in nature are ubiquitous, with illustrations like zebra stripe patterns on animals skin, Turing patterns in a coherent quantum field, or diffusive patterns in predator-prey models [1,2,3,4,5]. The spatial factors of species interplay has been recognized as a vital component in how ecological communities are created and ecological interplay occur over a broad limit of temporal and spatial scale [6]. Spatial population distribution is of major importance in the study of ecological systems [7,8,9]. Mechanisms and scenarios characterizing the spatial population distribution of ecological species in spatial habitat are a focus of special interest in population dynamics. The spatial population distribution is affected by the proliferation capacity of the species and interactions between individuals [10].

    Spatial pattern formation of predator-prey systems have started based upon the elementary work of A. M. Turing on morphogenesis [11]. The spatial predator-prey systems are studied to comprehend the role of random mobility of the prey and predator, inside their residence. A fully comprehensive elucidation of the spatial impact on ecological species interplays can be observed in the book, written by Okubo et.al. [12].

    Spatial mathematical model is an appropriate tool for investigating fundamental mechanism of complex spatiotemporal population dynamics. An appropriate mathematical structure to explain the spatial aspect of population dynamics is specified by reaction-diffusion equations. Reaction-diffusion models were initially applied to describe the ecological pattern formation by Segel and Jackson in 1972 [3], based on the primary work of Turing [11]. Over the last several decades, a lot of articles have been published on the spatial dynamics of predator-prey model based on reaction-diffusion equations and different types of patterns have emerged for these models [2,3,4,5,12,13,14,15,16,17,18,19].

    Cooperative behavior can stimulate a relation among the population density and per capita population growth rate [20,21]. Ecologists have accepted several mechanisms for stimulating cooperative behavior in prey, namely cooperating reproduction, foraging capacity, etc. The cooperative behavior in prey may be generated by predation or by procedure inborn to the prey lifespan history [21]. Theory has pervasively payed attention to cooperative behavior in preys [22,23,24,25,26,27,28,29,30] and cooperative behavior in predators is less studied and poorly understood [31,32,33], in particular when space is considered explicitly. A mathematical model of prey and predator population interplay with cooperative behavior in predators through the system of nonlinear ordinary differential equations has been studied in non-spatial domain by Alves et.al. [34]. Motivated from their work, we modify and extend the model in a spatial domain to study its spatial dynamics.

    Mechanisms of spatial pattern dynamics of predator-prey systems with cooperative behavior in predators have been comparatively new and to the best of our knowledge, not studied so far. The objective of this current investigation is to create deep intuition into methods of spatial pattern formation in predator-prey model with cooperative behavior in predators. Here, we investigate how distinct intensity of cooperation rate, basic reproduction numbers of the predator and diffusion coefficients affects the spatial patterns of predator-prey interaction.

    This paper is organized as follows: In section Ⅱ, we formulate a reaction-diffusion predator-prey system with zero flux boundary conditions and non-zero initial conditions. Furthermore, we analyze the non-spatial dynamics of the model with hunting cooperation in predators and give a survey of the linear stability analysis in section Ⅲ. In section Ⅳ, we obtain the sufficient condition for Turing bifurcation with zero-flux boundary condition. In section Ⅴ, we carry out a series of numerical simulations to reveal that there is a large variety of different spatiotemporal dynamics in the spatial model for different intensities of predator's hunting cooperation, basic reproduction number of the predator and the rates of diffusion coefficients. The paper ends with a discussion.

    By incorporating diffusion and Holling type I functional response in the general predator-prey system with cooperative behavior in predators [34], we obtain the following diffusive predator-prey model as

    UT=rU(1UK)(λ+aV)UV+D12U, (2.1)
    VT=e(λ+aV)UVmV+D22V, (2.2)

    where U(T) and V(T) are the densities of prey and predator population at time T and location (x1,x2), respectively. Here, r is the intrinsic growth rate of the prey and K is its carrying capacity. The parameter λ (λ>0) is the attack rate per predator and prey, a (a>0) describes the predator cooperation in hunting (aV is cooperation term). We consider a Holling type I functional response of the form

    (λ+aV)U (2.3)

    which depends on both prey and predator densities, thereby reflecting hunting cooperation (handling-driven). The parameter e is conversion efficiency and m is the per capita mortality rate of predators. The non-negative constants D1 and D2 are the diffusion coefficients for prey and predator densities respectively.

    We now non-dimensionalize the system Eqs (1) and (2) by introducing the non-dimensional variables

    X=eλmU,Y=λmV,t=mT,x=x1mD2,y=x2mD2, (2.4)

    and non-dimensional parameters

    σ=rm,ω=eλmK,α=amλ2,D=D1D2, (2.5)

    and obtain the modified model as

    Xt=σX(1Xω)(1+αY)XY+D2X, (2.6)
    Yt=(1+αY)XYY+2Y, (2.7)

    where the positive constant D is the ratio of diffusion coefficients of prey and predator densities and 2(=2x2+2y2) is the usual Laplacian operator in two dimensional space R=(x,y). To make certain that spatial patterns are governed by reaction-diffusion equations, system equations (6) and (7) is to be analyzed with the following non-zero initial conditions

    X(x,y,0)>0,Y(x,y,0)>0,(x,y)Ω=[0,L]×[0,L], (2.8)

    and zero-flux (Neumann) boundary conditions

    Xν=Yν=0, (2.9)

    where L denotes the size of the system in the direction of x and y. ν is outward unit normal on the boundary Ω. Condition (9) implies that no individual species leave the domain.

    In absence of diffusion, the equilibrium points of the system Eqs (6) and (7) are given by

    σX(1Xω)(1+αY)XY=0, (3.1)
    (1+αY)XYY=0. (3.2)

    The system Eqs (6) and (7) has following ecologically significant equilibrium solutions: (ⅰ) (0,0) (both extinct), (ⅱ) (ω,0) (predator extinct) and (ⅲ) (X,Y) (coexist), where X=1(1+αY) and Y is the positive solution of

    ωα2Y3+2αωY2+ω(1σα)Y+σ(1ω)=0. (3.3)

    The cubic Eq (12) have one and two positive real roots when (ω>1, σα>1 or σα<1) and (ω<1, σα>1) respectively.

    The variational matrix about the equilibrium point (X,Y) is given by

    [σY(1+αY)2XσωX(1+2Yα)Y(1+αY)X+2uvα1](X,Y). (3.4)

    (ⅰ) At (0,0), the variational matrix is

    [σ001],

    whose eigenvalues are 1 and σ(>0). Hence, the system is unstable at the origin.

    (ⅱ) At (ω,0), the variational matrix is

    [σω0ω1],

    whose eigenvalues are σ and ω1. Hence, the system is stable if ω<1.

    (ⅲ) At (X,Y), the variational matrix is

    [σY(1+αY)2σω(1+αY)2+1(1+αY)Y(1+αY)αY(1+αY)], (3.5)

    and, the corresponding characteristic equation is

    λ2+B1λ+B2=0, (3.6)

    where B1=Y(1+αY)+2σω(1+αY)αY(1+αY)σ,B2=Y(1+αY)(ασ+(1+αY)2)ω2ασY(1+αY)2ω.

    The equilibrium solution (X,Y) is locally asymptotically stable if and only if

    (σ+Y(1+α(1Y(2+Yα)+σ)))ω2σ<0

    and

    Y(1+Yα)((1+Yα)2+ασ)ω2Yασ>0.

    Interior equilibrium point (X,Y) of non-spatial system is spatially homogenous steady state, that is, constant in space and time for the reaction-diffusion system Eqs (6) and (7) (diffusive model). We assume that (X,Y) is stable in non-spatial system of (6–7) which means the spatially homogenous steady state is stable with respect to spatially homogenous perturbations. Though the diffusion is often considered as a stabilizing process, it is a well known fact that diffusion can make a spatially homogenous steady state linearly unstable (Turing instability) with respect to heterogenous perturbations in a system of two interacting species [3,11]. The condition for Turing instability may be obtained by introducing a small heterogenous perturbation of the homogenous steady state as follows:

    X(t,x,y)=X+ϵ1exp(λkt+i(kxx+kyy)), (4.1)
    Y(t,x,y)=Y+ϵ2exp(λkt+i(kxx+kyy)), (4.2)

    where ϵ1 and ϵ2 are two non-zero real numbers and k=(kx,ky) such that k2=(k2x+k2y) is the wave number.

    Substituting (16–17) into (6–7) and then linearizing it about interior equilibrium point (X,Y), we obtain the variational matrix as

    [σY(1+αY)2σω(1+αY)δk22+1(1+αY)Y(1+αY)αY(1+αY)k2]. (4.3)

    The corresponding characteristic equation is

    λ2+C1(k2)λ+C2(k2)=0, (4.4)

    where

    C1(k2)=(1+δ)k2+2σω(1+αY)σY(α1+αY1αY),

    and

    C2(k2)=δk4+(δωαY+(2σ+(1+αY)(Y+αY2σ)ω))(1+αY)k2+((1+αY)2+σα)ωY(1+αY)2σαY.

    By Routh-Hurwitz criterion, the system (6–7) will be stable about (X,Y) if C1(k2)>0 and C2(k2)>0. As the parameters, δ(=d1d2) and k2 are all positive and 2σω(1+αY)σY(α1+αY1αY)>0 (by the stability of the non-spatial model of (6–7)), hence, C1(k2)> always positive. Hence, the condition for diffusive instability is C2(k2)<0.

    The polynomial function C2(k2) has a minimum for some value of k2, say k2min, where

    k2min=δαωY+(ω(1+αY)2Y+σ(ω+αωY2))2δ(1+αY)ω. (4.5)

    Hence, the minimum value of k for which Turing instability will occur, is, C2(k2min)<0. Therefore, substitute k2min in C2(k2), we get the sufficient condition for Turing instability

    4δ(2ασ(1+αY)((1+αY)2+ασ)ω)Yω(1+αY)2+(δαωY+(ω(1+αY)2Y+σ(ω+αωY2)))2ω2(1+αY)2>0. (4.6)

    The interval of the wave number for which Turing instability take place is (k,k+) and in this interval, we have C2(k2)<0, where

    k=αY2(1+αY)A2δω(1+αY)12δB2+C (4.7)
    k+=αY2(1+αY)A2δω(1+αY)+12δB2+C (4.8)

    and

    A=2σ+(1+αY)(Y+αY2σ)ω,B=δαωY+(Y(1+αY)2ω+σ(ω+ωαY2)),C=4δω(2ασ(1+αY)((1+αY)2+ασ)ω)Y.

    We will now investigate the numerical results of spatiotemporal models, namely (6–7). For numerical simulation, we set σ and ω as σ=10.0, ω=0.8, and consider α and D, as controlling parameters. For these values of parameters, the positive equilibrium points are (0, 0), (0.8, 0), (0.6159, 1.4171) and (0.6431, 1.2614). The steady state (0.6159, 1.4171) is stable and (0.6431, 1.2614) is unstable. Throughout our study in the spatiotemporal domain, we have considered the stable steady state (0.6159, 1.4171). Please note that the non-dimensional parameter ω=eλmK, comprising of the dimensional carrying capacity, attack rate, per capita mortality rate of predators and the conversion efficiency, can also be interpreted as the basic reproduction number of the predator, which is defined as the average number of offspring produced by a single predator during its life time, when introduced into the prey population at carrying capacity.

    We now simulate the spatial model in two dimensional space with the help of finite difference scheme for spatial derivatives. The forward Euler's numerical method is used for the non-spatial part of model (6–7) and general finite difference scheme of five point is used for the spatial part. The reaction-diffusion partial differential equations, given by (6–7), is numerically solved by using splitting method [36,37]. The numerical values for the step sizes of time and space have been selected adequately small for avoiding the numerical artifacts. In this study, we have employed statistically uncorrelated Gaussian white noise perturbation in space, which is mathematically denoted in two dimensional case as

    X(xi,yj,0)=X+ϵij, (5.1)
    Y(xi,yj,0)=Y+ηij, (5.2)

    where ϵij and ηij are statistically uncorrelated Gaussian white noise perturbations with zero mean and fixed variance in two dimensional space.

    For spatial model, we perform all the numerical simulations of the system (6–7) over the non-zero initial condition and zero-flux boundary conditions, in two dimensional spatial domain. The domain size is 70×70 with time-step Δt=0.001 and space-step Δx=Δy=1.0. The parameter values of σ remain same and ω, α are used as the controlling parameter (just like the temporal case).

    Note: The Neumann zero-flux conditions are placed at boundary of the numerical domain in two dimensional problems. The size of the domain is chosen large enough so that the impact of the boundaries has been kept as small as possible during the simulation time.

    We now demonstrate diffusive induced instability (Turing instability) and the corresponding patterns formation for the system (6–7). Although, the sufficient conditions for Turing instability were obtained analytically in previous section, whether they are satisfied with our corresponding set of parameter values, is yet to be tested. In order to do so, we sketch the polynomial function C2(k2) for distinct values of D1, ω and α (other parameter values are fixed, namely, σ=10, D2=0.07). Figures 1, 2 and 3 shows the plot of C2(k2) against the wave number (k) for different values of D1, ω and α respectively. We observe that the sufficient condition of the diffusive instability min C2(k2)<0 holds, when D1 is adequately large, beginning from D1=2.2 (see Figure 1). The polynomial function C2(k2)<0 holds, the wave number (k) fit in the interval (k,k+) and the length of the interval increases with an increase in the diffusion coefficient D1. We next plot C2(k2) for different values of ω (other parameters remain fixed) and we observe that the conditions for the Turing instability holds for ω0.804 approximately, and the length of the interval (k,k+) decreases with an increase in ω (see Figure 2). We next plot C2(k2) for different values of α (other parameters remain fixed) and we observe that the conditions for the Turing instability holds for α0.45 approximately, and the length of the interval (k,k+) decreases with an increase in α (see Figure 3).

    Figure 1.  C2(k2) vs k for different values of D1.
    Figure 2.  C2(k2) vs k for different values of ω.
    Figure 3.  C2(k2) vs k for different values of α.

    In Figure 4, we illustrate the density distributions of prey and predator which covers two types of pattern namely spots and stripes. Figure 5 and Figure 6, shows the two dimensional stationary diffusive patterns of the model (6–7) at time t = 500 with ω = 0.800 and diffusion coefficients D1=3.1 and D2=0.07 for the prey and predator population respectively. In these figures hexagonal patterns (spots) prevail over the entire habitat eventually. In Figures 4, 5 and 6, it is observe that blue spots (minimum density of X) are distributed on a reddish background (maximum density of X), that is, the preys are segregated with low population density.

    Figure 4.  Spatial density distribution pattern formation of prey and predator (spots type).
    Figure 5.  Spatial density distribution pattern formation of prey and predator (mixed spots-stripes type).
    Figure 6.  Spatial density distribution pattern formation of prey and predator (stripes type).

    In theoretical ecology, intensive studies of the mechanisms and scenarios of pattern formation in models of interacting populations have always been an attraction as their perception help to enhance the understanding of real-world ecological systems. In this paper, we have considered a diffusive predator-prey model with hunting cooperation in predators and type I functional response, under non-zero initial conditions and zero-flux boundary conditions. We have provided the elaborate analysis of both temporal and spatiotemporal models and studied possible scenarios of pattern formation in the diffusive predator-prey model with hunting cooperation in predators. While studying the spatiotemporal model, we first obtain the condition for diffusive instability and identified the corresponding domain in the space of controlling parameters. The hunting cooperation coefficient (α), the basic reproduction number of the predator (ω) and the diffusion coefficient of the prey (D1) are the controlling parameters in our study. Using the parameter values from Turing domain, we investigate the properties of the system using extensive numerical simulations.

    By varying the values of cooperation coefficient, we get dissimilar types of diffusive patterns, namely, patchy pattern (spots), stripe pattern and mixed pattern (spot-stripe). From the point of view of population dynamics, one can observe that the spot formation for preys imply that the preys are scattered and isolated with low density and the remainder region is high density, which means that the preys may break out in the area and are safe. Similarly, spot formation in predators convey that with hunting cooperation, the predators are scattered and isolated but still survives. The methods and consequences in this study may amplify the systematic investigation of spatial pattern formation in the predator-prey systems, and may nicely enforce in some different research dimensions. Further analyze are important to study the patterns dynamics of some more diffusive ecological. It would be interesting to study the traveling waves in the spatial predator-prey models with hunting cooperation in predators with type Ⅱ, Ⅲ or type Ⅳ functional responses [35,36,37,38]. This article highlights a number of research areas for future consideration in spatial pattern formation.

    Teekam Singh gratefully acknowledges the School of Allied Sciences, Department of Mathematics, Graphic Era (deemed to be) University and Graphic Era Hill University Dehradun-248 002, Uttarakhand, India.

    The authors declare no conflict of interest.



    [1] J. L. Aragon, M. Torres, D. Gil, et al. Turing patterns with pentagonal symmetry, Phys. Rev. E, 65 (2002), 051913-051921. doi: 10.1103/PhysRevE.65.051913
    [2] V. Ardizzone, P. Lewandowski, M. H. Luk, et al. Formation and control of turing patterns in a coherent quantum fluid, Sci. Rep., 3 (2013), 3016-3021. doi: 10.1038/srep03016
    [3] L. A. Segel, J. L. Jackson, Dissipative structure: An explanation and an ecological example, J. Theor. Biol., 37 (1972), 545-559. doi: 10.1016/0022-5193(72)90090-2
    [4] S. Yuan, C. Xu, T. Zhang, Spatial dynamics in a predator-prey model with herd behavior, Chaos, 23 (2013), 033102-033112. doi: 10.1063/1.4812724
    [5] X. C. Zhang, G. Q. Sun, Z. Jin, Spatial dynamics in a predator-prey model with Beddington-DeAngelis functional response, Phys. Rev. E, 85 (2012), 021924-021937. doi: 10.1103/PhysRevE.85.021924
    [6] U. Dieckmann, R. Law, J. A. Metz, The Geometry of Ecological Interactions: Simplifying Spatial Complexity, Cambridge University Press, 2000.
    [7] S. A. Levin, L. A. Segel, Hypothesis for origin of planktonic patchiness, Nature, 259 (1976), 659-659.
    [8] P. K. Maini, D. L. Benson, J. A. Sherratt, Pattern formation in reaction-diffusion models with spatially inhomogeneous diffusion coefficients, Math. Med. Biol., 9 (1992), 197-213. doi: 10.1093/imammb/9.3.197
    [9] G. R. Shaver, Spatial Heterogeneity: Past, Present, and Future: Ecosystem Function in Heterogeneous Landscapes, Springer-Verlag, New York, 2005.
    [10] F. Vinatier, P. Tixier, P. F. Duyck, et al. Factors and mechanisms explaining spatial heterogeneity: A review of methods for insect populations, Methods Ecol. Evol., 2 (2011), 11-22. doi: 10.1111/j.2041-210X.2010.00059.x
    [11] A. M. Turing, The chemical basis of morphogenesis, Philos. Trans. R. Soc. B, 237 (1952), 37-72. doi: 10.1098/rstb.1952.0012
    [12] A. Okubo, S. A. Levin, Diffusion and Ecological Problems: Modern Perspectives, Springer Science and Business Media, 2013.
    [13] A. Goldbeter, Dissipative structures in biological systems: Bistability, oscillations, spatial patterns and waves, Philos. Trans. R. Soc. A, 376 (2018), 1-25.
    [14] X. Guan, W. Wang, Y. Cai, Spatiotemporal dynamics of a Leslie-Gower predator-prey model incorporating a prey refuge, Nonlinear Anal. Real World Appl., 12 (2011), 2385-2395. doi: 10.1016/j.nonrwa.2011.02.011
    [15] K. Konishi, N. Hara, Stabilization of a spatially uniform steady state in two systems exhibiting Turing patterns, Phys. Rev. E, 97 (2018), 052201-052213. doi: 10.1103/PhysRevE.97.052201
    [16] A. Morozov, S. V. Petrovskii, B. L. Li, Spatiotemporal complexity of patchy invasion in a predator-prey system with the Allee effect, J. Theor. Biol., 238 (2006), 18-35. doi: 10.1016/j.jtbi.2005.05.021
    [17] J. D. Murray, A pre-pattern formation mechanism for animal coat markings, J. Theor. Biol., 88 (1981), 161-199. doi: 10.1016/0022-5193(81)90334-9
    [18] W. Wang, Q. X. Liu, Z. Jin, Spatiotemporal complexity of a ratio-dependent predator-prey system, Phys. Rev. E, 75 (2007), 051913-051925. doi: 10.1103/PhysRevE.75.051913
    [19] T. Zhang, Y. Xing, H. Zang, et al. Spatiotemporal dynamics of a reaction-diffusion system for a predator-prey model with hyperbolic mortality, Nonlinear Dyn., 78 (2014), 265-277. doi: 10.1007/s11071-014-1438-6
    [20] F. Courchamp, L. Berec, J. Gascoigne, Allee Effects in Ecology and Conservation, Oxford University Press, 2008.
    [21] P. A. Stephens and W. J. Sutherland, Consequences of the Allee effect for behaviour, ecology and conservation, Trends Ecol. Evol., 14 (1999), 401-405. doi: 10.1016/S0169-5347(99)01684-5
    [22] P. Aguirre, E. González-Olivares, S. Torres, Stochastic predator-prey model with Allee effect on prey, Nonlinear Anal. Real World Appl., 14 (2013), 768-779. doi: 10.1016/j.nonrwa.2012.07.032
    [23] F. Courchamp, T. Clutton-Brock, B. Grenfell, Inverse density dependence and the Allee effect, Trends Ecol. Evol., 14 (1999), 405-410. doi: 10.1016/S0169-5347(99)01683-3
    [24] L. A. Dugatkin, Cooperation Among Animals: An Evolutionary Perspective, Oxford University Press, 1997.
    [25] E. González-Olivares, J. Mena-Lorca, A. Rojas-Palma, et al. Dynamical complexities in the Leslie-Gower predator-prey model as consequences of the Allee effect on prey, Appl. Math. Modell., 35 (2011), 366-381. doi: 10.1016/j.apm.2010.07.001
    [26] E. González-Olivares, H. Meneses-Alcay, B. González-Yañez, et al. Multiple stability and uniqueness of the limit cycle in a Gause-type predator-prey model considering the Allee effect on prey, Nonlinear Anal. Real World Appl., 12 (2011), 2931-2942. doi: 10.1016/j.nonrwa.2011.04.003
    [27] A. Morozov, S. V. Petrovskii, B. L. Li, Bifurcations and chaos in a predator-prey system with the Allee effect, Proc. Royal Soc. Lond. B, 271 (2004), 1407-1414. doi: 10.1098/rspb.2004.2733
    [28] F. Rao, Y. Kang, The complex dynamics of a diffusive prey-predator model with an Allee effect in prey, Ecol. Complex., 28 (2016), 123-144. doi: 10.1016/j.ecocom.2016.07.001
    [29] J. Wang, J. Shi, J. Wei, Dynamics and pattern formation in a diffusive predator-prey system with strong Allee effect in prey, J. Differ. Equations, 251 (2011), 1276-1304. doi: 10.1016/j.jde.2011.03.004
    [30] S. R. Zhou, Y. F. Liu, G. Wang, The stability of predator-prey systems subject to the Allee effects, Theor. Popul. Biol., 67 (2005), 23-31. doi: 10.1016/j.tpb.2004.06.007
    [31] A. Bompard, I. Amat, X. Fauvergue, et al. Host-parasitoid dynamics and the success of biological control when parasitoids are prone to Allee effects, PLoS One, 8 (2013), 76768-76779. doi: 10.1371/journal.pone.0076768
    [32] A. De-Roos, L. Persson, H. R. Thieme, Emergent Allee effects in top predators feeding on structured prey populations, Proc. Royal Soc. Lond. B, 270 (2003), 611-618. doi: 10.1098/rspb.2002.2286
    [33] A. Verdy, Modulation of predator-prey interactions by the Allee effect, Ecol. Modell., 221 (2010), 1098-1107. doi: 10.1016/j.ecolmodel.2010.01.005
    [34] M. T. Alves, F. M. Hilker, Hunting cooperation and Allee effects in predators, J. Theor. Biol., 419 (2017), 13-22. doi: 10.1016/j.jtbi.2017.02.002
    [35] S. R. Dunbar, Traveling wave solutions of diffusive Lotka-Volterra equations: A heteroclinic connection in R4, Trans. Amer. Math. Soc., 286 (1984), 557-594.
    [36] S. R. Dunbar, Traveling waves in diffusive predator-prey equations: periodic orbits and point-to-periodic heteroclinc orbits, SIAM J. Appl. Math., 46 (1986), 1057-1078. doi: 10.1137/0146063
    [37] R. Gardner, J. Smoller, The existence of periodic travelling waves for singularly perturbed predator-prey equations via the Conley Index, J. Differ. Equations, 47 (1983), 133-161. doi: 10.1016/0022-0396(83)90031-1
    [38] X. Lin, P. Weng, C. Wu, Traveling wave solutions for a predator-prey system with sigmoidal response function, J. Dyn. Differ. Equations, 23 (2011), 903-921. doi: 10.1007/s10884-011-9220-7
  • This article has been cited by:

    1. Danxia Song, Chao Li, Yongli Song, Stability and cross-diffusion-driven instability in a diffusive predator–prey system with hunting cooperation functional response, 2020, 54, 14681218, 103106, 10.1016/j.nonrwa.2020.103106
    2. Salih Djilali, Behzad Ghanbari, The influence of an infectious disease on a prey-predator model equipped with a fractional-order derivative, 2021, 2021, 1687-1847, 10.1186/s13662-020-03177-9
    3. Derya Avcı, Temperature profiles and thermal stresses due to heat conduction under fading memory effect, 2021, 136, 2190-5444, 10.1140/epjp/s13360-021-01319-w
    4. Pankaj Gulati, Sudipa Chauhan, Anuj Mubayi, Teekam Singh, Payal Rana, Dynamical analysis, optimum control and pattern formation in the biological pest (EFSB) control model, 2021, 147, 09600779, 110920, 10.1016/j.chaos.2021.110920
    5. Kuldeep Singh, Mukesh Kumar, Ramu Dubey, Teekam Singh, Untangling role of cooperative hunting among predators and herd behavior in prey with a dynamical systems approach, 2022, 162, 09600779, 112420, 10.1016/j.chaos.2022.112420
    6. Abdelheq Mezouaghi, Abdelkader Benali, Sunil Kumar, Salih Djilali, Anwar Zeb, Shahram Rezapour, Mathematical analysis of a fractional resource-consumer model with disease developed in consumer, 2021, 2021, 1687-1847, 10.1186/s13662-021-03642-z
    7. Zulqurnain Sabir, Muhammad Asif Zahoor Raja, Yolanda Guerrero Sánchez, Han Wang, Solving an Infectious Disease Model considering Its Anatomical Variables with Stochastic Numerical Procedures, 2022, 2022, 2040-2309, 1, 10.1155/2022/3774123
    8. TEEKAM SINGH, MUKESH KUMAR, SPATIOTEMPORAL DYNAMICAL ANALYSIS OF A PREDATOR–PREY SYSTEM WITH FEAR AND GROUP DEFENSE IN PREY, 2022, 30, 0218-3390, 387, 10.1142/S0218339022500140
    9. Salih Djilali, Carlo Cattani, Patterns of a superdiffusive consumer-resource model with hunting cooperation functional response, 2021, 151, 09600779, 111258, 10.1016/j.chaos.2021.111258
    10. Zulqurnain Sabir, Thongchai Botmart, Muhammad Asif Zahoor Raja, Wajaree Weera, António M. Lopes, An advanced computing scheme for the numerical investigations of an infection-based fractional-order nonlinear prey-predator system, 2022, 17, 1932-6203, e0265064, 10.1371/journal.pone.0265064
    11. Eric M. Takyi, Charles Ohanian, Margaret Cathcart, Nihal Kumar, Dynamical analysis of a predator-prey system with prey vigilance and hunting cooperation in predators, 2024, 21, 1551-0018, 2768, 10.3934/mbe.2024123
    12. Sushil Kumar Verma, Bipin Kumar, Bifurcation and pattern formation in a prey–predator model with cooperative hunting, 2024, 139, 2190-5444, 10.1140/epjp/s13360-024-05543-y
    13. Dingyong Bai, Jiaoxin Tang, Global Dynamics of a Predator–Prey System with Cooperative Hunting, 2023, 13, 2076-3417, 8178, 10.3390/app13148178
    14. Turki Aljrees, Teekam Singh, Neeraj Varshney, Mukesh Kumar, Kamred Udham Singh, Vrince Vimal, Influence of Spatial Dispersal among Species in a Prey–Predator Model with Miniature Predator Groups, 2023, 15, 2073-8994, 986, 10.3390/sym15050986
    15. Mohammad Partohaghighi, Ali Akgül, New Fractional Modelling and Simulations of Prey–Predator System with Mittag–Leffler Kernel, 2023, 9, 2349-5103, 10.1007/s40819-023-01523-5
    16. Ibtissam Benamara, Abderrahim El Abdllaoui, Radouane Yafia, Hemen Dutta, Qualitative analysis for a diffusive predator-prey model with hunting cooperation and holling type III functional response, 2023, 18, 0973-5348, 13, 10.1051/mmnp/2023010
    17. Jian Wang, Haixiao Wang, Jialin Chen, Wenjing Jiang, Runqiao Peng, Junseok Kim, An efficient image encryption algorithm based on a time-delay predator–prey model, 2025, 131, 0946-2171, 10.1007/s00340-025-08412-3
    18. Weirun Tao, Zhi-An Wang, Global well-posedness and Turing–Hopf bifurcation of prey-taxis systems with hunting cooperation, 2025, 0956-7925, 1, 10.1017/S0956792525000026
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5898) PDF downloads(666) Cited by(18)

Figures and Tables

Figures(6)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog