Complexity, selectivity and asymmetry in the conformation of the power phenomenon. Analysis of Chilean society

  • Received: 01 June 2014 Revised: 01 November 2014
  • Primary: 05C82, 91D30; Secondary: 90B15.

  • In this work we analyzed the relationships between powerful politicians and businessmen of Chile in order to study the phenomenon of social power. We developed our study according to Complex Network Theory but also using traditional sociological theories of Power and Elites. Our analyses suggest that the studied network displays common properties of Complex Networks, such as scaling in connectivity distribution, properties of small-world networks, and modular structure, among others. We also observed that social power (a proposed metric is presented in this work) is also distributed inhomogeneously. However, the most interesting observation is that this inhomogeneous power and connectivity distribution, among other observed properties, may be the result of a dynamic and unregulated process of network growth in which powerful people tend to link to similar others. The compatibility between people, increasingly selective as the network grows, could generate the presence of extremely powerful people, but also a constant inequality of power where the difference between the most powerful is the same as among the least powerful. Our results are also in accordance with sociological theories.

    Citation: Juan Pablo Cárdenas, Gerardo Vidal, Gastón Olivares. Complexity, selectivity and asymmetry in the conformation of the power phenomenon. Analysis of Chilean society[J]. Networks and Heterogeneous Media, 2015, 10(1): 167-194. doi: 10.3934/nhm.2015.10.167

    Related Papers:

    [1] Ceyu Lei, Xiaoling Han, Weiming Wang . Bifurcation analysis and chaos control of a discrete-time prey-predator model with fear factor. Mathematical Biosciences and Engineering, 2022, 19(7): 6659-6679. doi: 10.3934/mbe.2022313
    [2] Parvaiz Ahmad Naik, Muhammad Amer, Rizwan Ahmed, Sania Qureshi, Zhengxin Huang . Stability and bifurcation analysis of a discrete predator-prey system of Ricker type with refuge effect. Mathematical Biosciences and Engineering, 2024, 21(3): 4554-4586. doi: 10.3934/mbe.2024201
    [3] Yangyang Li, Fengxue Zhang, Xianglai Zhuo . Flip bifurcation of a discrete predator-prey model with modified Leslie-Gower and Holling-type III schemes. Mathematical Biosciences and Engineering, 2020, 17(3): 2003-2015. doi: 10.3934/mbe.2020106
    [4] Min Hou, Tonghua Zhang, Sanling Yuan . Stability and Hopf bifurcation of an intraguild prey-predator fishery model with two delays and Michaelis-Menten type predator harvest. Mathematical Biosciences and Engineering, 2024, 21(4): 5687-5711. doi: 10.3934/mbe.2024251
    [5] Mianjian Ruan, Xianyi Li, Bo Sun . More complex dynamics in a discrete prey-predator model with the Allee effect in prey. Mathematical Biosciences and Engineering, 2023, 20(11): 19584-19616. doi: 10.3934/mbe.2023868
    [6] Hongqiuxue Wu, Zhong Li, Mengxin He . Dynamic analysis of a Leslie-Gower predator-prey model with the fear effect and nonlinear harvesting. Mathematical Biosciences and Engineering, 2023, 20(10): 18592-18629. doi: 10.3934/mbe.2023825
    [7] Yajie Sun, Ming Zhao, Yunfei Du . Multiple bifurcations of a discrete modified Leslie-Gower predator-prey model. Mathematical Biosciences and Engineering, 2023, 20(12): 20437-20467. doi: 10.3934/mbe.2023904
    [8] Rajalakshmi Manoharan, Reenu Rani, Ali Moussaoui . Predator-prey dynamics with refuge, alternate food, and harvesting strategies in a patchy habitat. Mathematical Biosciences and Engineering, 2025, 22(4): 810-845. doi: 10.3934/mbe.2025029
    [9] Rongjie Yu, Hengguo Yu, Chuanjun Dai, Zengling Ma, Qi Wang, Min Zhao . Bifurcation analysis of Leslie-Gower predator-prey system with harvesting and fear effect. Mathematical Biosciences and Engineering, 2023, 20(10): 18267-18300. doi: 10.3934/mbe.2023812
    [10] Manoj K. Singh, Brajesh K. Singh, Poonam, Carlo Cattani . Under nonlinear prey-harvesting, effect of strong Allee effect on the dynamics of a modified Leslie-Gower predator-prey model. Mathematical Biosciences and Engineering, 2023, 20(6): 9625-9644. doi: 10.3934/mbe.2023422
  • In this work we analyzed the relationships between powerful politicians and businessmen of Chile in order to study the phenomenon of social power. We developed our study according to Complex Network Theory but also using traditional sociological theories of Power and Elites. Our analyses suggest that the studied network displays common properties of Complex Networks, such as scaling in connectivity distribution, properties of small-world networks, and modular structure, among others. We also observed that social power (a proposed metric is presented in this work) is also distributed inhomogeneously. However, the most interesting observation is that this inhomogeneous power and connectivity distribution, among other observed properties, may be the result of a dynamic and unregulated process of network growth in which powerful people tend to link to similar others. The compatibility between people, increasingly selective as the network grows, could generate the presence of extremely powerful people, but also a constant inequality of power where the difference between the most powerful is the same as among the least powerful. Our results are also in accordance with sociological theories.


    Population is a collection of individuals of the same species living together within a certain spatial range. The study of population dynamics mainly describes the dynamic relationship between population communities in predation systems and food web systems. Food web refers to the complex relationship network between predation, competition, cooperation and reciprocity between biological populations in biological systems. The study of food web population dynamics can give humans an important understanding of the basic nature of ecosystems, promoting the evolution of life, protecting the natural environment, and maintaining ecological balance at the macro level, and give certain guiding opinions on the protection of endangered species at the micro level. Therefore, studying the interactions between different biological populations through mathematical modeling is an important area of research for ecosystems.

    In 1976, May showed in [1] that the first-order discrete equation model, although simple, can produce a series of extremely complex dynamic behaviors, such as from stable points to unstable bifurcation levels, and eventually produce chaos. However, in continuous-time models, achieving such complex dynamic behavior requires differential equations of three dimensions and more than three dimensions to cause chaos [2]. It can be seen that discrete systems described by the difference equation are richer in dynamical phenomena. Consequently, the discrete dynamical system model has attracted the attention and research of more scholars [3,4,5,6,7,8,9,10,11].

    In [12], Gan et al. studied the stability in a simple food chain system with Michaelis-Menten functional responses and nonlocal delays, using the Lyapunov functional to derive sufficient conditions for global stability of positive steady state and semi-trivial steady state. In [13], Clark et al. described mathematical models of exploited fish stocks, assuming that certain stocks can be obtained through dynamic aggregation processes. The effects of aggregation on yield-effort relationships, abundance indices, and fishery dynamics are discussed, as well as various management approaches for these models. On the other hand, with the increasing demand for food and other resources, the exploitation of biological resources is also increasing. More ecologists are very interested in studying these types of models, such as predator-prey models, and consider the impact of exploiting (harvesting) resources to protect the sustainable use of biological resources [14,15,16,17]. To do this, they applied optimal capture and control strategies to achieve the recyclability of biological resources.

    In [20], we studied the behavioral analysis of a class of discrete dynamical system with linear harvest rates, and obtained many complex dynamic phenomena. Considering the finite nature of resources and space, the linear capture rate has no upper bound, so we will further study the Michaelis-Menten[13] type capture on the basis of the original, and this type of capture is a kind of harvest that gradually rises until the saturation state with the increase of the number of captured objects, which is bounded and more in line with the realistic ecological environment. So, in this paper we consider a discrete-time predator-prey model with Michaelis-Menten type harvesting in preys, which is given by

    $ {un+1=unexp[r1(1unK)r2vnc+unqEm1E+m2un],vn+1=vnexp[a+r2dunc+unbvnh2],
    $
    (1.1)

    where the meanings of all parameters of model (1.1) are shown in Table 1.

    Table 1.  The interpretation of parameters.
    parameters interpretation
    $ u, v $ the densities of prey and predator population, respectively
    $ r_{1} $ the intrinsic growth rates of prey
    $ a $ the intrinsic growth rates of predator
    $ K $ the environmental carrying capacity of prey
    $ r_2 $ the consumption rate of prey
    $ b $ the competition between individuals due to overcrowding of predator
    $ c $ the half saturation constant
    $ d $ the conversion rate of predator
    $ h_2 $ the capture rate of predator
    $ m_i, i=1, 2 $ suitable constants
    $ q $ the catchability co-efficients of prey
    $ E $ the degree of harvest effort

     | Show Table
    DownLoad: CSV

    The paper is organized as follows. In Section 2, we analyze the dynamics of system (1.1), including the existence and local stability of the equilibrium points, and the boundedness of system (1.1). In Section 3, using the bifurcation theory and the center manifold theorem, the flip bifurcation and Neimark-Sacker bifurcation are analyzed, and the conditions for determining the bifurcation direction and the stability of the bifurcation periodic solution are obtained. In Section 4, a feedback control strategy is used to control chaos in the system, and an optimal harvesting strategy is introduced to obtain the optimal value of the harvesting coefficient. In Section 5, we verify our analytical results through numerical simulations. In the last Section, the article is ended with a brief conclusion.

    Lemma 1 Solutions of system (1.1) with nonnegative initial conditions remain nonnegative. If $ u_0 = 0 $, then $ u_n = 0 $ for all $ n\geq0 $. If $ v_0 = 0 $, then $ v_n = 0 $ for all $ n\geq0 $. If $ u_0 > 0 $ and $ v_0\geq0 $, then $ u_n > 0 $ for all $ n\geq0 $. If $ u_0\geq0 $ and $ v_0 > 0 $, then $ v_n > 0 $ for all $ n\geq0 $.

    Proof. It can be directly demonstrated by system (1.1) structure.

    Lemma 2

    (I) System (1.1) always has a trivial equilibrium point $ E_{0}(0, 0) $. At this moment, the two species will go extinct.

    (II) If $ q < r_1m_1 $, then system (1.1) always has a positive semi-trivial equilibrium point $ E_{1}(\frac{-\alpha+\sqrt{\Delta }}{2}, 0) $, where $ \alpha = \frac{m_{1}E}{m_{2}}-K $, $ \beta = \frac{qEK-r_{1}EKm_{1}}{r_{1}m_{1}} $, $ \Delta = \alpha^2-4\beta > 0 $. At this time, the prey population reaches $ \frac{-\alpha+\sqrt{\Delta }}{2} $, and the predator tends to go extinct.

    (III) If $ a > h_{2} $, then system (1.1) always has a positive semi-trivial equilibrium point $ E_{2}(0, \frac{a-h_{2}}{b}) $. At this time, the prey population tend to go extinct, and the predator population converge to $ \frac{a-h_{2}}{b} $.

    (IV) System (1.1) has a positive nontrivial equilibrium point $ E^{*}(u^*, \frac{1}{b}(a+\frac{r_2du^*}{u^*+c}-h_2)) $, where $ a > h_2 $, and $ u $ is the positive solution to the quartic equation of one variabie

    $ A_1u^4+A_2u^3+A_3u^2+A_4u+A_5 = 0, $

    where

    $ A1=bm2,A2=bm2(r1K2c)bm1E,A3=r1bK(m1E+2cm2)bc(2m1E+cm2)+r2Km2(h2a)dr22Km2bqEK,A4=r1bcK(2m1E+cm2)bc2m1E+r2Km1E(h2a)+r2cKm2(h2a)EK(dr22m1+2bcq),A5=cEKm1(r1bc+r2h2r2a)+bc2qEK.
    $

    When system (1.1) has a positive interior equilibrium point $ E^{*} $, two species coexist, i.e. two species do not go extinct.

    Proof. Direct computation.

    Lemma 3 [18] The equilibrium point $ (u, v) $ is called

    (I) Sink (locally asymptotically stable) if $ |\lambda_{1}| < 1 $ and $ |\lambda_{2}| < 1 $;

    (II) Source (locally unstable) if $ |\lambda_{1}| > 1 $ and $ |\lambda_{2}| > 1 $;

    (III) Saddle if $ |\lambda_{1}| > 1 $ and $ |\lambda_{2}| < 1 $ (or $ |\lambda_{1}| < 1 $ and $ |\lambda_{2}| > 1 $);

    (IV) Non-hyperbolic if $ |\lambda_{1}| = 1 $ or $ |\lambda_{2}| = 1 $.

    Jacobian matrix can be evaluated at $ E_{0}(0, 0) $ as

    $ J(E0)=(er1qm100eah2).
    $
    (2.1)

    The eigenvalues of $ J(E_{0}) $ are $ \lambda_{1} = e^{r_{1}-\frac{q}{m_{1}}} $ and $ \lambda_{2} = e^{a-h_{2}} $. The results regarding dynamical behaviors are listed in Table 2.

    Table 2.  Properties of equilibrium point $ E_{0}(0, 0) $.
    Conditions Eigenvalues Properties
    $ \lambda_{1}=e^{r_{1}-\frac{q}{m_{1}}} $ $ \lambda_{2}=e^{a-h_{2}} $
    $ r_{1}m_{1} < q $ $ a < h_{2} $ $ |\lambda_{1}| < 1 $ $ |\lambda_{2}| < 1 $ Sink
    $ a > h_{2} $ $ |\lambda_{2}| > 1 $ Saddle
    $ a=h_{2} $ $ |\lambda_{2}|=1 $ Non-hyperbolic
    $ r_{1}m_{1} > q $ $ a < h_{2} $ $ |\lambda_{1}| > 1 $ $ |\lambda_{2}| < 1 $ Saddle
    $ a > h_{2} $ $ |\lambda_{2}| > 1 $ Source
    $ a=h_{2} $ $ |\lambda_{2}|=1 $ Non-hyperbolic
    $ r_{1}m_{1}=q $ $ a < h_{2} $ $ |\lambda_{1}|=1 $ $ |\lambda_{2}| < 1 $ Non-hyperbolic
    $ a > h_{2} $ $ |\lambda_{2}| > 1 $ Non-hyperbolic
    $ a=h_{2} $ $ |\lambda_{2}|=1 $ Non-hyperbolic

     | Show Table
    DownLoad: CSV

    From Table 2, we can get the following theorem.

    Theorem 1. When $ r_{1}m_{1} < q $ and $ a < h_{2} $ are satisfied, the trivial equilibrium point $ E_{0}(0, 0) $ is locally asymptotically stable.

    The Jacobian matrix computed at $ E_{1}(\frac{-\alpha+\sqrt{\Delta }}{2}, 0) $ is

    $ J(E1)=(1r1(Δ2α)2K+2qEm2(Δ2α)4m21E2+m22(Δ2α)2+4m1m2E(Δ2α)r2(Δ2α)(Δ2α)+2c0exp[a+r2d(Δ2α)(Δ2α)+2ch2]).
    $

    The eigenvalues of the Jacobian are $ \lambda_{1} = 1-\frac{r_{1}(\Delta^2-\alpha)}{2K}+\frac{2qEm_{2}(\Delta^2-\alpha)}{4m_{1}^2E^2+m_{2}^2(\Delta^2-\alpha)^2+4m_{1}m_{2}E(\Delta^2-\alpha)} $ and $ \lambda_{2} = {\rm exp}[a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c}-h_{2}] $. The properties of semi-trivial equilibrium point $ E_{1}(\frac{-\alpha+\sqrt{\Delta }}{2}, 0) $ are summarized in Table 3.

    Table 3.  Properties of semi-trivial equilibrium point $ E_{1}(\frac{-\alpha+\sqrt{\Delta }}{2}, 0) $.
    Conditions Eigenvalues Properties
    $ \lambda_{1} $ $ \lambda_{2} $
    $ 0 < \frac{r_{1}(\Delta^2-\alpha)}{2K}-\frac{2qEm_{2}(\Delta^2-\alpha)}{4m_{1}^2E^2+m_{2}^2(\Delta^2-\alpha)^2+4m_{1}m_{2}E(\Delta^2-\alpha)} < 2 $ $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c}=h_{2} $ $ |\lambda_{1}| < 1 $ $ |\lambda_{2}|=1 $ Non-hyperbolic
    $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c} < h_{2} $ $ |\lambda_{2}| < 1 $ Sink
    $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c} > h_{2} $ $ |\lambda_{2}| > 1 $ Saddle
    $ \frac{r_{1}(\Delta^2-\alpha)}{2K}-\frac{2qEm_{2}(\Delta^2-\alpha)}{4m_{1}^2E^2+m_{2}^2(\Delta^2-\alpha)^2+4m_{1}m_{2}E(\Delta^2-\alpha)} > 2 $ $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c}=h_{2} $ $ |\lambda_{1}| > 1 $ $ |\lambda_{2}|=1 $ Non-hyperbolic
    $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c} < h_{2} $ $ |\lambda_{2}| < 1 $ Saddle
    $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c} > h_{2} $ $ |\lambda_{2}| > 1 $ Source
    $ \frac{r_{1}(\Delta^2-\alpha)}{2K}-\frac{2qEm_{2}(\Delta^2-\alpha)}{4m_{1}^2E^2+m_{2}^2(\Delta^2-\alpha)^2+4m_{1}m_{2}E(\Delta^2-\alpha)}=2 $ $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c}=h_{2} $ $ |\lambda_{1}|=1 $ $ |\lambda_{2}|=1 $ Non-hyperbolic
    $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c} < h_{2} $ $ |\lambda_{2}| < 1 $ Non-hyperbolic
    $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c} > h_{2} $ $ |\lambda_{2}| > 1 $ Non-hyperbolic

     | Show Table
    DownLoad: CSV

    From Table 3, we can have the following theorem.

    Theorem 2. When $ \frac{r_{1}(\Delta^2-\alpha)}{2K}-\frac{2qEm_{2}(\Delta^2-\alpha)}{4m_{1}^2E^2+m_{2}^2(\Delta^2-\alpha)^2+4m_{1}m_{2}E(\Delta^2-\alpha)} > 2 $ and $ a+\frac{r_{2}d(\Delta^2-\alpha)}{(\Delta^2-\alpha)+2c} > h_{2} $ are satisfied, the semi-trivial equilibrium point $ E_{1}(\frac{-\alpha+\sqrt{\Delta }}{2}, 0) $ is locally unstable.

    Jacobian matrix can be evaluated at $ E_{2}(0, \frac{a-h_{2}}{b}) $ as

    $ J(E2)=(exp[r1r2(ah2)bcqm1]0r2d(ah2)bc1(ah2)).
    $
    (2.2)

    The eigenvalues of the Jacobian are $ \lambda_{1} = {\rm exp}[r_{1}-\frac{r_{2}(a-h_{2})}{bc}-\frac{q}{m_{1}}] $ and $ \lambda_{2} = 1-(a-h_{2}) $ at semi-trivial equilibrium point $ E_{2}(0, \frac{a-h_{2}}{b}) $. The results of dynamical behaviors are listed in Table 4.

    Table 4.  Properties of semi-trivial equilibrium point $ E_{2}(0, \frac{a-h_{2}}{b}) $.
    Conditions Eigenvalues Properties
    $ \lambda_{1}={\rm exp}[r_{1}-\frac{r_{2}(a-h_{2})}{bc}-\frac{q}{m_{1}}] $ $ \lambda_{2}=1-(a-h_{2}) $
    $ r_{1} > \frac{r_{2}(a-h_{2})}{bc}+\frac{q}{m_{1}} $ $ 0 < a-h_{2} < 2 $ $ |\lambda_{1}| > 1 $ $ |\lambda_{2}| < 1 $ Saddle
    $ a-h_{2} > 2 $ $ |\lambda_{2}| > 1 $ Source
    $ a-h_{2}=2 $ $ |\lambda_{2}|=1 $ Non-hyperbolic
    $ r_{1} < \frac{r_{2}(a-h_{2})}{bc}+\frac{q}{m_{1}} $ $ 0 < a-h_{2} < 2 $ $ |\lambda_{1}| < 1 $ $ |\lambda_{2}| < 1 $ Sink
    $ a-h_{2} > 2 $ $ |\lambda_{2}| > 1 $ Saddle
    $ a-h_{2}=2 $ $ |\lambda_{2}|=1 $ Non-hyperbolic
    $ r_{1}=\frac{r_{2}(a-h_{2})}{bc}+\frac{q}{m_{1}} $ $ 0 < a-h_{2} < 2 $ $ |\lambda_{1}|=1 $ $ |\lambda_{2}| < 1 $ Non-hyperbolic
    $ a-h_{2} > 2 $ $ |\lambda_{2}| > 1 $ Non-hyperbolic
    $ a-h_{2}=2 $ $ |\lambda_{2}|=1 $ Non-hyperbolic

     | Show Table
    DownLoad: CSV

    From Table 4, we can get the following theorem.

    Theorem 3 The semi-trivial equilibrium point $ E_{2}(0, \frac{a-h_{2}}{b}) $ is always locally asymptotically stable when $ r_{1} < \frac{r_{2}(a-h_{2})}{bc}+\frac{q}{m_{1}} $ and $ 0 < a-h_{2} < 2 $ are satisfied.

    $ J|_{(u, v)} $ evaluated at the positive equilibrium point $ E^{*}(u^{*}, v^{*}) $ is

    $ J(E)=(1r1uK+r2uv(u+c)2+qEm2u(m1E+m2u)2r2uc+ur2dcv(c+u)21bv).
    $
    (2.3)

    Then characteristic equation of $ J(E^{*}) $ is given by

    $ λ2TrAλ+DetA=0,
    $
    (2.4)

    where

    $ T=TrA=2r1uK+r2uv(u+c)2+qEm2u(m1E+m2u)2bv,D=DetA=[1r1uK+r2uv(u+c)2+qEm2u(m1E+m2u)2](1bv)+r22dcuv(c+u)3.
    $

    Lemma 4 [18] Suppose that $ F(\lambda) = \lambda^{2}-T\lambda+D, $ and $ F(1) > 0 $, $ \lambda_{1} $ and $ \lambda_{2} $ are roots of $ F(\lambda) = 0 $. Then the following results hold true:

    (I) $ \left|\lambda_{1}\right| < 1 $ and $ \left|\lambda_{2}\right| < 1 $ if and only if $ F(-1) > 0 $ and $ D < 1; $

    (II) $ \left|\lambda_{1}\right| < 1 $ and $ \left|\lambda_{2}\right| > 1 $ (or $ \left|\lambda_{1}\right| > 1 $ and $ \left|\lambda_{2}\right| < 1 $) if and only if $ F(-1) < 0; $

    (III) $ \left|\lambda_{1}\right| > 1 $ and $ \left|\lambda_{2}\right| > 1 $ if and only if $ F(-1) > 0 $ and $ D > 1; $

    (IV) $ \lambda_{1} = -1 $ and $ \left|\lambda_{2}\right|\neq1 $ if and only if $ F(-1) = 0 $ and $ D\neq0, 2; $

    (V) $ \lambda_{1} $ and $ \lambda_{2} $ are complex and $ \left|\lambda_{1}\right| = \left|\lambda_{2}\right| = 1 $ if and only if $ T^{2}-4D < 0 $ and $ D = 1. $

    Lemma 5 [11] Let $ E^{*}(u^{*}, v^{*}) $ be the unique positive equilibrium point of system (1.1), then the following propositions hold:

    $ (1.) $ It is a sink if and only if

    $ |T| < D+1 \; \; \; \; {\rm and}\; \; \; \; D < 1. $

    $ (2.) $ It is a source if and only if

    $ |T| < D+1 \; \; \; \; {\rm and}\; \; \; \; D > 1, \; \; \; \; {\rm or}\; \; \; \; |T| > D+1. $

    $ (3.) $ It is a saddle if and only if

    $ T^2 > 4D\; \; \; \; {\rm and}\; \; \; \; |T| > |D+1|. $

    $ (4.) $ It is non-hyperbolic if and only if

    $ |T| = |D+1|, \; \; \; \; {\rm or}\; \; \; D = 1\; \; \; \; {\rm and}\; \; \; \; |T|\leq2. $

    To sum up, we have the following theorem.

    Theorem 4 System (1.1) at the the positive equilibrium point $ E^*(u^{*}, v^{*}) $ is local asymptotically stable when the conditions

    $ 4-\frac{r_{1}u^*(2+bv^*)}{K}+\frac{r_{2}(2-bv^*) u^*v^*}{(c+u^*)^2}+\frac{qEm_{2}u^*(2-bv^*) }{(m_{1}E+m_{2}u^*)^2}-2bv^*+\frac{r_{2}^2dcu^*v^*}{(u^*+c)^3} > 0 $

    and

    $ -\frac{r_{1}u^*(1+bv^*)}{K}+\frac{r_{2}(1-bv^*) u^*v^*}{(c+u^*)^2}+\frac{qEm_{2}u^*(1-bv^*) }{(m_{1}E+m_{2}u^*)^2}-bv^*+\frac{r_{2}^2dcu^*v^*}{(u^*+c)^3} < 0 $

    hold.

    Proof. According to Lemma 3 and Lemma 4, $ E^*(u^{*}, v^{*}) $ is local asymptotically stable if and only if $ F(1) > 0, \; F(-1) > 0 $ and $ D < 0 $, the conclusion of Theorem 4 obtained by calculation holds.

    Lemma 6 [19] Assume that $ u_{t} $ satisfies $ u_{0} > 0 $, and $ u_{t+1}\leq u_{t} \exp[B(1-C u_{t})] $ for $ t\in[t_{1}, \infty) $, where $ C $ is a positive constant. Then $ \lim\limits_{t \to \infty} \sup\ u_{t} \leq \frac{1}{B C} \exp (B-1) $.

    Theorem 5 Every positive solution $ \left\{\left(u_{n}, v_{n}\right)\right\} $ of system (1.1) is uniformly bounded.

    Proof. Suppose that $ \left\{\left(u_{n}, v_{n}\right)\right\} $ be an arbitrary positive solution corresponding to system (1.1). Then, from first part of system (1.1), one has

    $ un+1=unexp[r1(1unK)r2vnc+unqEm1E+m2un]unexp[r1(1unK)r2vnc+un]unexp[r1(1unK)],
    $

    for all $ n = 0, 1, 2, \cdots $. Suppose that $ u_{0} > 0 $, then according to Lemma 6, we gain

    $ \lim\limits_{n \to \infty} \sup\ u_{n} \leq \frac{K}{r_1} \exp (r_1-1): = M_{1}. $

    From the second part of system (1.1), we have

    $ vn+1=vnexp[a+r2dunc+unbvnh2]vnexp[a+r2dunc+unbvn]vnexp[a+r2dM1c+M1bvn].
    $

    Assume that $ v_{0} > 0 $, then using Lemma 6, we obtain

    $ \lim\limits_{n \to \infty} \sup\ v_{n} \leq b \exp (\frac{a(M_1+c)+r_2dM_1}{M_1+c}-1): = M_{2}. $

    That is to say that $ \lim\limits_{n \to \infty} \sup\ (u_{n}, v_{n}) \leq M, $ where $ M = \max\left\{M_{1}, M_{2}\right\} $. This completes the proof.

    The characteristic equation related to system (1.1) at the unique positive interior equilibrium point $ E^{*}(u^*, v^*) $ is

    $ F(\lambda) = \lambda^{2}-T(u^{*}, v^{*}) \lambda+D(u^{*}, v^{*}) = 0, $

    where

    $ T(u,v)=1r1uK+Φ+Ψ,D(u,v)=Ψ[1r1uK+Φ]+Θ,Ψ:=1bv,Θ:=r2dcuv(u+c)3,Φ:=r2uv(u+c)2+qEm2u(m2u+m1E)2.
    $

    Assume that $ T^{2}(u^{*}, v^{*}) > 4D(u^{*}, v^{*}) $, that is,

    $ (1r1uK+Φ+Ψ)2>4Ψ[1r1uK+Θ]+4Θ
    $
    (3.1)

    and $ T(u^{*}, v^{*})+D(u^{*}, v^{*}) = -1, $ that is to say

    $ r1=Ku(1+Ψ)(2+2Ψ+Φ(1+Ψ)+Θ).
    $
    (3.2)

    Then eigenvalue of $ F(\lambda) = 0 $ are $ \lambda_{1} = -1 $ and $ \lambda_{2} = 2+\Phi+\Psi-\frac{r_1u^{*}}{K} $. The condition $ \left|\lambda_{2}\right|\neq1 $ indicates that

    $ Ψ[1r1uK+Φ]+Θ±1.
    $
    (3.3)

    Consider the following set

    $ A_{F} = \left\{(a, b, c, d, K, r_1, r_2, m_1, m_2, h_2, q, E) \in \mathbb{R}_{+}^{12}:(3.1), \text { (3.2) and (3.3) are satisfied }\right\} \text {. } $

    When the perturbation parameter changes within a small field of $ A_{F} $, the system (1.1) will have flip bifurcation at $ E^* $. Let parameters $ (a, b, c, d, K, r_1, r_2, m_1, m_2, h_2, q, E) \in A_{F} $ and consider the following systems:

    $ {un+1=unexp[(r1+r)(1unK)r2vnc+unqEm1E+m2un],vn+1=vnexp[a+r2dunc+unbvnh2],
    $

    where $ r^* $ is a small perturbation parameter and $ \left|r^*\right|\ll1. $

    Let $ x = u-u^{*} $ and $ y = v-v^{*}. $ Then we gain

    $ (xn+1yn+1)=(W11W12W21W22)(xy)+(f(x,y,r)g(x,y,r)),
    $
    (3.4)

    where

    $ f(x,y,r)=W13x2+W14xy+W15y2+W16x3+W17x2y+W18xy2+W19y3+Z1xr+Z2yr+Z3r2+Z4xyr+Z5x2r+Z6y2r+Z7xr2+Z8yr2+Z9r3+O((|x|,|y|,|r|)4),g(x,y,r)=W23x2+W24xy+W25y2+W26x3+W27x2y+W28xy2+W29y3+O((|x|,|y|,|r|)4),W11=1r1uK+r2uv(u+c)2+qEm2(m1E+m2u)2=1+Ω,W12=r2u(c+u),
    $
    $ W13=Ω2r12Kr1uΩ2K+r2v(cu)2(u+c)3+r2uvΩ2(u+c)2+qEm2(m1Em2u)2(m1E+m2u)3+qEm2uΩ2(m1E+m2u)2,W14=r2c+u+r1r2u(u+c)Kr2qEm2(c+u)(m1E+m2u)2,W15=r21u2(c+u)2,W16=Ωu+Ω26r1Ω3Kr1uΩΩu6Kr2cv2(u+c)+r2cΩv6(c+u)3c2u6(c+u)4qE2m1m22(m1E+m2u)4uΩ6(c+u)3+r2v(cu)Ω6(u+c)3+r2uvΩΩu6(c+u)2+qEm2(m1Em2u)Ω6(m1E+m2u)3+qEm2uΩΩu6(m1E+m2u)2+qE2m1m2Ω6(m1E+m2u)3qEm22uΩ6(m1E+m2u)3qEm22(m1Em2u)6(m1E+m2u)4,W17=r1r2(1+u)2K(c+u)+r2u2(c+u)2+r2(cu)r22uv2(c+u)3r21r2u22K2(c+u)r1r2u2K(c+u)2+r1r22u2v(u+c)3+2r22u2vr22v(cu)2(c+u)4r32u2v22(c+u)5r2qEm22(u+c)(m1E+m2u)2+r2qEm1u(1+u)2(c+u)2(m2u+m1E)2r22qEm2uv(1+u)2(c+u)3(m2u+m1E)2r2qEm2(m1Em2u)2(c+u)(m2u+m1E)3+r1r2qEm2(1+u)u2K(c+u)(m1Em2u)2r2q2m22E2u2(c+u)(m2u+m1E)4,W18=r222(c+u)2r22u(c+u)3+r32uv2(u+c)4r1r22u2K(c+u)2+r22qEm22(u+c)2(m1E+m2u)2,W19=r32u6(c+u)3,Z1=12uKr1(Ku)uK2+r2uv(Ku)K(u+c)2+qEm2(Ku)K(m1E+m2u)2,Z2=r2(Ku)uK(c+u),Z3=u(Ku)22K2,Z6=r22u(Ku)2K(u+c)2,Z4=r2(2uK)K(c+u)+r1r2u(Ku)K2(u+c)r2qEm2(Ku)K(c+u)(m1E+m2u)2,Z5=u+12K+r21(Ku)u22K3+r1u(3u+1)r1K(1+u)2K2+r2v(cu)(Ku)2K(c+u)3+qEm2(uK)(m1Em2u)2K(m1E+m2u)3+r2(K3u)uv2K(c+u)2+r2qEm2uv(Ku)(1+u)2K(c+u)2(m1E+m2u)2r1r2(Ku)u2vK2(c+u)2+r22(Ku)u2v22K(c+u)4r1qEm2u(Ku)(1+u)2K2(m1E+m2u)2+q2E2m22(Ku)2K(m1E+m2u)4+qEm2(K2uu2)2K(m1E+m2u)2,Z7=12u2K3(Ku)u2K2r1u(Ku)22K3+r2uv(Ku)22K2(u+c)2+qEm2(Ku)22K2(m1E+m2u)2,Z8=r2u(Ku)22K2(c+u),Z9=u(c+u)36K3,W21=r2dcv(c+u)2,W22=1bv,
    $
    $ W23=r2dcv(c+u)3+r22d2c22(u+c)4,W24=r2dc(1bv)(u+c)2,W25=b(bv2)2,W26=r2dcv(c+u)4r22d2c2v2(c+u)5+r32d3c3v36(c+u)6,W27=r2dc(bv1)(c+u)3+r22d2c2v(2v)2(c+u)4,W28=r2bdc(c+u)2+r2b2dcv2(c+u)2,W29=b2(3bv)6.
    $

    Construct a nonsingular matrix $ D_{1} $ and consider the following translation:

    $ (xy)=D1(uv),
    $
    (3.5)

    where

    $ D_{1} = \left(W12W121W11λ2W11
    \right). $

    Taking $ D_{1}^{-1} $ on both sides of Eq (3.5), we obtain

    $ (uv)=(100λ2)(uv)+(f1(x,y,r)g1(x,y,r)),
    $
    (3.6)

    where

    $ f1(x,y,r)=[W13(λ2W11)W12W23]x2W12(λ2+1)+[W14(λ2W11)W12W24]xyW12(λ2+1)+Z1(λ2W11)xrW12(λ2+1)+[W15(λ2W11)W12W25]y2W12(λ2+1)+[W16(λ2W11)W12W26]x3W12(λ2+1)+Z2(λ2W11)yrW12(λ2+1)+[W17(λ2W11)W12W27]x2yW12(λ2+1)+[W18(λ2W11)W12W28]xy2W12(λ2+1)+Z3(λ2W11)r2W12(λ2+1)+[W19(λ2W11)W12W29]y3W12(λ2+1)+Z4(λ2W11)xyrW12(λ2+1)+Z5(λ2W11)x2rW12(λ2+1)+Z6(λ2W11)y2rW12(λ2+1)+Z7(λ2W11)xr2W12(λ2+1)+Z8(λ2W11)yr2W12(λ2+1)+Z9(λ2W11)r3W12(λ2+1)+O((|x|,|y|,|r|)4),g1(x,y,r)=[W13(λ2+W11)+W12W23]x2W12(λ2+1)+[W14(λ2+W11)+W12W24]xyW12(λ2+1)+Z1(λ2+W11)xrW12(λ2+1)+[W15(λ2+W11)+W12W25]y2W12(λ2+1)+[W16(λ2+W11)+W12W26]x3W12(λ2+1)+Z2(λ2+W11)yrW12(λ2+1)+[W17(λ2+W11)+W12W27]x2yW12(λ2+1)+[W18(λ2+W11)+W12W28]xy2W12(λ2+1)+Z3(λ2+W11)r2W12(λ2+1)+[W19(λ2+W11)+W12W29]y3W12(λ2+1)+Z4(λ2+W11)xyrW12(λ2+1)+Z5(λ2+W11)x2rW12(λ2+1)+Z6(λ2+W11)y2rW12(λ2+1)+Z7(λ2+W11)xr2W12(λ2+1)+Z8(λ2+W11)yr2W12(λ2+1)+Z9(λ2+W11)r3W12(λ2+1)+O((|x|,|y|,|r|)4),x=W12(u+v),y=(λ2+W11)v(1+W11)u.
    $

    The center manifold theorem $ W^{c} $ is applied in a small field of $ r^* = 0 $ at the equilibrium point $ E_{0} $. Then there exists $ W^{c}(0) $ as follows:

    $ W^{c}(0) = \left\{(x, y, r^*)\in \mathbb{R}^{3}: y(x, r^*) = e_{0} r^*+e_{1} x^{2}+e_{2} x r^*+e_{3} r^{*2}+O(\left(|x|+\left|r^*\right|\right)^{3})\right\} $

    and satisfies

    $ H(y(x, r^*)) = y\left(-u+f_{1}(x, y(x, r^*), r^*)\right)-\lambda_{2}y(x, r^*)-g_{1}(x, y(x, r^*), r^*) = 0, $

    and we have

    $ e0=0,e1=[W13(1+W11)+W12W23]W12[W14(1+W11)+W12W24](1+W11)1λ22+[W15(1+W11)+W12W25](1+W11)2(1λ22)W12,e2=[W12Z1Z2(1+W11)](1+W11)(1λ2)2,e3=Z3(1+W11)W12(1λ2)2.
    $

    Therefore, consider the following map on the center manifold $ W^{c}(0) $:

    $ G:x\rightarrow -x+s_{1}x^{2}+s_{2}xr^*+s_{3}x^{2} r^*+n_{4}xr^{* 2}+s_{5}x^{3}+O((|x|+\left|r^*\right|)^{4}), $

    where

    $ s1=[W13(λ2W11)W12W23]W121+λ2[W14(λ2W11)W12W24](1+W11)1+λ2+[W15(λ2W11)W12W25](1+W11)2W12(1+λ2),s2=Z1(λ2W11)1+λ2Z2(λ2W11)(1+W11)W12(1+λ2),s3=[W13(λ2W11)W12W23]2e2W121+λ2+[W14(λ2W11)W12W24](λ2W11)e21+λ22[W15(λ2W11)W12W25](1+W11)(λ2W11)e2W12(1+λ2)+Z1(λ2W11)e11+λ2+(λ2W11)2e1W12(1+λ2)+W12(λ2W11)Z51+λ2+Z6(λ2W11)(1+W11)2(1+λ2)W12[W14(λ2W11)W12W24]e2(1+W11)+Z4(λ2W11)(1+W11)1+λ2,s4=[W13(λ2W11)W12W23]2e2W121+λ2+[W14(λ2W11)W12W24](λ2W11)e3λ2+1[W14(λ2W11)W12W24](1+W11)e3λ2+1+(λ2W11)(Z1e2+Z7)λ2+1[W15(λ2W11)W12W25](1+W11)(λ2W11)e3λ2+1+(λ2W11)2Z2e2W12(λ2+1)Z8(λ2W11)(W11+1)W12(λ2+1),
    $
    $ s5=[W16(λ2W11)W12W26]W2121+λ2[W17(λ2W11)W12W27](1+W11)W12λ2+12[W15(λ2W11)W12W25](1+W11)(λ2W11)e1(1+λ2)W12+[W13(λ2W11)W12W23]2e2W121+λ2[W14(λ2W11)W12W24](1+W11)e1λ2+1+[W14(λ2W11)W12W24](λ2W11)e1λ2+1+[W18(λ2W11)W12W28](1+W11)2λ2+1[W19(λ2W11)W12W29](1+W11)3W12(λ2+1).
    $

    By flip bifurcation, we define two non-zero real numbers $ \delta_1 $ and $ \delta_2 $, where

    $ \delta_{1} = \left.\left(\frac{\partial^{2} G}{\partial x \partial r^*}+\frac{1}{2} \frac{\partial G}{\partial r^*} \frac{\partial^{2} G}{\partial x^{2}}\right)\right|_{(0, 0)} = s_{2}, \; \; \delta_{2} = \left.\left(\frac{1}{6} \frac{\partial^{3} G}{\partial x^{3}}+\left(\frac{1}{2} \frac{\partial^{2} G}{\partial x^{2}}\right)^{2}\right)\right|_{(0, 0)} = s_{1}^{2}+s_{5}. $

    Based on the above analysis, the following theorem can be obtained.

    Theorem 6 System (1.1) undergoes a flip bifurcation at the positive internal equilibrium point $ E^{*}(u^{*}, v^{*}) $ if $ \delta_1\neq0 $, $ \delta_2\neq0 $ are satisfied and when parameter $ r^* $ changes within a small field of $ r_1 $. Moreover, if $ \delta_{2} > 0 $ (resp., $ \delta_{2} < 0 $), then the period-two orbits that bifurcate from equilibrium point $ E^{*}(u^{*}, v^{*}) $ are stable (resp., unstable).

    Consider the characteristic equation at $ E^* $, then $ \mathbb{F}(\lambda) = 0 $ has two complex conjugate roots with modulus one if the following conditions are satisfied:

    $ (1r1uK+Φ)Ψ+Θ=1
    $
    (3.7)

    and

    $ |T|=|1r1uK+Φ+Ψ|=|1+D|<2.
    $
    (3.8)

    Let

    $ A_{NS} = \{(a, b, c, d, K, r_1, r_2, m_1, m_2, h_2, q, E) \in \mathbb{R}_{+}^{12}: (3.7)\; {\rm and}\; (3.8)\; {\rm are\; satisfied}\}. $

    When the parameter changes in a small field of $ A_{NS} $, system (1.1) will have Neimark-Sacker bifurcation at the unique positive equilibrium point $ E^{*}(u^{*}, v^{*}). $ Select parameter $ (a, b, c, d, K, r_1, r_2, \\ m_1, m_2, h_2, q, E)\in A_{NS} $ and analyze the following system:

    $ {un+1=unexp[(r1+¯r)(1unK)r2vnc+unqEm1E+m2un],vn+1=vnexp[a+r2dunc+unbvnh2],
    $

    where $ \overline{r} $ is a small perturbation parameter and $ \left|\overline{r}\right|\ll1. $

    Let $ x = u-u^{*} $ and $ y = v-v^{*}. $ Then we gain

    $ {xn+1=W11x+W12y+W13x2+W14xy+W15y2+W16x3+W17x2y+W18xy2+W19y3+O((|x|+|y|)4),yn+1=W21x+W22y+W23x2+W24xy+W25y2+W26x3+W27x2y+W28xy2+W29y3+O((|x|+|y|)4),
    $
    (3.9)

    where $ W_{ij}(i = 1, 2, \; 1\leq j\leq 9) $ are given in (3.6) by substituting $ r_{1} $ for $ r_{1}+\overline{r} $.

    The characteristic equation of system (3.9) at $ (x, y) = (0, 0) $ is as follows:

    $ λ2T(¯r)λ+D(¯r)=0,
    $

    where

    $ T(¯r)=1(r1+¯r)uK+Φ+Ψ,D(¯r)=(1(r1+¯r)uK+Φ)Ψ+Θ.
    $

    Since parameters $ (a, b, c, d, K, r_1, r_2, m_1, m_2, h_2, q, E)\in A_{NS} $, the roots of the characteristic equation are

    $ λ1,2=T(¯r)2±i24D(¯r)D2(¯r)
    $

    and we have

    $ |\lambda_{1, 2}| = \sqrt{D(\overline{r})}. $

    Suppose that

    $ L = \frac{d|\lambda_{1, 2}|}{d\overline{r}}\Big|_{\overline{r} = 0} = -\frac{\Psi u^*}{2K\sqrt{D(0)}}\neq0. $

    In addition, it is required that $ \overline{r} = 0, \; \lambda^{l}_{1, 2} \neq 1\; (l = 1, 2, 3, 4) $ which is equal to $ T(0)\neq $ -2, 0, -1, 2. Because $ (a, b, c, d, K, r_1, r_2, m_1, m_2, h, q, E)\in A_{NS} $, thus $ T(0)\neq $ -2, 2. We only require $ T(0) \neq 0, -1, $ so that

    $ 1r1uK+Φ+Ψ0and2r1uK+Φ+Ψ0.
    $
    (3.10)

    Let $ \eta = \frac{T(0)}{2}, \; \omega = \frac{\sqrt{4D(0)-T^{2}(0)}}{2} $, we use the following transformation:

    $ [xy]=D2[uv]=[W120ηW11ω][uv],
    $

    and system (3.10) becomes into

    $ [ut+1vt+1]=[ηωωη][utvt]+[¯f(u,v)¯g(u,v)],
    $

    where

    $ ¯f(u,v)=W13W12x2+W14W12xy+W15W12y2+W16W12x3+W17W12x2y+W18W12xy2+W19W12y3+O((|x|+|v|)4),¯g(u,v)=[W13(ηW11)ωW12W23ω]x2+[W14(ηW11)ωW12W24ω]xy+[W15(ηW11)ωW12W25ω]y2+[W16(ηW11)ωW12W26ω]x3+[W17(ηW11)ωW12W27ω]x2y+[W18(ηW11)ωW12W28ω]xy2+[W19(ηW11)ωW12W29ω]y3+O((|x|+|y|)4),x=W12u,y=(ηW11)uωv.
    $

    System (3.9) undergoes the Neimark-Sacker bifurcation if the following quantity $ \Lambda $ is not zero

    $ Λ=Re[(12λ1)λ221λ1P11P12]12|P11|2|P21|2+Re(λ2P22),
    $
    (3.11)

    where

    $ P11=14[(¯fuu+¯fvv)+i(¯guu+¯gvv)],P12=18[(¯fuu¯fvv+2¯guv)+i(¯guu¯gvv2¯fuv)],P21=18[(¯fuu¯fvv2¯guv)+i(¯guu¯gvv+2¯fuv)],P22=116[(¯fuuu+¯fuvv+¯guuv+¯gvvv)+i(¯guuu+¯guvv¯fuuv¯fvvv)].
    $

    If $ \Lambda\neq0 $, Neimark-Sacker bifurcation will occur in system (1.1), and the following theorem holds:

    Theorem 7 System (1.1) undergoes a Neimark-Sacker bifurcation at the positive equilibrium point $ E^{*}(u^{*}, v^{*}) $ if conditions (3.10) are satisfied and $ \Lambda\neq0 $. Moreover, if $ \Lambda < 0 $(resp., $ \Lambda > 0 $), an attracting (resp., repelling) invariant closed curve bifurcates from the steady state for $ r_1 > \overline{r} $ (resp., $ r_1 < \overline{r} $).

    In this section, we will adopt the feedback control method[23,24,25] to stabilize the chaotic orbit at an unstable equilibrium point by adding a feedback control term to the system (1.1). Therefore, system (1.1) makes the following form:

    $ {un+1=unexp[r1(1unK)r2vnc+unqEm1E+m2un]h(un,vn)=˜f(un,vn),vn+1=vnexp[a+r2dunc+unbvnh2]=˜g(un,vn),
    $
    (4.1)

    where $ h(u_{n}, v_{n}) = q_{1}(u_{n}-u^{*}) + q_{2}(v_{n}-v^{*}) $ is feedback controlling force, $ q_{1} $ and $ q_{2} $ are feedback gains. Furthermore, $ \widetilde{f}(u^{*}, v^{*}) = u^{*} $, and $ \widetilde{g}(u^{*}, v^{*}) = v^{*} $.

    The Jacobian matrix corresponding to system (4.1) at interior equilibrium point $ (u^{*}, v^{*}) $ is as follows:

    $ J(u^{*}, v^{*}) = \left[W11q1W12q2W21W22
    \right]. $

    Thus, the characteristic equation related to $ A(u^{*}, v^{*}) $ is:

    $ λ2(W11+W22q1)λ+(W11q1)W22(W12q2)W21=0.
    $
    (4.2)

    Let $ \lambda_{1} $ and $ \lambda_{2} $ be the eigenvalues of characteristic equation (4.2), then

    $ λ1+λ2=W11+W22q1,λ1λ2=(W11q1)W22(W12q2)W21.
    $
    (4.3)

    Next, we must solve equations $ \lambda_{1} = \pm1 $ and $ \lambda_{1}\lambda_{2} = 1 $ to gain the critical stability line. At the same time, it also ensures that the absolute value $ \lambda_{1} $ and $ \lambda_{2} $ are less than one.

    Suppose that $ \lambda_{1}\lambda_{2} = 1 $, then we gain

    $ L^{'}:W_{11}W_{22}-W_{12}W_{21}-1 = W_{22}q_{1}-W_{21}q_{2}. $

    Assume that $ \lambda_{1} = 1 $, then we have

    $ L^{''}:W_{11}+W_{22}-W_{11}W_{22}+W_{12}W_{21}-1 = (1-W_{22})q_{1}+W_{21}q_{2}. $

    Assume that $ \lambda_{1} = -1 $, then we obtain

    $ L^{'''}:W_{11}+W_{22}+W_{11}W_{22}-W_{12}W_{21}+1 = (1+W_{22})q_{1}-W_{21}q_{2}. $

    Thus, the stable eigenvalues lie within the triangular region with the boundaries of the straight lines $ L^{'}, \; L^{''}, \; L^{'''} $. In addition, when the control parameters $ q_{1} $ and $ q_{2} $ take values in the triangular region, system (4.1) will not create chaos phenomena.

    For the sustainable use of biological resources and the protection of the natural environment on which human beings depend. Therefore, the development of renewable resources must be reasonable and proportionate. Under the premise of achieving sustainable development of biological resources, pursue maximum yield or best economic benefits. The biological and economic equilibrium combines to form the bioeconomic equilibrium. Biological equilibrium[14,15,16,17] can be obtained by solving $ u_{n+1} = u_{n}, \; v_{n+1} = v_{n} $ and when the economic rent is equal to zero (meaning that the total income equals the total cost), the economic equilibrium can be obtained. If $ h_1, \; p $ are the cost of harvest per unit and unit price of the prey population, respectively, then the total cost is $ TC = h_1E $ and total income as $ TR = \frac{pqE}{m_1E+m_2u_n} $. Then the economic rent at the moment $ t $ can be expressed as $ \Xi = TR-TC = (\frac{pq}{m_1E+m_2u_n}-h_1)E $. The bioeconomic equilibrium can be obtained by solving the following simultaneous equations:

    $ {r1(1unK)r2vnc+unqEm1E+m2un=0,a+r2dunc+unbvnh2=0,pqm1E+m2unh1=0.
    $
    (4.4)

    At present, if $ \frac{pq}{m_1E+m_2u_n} < h_1 $, then we stop capturing because the cost of harvesting is greater than the revenue which also means losses. Similarly, if $ \frac{pq}{m_1E+m_2u_n} > h_1 $, then we will continue to capture because of the harvest cost less than revenue which means profit. In order to find the bioeconomic equilibrium point $ (u^*, v^*, E^*) $ from system (4.4), we can perform the following steps: first, we solve the value of $ u^* $ from the third equation, and then substitute the value of $ u^* $ into the second equation to get the value of $ v^* $. Finally, substitute the values of $ u^* $ and $ v^* $ into the first equation to get the value of $ E^* $.

    Next, we aim to maximize net income while maintaining ecological balance. Define the net income function $ J = \sum {\rm exp}(-\delta t)(\frac{pq}{m_1E+m_2u_n}-h_1)E $, where $ \delta $ is the discount rate and $ {\rm exp}(-\delta t) $ is the discount factor. At the same time, we use the discrete Pontryagin maximum principle[21] to acquire the optimal capture effort. So the optimal capture problem is

    $ \max \sum\limits_{t = 1}^{n}{{\rm exp}(-\delta t)\left(\frac{pq}{m_1E_t+m_2u_t}-h_1\right)E_t}, $
    $ s.t. {ut+1=utexp[r1(1utK)r2vtc+utqEtm1Et+m2ut],vt+1=vtexp[a+r2dutc+utbvth2],u1=u0,v1=v0and0EtEmax,fort=0,1,,N1,
    $

    where $ u_t, \; v_t $ are state variables and $ E_t $ is the control variable. The Hamiltonian function of the correlation at this moment is

    $ Ht=exp(δt)(pqm1Et+m2uth1)Et+λ1(t+1)[utexp(r1(1utK)r2vtc+utqEtm1Et+m2ut)]+λ2(t+1)[vtexp(a+r2dutc+utbvth2)].
    $

    Where $ \lambda_{1(t+1)} $ and $ \lambda_{2(t+1)} $ are adjoint variables. In addition, the necessary condition for the optimal problem is that $ \frac{\partial H_t}{\partial u_t} = 0 $, $ \frac{\partial H_t}{\partial v_t} = 0 $ and $ \frac{\partial H_t}{\partial e_t} = 0 $ are valid at the same time. Optimal harvest $ E^{*}_{t} $ is available at optimal population size level $ (u^{*}_{t}, v^{*}_{t}) $.

    $ Htut=exp(δt)pqm2Et(m1Et+m2ut)2+λ1(t+1)[1r1utK+r2utvt(c+ut)2+qm2utEt(m1Et+m2ut)2]+λ2(t+1)r2dcvt(c+ut)2=0,Htvt=λ1(t+1)r2utc+ut+λ2(t+1)(1bvt)=0,HtEt=exp(δt)(pqm2Et(m1Et+m2ut)2h1)λ1(t+1)qm2u2t(m1Et+m2ut)2=0.
    $

    As a consequence, we first solve the value of $ \lambda_{1(t+1)} = \frac{{\rm exp}(-\delta t)[pqm_2E^{*}_{t}-h_1(m_1E^{*}_{t}+m_2u^{*}_{t})^2]}{qm_2u^{*2}_{t}} $ from $ \frac{\partial H_t}{\partial E^{*}_{t}} = 0 $, and then substitute the value of $ \lambda_{1(t+1)} $ into the second equation $ \frac{\partial H_t}{\partial v^{*}_{t}} = 0 $ to get the value of $ \lambda_{2(t+1)} = \frac{{\rm exp}(-\delta t)[pqm_2E^{*}_{t}-h_1(m_1E^{*}_{t}+m_2u^{*}_{t})^2]r_2}{qm_2u^{*}_{t}(c+u^{*}_{t})(1-bv^{*}_{t})} $. Finally, substitute the values of $ \lambda_{1(t+1)} $ and $ \lambda_{2(t+1)} $ into the first equation $ \frac{\partial H_t}{\partial u^{*}_{t}} = 0 $ to obtain the value of $ E^{*}_{t} $.

    This section will show the bifurcation diagram, phase diagram and maximum Lyapunov exponent diagram of system (1.1) to verify the correctness of theoretical analysis.

    Suppose that the parameters $ (a, \; b, \; c, \; d, \; E, \; K, \; r_2, \; q, \; m_1, \; m_2, \; h_2) = (0.4, \; 0.2, \; 2, \; 0.1, \; 0.4, \; 5, \; 0.5, \; 0.1, \; 0.1, \; 0, \; 0.01)\in A_F $, $ r_1 $ as the bifurcation parameter, and the initial value is (3, 1). Meanwhile, when $ r_1 < 1.5 $, the interior equilibrium point does not exist; when $ r_1 > 1.5 $, there is a unique interior equilibrium point, and when $ r_1 = 1.5 $, the system (1.1) has a transcritical bifurcation at the boundary equilibrium point $ E_2 $. According to Theorem 6, when $ r_1 = 3.32 $, the system (1.1) will have a flip bifurcation at the interior equilibrium point (3.188, 2.104). The bifurcation diagram and the maximum Lyapunov exponent diagram are shown from Figure 1. Combined with Figures 1 and 2, when $ r_1 < 3.32 $, the interior equilibrium point is stable. However, when $ r_1 > 3.32 $, the interior equilibrium loses its stability, and orbits with periods of 2, 4, 8 appear. As the $ r_1 $ increases, the maximum Lyapunov exponent value is greater than zero, and it can be seen from Figure 1(c) and Figure 2(c), (f) that system (1.1) will generate chaos.

    Figure 1.  Flip bifurcation and MLE diagram.
    Figure 2.  Phase and solution diagram related to Figure 1 when $ r_1 $ takes different values.

    Assume that the parameters $ (a, \; b, \; c, \; d, \; E, \; K, \; r_2, \; q, \; m_1, \; m_2, \; h_2) = (1.7, \; 3.5, \; 1.2, \; 0.3, \; 2, \; 3, \; 2.6, \; 0.1, \; 5, \; 2, \; 0.01) $, $ r_1 $ as the bifurcation parameter, and the initial value is (2, 3). The bifurcation diagram and the maximum Lyapunov exponent diagram are shown from Figure 3. Combined with Figures 4 and 5, it is clear from the figure that for smaller values of $ r_1 $ the system (1.1) is stable, with the increase of $ r_1 $ the system (1.1) stability disappears and a flip bifurcation with a period of 2 occurs, which subsequently disappears and tends to stabilize. Furthermore, when $ 2.43 < r_1 < 2.59 $, the equilibrium point is stable. However, when $ r_1 > 2.59 $, the system loses its stability, and a stable invariant loop appears. At this moment, the system (1.1) produces Neimark-Sacker bifurcation and periodic solution. When $ r_1 $ increases, system (1.1) produces quasi-periodic solutions and chaotic phenomena. As the $ r_1 $ continues to increase, the maximum Lyapunov exponent value is greater than zero, the system (1.1) will generate chaos.

    Figure 3.  Mixed bifurcation and MLE diagram.
    Figure 4.  Phase diagram related to Figure 3 when $ r_1 $ takes various values.
    Figure 5.  Solution diagram corresponding to Figure 4 when $ r_1 $ takes various values.

    Considering the parameter values $ (a, \; b, \; c, \; d, \; E, \; K, \; r_2, \; q, \; m_1, \; m_2, \; h_2) = (1.7, \; 3, \; 1.2, \; 0.3, \; 1.5, \; 3, \; 1.9, \; 0.1, \; 1, \; 2, \; 0.01)\in A_{NS} $ with the initial value is (2, 3), and $ r_1 $ as the bifurcation parameter. According to Theorem 7, when $ r_1 = 2.518 $, the system (1.1) has Neimark-Sacker bifurcation at the interior equilibrium point (2.52, 0.69). Figure 6 is the bifurcation and MLE graph corresponding to $ u $ and $ v $ with $ r_1 \in [2.2, \; 3.2] $, and Figure 7 is the phase graph related to Figure 6(a). It can be seen from Figures 6 and 7 that when $ r_1 < 2.518 $, the equilibrium point is stable; when $ r_1 > 2.518 $, the equilibrium point loses its stability, and a stable invariant loop appears. At this moment, system (1.1) arises a periodic solution. When $ r_1 $ increases, system (1.1) generates quasi-periodic solutions and chaotic phenomena. Furthermore, as the $ r_1 $ continues to increase, the maximum Lyapunov exponent value is greater than zero, the system (1.1) will generate chaos.

    Figure 6.  Neimark-Sacker bifurcation and MLE diagram.
    Figure 7.  Phase diagram related to Figure 6 when $ r_1 $ takes various values.

    To verify the chaotic control theory, we analyze Figure 6 and its numerical simulation parameters. In Figure 6(c) when the bifurcation parameter $ r_1 = 3.1 $, the maximum Lyapunov exponent value is greater than zero, system (1.1) will produce chaos. When the $ q_{1} $ and $ q_{2} $ are controlled in the triangular region surrounded by three straight lines $ L^{'} $, $ L^{''} $, and $ L^{'''} $ (see Figure 8), the chaos generated by system (4.1) will be controlled near the equilibrium point and become asymptotically stable state.

    Figure 8.  The bounded region for the eigenvalues related to the controlled system (4.1) in the $ (q_{1}, q_{2}) $ plane.

    Considering the parameter values $ (a, \; b, \; c, \; d, \; r_1, \; K, \; r_2, \; q, \; m_1, \; m_2, \; h_2) = (0.95, \; 1.5, \; 1.2, \; 0.3, \; 0.7, \\\; 5, \; 0.5, \; 0.15, \; 1, \; 1, \; 0.1) $ with the initial value is (3, 2), and $ E $ as the bifurcation parameter. At this time, the bifurcation phenomenon of system (1.1) will not occur. As the degree of capture effort $ E $ increases, the population density of prey and predator will continue to decrease and will not tend to 0 (see Figure 9).

    Figure 9.  Bifurcation diagram.

    In this paper, we study the stability and bifurcation of equilibrium points in a discrete predator-prey model with Michaelis-Menten type harvesting. The stability analysis indicates that the model has a trivial equilibrium point, two positive boundary equilibrium points and the boundary equilibrium point is always unstable. The bifurcation analysis shows that when $ r_1 = 1.5 $, the boundary equilibrium point $ E_2 $ will have a transcritical bifurcation, and when the coexistence equilibrium $ E^* $ exists and loses stability, system (1.1) will have a flip bifurcation (see Figure 1). System (1.1) has, in addition, Neimark-Sacker bifurcation occur at the interior equilibrium point $ E^* $ when bifurcation parameter $ r_1 $ changes in $ A_{NS} $ small ranges (see Figure 6). Numerical simulation reveals that when the internal growth rate of prey $ r_1 $ gradually increases, system (1.1) will produce periodic, quasi-periodic windows and chaos.

    Finally, we analyze chaos control theory and the existence of bioeconomic equilibrium points. In order to maximize profit in a finite time, we built an optimal control problem with the harvest effort as the control parameter, and theoretically obtained the optimal value of the control variable (harvest effort). As a result, we detected that harvesting efforts for prey and predator populations had a specific value that maximizes net income.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This work was supported by NNSF of China (No.12161079), the Northwest Normal University Graduate Research Grant Project (No.2022KYZZ-S114), and the Gansu Province Innovation Star Project (No.2023CXZX-325).

    The authors declare there is no conflict of interest regarding the publication of this paper.

    [1] P. S. Adler and S. W. Kwon, Social Capital: Prospects for a new concept, The Academy of Management Review, 27 (2002), 17-40. doi: 10.2307/4134367
    [2] J. Alonso, Pareto, $1^{st}$ edition, Edicol, México, 1977.
    [3] preprint, arXiv:1305.0215v3.
    [4] F. Alvaredo, A. B. Atkinson, T. Piketty and E. Saez, The top 1 percent in international and historical perspective, Journal of Economic Perspectives, 27 (2013), 3-20. doi: 10.1257/jep.27.3.3
    [5] G. Álvarez, Caos/Complejidad, Fractales e Identidades sociales, Razón y Palabra, 79 (2012). Available from: http://www.redalyc.org/articulo.oa?id=199524411055.
    [6] R. Aron, La Sociología Alemana Contemporánea, Paidós, Buenos Aires, 1965.
    [7] P. Ball, Critical Mass, Farrar, Straus and Giroux, New York, 2004.
    [8] A.-L. Barabási and R. Albert, Emergence of scaling in random networks, Science, 286 (1999), 509-512. doi: 10.1126/science.286.5439.509
    [9] J. A. Barbier, Elite and cadres in bourbon Chile, The Hispanic American Historical Review, 52 (1972), 416-435. doi: 10.2307/2512001
    [10] J. Bascompte and B. Luque, Evolución y Complejidad, Publicacions de la Universitat de Valéncia, Valéncia, 2012.
    [11] J. Bascompte and P. Jordano, Plant-animal mutualistic networks: the architecture of biodiversity, Anual Review of Ecology, Evolution and Systematics, 38 (2007), 567-593. doi: 10.1146/annurev.ecolsys.38.091206.095818
    [12] R. Bendix, Max Weber and Intellectual Portrait, Doubleday & Company, Inc., New York, 1962.
    [13] L. von Bertalanffy, Teoría General de los Sistemas, México, 1991. Available from: http://es.scribd.com/doc/135617671/TGS-Bertalanffy-pdf?
    [14] V. D. Blondel, J.-L. Guillaume, R. Lambiotte and E. Lefebvre, Fast unfolding of communities in large networks, Journal of Statistical Mechanics: Theory and Experiments, 2008 (2008), P10008.
    [15] M. Buchanan, Ubiquity, Three River Press, New York, 2000.
    [16] J. P. Cárdenas, M. Mouronte, J. C. Losada and R. M. Benito, Compatibility as underlying mechanism behind the evolution of networks, Physica A, 389 (2010), 1789-1798.
    [17] R. Dahl, La Poliarquía: Partcipación y Oposición, Tecnos, 2009.
    [18] R. Dahrendorf, La Ley y el Orden, $1^{st}$ edition, Editorial Civitas, 1994.
    [19] S. Deploigé, El Conflicto de la Moral y de la Sociología, McGraw-Hill, México, 1920.
    [20] S. N. Dorogovtsev and J. F. F. Mendes, Evolution of Networks. From Biological Nets to the Internet and WWW, Oxford University Press, New York, 2003. doi: 10.1093/acprof:oso/9780198515906.001.0001
    [21] E. Durkheim, Las Reglas del Método Sociológico, Dedalo, Buenos Aires, 1964.
    [22] P. Erdős and A. Rényi, On random graphs, Publicationes Mathematicae (Debrecen), 6 (1959), 290-297.
    [23] M. Foucault, Microfísica del Poder, $2^{nd}$ edition, La Piqueta, Madrid, 1980.
    [24] O. González del Cardenal, El Poder y la Conciencia. Rostros Personales Frente a Poderes An\'onimos, $2^{nd}$ edition, Espasa Calpe, Madrid, 1985.
    [25] M. S. Granovetter, The strength of weak ties, American Journal of Sociology, 78 (1973), 1360-1380. doi: 10.1086/225469
    [26] F. Harary, R. Norman and D. Cartwright, Structural Models, Wiley, New York, 1965.
    [27] P. Klimek, R. Lambiotte and S. Thurner, Opinion formation in laggard societies, Europhysics Letters, 82 (2008), 28008, 6pp. doi: 10.1209/0295-5075/82/28008
    [28] W. Kornhauser, The Politics of Mass Society, Routledge, New York, 1959.
    [29] E. Laszlo, La Gran Bifurcación. Crisis y Oportunidad; Anticipación del Nuevo Paradigma que Está Tomando Forma, Gedisa, Barcelona, 1990.
    [30] G. Lenski, Poder y Privilegio. Teoría de la Estratificación Social, Paidós, Buenos Aires, 1969.
    [31] R. López, E. Figueroa and P. Gutiérrez, La parte del León: Nuevas estimaciones de la participación de los súper ricos en el ingreso de Chile, Universidad de Chile, Departamento de Economía. Serie de Documentos de Trabajo, 2013. Available from: http://www.econ.uchile.cl/uploads/publicacion/306018fadb3ac79952bf1395a555a90a86633790.pdf.
    [32] C. Maldonado, Complejidad y ciencias sociales. el problema de la medición de los sistemas sociales humanos, Complejidad de las ciencias y ciencias de la complejidad, (2006), 1-145.
    [33] B. Marmuc, Ideología, matemáticas y ciencias sociales, W. Pareto y G. Sorel y la ambigüedad en la comparación de las desigualdades, Revista de Metodología de Ciencias Sociales, 5 (2013), 11-28.
    [34] A. Mayol, El Derrumbe del Modelo, LOM, Santiago, 2013.
    [35] M. Mitchell, Complexity. A Guided Tour, Oxford University Press, New York, 2003.
    [36] R. Michels, Los Partidos Políticos, Amorrortu, Buenos Aires, 1976.
    [37] W. Mills, La Elite del Poder, $9^{th}$ edition, Fondo de Cultura Económica, México, 1988.
    [38] J. Molina-Cano, Las nociones de mando y obediencia en la teoría política de Julien Freund, Díkaion, 23 (2009). Available from: http://dikaion.unisabana.edu.co/index.php/dikaion/article/view/1550/2126.
    [39] G. Mosca, La Clase Política, $2^{nd}$ edition, Fondo de Cultura Económica, México, 1980.
    [40] Reproduction of the original version. Available from: http://americo.usal.es/iberoame/sites/default/files/laclasepolitica.pdf.
    [41] J. Nahapiet and S. Ghoshal, Social capital, intellectual capital, and the organization advantage, The Academy of Management Review, 23 (1998), 242-266.
    [42] P. Navarro, El Fenómeno de la Complejidad Social Humana, Curso de Doctorado Interdisciplinar en Sistemas Complejos, 1996. Available from: http://home.dsoc.uevora.pt/ eje/complexidade_social.html.
    [43] M. E. J. Newman, The structure and functions of complex networks, SIAM Review, 45 (2003), 167-256. doi: 10.1137/S003614450342480
    [44] M. E. J. Newman, A.-L. Barabási and D. Watts, The Structure and Dynamics of Networks, Princeton University Press, New York, 2006.
    [45] M. E. J. Newman, Assortativity mixing in networks, Physical Review Letters, 89 (2002), 208701. doi: 10.1103/PhysRevLett.89.208701
    [46] Available from: http://webiigg.sociales.uba.ar/iigg/jovenes_investigadores/5jornadasjovenes/EJE9/Mesa%0Teoria%20Sociologica%20Clasica%20I/NOCERA_Pablo_d.PDF.
    [47] 1999. Available from: http://dbpubs.stanford.edu:8090/pub/showDoc.Fulltext?lang=en&doc=1999-66&format=pdf.
    [48] W. Pareto, Reseña de su Obra, Socio-Ciencia, 2009. Available from: http://socio-ciencia.blogspot.com/2009/10/wlfredo-pareto-una-resena-de-su-obra.html.
    [49] T. Parsons and E. Shils, Toward a General Theory of Action, Harper & Row Publishers, New York, 1962. doi: 10.4159/harvard.9780674863507
    [50] PNUD-Chile, La Élite Chilena y la Difícil Conducción del Desarrollo, PNUD, 2004. Available from: http://www.centrodesarrollohumano.org/pmb/opac_css/index.php?lvl=notice_display&id=923#.U3DOTtyodrQ.
    [51] H. Rheingold, Multitudes Inteligentes, Gedisa, Barcelona, 2004.
    [52] G. Ritzer, Teoría Sociológica Clásica, $1^{st}$ edition, Imprenta San Francisco de Sales, Madrid, 1920.
    [53] G. Ritzer, Teoría Sociológica Contemporánea, McGraw-Hill, México, 1994.
    [54] G. Ritzer, Teoría Sociológica Clásica, McGraw-Hill, México, 1994.
    [55] P. Rodrigo and D. Arenas, La nueva gobernanza política y las colaboraciones intersectoriales para el desarrollo sostenible, Innovar, 24 (2014), 197-210. doi: 10.15446/innovar.v24n53.43945
    [56] G. Runciman, Crítica de la Filosofía de las Ciencias Sociales de Max Weber, Fondo de Cultura Económica, México, 1976.
    [57] F. Sanchez López, Sociología de la Acción. Introducción a la obra de Talcott Parsons, Instituto Balmes de Sociología, Consejo Superior de Investigaciones Científicas, Madrid, 1964.
    [58] R. Sole, Redes Complejas. Del Genoma a Internet, Tusquets Eds., Barcelona, 2009.
    [59] N. N. Taleb, The Black Swan, Random House, New York, 2010.
    [60] R. Thom, Esbozo de una Semiofísica. Física Aristotélica y Teoría de las Catástrofes, Gedisa, Barcelona, 1990.
    [61] J. Touchard, Historia de las Ideas Políticas, Tecnos, Madrid, 1987.
    [62] 2011. Available from: http://www.ugm.cl/info-alumnos-y-alumnas/chile-ante-una-encrucijada/.
    [63] D. J. Watts and S. H. Strogatz, Collective dynamics of small-world networks, Nature, 393 (1998), 440-442.
    [64] M. Weber, Economía y Sociedad, Fondo de Cultura Económica, México, 1964.
  • This article has been cited by:

    1. Yao Shi, Xiaozhen Liu, Zhenyu Wang, Bifurcation Analysis and Chaos Control of a Discrete Fractional-Order Modified Leslie–Gower Model with Nonlinear Harvesting Effects, 2024, 8, 2504-3110, 744, 10.3390/fractalfract8120744
    2. Xiongxiong Du, Xiaoling Han, Dynamics analysis of a discrete diffusion predator–prey model with prey refuge, 2025, 03770427, 116706, 10.1016/j.cam.2025.116706
  • Reader Comments
  • © 2015 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(4237) PDF downloads(72) Cited by(6)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog