Research article

Stability and bifurcation analysis of a discrete-time plant-herbivore model with harvesting effect

  • The dynamics of plant-herbivore interactions are essential for understanding ecosystem stability and resilience. This article investigated the effects of incorporating a harvesting effect on the dynamics of a discrete-time plant-herbivore system. An analysis was performed to determine the existence and stability of fixed points. In addition, studies have shown that the system experienced transcritical, period-doubling, and Neimark-Sacker bifurcations. Moreover, we provided numerical simulations to substantiate our theoretical results. Our research indicated that harvesting in excessive amounts may have negative effects on the populations of both plants and herbivores. However, when harvesting was done at moderate levels, it promoted the coexistence and stability of both populations. The findings of our analysis provided a deep understanding of the intricate dynamics of ecological systems and underscored the need to use sustainable harvesting methods for the management and preservation of ecosystems.

    Citation: Mohammed Alsubhi, Rizwan Ahmed, Ibrahim Alraddadi, Faisal Alsharif, Muhammad Imran. Stability and bifurcation analysis of a discrete-time plant-herbivore model with harvesting effect[J]. AIMS Mathematics, 2024, 9(8): 20014-20042. doi: 10.3934/math.2024976

    Related Papers:

    [1] Jie Liu, Qinglong Wang, Xuyang Cao, Ting Yu . Bifurcation and optimal harvesting analysis of a discrete-time predator–prey model with fear and prey refuge effects. AIMS Mathematics, 2024, 9(10): 26283-26306. doi: 10.3934/math.20241281
    [2] 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
    [3] Ashraf Adnan Thirthar, Salam J. Majeed, Kamal Shah, Thabet Abdeljawad . The dynamics of an aquatic ecological model with aggregation, Fear and Harvesting Effects. AIMS Mathematics, 2022, 7(10): 18532-18552. doi: 10.3934/math.20221018
    [4] Xiongxiong Du, Xiaoling Han, Ceyu Lei . Dynamics of a nonlinear discrete predator-prey system with fear effect. AIMS Mathematics, 2023, 8(10): 23953-23973. doi: 10.3934/math.20231221
    [5] Abdul Qadeer Khan, Fakhra Bibi, Saud Fahad Aldosary . Bifurcation analysis and chaos in a discrete Hepatitis B virus model. AIMS Mathematics, 2024, 9(7): 19597-19625. doi: 10.3934/math.2024956
    [6] A. Q. Khan, Ibraheem M. Alsulami . Discrete Leslie's model with bifurcations and control. AIMS Mathematics, 2023, 8(10): 22483-22506. doi: 10.3934/math.20231146
    [7] Kottakkaran Sooppy Nisar, G Ranjith Kumar, K Ramesh . The study on the complex nature of a predator-prey model with fractional-order derivatives incorporating refuge and nonlinear prey harvesting. AIMS Mathematics, 2024, 9(5): 13492-13507. doi: 10.3934/math.2024657
    [8] Xin-You Meng, Fan-Li Meng . Bifurcation analysis of a special delayed predator-prey model with herd behavior and prey harvesting. AIMS Mathematics, 2021, 6(6): 5695-5719. doi: 10.3934/math.2021336
    [9] 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
    [10] A. Q. Khan, Ibraheem M. Alsulami, S. K. A. Hamdani . Controlling the chaos and bifurcations of a discrete prey-predator model. AIMS Mathematics, 2024, 9(1): 1783-1818. doi: 10.3934/math.2024087
  • The dynamics of plant-herbivore interactions are essential for understanding ecosystem stability and resilience. This article investigated the effects of incorporating a harvesting effect on the dynamics of a discrete-time plant-herbivore system. An analysis was performed to determine the existence and stability of fixed points. In addition, studies have shown that the system experienced transcritical, period-doubling, and Neimark-Sacker bifurcations. Moreover, we provided numerical simulations to substantiate our theoretical results. Our research indicated that harvesting in excessive amounts may have negative effects on the populations of both plants and herbivores. However, when harvesting was done at moderate levels, it promoted the coexistence and stability of both populations. The findings of our analysis provided a deep understanding of the intricate dynamics of ecological systems and underscored the need to use sustainable harvesting methods for the management and preservation of ecosystems.



    A plant-herbivore system is an ecological interaction in which plants function as main producers and herbivores consume them as their major source of living. The creation of this connection is vital for the efficient operation of the ecosystem and the conservation of biodiversity. Plants use photosynthesis to capture solar energy and convert it into organic matter. They serve as the fundamental basis of food webs, providing herbivores and other trophic levels with the energy and nutrition they need to survive. Herbivores are living things that eat plant tissues such as leaves, stems, fruits, and roots. Plant-herbivore interactions have a profound influence on ecosystem processes and resources. It appears that variations in the number of plants may have a substantial impact on the cycling of nutrients, the structure of the soil, and the availability of habitat, which in turn can have an effect on the abundance and distribution of other organisms within the ecosystem. Herbivores, in addition to being prey for predators, have the ability to impact the interactions that occur between predators and prey, as well as the trophic cascades that occur between different groupings of animals. The interaction between plants and herbivores is a fascinating and nuanced element of ecology, characterized by complex, frequently nonlinear relationships that may result in threshold effects. Minor changes on small scales may have significant implications on bigger scales [1,2,3,4]. Plant-herbivore interactions are important, but their dynamics may be challenging to understand due to complex ecological relationships [5,6,7,8,9,10,11].

    Plant-herbivore interactions are often described through modified versions of predator-prey models due to their similar dynamics and ecological principles. Both types of models aim to represent the relationships between predators (also known as herbivores) and their prey (also known as plants) at different trophic levels within an ecosystem. Although plant-herbivore interactions and predator-prey interactions have certain similarities, such as population control and feeding relationships, there are key differences that arise from the unique behaviors of each kind of interaction. In the context of plants and herbivores, there is no occurrence of a predator (herbivore) chasing its victim (plant) [12]. Contrary to predators, herbivores do not engage in the aggressive pursuit or killing of plants. Herbivory, on the other hand, refers to consuming leaves, stems, fruits, or roots using specialized mouthparts or feeding structures to obtain nutrients from plant tissues.

    Various studies have explored the qualitative behavior of plant-herbivore models, examining phenomena such as bistability, bifurcations, and chaos control. These investigations have provided valuable insights into the dynamic complexities of plant-herbivore interactions, contributing to our understanding of ecosystem dynamics. Researchers have investigated plant-herbivore interactions using both continuous-time differential equations and discrete-time difference equations. Kartal [13] investigated the dynamical behavior of a plant-herbivore system, including both differential and difference equations. Beso et al. [14] investigated stability and various types of bifurcations in a plat-herbivore system with a strong Allee effect. Din [15] investigated the global behavior of a discrete-time plant-herbivore system. Khan et al. [16] conducted bifurcation analysis of a discrete-time plant-herbivore system. Hamada [17] investigated stability, bifurcation, and chaos in a discrete-time plant-herbivore system obtained from a continuous-time plant-herbivore system by applying the piecewise constant argument method. Similarly, for some other discussions related to the population dynamics of plant-herbivore systems, we refer the interested reader to [18,19,20,21,22,23] and references therein.

    The apple twig borer (ATB), a pest infesting grapevines, serves as a case study illustrating the ecological dynamics of plant-herbivore systems. Previous studies [24,25] proposed fundamental frameworks based on the life cycle of adult ATBs, laying the groundwork for understanding the dynamics of plant-herbivore interactions. Understanding these dynamics is crucial for land management, environmental protection, and animal husbandry, prompting extensive research into dynamical system models to elucidate population and ecosystem behaviors. In [11], Din et al. investigated the stability, Neimark-Sacker bifurcation, and chaos in the following discrete-time plant-herbivore system with Holling type-Ⅱ functional response:

    {xn+1=xnr(1+yn)+κxn,yn+1=σ(1+xn)yn, (1.1)

    where xn and yn, respectively, represent the population densities of grapevine and ATB. Moreover, r,κ, and σ are positive constants.

    In population dynamics, harvesting refers to the intentional removal of individuals from a population for human use or consumption. This practice is common in various natural resource management contexts, including fisheries [26,27], forestry [28,29], and wildlife management [30,31,32]. Harvesting may be detrimental to the health of ecosystems and biodiversity if it is carried out in an excessive or indiscriminate manner. Understanding the consequences of such efforts is necessary for the implementation of sustainable management systems. Within the framework of plant-herbivore interactions, the act of harvesting may directly influence the populations of both plants and herbivores, resulting in changes to the composition, abundance, and distribution of these populations. The harvesting of plants has the potential to reduce population densities, disrupt reproductive cycles, and alter the structure of communities by favoring some species over others. There is also the possibility that these effects might occur simultaneously. Similarly, herbivores may experience changes in population dynamics, resource availability, and foraging behavior as a result of the pressure exerted by harvesting. Furthermore, the removal of individuals from both the plant and herbivore populations has the potential to upset the delicate balance that exists between them, which may have a domino effect on higher trophic levels as well as the general functioning of the ecosystem. In the study conducted by Virtala [33], the optimum harvesting in a plant-herbivore system was investigated. The study conducted by Asfaw et al. [34] investigated a plant-herbivore system that included herbivore harvesting effects. They provided evidence that the harvest rate of the herbivore population has a significant influence on the dynamics of herbivores.

    When it comes to effectively managing plant-herbivore systems, there are a few different harvesting strategies to choose from. These include proportional, constant, nonselective, and selective harvesting [35,36,37,38]. Each approach has its own set of repercussions, both for the coexistence of organisms and for the dynamics of the ecosystem. The term selective harvesting refers to a technique that focuses on certain individuals or species, taking into consideration specific characteristics such as size, age, or quality. This strategy makes it possible to exert conservation and management activities in a focused manner. In contrast, nonselective harvesting entails the extraction of individuals without taking into account their distinct features or characteristics. This might potentially have unanticipated consequences for the structure of populations and the operation of ecosystems. Constant harvesting is the consistent and regular extraction of a set quantity from a population, which consistently affects the population dynamics independent of variations in population density. In contrast, proportionate harvesting maintains a harvest rate that is directly proportional to the size of the present population, ensuring that the intensity of harvesting increases or decreases with the abundance of the population. This adaptable strategy enables flexible administration by shifting population dynamics and fostering sustainable use of resources and stability in the population. Proportionate harvesting may provide more adaptability to changes in the environment and variations in population size, especially in ecosystems that are dynamic or changing.

    Therefore, inspired by the preceding discussion, we are interested to inquire: What are the consequences for the dynamic properties when a harvesting impact is applied to the plant population in system (1.1)? Therefore, we modified system (1.1) by integrating the harvesting impact into the plant population. The result is the following modified system:

    {xn+1=xnr(1+yn)+κxnhxn,yn+1=σ(1+xn)yn, (1.2)

    where h>0 is the harvesting rate. It represents the intensity of harvesting. Ensuring nonnegative solutions is of utmost importance in ecological models, such as the plant-herbivore model, to maintain biological relevance. Negative solutions are not ecologically meaningful since populations cannot have negative densities. Ensuring solution positivity is crucial for aligning model predictions with ecological reality, which in turn facilitates the correct understanding and management of ecosystems. We assume that the initial values x0 and y0 are positive. Clearly, from the second equation of system (1.2), it is evident that yn+1>0 for all n. In the first equation, the term hxn can produce negative values of xn+1. One should select h sufficiently small so that xn+1 is positive. It can be checked that if h is sufficiently small enough and h<1r(1+y0)+κx0, then xn+1>0 for all n. Another approach, given in Section 3.6 of [39], is that xn+1 can be redefined as follows:

    xn+1=max{0,xnr(1+yn)+κxnhxn}.

    Moreover, in all our simulations performed here, solutions remained nonnegative. Through a combination of analytical techniques and numerical simulations, we seek to address the following research questions:

    ● What are the effects of plant harvesting on the fixed points of the plant-herbivore system (1.2)?

    ● How does plant harvesting influence the stability properties of fixed points, and under what conditions do bifurcation phenomena occur?

    ● What insights can be gained from bifurcation analysis regarding the long-term dynamics of the modified plant-herbivore system?

    For the detailed analysis of stability and bifurcation in discrete-time systems, we refer the readers to [40,41,42,43,44] and references therein. The subsequent sections of the paper are arranged as follows: Section 2 is dedicated to investigating the existence and stability of fixed points. Section 3 examines the study of bifurcations that include period-doubling (PD), transcritical (TC), and Neimark-Sacker (NS) occurring at the positive fixed point (FP). In Section 4, numerical simulation results are presented to support the theoretical analysis and display the new and rich dynamic behavior. Moreover, the influence of harvesting on system dynamics is presented in Section 5. Finally, a brief conclusion is presented in Section 6.

    Exploring the stability of FPs is very important in plant-herbivore environments. These FPs are states of equilibrium where both plant and herbivore populations are balanced. By examining their stability, we can predict long-term behavior in ecological systems, improving our understanding of the multiple components that affect ecosystem dynamics.

    The FPs of system (1.2) can be determined by solving the subsequent nonlinear equations:

    {x=xr(1+y)+κxhx=f(x,y),y=σ(1+x)y=g(x,y), (2.1)

    for x and y. The system (1.2) possesses three FPs: E0=(0,0),E1=(1rhrκ+κh,0), and

    E2=(1σσ,1r(κκσ+11+hr)).

    E0=(0,0) is the trivial FP that exists always. E1=(1rhrκ+κh,0) is a boundary FP that exists if r+rh<1. Moreover, E2=(1σσ,1r(κκσ+11+hr)) is coexistence FP, which exists if r<1, κ1+κr<σ<1, and h<σκκσ+σr1.

    The Jacobian matrix of the system

    {xn+1=f(xn,yn)yn+1=g(xn,yn)

    is the matrix given below:

    J(x,y)=[fxfygxgy].

    Thus, the Jacobian matrix J(x,y) of the system (1.2) evaluated at any FP (x,y) is as follows:

    J(x,y)=[r+ryh(r+κx+ry)2(r+κx+ry)2rx(r+κx+ry)2σyσ(1+x)].

    The eigenvalues ξ1,2 of the Jacobian matrix J are helpful in determining the stability of FPs. The FP (x,y) is called a sink if |ξ1,2|<1 and a source if |ξ1|>1 and |ξ2|>1. Furthermore, the FP (x,y) is classified as a saddle point (SP) if |ξ1|>1 and |ξ2|<1 (or |ξ1|<1 and |ξ2|>1). In the case of a non-hyperbolic point (NHBP) (x,y), either |ξ1|=1 or |ξ2|=1. However, if the eigenvalues are in complicated form, making direct use of this definition challenging, the following result proves to be very useful in determining the stability of the FP.

    Lemma 2.1. [45] Let Λ(ξ)=ξ2+K1ξ+K0. Assume that Λ(1)>0. If ξ1 and ξ2 are solutions of Λ(ξ)=0, then

    (1) |ξ1|<1 together with |ξ2|<1 if Λ(1)>0K0<1,

    (2) |ξ1|<1|ξ2|>1 (or |ξ1|>1|ξ2|<1) if Λ(1)<0,

    (3) |ξ1,2|>1 if Λ(1)>0K0>1,

    (4) |ξ2|1ξ1=1 if Λ(1)=0K10,2,

    (5) ξ1,ξ2C along with |ξ1,2|=1 if K124K0<0K0=1.

    A sink reflects the long-term coexistence of herbivores and plants, suggesting a balanced environment in which populations maintain sustainable levels throughout time. An SP suggests intermittent stability, with the system alternating between plenty and shortage, similar to natural population fluctuations. An unstable source suggests vulnerability to unexpected changes, stressing the potential for population collapses or rapid expansion that disrupts ecological equilibrium. Furthermore, non-hyperbolic points may indicate intricate and nuanced interactions, demonstrating the complexity of ecological dynamics and the difficulty of predicting population behavior. Through computations, it is obtained that

    J(E0)=[h+1r00σ].

    One can easily obtain the following result for the topological classification of E0:

    Theorem 2.2. E0 is a

    (1) sink if σ<1 and 1rr<h<1+rr,

    (2) source if one of the below-listed parametric conditions is fulfilled:

    (i) σ>1,r<1, and h<1rr,

    (ii) σ>1 and h>1+rr,

    (3) SP if one of the listed below parametric conditions is fulfilled:

    (i) σ<1,r<1, and h<1rr,

    (ii) σ<1 and h>1+rr,

    (iii) σ>1 and 1rr<h<1+rr,

    (4) NHBP if one of the listed below parametric conditions is fulfilled:

    (i) σ=1,

    (ii) h=1+rr,

    (iii) r<1, h=1rr.

    The eigenvectors corresponding to eigenvalues ξ1=h+1r and ξ2=σ of J(E0) are v1=<1,0> and v2=<0,1>. The stable Ws and unstable Wu manifolds in the saddle case (3-ⅲ) are given by

    Ws=Span({v1}), Wu=Span({v2}).

    Remark 2.3. At NHBP, the system experiences different types of bifurcations depending on the nature of the eigenvalues of the Jacobian matrix. If (4-ⅰ) is true, it follows that one of the eigenvalues of the matrix J(E0) is 1. Consequently, a TC bifurcation occurs at E0. Similarly, a TC bifurcation occurs at E0 if (4-ⅲ) is true. Moreover, if (4-ⅱ) is true, it follows that one of the eigenvalues of the matrix J(E0) is 1. Consequently, a PD bifurcation occurs at E0.

    Next, we obtain

    J(E1)=[r+h2r+h(1+2r)(1+h)r(1+r+hr)κ0σσ(1+r+hr)κ(1+h)].

    Since E1 exists if r<11+h. Thus, we obtain the following result:

    Theorem 2.4. E1 is a

    (1) sink if h1(h+1)2<r<1h+1 and 0<σ<κ+κhκ+κhrhr+1,

    (2) source if r<h1(h+1)2,h>1 and σ>κ+κhκ+κhrhr+1,

    (3) SP if any of the following criteria is met:

    (i) r<h1(h+1)2,h>1, and 0<σ<κ+κhκ+κhrhr+1,

    (ii) σ>κ+κhκ+κhrhr+1 and h1(h+1)2<r<1h+1,

    (4) NHBP if one of the listed below criteria is met:

    (i) r=h1(h+1)2,

    (ii) σ=κ+κhκ+κhrhr+1.

    The eigenvectors corresponding to eigenvalues ξ1=r+h2r+h(1+2r) and ξ2=σσ(1+r+hr)κ(1+h) are v1=<1,0> and

    v2=0,r(1+h)2(1+r+rh)κh(1+h)σ(1+κ+κh)+σr(1+h)+κr(1+h)3.

    The stable Ws and unstable Wu manifolds in the saddle case (3-ⅱ) are given by

    Ws=Span({v1}), Wu=Span({v2}).

    Remark 2.5. If (4-i) is true, then it follows that one of the eigenvalues of the matrix J(E1) is 1. Consequently, a PD bifurcation occurs at E1. Moreover, if (4-ii) is true, it follows that one of the eigenvalues of the matrix J(E1) is 1. Consequently, a TC bifurcation occurs at E1.

    The Jacobian matrix at positive FP is

    J(E2)=[1+κ(1+σ)(1+h)2σ(1+σ)(1+h)2rσσ(κκσ+11+hr)r1]. (2.2)

    The characteristic polynomial of J(E2) is

    Λ(ξ)=ξ2+K1ξ+K0,

    where

    K1=2κ(1+σ)(1+h)2σ,K0=2σ+hσhκ(2+σ)(1+σ)(1+h)2σ+(1+σ)(1+h)2r.

    Through computations, it is obtained that

    Λ(0)=2σ+hσhκ(2+σ)(1+σ)(1+h)2σ+(1+σ)(1+h)2r,Λ(1)=5σ+hσhκ(3+σ)(1+σ)(1+h)2σ+(1+σ)(1+h)2r,Λ(1)=(1σ)(1+h)(σ+(1+h)(κ(1+σ)σr))σ.

    After applying Lemma 2.1, we obtain the following result for the classification of the positive FP, E2, of system (1.2).

    Theorem 2.6. The positive FP

    (1) E2 is a sink if r<1σ(1+h)(σ+κσ+κσh2κ2h) and

    r>1(σ1)(1+h)2(5+σh+σh+κ(σ3)(σ1)(1+h)2σ),

    (2) E2 is a source if r>1σ(1+h)(σ+κσ+κσh2κ2h) and

    r>1(σ1)(1+h)2(5+σh+σh+κ(σ3)(σ1)(1+h)2σ),

    (3) E2 is an SP if r<1(σ1)(1+h)2(5+σh+σh+κ(σ3)(σ1)(1+h)2σ),

    (4) E2 is NHBP and experiences PD bifurcation if σκ(1+h)22+κ(1+h)2,κ(1+h)24+κ(1+h)2 and

    r=1(σ1)(1+h)2(5+σh+σh+κ(σ3)(σ1)(1+h)2σ),

    (5) E2 is NHBP and experiences NS bifurcation if 0<κ<4(1σ)(1+h)2 and

    r=1σ(1+h)(σ+κσ+κσh2κ2κh).

    This section aims to thoroughly investigate bifurcation phenomena, including TC bifurcation at E1 as well as PD and NS bifurcations at E2. For a comprehensive understanding of bifurcation analysis, we suggest referring to the sources [46,47,48,49]. Bifurcation is a crucial factor in determining the dynamics of the system, revealing situations where even little changes in parameters may lead to significant changes in plant-herbivore interactions. Moreover, exploring bifurcation processes improves our comprehension of ecosystem dynamics, facilitating the development of effective strategies for preserving and controlling plant and herbivore populations in a sustainable manner.

    We start our examination of the TC bifurcation at E1 by considering condition (4-ⅱ) outlined in Theorem 2.4. We can rewrite it as r=κhκ+σ+κσ+hκσ(1+h)σ. Introducing a sufficiently small perturbation ε into the bifurcation parameter r around R10=κhκ+σ+κσ+hκσ(1+h)σ, the system (1.2) takes the subsequent form:

    {xn+1=xn(R10+ε)(1+yn)+κxnhxn,yn+1=σ(1+xn)yn. (3.1)

    We shift E1 to (0,0) by taking un=xn1(R10+ε)h(R10+ε)κ+κh, vn=yn. Consequently, the system (3.1) becomes

    [un+1vn+1]=[1+(1+h)2κ(1+σ)σ(1+h)(1+σ)((1+h)κ(1+σ)+σ)σ201][unvn]+[F(un,vn,ε)G(un,vn,ε)], (3.2)

    where

    F(un,vn,ε)=a1un2+a2vn2+a3unvn+a4unε+a5vnε+a6un3+a7vn3+a8un2vn+a9un2ε+a10unvn2+a11vn2ε+a12vnε2+a13unvnε+O((|un|+|vn|+|ε|)4),G(un,vn,ε)=b1unvn+b2vnε+O((|un|+|vn|+|ε|)4),a1=(1+h)2κ((1+h)κ(1+σ)+σ)σ, a2=(1+h)(1+σ)((1+h)κ(1+σ)+σ)2σ3,a3=(1+h)((1+h)κ(1+σ)+σ)(2(1+h)κ(1+σ)+σ)σ2, a4=(1+h)2,a5=(1+h)(2(1+h)κ(1+σ)+σ)κσ, a6=(1+h)3κ2((1+h)κ(1+σ)+σ)σ,a7=(1+h)(1+σ)((1+h)κ(1+σ)+σ)3σ4,a8=(1+h)2κ((1+h)κ(1+σ)+σ)(3(1+h)κ(1+σ)+2σ)σ2, a9=(1+h)3κ,a10=(1+h)((1+h)κ(1+σ)+σ)2(3(1+h)κ(1+σ)+σ)σ3,a11=(1+h)((1+h)κ(1+σ)+σ)(3(1+h)κ(1+σ)+σ)κσ2, a12=(1+h)2κ,a13=(1+h)2(4(1+h)κ(1+σ)+3σ)σ, b1=σ, b2=σκ.

    Next, the system (3.2) is diagonalized through the consideration of the following transformation:

    [unvn]=[11κ+hκ+1σ110][enfn], (3.3)

    Upon applying the mapping (3.3), the system (3.2) undergoes the alteration as follows:

    [en+1fn+1]=[100ξ][enfn]+[Γ(en,fn,ε)Υ(en,fn,ε)], (3.4)

    where

    ξ=1+(1+h)2κ(1+σ)σ,
    Γ(en,fn,ε)=c1en2+c2enfn+c3enε+O((|en|+|fn|+|ε|)4),Υ(en,fn,ε)=d1en2+d2fn2+d3enfn+d4enε+d5fnε+d6fn3+d7en2ε+d8enfn2+d9fn2ε+d10enε2+d11enfnε+O((|en|+|fn|+|ε|)4),c1=1(1+1κ+hκ)σ, c2=σ, c3=σκ, d1=((1+h)κ(1+σ)+σ)2(1+h)2κ2σ,d2=(1+h)2κ((1+h)κ(1+σ)+σ)σ, d3=h+(1+h)2κ(1+σ)σ+σ+σκ+hκ,d4=1σκ+(1+h)2(1+σ)σσ(1+h)κ2, d5=(1+h)2,d6=(1+h)3κ2((1+h)κ(1+σ)+σ)σ, d7=(1+h)((1+h)κ(1+σ)+σ)κσ,d8=(1+h)2κ((1+h)κ(1+σ)+σ)σ, d9=(1+h)3κ,d10=(1+h)2κ, d11=(1+h)2(2(1+h)κ(1+σ)+σ)σ.

    Now, we use the center manifold theory to obtain the center manifold MC of (3.2) at (0,0) in a close neighborhood of ε=0. The center manifold MC can be obtained using

    MC={(en,fn,ε)R+3|fn=p1en2+p2enε+p3ε2+O(3)}.

    Through calculations, we obtain p1=d11ξ,p2=d41ξ,p3=0. Consequently, the system (3.2) is restricted to MC as follows:

    ω:=en+1=en+c1en2+c3enε+(c2d11ξ)en3+(c2d41ξ)en2ε+O(4). (3.5)

    Through simple computations, we obtain

    ω(0,0)=0, ωen(0,0)=1, ωε(0,0)=0,ωenen(0,0)=2c1=22(1+1κ+hκ)σ0,ωenε(0,0)=c3=σκ0.

    Thus, the system (1.2) experiences TC bifurcation at E1. The next result gives the conditions for the existence and direction of TC bifurcation in the system (1.2) at E1.

    Theorem 3.1. If condition (4-ii) in Theorem 2.4 is fulfilled, then (1.2) undergoes TC bifurcation at E1 when r differs in a close neighborhood of R10=κhκ+σ+κσ+hκσ(1+h)σ.

    The boundary FP is unstable for r<R10. It collides with positive FP at r=R10, and subsequently they exchange stability. The result shows that small variations in environmental or biological parameters may cause major modifications in plant-herbivore population dynamics, culminating in a TC bifurcation at the border fixed point. At the transition, essential ecosystem dynamics shift, causing the system to shift from a state of stable coexistence between plants and herbivores to one in which plants can live on their own.

    Next, we investigate the PD bifurcation at E2 by considering condition (4) presented in Theorem 2.6. Biologically, PD bifurcations indicate significant alterations in plant-herbivore dynamics. Consider a situation in which plant and herbivore populations vary regularly over time. During a PD bifurcation, these variations become more intense, and the space between peak and trough densities widens, perhaps doubling. This may have a severe environmental effect. If environmental circumstances unexpectedly favor plant growth, herbivore populations may increase, thereby depleting plant resources. Plant populations rebound when herbivore numbers drop. However, when plant numbers recover, herbivore populations rebound, resulting in increased grazing pressure and a consequent drop in plant populations. Introducing a sufficiently small perturbation ε into the bifurcation parameter r around R1=1(σ1)(1+h)2(5+σh+σh+κ(σ3)(σ1)(1+h)2σ), the system (1.2) takes the subsequent form:

    {xn+1=xn(R1+ε)(1+yn)+κxnhxn,yn+1=σ(1+xn)yn. (3.6)

    To translate the point E2 at (0,0), we use the change of variables un=xn1σσ, vn=yn1(R1+ε)(κκσ+11+h(R1+ε)). Consequently, the system (3.6) is expressed as below:

    [un+1vn+1]=[1+κ(1+σ)(1+h)2σκ(34σ+σ2)(1+h)2+σ(5+σh+σh)σ22σ(2σ+κ(1+σ)(1+h)2)κ(34σ+σ2)(1+h)2+σ(5+σh+σh)1][unvn]+[Π1(un,vn,ε)Π2(un,vn,ε)], (3.7)

    where

    Π1(un,vn,ε)=a1un2+a2vn2+a3unvn+a4vnε+a5un3+a6vn3+a7un2vn+a8unvn2+a9vn2ε+a10unvnε+O(4),Π2(un,vn,ε)=b1unvn+b2unε+b3unε2+O(4),a1=κ(1+h)2(σ+κ(1+σ)(1+h))σ, a2=(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2(1+σ)σ3(1+h),a3=(σ+2κ(1+σ)(1+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))(1+σ)σ2,a4=(1+σ)(1+h)2σ, a5=κ2(1+h)3(σ+κ(1+σ)(1+h))σ,a6=(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))3(1+σ)2σ4(1+h)2,a7=κ(1+h)(2σ+3κ(1+σ)(1+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))(1+σ)σ2,a8=(σ+3κ(1+σ)(1+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2(1+σ)2σ3(1+h),a9=2(1+h)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))σ2,a10=(1+h)2(σ+2κ(1+σ)(1+h))σ,b1=σ, b2=(1+σ)2σ2(1+h)3(σ+κ(1+σ)(1+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2,b3=(1+σ)3σ3(1+h)5(σ+κ(1+σ)(1+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))3.

    Subsequently, the system (3.7) goes through diagonalization after a consideration of the following transformation:

    [unvn]=[κ(34σ+σ2)(1+h)2+σ(5+hσ(1+h))σ(2σ+κ(1+σ)(1+h)2)κ(34σ+σ2)(1+h)2+σ(5+σh+σh)2σ211][enfn]. (3.8)

    After implementing the map (3.8), the system (3.7) takes the subsequent form:

    [en+1fn+1]=[100ξ][enfn]+[Γ(en,fn,ε)Υ(en,fn,ε)], (3.9)

    where

    ξ=3+κ(1+σ)(1+h)2σ,
    Γ(en,fn,ε)=c1en2+c2fn2+c3enfn+c4enε+c5fnε+c6en3+c7fn3+c8en2fn+c9en2ε+c10enfn2+c11fn2ε+c12enε2+c13fnε2+c14enfnε+O(4),Υ(en,fn,ε)=d1en2+d2fn2+d3enfn+d4enε+d5fnε+d6en3+d7fn3+d8en2fn+d9en2ε+d10enfn2+d11fn2ε+d12enε2+d13fnε2+d14enfnε+O(4),c1=(κ(34σ+σ2)(1+h)2+σ(σ5h+σh))(κ(σ1)2(1+h)3+2σ(σ3+h+σh))(1+σ)(1+h)(2σ+κ(1+σ)(1+h)2)(4σ+κ(1+σ)(1+h)2),
    c2=((2σ+κ(1+σ)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))(κ2(1+σ)2(1+h)4+κ(1+σ)σ(1+h)2(5+h)+σ2(5+σ+h+σh)))/(2(1+σ)σ3(1+h)(4σ+κ(1+σ)(1+h)2)),c3=(16σ+κ(σ1)(1+h)2(5+σ3h+σh))(κ(34σ+σ2)(1+h)2+σ(σ5h+σh))2(σ1)σ(1+h)(4σ+κ(σ1)(1+h)2),c4=(1+σ)σ(1+h)24σ+κ(1+σ)(1+h)2,c5=(σ1)(1+h)2(2σ+κ(σ1)(1+h)2)(κ(σ1)2(1+h)2+σ(3+σh+σh))2(4σ+κ(1+σ)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh)),c6=8σ(1+h)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2(1+σ)2(1+h)2(2σ+κ(1+σ)(1+h)2)2(4σ+κ(1+σ)(1+h)2),
    c7=(((2σ+κ(σ1)(1+h)2)3(κ(σ1)(1+h)2+σ(3+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2)/(4(σ1)2σ5(1+h)2(4σ+κ(σ1)(1+h)2))),c8=4(σ(h5)+κ(σ1)(h2)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(σ5h+σh))2(σ1)2σ(1+h)2(2σ+κ(σ1)(1+h)2)(4σ+κ(σ1)(1+h)2),c9=2σ(3+h)(1+h)4σ+κ(1+σ)(1+h)2,c10=((2σ+κ(1+σ)(1+h)2)(κ(1+σ)(5+h)(1+h)22σ(7+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2)/(2(1+σ)2σ3(1+h)2(4σ+κ(1+σ)(1+h)2)),c11=(1+h)(2σ+κ(1+σ)(1+h)2)(2κ(1+σ)(1+h)2+σ(5+h))σ(4σ+κ(1+σ)(1+h)2),
    c12=(1+σ)3σ2(1+h)5(σ+κ(1+σ)(1+h))(4σ+κ(1+σ)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2,c13=(1+σ)3σ(1+h)5(σ+κ(1+σ)(1+h))(2σ+κ(1+σ)(1+h)2)2(4σ+κ(1+σ)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2,c14=(1+h)(16σ2+2κ2(1+σ)2(1+h)4+κ(1+σ)σ(1+h)2(9+h))σ(4σ+κ(1+σ)(1+h)2),d1=2σ(1+σ3h+σh)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))(1+σ)(1+h)(2σ+κ(1+σ)(1+h)2)(4σ+κ(1+σ)(1+h)2),d2=(((κ3(1+σ)3(1+h)6+4κ(1+σ)σ2(1+h)2(4+h)+κ2(1+σ)2σ(1+h)4(7+h)2σ3(7+σ3h+σh))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh)))/(2(1+σ)σ3(1+h)(4σ+κ(1+σ)(1+h)2))),d3=((κ(34σ+σ2)(1+h)2+σ(5+σh+σh))(16σ2+κ2(1+σ)2(3+h)(1+h)4+κ(1+σ)σ(1+h)2(15+σ+h+σh)))/((1+σ)σ(1+h)(2σ
    +κ(1+σ)(1+h)2)(4σ+κ(1+σ)(1+h)2)),d4=2(σ1)σ(1+h)2(κσ(2σ3+σ2)(1+h)2+κ2(σ1)2(1+h)4+σ2(3+σh+σh))(2σ+κ(σ1)(1+h)2)(4σ+κ(σ1)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(σ5h+σh)),d5=(1+σ)σ(1+h)24σ+κ(1+σ)(1+h)2,d6=8σ(1+h)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2(1+σ)2(1+h)2(2σ+κ(1+σ)(1+h)2)2(4σ+κ(1+σ)(1+h)2),
    d7=((2σ+κ(1+σ)(1+h)2)3(κ(1+σ)(1+h)2+σ(3+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2)/(4(1+σ)2σ5(1+h)2(4σ+κ(1+σ)(1+h)2)),d8=4(σ(h5)+κ(σ1)(h2)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(σ5h+σh))2(1+σ)2σ(1+h)2(2σ+κ(1+σ)(1+h)2)(4σ+κ(1+σ)(1+h)2),d9=2σ(3+h)(1+h)4σ+κ(1+σ)(1+h)2,d10=(((2σ+κ(1+σ)(1+h)2)(κ(1+σ)(5+h)(1+h)22σ(7+h))(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2)/(2(1+σ)2σ3(1+h)2(4σ+κ(σ1)(1+h)2))),d11=(1+h)(2σ+κ(1+σ)(1+h)2)(2κ(1+σ)(1+h)2+σ(5+h))σ(4σ+κ(1+σ)(1+h)2),d12=2(1+σ)3σ3(1+h)5(σ+κ(1+σ)(1+h))(2σ+κ(σ1)(1+h)2)(4σ+κ(σ1)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(σ5h+σh))2,d13=(1+σ)3σ2(1+h)5(σ+κ(1+σ)(1+h))(4σ+κ(1+σ)(1+h)2)(κ(34σ+σ2)(1+h)2+σ(5+σh+σh))2,d14=(1+h)(16σ2+2κ2(1+σ)2(1+h)4+κ(1+σ)σ(1+h)2(9+h))σ(4σ+κ(1+σ)(1+h)2).

    Now, we use the center manifold theory to obtain the center manifold MC of (3.9) at (0,0) in a close neighborhood of ε=0. Its approximation is as follows:

    MC={(en,fn,ε)R+3|fn=p1en2+p2enε+p3ε2+O(3)},

    where

    p1=d11ξ, p2=d41+ξ, p3=0.

    Consequently, the system (3.9) is restricted to MC as follows:

    ω:=en+1=en+c1en2+c4enε+(c6+c3d11ξ)en3+(c9+c5d11ξc3d41+ξ)en2ε+(c12c5d41+ξ)enε2+O(4). (3.10)

    The mapping (3.10) undergoes PD bifurcation if the following are nonzero:

    l1=ωεωenen+2ωenε|(0,0)=2c4, (3.11)
    l2=12(ωenen)2+13ωenenen|(0,0)=2(c12+c6+c3d11ξ). (3.12)

    Thus, we obtain the following result:

    Theorem 3.2. If condition (4) of Theorem 2.6 is fulfilled, then (1.2) undergoes PD bifurcation at E2 if l1,l2 provided in (3.11) and (3.12) are nonzero and r fluctuates in a small neighborhood of

    R1=1(σ1)(1+h)2(5+σh+σh+κ(σ3)(σ1)(1+h)2σ).

    Moreover, if  l2>0 (respectively, l2<0), then an orbit of period-2 emanates from E2, which is stable (respectively, unstable).

    The result illustrates how minor fluctuations in parameters can lead to substantial shifts in the system's dynamics, specifically causing a doubling of the population oscillation periods. This happens when the herbivore population increases and feeds more on the plants, thus leading to a reduction in the plant population. With fewer plants available, the herbivore population begins to decrease due to the lack of food. When the herbivore numbers decrease, the plant population is given the chance to increase once again, and the cycle begins again. This cycle reflects the delicate balance and complex interactions that exist within the environment.

    Next, we investigate NS bifurcation at E2 by considering condition (5) presented in Theorem 3.3. Biologically, NS bifurcations reveal complex dynamics within a plant-herbivore system, often marked by unexpected changes and unpredictable patterns. Imagine a situation in which the changes in the populations of plants and herbivores cannot be accurately predicted. NS bifurcation may cause population numbers to fluctuate significantly due to even minor alterations in environmental circumstances or plant-herbivore interactions, making it difficult to predict these changes. The instability has a dramatic effect on the dynamics of ecosystems, which presents issues for ecologists attempting to forecast how populations will respond to changes in the environment. For example, sudden changes in the environment may cause unpredictable and disorderly variations in populations of plants and herbivores, leading to disruptions in the interactions between different trophic levels and unpredictable changes in how the ecosystem functions. Introducing a sufficiently small perturbation ε into the bifurcation parameter r around R2=1σ(1+h)(σ+κσ+κσh2κ2κh), the system (1.2) takes the subsequent form:

    {xn+1=xn(R2+ε)(1+yn)+κxnhxn,yn+1=σ(1+xn)yn. (3.13)

    We use the translation un=xn1σσ, vn=yn1(R2+ε)(κκσ+11+h(R2+ε)) to translate the point E2 to origin. Consequently, the system (3.13) becomes

    [un+1vn+1]=[1+κ(1+σ)(1+h)2σ(1+σ)(1+h)(κ(2+σ)(1+h)+σ(1+ε+hε))σ2σ(κσε)κ2κσ+11+h+ε1][unvn]+[Π1(un,vn)Π2(un,vn)], (3.14)

    where

    Π1(un,vn)=a1un2+a2vn2+a3unvn+a4un3+a5vn3+a6un2vn+a7unvn2+O(4),Π2(un,vn)=σunvn,a1=κ(1+h)2(σ+κ(1+σ)(1+h))σ, a2=(1+σ)(1+h)(κ(2+σ)(1+h)+σ(1+ε+hε))2σ3,a3=(1+h)(σ+2κ(1+σ)(1+h))(κ(2+σ)(1+h)+σ(1+ε+hε))σ2,a4=κ2(1+h)3(σ+κ(1+σ)(1+h))σ, a5=(1+σ)(1+h)(κ(2+σ)(1+h)+σ(1+ε+hε))3σ4,a6=κ(1+h)2(2σ+3κ(1+σ)(1+h))(κ(2+σ)(1+h)+σ(1+ε+hε))σ2,a7=(1+h)(σ+3κ(1+σ)(1+h))(κ(2+σ)(1+h)+σ(1+ε+hε))2σ3.

    The characteristic equation of the linearized part of (3.14) computed at (0,0) is

    ξ2α1(ε)ξ+α2(ε)=0, (3.15)

    where

    α1(ε)=2+κ(1+σ)(1+h)2σ, α2(ε)=1+(1+σ)(1+h)2ε.

    The solutions of (3.15) are given by

    ξ1,2=α1(ε)2±i24α2(ε)α12(ε). (3.16)

    We then obtain |ξ1,2|=α2(ε). Since σ<1, then

    (d|ξ1|dε)ε=0=(d|ξ2|dε)ε=0=12(1σ)(1+h)2<0.

    Moreover, it is necessary that ξ1,2i1 when ε=0 for i=1,2,3,4, which is equivalent to α1(0)2,0,1,2. Through simple computations, we obtain

    α1(0)=2κ(1σ)(1+h)2σ.

    Clearly, α1(0)2. Thus, we only require that

    κ4σ(1σ)(1+h)2,2σ(1σ)(1+h)2,σ(1σ)(1+h)2. (3.17)

    To determine the canonical form of (3.14) at ε=0, we employ the map

    [unvn]=[(1+σ)(1+h)(σ+κ(2+σ)(1+h))σ20κ(1+σ)(1+h)22σ124(2+κ(1+σ)(1+h)2σ)2][enfn]. (3.18)

    After applying the map (3.18), the system (3.14) converts to

    [en+1fn+1]=[1+κ(1+σ)(1+h)22σ124(2+κ(1+σ)(1+h)2σ)2124(2+κ(1+σ)(1+h)2σ)21+κ(1+σ)(1+h)22σ][enfn]+[χ(en,fn)Υ(en,fn)], (3.19)

    where

    χ(en,fn)=c1en2+c2fn2+c3enfn+c4en3+c5fn3+c6en2fn+c7enfn2+O(4),Υ(en,fn)=d1en2+d2fn2+d3enfn+d4en3+d5fn3+d6en2fn+d7enfn2+O(4),c1=κ(1+σ)(1+h)3(σ+κ(2+σ)(1+h))(2σ+κ(1+σ)(1+h))4σ3,c2=(σ+κ(2+σ)(1+h))(4(2+κ(1+σ)(1+h)2σ)2)4σ,c3=(1+h)(σ+κ(2+σ)(1+h))(σ+κ(1+σ)(1+h))4(2+κ(1+σ)(1+h)2σ)22σ2,c4=κ2(1+σ)2(1+h)5(σ+κ(2+σ)(1+h))2(2σ+κ(1+σ)(1+h))8σ5,c5=(σ+κ(2+σ)(1+h))2(4(2+κ(1+σ)(1+h)2σ)2)3/28σ2,c6=κ(σ1)(1+h)3(σ+κ(σ2)(1+h))2(4σ+3κ(σ1)(1+h))4(2+κ(σ1)(1+h)2σ)28σ4,c7=κ(σ1)(1+h)3(σ+κ(σ2)(1+h))2(8σ2+3κ2(σ1)2(1+h)3+2κ(σ1)σ(7+8h+h2))8σ5,
    d1=κ(1+σ)2(1+h)3(σ+κ(2+σ)(1+h))(4σ2+2κσ(1+h)2+κ2(1+σ)(1+h)3)4σ44(2+κ(1+σ)(1+h)2σ)2,d2=κ(1+σ)(1+h)2(σ+κ(2+σ)(1+h))4(2+κ(1+σ)(1+h)2σ)24σ2,d3=(1+σ)(1+h)(σ+κ(2+σ)(1+h))(2σ2κσ(1+h)2κ2(1+σ)(1+h)3)2σ3,d4=κ3(1+σ)3(1+h)7(σ+κ(2+σ)(1+h))2(2σ+κ(1+σ)(1+h))8σ64(2+κ(1+σ)(1+h)2σ)2,d5=κ2(1+σ)2(1+h)4(σ+κ(2+σ)(1+h))2(4σ+κ(1+σ)(1+h)2)8σ5,d6=κ2(1+σ)2(1+h)5(σ+κ(2+σ)(1+h))2(4σ+3κ(1+σ)(1+h))8σ5,d7=κ2(σ1)2(1+h)5(σ+κ(σ2)(1+h))2(8σ2+3κ2(σ1)2(1+h)3+2κ(σ1)σ(7+8h+h2))8σ64(2+κ(σ1)(1+h)2σ)2.

    In order to investigate the direction of the NS bifurcation, we consider the first Lyapunov exponent, which is derived as follows:

    L=([Re((12ξ1)ξ221ξ1τ20τ11)12|τ11|2|τ02|2+Re(ξ2τ21)])ε=0, (3.20)

    where

    τ20=18[χenenχfnfn+2Υenfn+i(ΥenenΥfnfn2χenfn)],τ11=14[χenen+χfnfn+i(Υenen+Υfnfn)],τ02=18[χenenχfnfn2Υenfn+i(ΥenenΥfnfn+2χenfn)],τ21=116[χenenen+χenfnfn+Υenenfn+Υfnfnfn+i(Υenenen+Υenfnfnχenenfnχfnfnfn)].

    Based on the aforementioned study, the following conclusion can be drawn:

    Theorem 3.3. Assume that condition (5) of Theorem 2.6 is fulfilled. If condition (3.17) is satisfied and L presented in (3.20) is nonzero, then (1.2) undergoes NS bifurcation at E2 when the parameter r varies in a small neighborhood of R2=1σ(1+h)(σ+κσ+κσh2κ2κh). Furthermore, if L<0 (or L>0), the NS bifurcation at E2 is classified as supercritical (or subcritical), and a single closed invariant curve bifurcates from E2, which is attracting (or repelling).

    The result shows that small changes in environmental or biological conditions may generate significant changes in plant-herbivore population dynamics, resulting in an NS bifurcation. This indicates that the system shifts from steady-state oscillations to quasiperiodic oscillations, in which populations no longer follow simple cycles but instead show increasingly complicated and interconnected patterns.

    The purpose of this section is to verify our theoretical results through numerical simulations. Calculations are done using MATHEMATICA, while MATLAB is used for plotting graphs.

    Assume that h=0.15,κ=0.01,σ=0.04 and x0=23.4,y0=0.4. For these parametric values, the boundary FP, E1, is obtained as (24,0). The system (1.2) experiences TC bifurcation at E1 when r0.629565. The eigenvalues of J(E1) are ξ1=1 and ξ2=0.6826. Moreover, some careful calculations give

    ω(en,ε)=en0.039995en20.0050397en34enε0.56756en2ε.ω(0,0)=0, ωen(0,0)=1, ωε(0,0)=0,ωenen(0,0)=0.07998990, ωenε(0,0)=40.

    It verifies Theorem 3.1. We present phase portraits in Figure 1 to graphically verify the TC bifurcation at E1. Figure 1a illustrates that E1 is unstable while E2 is stable. In Figure 1b, it is shown that E1 and E2 collide at the critical value r=0.629565 and, subsequently, they exchange stability. By choosing r=0.65, Figure 1c demonstrates that E1 becomes stable while E2 turns unstable. This phenomenon is known as TC bifurcation.

    Figure 1.  Phase portraits of system (1.2) for some values of r. Fixed parameter values are h=0.15,κ=0.01,σ=0.04, and initial conditions are x0=23.4,y0=0.4.

    Our modified system (1.2) is a generalization of the system (1.1) studied in [11]. We can obtain the original system (1.1) by substituting h=0 into the modified system (1.2). In this analysis, we consider the same parameter values κ=0.01,σ=0.04 and initial conditions x0=23.4,y0=0.4, which were used in [11]. Moreover, we assume that h=0.15. We vary r in a bigger interval [0.2,0.9] instead of [0.2,0.7] to address additional important dynamics that are not reported in the numerical analysis presented in [11]. Our investigation shows that harvesting is helpful in stabilizing the plant-herbivore system. The system (1.2) undergoes NS bifurcation for r0.379565 at FP E2=(24,0.658648). The eigenvalues of J(E2) are ξ1,2=0.8413±0.540569i with |ξ1,2|=1. Moreover, some careful calculations give

    τ20=0.0569996+0.0006557i,τ11=0.0294408+0.0620942i,τ02=0.15511+0.0988851i,τ21=0.006887+0.0094604i.

    Thus, it is obtained that L=0.016401<0, which verifies Theorem 3.3. Thus, NS bifurcation is supercritical, and a closed invariant attracting curve will emerge from E2. Bifurcation diagrams are depicted in Figures 2a and 2b. These illustrate that r stabilizes the system (1.2). The corresponding maximum Lyapunov exponent (MLE) graph is given in Figure 2c.

    Figure 2.  Bifurcation diagrams and MLE graph of system (1.2) varying r. Fixed parameter values are h=0.15,κ=0.01,σ=0.04, and initial conditions are x0=23.4,y0=0.4.

    Next, Figure 3 shows phase portraits of (1.2) for different values of r. It can be seen that E2 is a sink for r>0.379565, yet it becomes unstable at r0.379565 due to the occurrence of NS bifurcation. For r0.379565, an invariant curve emanates from E2, with its radius expanding as r is decreasing.

    Figure 3.  Phase portraits of (1.2) for different values of r and using h=0.15,κ=0.01,σ=0.04,x0=23.4,y0=0.4.

    In [11], for the same parametric values and initial conditions, the positive FP is a sink for 0.51<r<0.76. Their investigation shows that the system experiences NS bifurcation at r0.51. Although it is not highlighted in their paper, positive FP vanishes at r0.76 and a boundary FP appears due to TC bifurcation at the boundary FP. Later, this boundary FP also disappears for r1 and the system has only a trivial FP, E0, which is then stable for r>1. The same phenomenon is observed in our modified system (1.2). By adding the harvesting effect, the positive FP of the modified system (1.2) is a sink for 0.379565<r<0.629565. The system experiences NS bifurcation at positive FP, E2, for r0.379565. The positive FP E2 disappears for r0.629565. The system (1.2) is experiencing TC bifurcation at boundary FP E1 for r0.629565. Before r0.629565, E2 is the only nontrivial FP, which is a sink for 0.379565<r<0.629565. After r0.629565, E1 is the only nontrivial FP, which is a sink for 0.629565<r<0.869565. Later, the boundary FP E1 also disappears for r0.869565. The system (1.2) is experiencing TC bifurcation at trivial FP E0 for r0.869565. For 0.629565<r<0.869565, the system possesses one boundary FP E1 and one trivial FP E0. Here, E1 is stable, but E0 is unstable. For r>0.869565, the boundary FP, E1, disappears, but the trivial FP E0 exists, and it also becomes stable.

    In the previous section, it was shown that the harvesting effect has a stabilizing impact on plant-herbivore systems. One natural question arises: What will happen if we harvest without any set upper and lower limits? Will it still stabilize the system or destabilize if we significantly increase or decrease the harvesting limit? Ecologically, threshold harvesting is beneficial for both plants and herbivores. In this numerical analysis, it is shown that a balanced amount of harvesting is advantageous for both plants and herbivores. For this purpose, assume that r=0.022,κ=0.01,σ=0.5, and vary h. For these parametric values, the NS bifurcation value is h1=18.2308 and the PD bifurcation value is h2=18.5673. The FP E2 is a sink if 18.2308<h<18.5673. The FP E2 becomes unstable for h18.2308 due to NS bifurcation. Moreover, it is unstable for h18.5673 due to NS bifurcation.

    The positive FP is obtained as E2=(1,0.909091) for h=h1. The eigenvalues of J(E2) are ξ1,2=0.849112±0.528212i with |ξ1,2|=1. Moreover, some careful calculations give

    τ20=5.44983+14.6964i,τ11=6.73996+30.7151i,τ02=1.40817+16.4319i,τ21=4.7593+13.5547i.

    Thus, it is obtained that L=93.568<0, which verifies Theorem 3.3. Thus, NS bifurcation is supercritical, and a closed invariant attracting curve will emerge from E2. The bifurcation diagrams of system (1.2) are given in Figures 4a and 4b by using initial conditions x0=1.1,y0=0.9, and varying h in [18.20,18.30]. The corresponding MLE graph is given in Figure 4e. Next, the positive FP is obtained as E2=(1,0.86844) for h=h2. The eigenvalues of J(E2) are ξ1=1 and ξ2=0.828791. Moreover, some careful calculations give l1=10.61680 and l2=6.906680, which verifies Theorem 3.2. Since l2>0, a stable orbit of period-2 emanates from E2. The bifurcation diagrams are given in Figures 4c and 4d by using initial conditions x0=1.1,y0=0.85, and varying h in [18.50,18.65]. The corresponding MLE graph is given in Figure 4f. Figures depict phase portraits of (1.2) for different values of r. These illustrate that E2 is a sink if 18.2308<h<18.5673. The system undergoes NS bifurcation at h18.2308 and PD bifurcation at h18.5673.

    Figure 4.  Bifurcation diagrams and MLE graphs of system (1.2) varying h and fixing r=0.022,κ=0.01,σ=0.5. The initial conditions are (a, b, e) x0=1.1,y0=0.9, (c, d, f) x0=1.1,y0=0.85.

    Figures 4 and 5 illustrate that the higher or lower harvesting effect destabilizes the system, resulting in more complicated dynamical behaviors. Thus, a balanced amount of harvesting is advantageous for both plants and herbivores. Ecologically, under moderate harvesting, there is a stable coexistence of both plants and herbivores, represented by the positive FP. As harvesting increases, herbivores struggle to find food, eventually leading to extinction. Continued harvesting eventually causes the plant population to decline as well, resulting in the extinction of plants. This indicates that moderate harvesting can sustain both populations, but excessive harvesting leads to the collapse of the entire ecosystem.

    Figure 5.  Phase portraits of system (1.2) for some values of r and fixing r=0.022,κ=0.01,σ=0.5. The initial conditions are (a-c) x0=1.1,y0=0.9, (d-f) x0=1.1,y0=0.85.

    The aim of this section is to investigate the influence of the harvesting effect on the dynamics of system (1.2). Incorporating a harvesting effect into a plant-herbivore system yields insightful findings that underscore the intricate dynamics of species interactions in ecological systems. While the harvesting effect is directly applied to the plant population, its influence extends beyond that, indirectly impacting herbivore dynamics through the interaction between plant and herbivore populations. One can observe in Figure 4 that system (1.2) undergoes PD and NS bifurcations at the positive FP E2 by varying the harvesting effect parameter h. These show that the harvesting effect can stabilize or destabilize the system. Notably, a moderate level of the harvesting effect appears to be advantageous for both plant and herbivore populations, as evidenced by the occurrence of these bifurcations. Furthermore, the positive FP E2=(x2,y2) is dependent on the harvesting effect parameter h. Through simple calculations, one can obtain that dx2dh=0 and dy2dh=1r(1+h)2<0. Ecologically, it represents that when the harvesting effect strengthens in the plant population, it might become slightly more difficult for herbivores to survive. A decrease in plant availability could lead to reduced reproduction rates or increased mortality among herbivores, ultimately resulting in a decrease in the population growth rate of herbivores.

    This study explores the implications of adding a harvesting impact to the dynamics of a plant-herbivore system. Our findings highlight the intricate interplay between species interactions in ecological systems, emphasizing the significance of the harvesting effect. The harvesting effect affects the plant population directly, but it also has an impact on herbivore dynamics because of the complex interactions that occur between herbivores and plant populations. The existence and stability of FPs are examined. Furthermore, a comprehensive analysis of local bifurcations at the positive FP is conducted. The investigation reveals that the system (1.2) undergoes PD, TC, and NS bifurcations. These bifurcations demonstrate the system's sensitivity to small parameter changes, which result in significant changes in population dynamics. The role of transcritical bifurcation is to emphasize the critical transition from coexistence to dominance of plant populations; PD bifurcations show the formation of complex oscillatory patterns. NS bifurcations, in turn, indicate the transition from steady-state to quasiperiodic oscillations, displaying the complexity and adaptability of ecological interactions.

    The plant-herbivore system without harvesting effect was investigated in [11]. Our plant-herbivore system is more general, and we can obtain the results of [11] by just replacing h=0 into our results. Although we obtain similar results as in [11], the following are the main differences observed in both plant-herbivore systems:

    ● With the same parametric values and initial conditions, the positive FP of system (1.1) is a sink inside the range 0.51<r<0.76. However, when the harvesting effect is included, the positive FP of system (1.2) is a sink inside the range 0.379565<r<0.629565 for h=0.15. This illustrates that harvesting helps to stabilize the plant-herbivore system for lower values of r.

    ● In the absence of harvesting, the system (1.1) does not experience PD bifurcation. However, when the harvesting is added, the system (1.2) experiences PD bifurcation at the positive FP.

    An interesting finding in our paper is that a moderate amount of the harvesting effect benefits both plant and herbivore populations, emphasizing the intricate impact of harvesting on ecosystem dynamics. These results have significant biological consequences, as they reveal the intricate relationship between human involvement and the capacity of ecosystems to recover. They also highlight the need to adopt sustainable harvesting methods to safeguard the long-term well-being and stability of ecological systems. In the future, one can try to investigate the global dynamics and codimension-two bifurcations for the plant-herbivore system (1.2).

    Mohammed Alsubhi: Formal Analysis, Validation, Writing-original draft, Methodology; Rizwan Ahmed: Conceptualization, Investigation, Writing-review & editing, Software, Supervision; Ibrahim Alraddadi: Conceptualization, Validation, Writing-review & editing, Visualization, Supervision, Methodology; Faisal Alsharif: Formal Analysis, Software, Visualization, Methodology; Muhammad Imran: Formal Analysis, Investigation, Writing-original draft, Software. All authors have read and approved the final version of the manuscript for publication.

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

    The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia for funding this research work through the project number (445-9-923).

    All authors declare no conflicts of interest in this paper.



    [1] D. Choquenot, J. Parkes, Setting thresholds for pest control: How does pest density affect resource viability?, Biol. Conserv., 99 (2001), 29–46. https://doi.org/10.1016/S0006-3207(00)00186-5 doi: 10.1016/S0006-3207(00)00186-5
    [2] L. Edelstein-Keshet, Mathematical theory for plant-herbivore systems, J. Math. Biol., 24 (1986), 25–58. https://doi.org/10.1007/bf00275719 doi: 10.1007/bf00275719
    [3] E. P. Holland, R. P. Pech, W. A. Ruscoe, J. P. Parkes, G. Nugent, R. P. Duncan, Thresholds in plant-herbivore interactions: predicting plant mortality due to herbivore browse damage, Oecologia, 172 (2013), 751–766. https://doi.org/10.1007/s00442-012-2523-5 doi: 10.1007/s00442-012-2523-5
    [4] Z. Feng, Z. Qiu, R. Liu, D. L. DeAngelis, Dynamics of a plant-herbivore-predator system with plant-toxicity, Math. Biosci., 229 (2011), 190–204. https://doi.org/10.1016/j.mbs.2010.12.005 doi: 10.1016/j.mbs.2010.12.005
    [5] K. C. Abbott, G. Dwyer, Food limitation and insect outbreaks: complex dynamics in plant-herbivore models, J. Anim. Ecol., 76 (2007), 1004–1014. https://doi.org/10.1111/j.1365-2656.2007.01263.x doi: 10.1111/j.1365-2656.2007.01263.x
    [6] G. Sui, M. Fan, I. Loladze, Y. Kuang, The dynamics of a stoichiometric plant-herbivore model and its discrete analog, Math. Biosci. Eng., 4 (2007), 29–46. https://doi.org/10.3934/mbe.2007.4.29 doi: 10.3934/mbe.2007.4.29
    [7] Y. Kang, D. Armbruster, Y. Kuang, Dynamics of a plant-herbivore model, J. Biol. Dynam., 2 (2008), 89–101. https://doi.org/10.1080/17513750801956313 doi: 10.1080/17513750801956313
    [8] Q. Din, A. A. Elsadany, H. Khalil, Neimark-Sacker bifurcation and chaos control in a fractional-order plant-herbivore model, Discrete Dyn. Nat. Soc., 2017 (2017), 6312964. https://doi.org/10.1155/2017/6312964 doi: 10.1155/2017/6312964
    [9] M. S. Khan, M. Samreen, M. Ozair, T. Hussain, E. M. Elsayed, J. F. Gomez-Aguilar, On the qualitative study of a two-trophic plant-herbivore model, J. Math. Biol., 85 (2022), 34. https://doi.org/10.1007/s00285-022-01809-0 doi: 10.1007/s00285-022-01809-0
    [10] E. Beso, S. Kalabusic, E. Pilav, Food-limited plant-herbivore model: bifurcations, persistence, and stability, Math. Biosci., 370 (2024), 109157. https://doi.org/10.1016/j.mbs.2024.109157 doi: 10.1016/j.mbs.2024.109157
    [11] M. S. Shabbir, Q. Din, M. D. la Sen, J. F. Gómez-Aguilar, Exploring dynamics of plant-herbivore interactions: bifurcation analysis and chaos control with Holling type-Ⅱ functional response, J. Math. Biol., 88 (2024), 8. https://doi.org/10.1007/s00285-023-02020-5 doi: 10.1007/s00285-023-02020-5
    [12] Z. Feng, D. L. DeAngelis, Mathematical models of plant-herbivore interactions, Chapman and Hall/CRC, 2017. https://doi.org/10.1201/9781315154138
    [13] S. Kartal, A. Debbouche, Dynamics of a plant-herbivore model with differential-difference equations, Cogent Mathematics, 3 (2016), 1136198. https://doi.org/10.1080/23311835.2015.1136198 doi: 10.1080/23311835.2015.1136198
    [14] E. Beso, S. Kalabusic, E. Pilav, A. Bilgin, Dynamics of a plant-herbivore model subject to Allee effects with logistic growth of plant biomass, Int. J. Bifurcat. Chaos, 33 (2023), 2330026. https://doi.org/10.1142/s0218127423300264 doi: 10.1142/s0218127423300264
    [15] Q. Din, Global behavior of a plant-herbivore model, Adv. Differ. Equ., 2015 (2015), 119. https://doi.org/10.1186/s13662-015-0458-y doi: 10.1186/s13662-015-0458-y
    [16] A. Q. Khan, J. Ma, D. Xiao, Bifurcations of a two-dimensional discrete time plant-herbivore system, Commun. Nonlinear Sci., 39 (2016), 185–198. https://doi.org/10.1016/j.cnsns.2016.02.037 doi: 10.1016/j.cnsns.2016.02.037
    [17] M. Y. Hamada, Dynamical analysis of a discrete-time plant-herbivore model, Arab. J. Math., 13 (2024), 121–131. https://doi.org/10.1007/s40065-023-00442-z doi: 10.1007/s40065-023-00442-z
    [18] T. Saha, M. Bandyopadhyay, Dynamical analysis of a plant-herbivore model bifurcation and global stability, J. Appl. Math. Comput., 19 (2005), 327–344. https://doi.org/10.1007/bf02935808 doi: 10.1007/bf02935808
    [19] Y. Li, Z. Feng, R. Swihart, J. Bryant, N. Huntly, Modeling the impact of plant toxicity on plant-herbivore dynamics, J. Dyn. Diff. Equat., 18 (2006), 1021–1042. https://doi.org/10.1007/s10884-006-9029-y doi: 10.1007/s10884-006-9029-y
    [20] C. Castillo-Chavez, Z. Feng, W. Huang, Global dynamics of a plant-herbivore model with toxin-determined functional response, SIAM J. Appl. Math., 72 (2012), 1002–1020. https://doi.org/10.1137/110851614 doi: 10.1137/110851614
    [21] E. M. Elsayed, Q. Din, Period-doubling and Neimark-Sacker bifurcations of plant-herbivore models, Adv. Differ. Equ., 2019 (2019), 271. https://doi.org/10.1186/s13662-019-2200-7 doi: 10.1186/s13662-019-2200-7
    [22] Q. Din, M. S. Shabbir, M. A. Khan, K. Ahmad, Bifurcation analysis and chaos control for a plant-herbivore model with weak predator functional response, J. Biol. Dynam., 13 (2019), 481–501. https://doi.org/10.1080/17513758.2019.1638976 doi: 10.1080/17513758.2019.1638976
    [23] S. Kalabusic, E. Pilav, Bifurcations, permanence and local behavior of the plant-herbivore model with logistic growth of plant biomass, Qual. Theory Dyn. Syst., 21 (2022), 26. https://doi.org/10.1007/s12346-022-00561-6 doi: 10.1007/s12346-022-00561-6
    [24] L. J. Allen, M. J. Strauss, H. G. Thorvilson, W. N. Lipe, A preliminary mathematical model of the apple twig borer (Coleoptera: Bostrichidae) and grapes on the texas high plains, Ecol. Model., 58 (1991), 369–382. https://doi.org/10.1016/0304-3800(91)90046-4 doi: 10.1016/0304-3800(91)90046-4
    [25] L. J. Allen, M. K. Hannigan, M. J. Strauss, Mathematical analysis of a model for a plant-herbivore system, Bull. Math. Biol., 55 (1993), 847–864. https://doi.org/10.1016/S0092-8240(05)80192-2 doi: 10.1016/S0092-8240(05)80192-2
    [26] H. P. Benoit, D. P. Swain, Impacts of environmental change and direct and indirect harvesting effects on the dynamics of a marine fish community, Can. J. Fish. Aquat. Sci., 65 (2008), 2088–2104. https://doi.org/10.1139/f08-112 doi: 10.1139/f08-112
    [27] S. A. Khamis, J. M. Tchuenche, M. Lukka, M. Heilio, Dynamics of fisheries with prey reserve and harvesting, Int. J. Comput. Math., 88 (2011), 1776–1802. https://doi.org/10.1080/00207160.2010.527001 doi: 10.1080/00207160.2010.527001
    [28] C. K. Yosi, R. J. Keenan, J. C. Fox, Forest dynamics after selective timber harvesting in Papua New Guinea, Forest Ecol. Manag., 262 (2011), 895–905. https://doi.org/10.1016/j.foreco.2011.06.007 doi: 10.1016/j.foreco.2011.06.007
    [29] D. N. Rasquinha, D. R. Mishra, Impact of wood harvesting on mangrove forest structure, composition and biomass dynamics in india, Estuar. Coast. Shelf Sci., 248 (2021), 106974. https://doi.org/10.1016/j.ecss.2020.106974 doi: 10.1016/j.ecss.2020.106974
    [30] R. Ahmed, Complex dynamics of a fractional-order predator-prey interaction with harvesting, Open Journal of Discrete Applied Mathematics, 3 (2020), 24–32. https://doi.org/10.30538/psrp-odam2020.0040 doi: 10.30538/psrp-odam2020.0040
    [31] Y. Tian, H. M. Li, The study of a predator-prey model with fear effect based on state-dependent harvesting strategy, Complexity, 2022 (2022), 9496599. https://doi.org/10.1155/2022/9496599 doi: 10.1155/2022/9496599
    [32] M. Imran, M. B. Almatrafi, R. Ahmed, Stability and bifurcation analysis of a discrete predator-prey system of Ricker type with harvesting effect, Commun. Math. Biol. Neurosci., 2024 (2024), 11. https://doi.org/10.28919/cmbn/8313 doi: 10.28919/cmbn/8313
    [33] M. Virtala, Optimal harvesting of a plant-hervibore system: lichen and reindeer in northern Finland, Ecol. Model., 60 (1992), 233–255. https://doi.org/10.1016/0304-3800(92)90035-d doi: 10.1016/0304-3800(92)90035-d
    [34] M. D. Asfaw, S. M. Kassa, E. M. Lungu, Co-existence thresholds in the dynamics of the plant-herbivore interaction with Allee effect and harvest, Int. J. Biomath., 11 (2018), 1850057. https://doi.org/10.1142/s1793524518500572 doi: 10.1142/s1793524518500572
    [35] M. Xiao, J. Cao, Hopf bifurcation and non-hyperbolic equilibrium in a ratio-dependent predator-prey model with linear harvesting rate: analysis and computation, Math. Comput. Model., 50 (2009), 360–379. https://doi.org/10.1016/j.mcm.2009.04.018 doi: 10.1016/j.mcm.2009.04.018
    [36] L. Ji, C. Wu, Qualitative analysis of a predator-prey model with constant-rate prey harvesting incorporating a constant prey refuge, Nonlinear Anal.-Real, 11 (2010), 2285–2295. https://doi.org/10.1016/j.nonrwa.2009.07.003 doi: 10.1016/j.nonrwa.2009.07.003
    [37] D. Jana, R. Agrawal, R. K. Upadhyay, G. Samanta, Ecological dynamics of age selective harvesting of fish population: maximum sustainable yield and its control strategy, Chaos Soliton. Fract., 93 (2016), 111–122. https://doi.org/10.1016/j.chaos.2016.09.021 doi: 10.1016/j.chaos.2016.09.021
    [38] A. Xiao, C. Lei, Dynamic behaviors of a non-selective harvesting single species stage-structured system incorporating partial closure for the populations, Adv. Differ. Equ., 2018 (2018), 245. https://doi.org/10.1186/s13662-018-1709-5 doi: 10.1186/s13662-018-1709-5
    [39] L. J. S. Allen, An introduction to mathematical biology, Pearson/Prentice Hall, 2007.
    [40] Q. Din, Neimark-Sacker bifurcation and chaos control in Hassell-Varley model, J. Differ. Equ. Appl., 23 (2017), 741–762. https://doi.org/10.1080/10236198.2016.1277213 doi: 10.1080/10236198.2016.1277213
    [41] Q. Din, M. I. Khan, A discrete-time model for consumer-resource interaction with stability, bifurcation and chaos control, Qual. Theor. Dyn. Syst., 20 (2021), 56. https://doi.org/10.1007/s12346-021-00488-4 doi: 10.1007/s12346-021-00488-4
    [42] A. A. Khabyah, R. Ahmed, M. S. Akram, S. Akhtar, Stability, bifurcation, and chaos control in a discrete predator-prey model with strong Allee effect, AIMS Mathematics, 8 (2023), 8060–8081. https://doi.org/10.3934/math.2023408 doi: 10.3934/math.2023408
    [43] R. Ahmed, M. B. Almatrafi, Complex dynamics of a predator-prey system with Gompertz growth and herd behavior, Int. J. Anal. Appl., 21 (2023), 100. https://doi.org/10.28924/2291-8639-21-2023-100 doi: 10.28924/2291-8639-21-2023-100
    [44] R. Ahmed, M. Rafaqat, I. Siddique, M. A. Arefin, Complex dynamics and chaos control of a discrete-time predator-prey model, Discrete Dyn. Nat. Soc., 2023 (2023), 8873611. https://doi.org/10.1155/2023/8873611 doi: 10.1155/2023/8873611
    [45] A. C. J. Luo, Regularity and complexity in dynamical systems, New York: Springer, 2012. https://doi.org/10.1007/978-1-4614-1524-4
    [46] J. Guckenheimer, P. Holmes, Nonlinear oscillations, dynamical systems, and bifurcations of vector fields, New York: Springer, 1983. https://doi.org/10.1007/978-1-4612-1140-2
    [47] S. Wiggins, Introduction to applied nonlinear dynamical systems and chaos, New York: Springer, 2003. https://doi.org/10.1007/b97481
    [48] W. Yao, X. Li, Complicate bifurcation behaviors of a discrete predator-prey model with group defense and nonlinear harvesting in prey, Appl. Anal., 102 (2023), 2567–2582. https://doi.org/10.1080/00036811.2022.2030724 doi: 10.1080/00036811.2022.2030724
    [49] P. A. Naik, M. Amer, R. Ahmed, S. Qureshi, Z. Huang, Stability and bifurcation analysis of a discrete predator-prey system of Ricker type with refuge effect, Math. Biosci. Eng., 21 (2024), 4554–4586. https://doi.org/10.3934/mbe.2024201 doi: 10.3934/mbe.2024201
  • This article has been cited by:

    1. Seval Işık, Figen Kangalgil, MD. Jasim Uddin, Sarker MD. Sohel Rana, An Investigation of the Discrete-Time Model Subject to Immigration, Harvesting, and Allee Effect, 2025, 35, 0218-1274, 10.1142/S0218127425500221
  • Reader Comments
  • © 2024 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(1205) PDF downloads(70) Cited by(1)

Figures and Tables

Figures(5)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog