Processing math: 100%
Research article

On generalized discrete Ricker map

  • Received: 22 August 2024 Revised: 23 September 2024 Accepted: 08 October 2024 Published: 15 October 2024
  • MSC : 37N25, 37Gxx, 37D45, 37M10

  • In recent years, conventional Ricker maps have enjoyed widespread applications across crucial domains such as modeling and security. However, their limitation to a single changeable parameter poses constraints on their adaptability. This paper introduces a generalized form of the Ricker map, incorporating arbitrary powers, thus offering enhanced versatility compared to the traditional Ricker map. By introducing an additional parameter (arbitrary power), the map gains increased degrees of freedom, thereby accommodating a broader spectrum of applications. Consequently, the conventional Ricker map emerges as merely a special case within each proposed framework. This newfound parameter enhances system flexibility and elucidates the conventional system's performance across diverse contexts. Through numerous illustrations, we meticulously investigate the impact of the arbitrary power and equation parameters on equilibrium points, their positions, basin of attraction, stability conditions, and bifurcation diagrams, including the emergence of chaotic behavior.

    Citation: H. El-Metwally, Ibraheem M. Alsulami, M. Y. Hamada. On generalized discrete Ricker map[J]. AIMS Mathematics, 2024, 9(10): 29235-29249. doi: 10.3934/math.20241417

    Related Papers:

    [1] Vinoth Seralan, R. Vadivel, Dimplekumar Chalishajar, Nallappan Gunasekaran . Dynamical complexities and chaos control in a Ricker type predator-prey model with additive Allee effect. AIMS Mathematics, 2023, 8(10): 22896-22923. doi: 10.3934/math.20231165
    [2] Wei Li, Qingkai Xu, Xingjian Wang, Chunrui Zhang . Dynamics analysis of spatiotemporal discrete predator-prey model based on coupled map lattices. AIMS Mathematics, 2025, 10(1): 1248-1299. doi: 10.3934/math.2025059
    [3] Xinna Mao, Hongwei Feng, Maryam A. Al-Towailb, Hassan Saberi-Nik . Dynamical analysis and boundedness for a generalized chaotic Lorenz model. AIMS Mathematics, 2023, 8(8): 19719-19742. doi: 10.3934/math.20231005
    [4] Junhong Li, Ning Cui . A hyperchaos generated from Rabinovich system. AIMS Mathematics, 2023, 8(1): 1410-1426. doi: 10.3934/math.2023071
    [5] Mohamed Lamine Sahari, Abdel-Kaddous Taha, Louis Randriamihamison . Exploring three periodic point dynamics in $ 2 $D spatiotemporal discrete systems. AIMS Mathematics, 2025, 10(3): 5021-5051. doi: 10.3934/math.2025230
    [6] Zhongcai Zhu, Xue He . Rich and complex dynamics of a time-switched differential equation model for wild mosquito population suppression with Ricker-type density-dependent survival probability. AIMS Mathematics, 2023, 8(12): 28670-28689. doi: 10.3934/math.20231467
    [7] Abdul Qadeer Khan, Ayesha Yaqoob, Ateq Alsaadi . Neimark-Sacker bifurcation, chaos, and local stability of a discrete Hepatitis C virus model. AIMS Mathematics, 2024, 9(11): 31985-32013. doi: 10.3934/math.20241537
    [8] A. E. Matouk . Chaos and hidden chaos in a 4D dynamical system using the fractal-fractional operators. AIMS Mathematics, 2025, 10(3): 6233-6257. doi: 10.3934/math.2025284
    [9] Yongfeng Li, Song Huang, Zhongyi Xiang . A state-dependent impulsive system with ratio-dependent action threshold for investigating SIR model. AIMS Mathematics, 2024, 9(2): 4781-4804. doi: 10.3934/math.2024231
    [10] Figen Kangalgil, Seval Isșık . Effect of immigration in a predator-prey system: Stability, bifurcation and chaos. AIMS Mathematics, 2022, 7(8): 14354-14375. doi: 10.3934/math.2022791
  • In recent years, conventional Ricker maps have enjoyed widespread applications across crucial domains such as modeling and security. However, their limitation to a single changeable parameter poses constraints on their adaptability. This paper introduces a generalized form of the Ricker map, incorporating arbitrary powers, thus offering enhanced versatility compared to the traditional Ricker map. By introducing an additional parameter (arbitrary power), the map gains increased degrees of freedom, thereby accommodating a broader spectrum of applications. Consequently, the conventional Ricker map emerges as merely a special case within each proposed framework. This newfound parameter enhances system flexibility and elucidates the conventional system's performance across diverse contexts. Through numerous illustrations, we meticulously investigate the impact of the arbitrary power and equation parameters on equilibrium points, their positions, basin of attraction, stability conditions, and bifurcation diagrams, including the emergence of chaotic behavior.



    In recent decades, the study of nonlinear dynamics in population models has made significant progress, unveiling complex behaviors such as chaos and bifurcations. [11] demonstrated that even simple mathematical models, including population dynamics, can lead to highly complex dynamics, sometimes resulting in chaotic behavior in seemingly straightforward systems. Similarly, the pioneering work of [14] laid the foundation for understanding stock-recruitment relationships in fisheries, with the Ricker map emerging as a central tool in discrete population modeling.

    The investigation of stability in population models has continued to evolve. [1,12] contributed to this by analyzing global stability, shedding light on how various conditions can either stabilize or destabilize population dynamics. This line of inquiry was further advanced by [19], who explored global stability in various population models, reinforcing the importance of stability analysis in ecological studies. More recently, [2] revisited the stability and instability in one-dimensional population models, broadening our understanding of discrete-time dynamics.

    The Ricker map has frequently been employed in predator-prey models, serving as a cornerstone for research into complex population dynamics. Particularly, Ricker-type predator-prey models have been extensively studied in the context of bifurcations and chaotic behavior. [9,10] conducted detailed bifurcation analyses within these models, revealing intricate predator-prey interactions across various parameter regimes. Their work underscores how different parameter settings can significantly influence population persistence and the stability of predator-prey relationships [5,8,9].

    Beyond ecological studies, mathematicians like [4,15] have deepened our understanding of nonlinear dynamics and chaos, providing a theoretical framework applicable to both biological and physical systems. Works such as [17,18] offer comprehensive introductions to nonlinear dynamical systems and chaos, providing valuable insights into the behavior of both discrete and continuous systems.

    Building on foundational concepts in population dynamics, [3,6,7] expanded the study of stability, bifurcation, and chaos to piecewise and discrete systems, demonstrating their wide applicability across various fields. Radwan et al. extended this understanding by investigating generalized discrete logistic maps, providing additional insights into the complexity of long-term population dynamics [13].

    The study of Ricker-type models has seen significant advancements and generalizations. In particular, [16] generalized the conventional Ricker map by introducing an arbitrary power to the population variable x, resulting in the following form:

    xn+1=xδnerxn. (1.1)

    Building upon Stewart's work, we further extend the Ricker map (1.1) by introducing an additional parameter to increase its generality. The new generalized Ricker map with arbitrary powers is given by:

    xn+1=xαner(1xβn), (1.2)

    where α, β, and r are positive parameters. This extension allows for a more flexible analysis of population dynamics under various conditions.

    In the subsequent sections, we investigate the behavior and properties of the proposed generalized Ricker map. Section 2 focuses on the existence and stability of equilibria, offering both theoretical and numerical insights. Section 3 provides bifurcation analyses and numerical simulations, exploring how parameter changes affect the system's dynamics. Finally, Section 4 concludes with a summary of the findings and a comparison of the results, discussing the broader implications of the generalized Ricker map in population dynamics.

    The following passage sets the foundation for our inquiry into the existence and stability of equilibria within the generalized Ricker map represented by Eq (1.2):

    f(x,r,α)=xαer(1xβ), (2.1)

    where x0, α>0, β>0, and r>0. Our investigation is structured around the parameter α, which guides our analysis into two distinct cases: the first case where α>1, and the second where 0<α1.

    The function f may exhibit either one or two positive fixed points, contingent upon the number of intersections between its graph and y=x. In Figure 1, the graph of f for two distinct cases is illustrated. In Figure 1(a,c), there are two positive equilibria, while in Figure 1(b), there is only one positive fixed point. The next theorem will provide us with the conditions to distinguish between these different cases.

    Figure 1.  Number of intersections between graph of f and y=x.

    Theorem 1. For the difference Eq (1.2) with α>1, the following statements are true:

    (ⅰ) If α=rβ+1, the Eq (1.2) has two equilibria 0 and one positive equilibrium point ˉx=1;

    (ⅱ) If αrβ+1, the Eq (1.2) has three equilibria 0 and two positive equilibria ˉx=1 and ˜x.

    Proof. To establish the validity of the theorem, we delve into the solutions of the equation:

    x=xαer(1xβ). (2.2)

    It is evident that x=0 and x=1 satisfy Eq (2.2). To explore further positive solutions, we introduce the functions:

    f(x)=xαer(1xβ), (2.3)

    and

    h(x)=1x(α1)er(1xβ). (2.4)

    The roots of h(x) coincide with the fixed points of f(x).

    The derivative of h(x) is:

    h(x)=[rβxβα+1]x(α2)er(1xβ),

    which equals zero when x=βα1rβ.

    Given that

    h(x)=(α1)α+β3βr3αββ3α+ββerβα+1β>0

    for α>1. Now, at x=βα1rβ, function h has a minimum. For x(0,βα1rβ), rβxβα+1<0, making h decreasing on (0,βα1rβ). Conversely, for x(βα1rβ,), rβxβα+1>0, indicating h increases on (βα1rβ,). Hence, at x=βα1rβ, there is a global minimum of h. Notably, h(0)=1 and limxh(x)=1. If h(βα1rβ)=0, then h(x)>0 for all xβα1rβ, indicating h possesses one zero only at x=1. This condition holds true when

    βα1rβ=1

    which is equivalent to α=rβ+1. When α=rβ+1 holds, it implies that h(x) equals zero at x=βα1rβ=1, thereby completing the proof of part (ⅰ).

    A similar argument demonstrates that h(βα1rβ)<0 when αrβ+1. It's noteworthy that 0<˜x<βα1rβ<1 when α<rβ+1. Conversely, 0<1<βα1rβ<˜x when α>rβ+1.

    After defining the equilibria of Eq (1.2), the subsequent focus lies in determining their local asymptotic stability. This entails analyzing the behavior of the system in the vicinity of each equilibrium point to discern whether small disturbances lead to a return to equilibrium over time.

    Theorem 2. Consider the difference Eq (1.2) with α>1, β>0, and r>0. It can be deduced that

    (ⅰ) The equilibrium at 0 is locally asymptotically stable;

    (ⅱ) The equilibrium at 1 is locally asymptotically stable when α1β<r<α+1β;

    (ⅲ) The equilibrium at ˜x is locally asymptotically stable when 1<βα1rβ<˜x<βα+1rβ;

    Proof. Starting with Eq (2.1), we express the derivative f(x) as follows:

    f(x)=(αrβxβ)x(α1)er(1xβ).

    Upon evaluation, we find that f(0)=0<1, indicating the local asymptotic stability of the equilibrium at 0. This concludes the proof of part (ⅰ).

    For part (ⅱ), we examine f(1)=αrβ. The condition 1<f(1)<1 holds if and only if α1β<r<α+1β.

    Turning to part (ⅲ), let us analyze f(˜x)=(αrβ˜xβ)˜x(α1)er(1˜xβ), with the constraint ˜x(α1)er(1˜xβ)=1. This yields f(˜x)=αrβ˜xβ.

    Now, the condition 1<f(˜x)<1 translates to βα1rβ<˜x<βα+1rβ. However, this condition cannot hold when 0<˜x<1.

    To clarify this, note that ˜x and 1 are roots of the function h(x) in Eq (2.3). Moreover, this function has a global minimum value at βα1rβ. Therefore, if this minimum value falls within the interval (0,1), for ˜x to be stable, it should satisfy the inequality βα1rβ<˜x<1. However, such a scenario contradicts the nature of the function h(x), which is increasing on the interval (βα1rβ,). As a result, the stability condition for ˜x can be satisfied when 1<βα1rβ<˜x<βα+1rβ.

    From this theorem, it is evident that the equilibrium point x=1 may lose its stability via a fold bifurcation when the slope of the tangent at this point is 1. This occurs when α=rβ+1. On the other hand, a period-doubling bifurcation can arise when the slope of the tangent is 1, which happens when α=rβ1.

    As for the other positive equilibrium point, the potential bifurcation at this point is a period-doubling bifurcation, occurring at ˜x=βα+1rβ.

    Two Lemmas will now be introduced, playing a pivotal role in the analysis of the basin of attraction encompassing equilibrium points within the difference equation referenced as (1.2).

    Lemma 1. The function f(x)=xαer(1xβ) is increasing on (0,βαrβ) and decreasing on (βαrβ,).

    Proof. By computing the derivative of f(x):

    f(x)=(αrβxβ)x(α1)er(1xβ).

    The sign of f(x) is determined by the sign of αrβxβ.

    Lemma 2. Consider the difference equation defined by (1.2) with rβ+1>α>1. Under these conditions, there exists ˆx˜x such that f(ˆx)=˜x.

    Proof. The function f(x)=xαer(1xβ) achieves a global maximum at x=βαrβ, where f(βαrβ)>˜x. This is due to the fact that f is increasing on the interval (0,βαrβ) as demonstrated in Lemma 1. Furthermore, f decreases on (βαrβ,), and limxf(x)=0. Consequently, by the intermediate value theorem, there must exist ˆx in the interval (βαrβ,) such that f(ˆx)=˜x.

    Theorem 3. Let α>1 and rβ+1>α. Consider the difference equation defined by (1.2). Suppose there exists ˆx˜x such that f(ˆx)=˜x. Then the basin of attraction of the equilibrium at 0 is the set [0,˜x)(ˆx,).

    Proof. Let x0[0,˜x). We establish by induction that {xn} remains within [0,˜x). For n=0, this holds since x0[0,˜x). Assuming it holds for n=k (0xk<˜x), we have to prove that 0xk+1<˜x, as a result of that, we obtain {xn}[0,˜x) for all integers n. Given that f(x) is increasing on the interval (0,βαrβ) according to Lemma (1), and both 0 and ˜x are equilibrium points, we can establish the following inequality: f(0)f(xk)<f(˜x), which is equivalent to 0xk+1<˜x. Furthermore, since f(x)<x on the interval (0,˜x), it follows that xn+1=f(xn)<xn, indicating that the sequence xn is decreasing. As a consequence, since the sequence xn is both bounded and decreasing, it is guaranteed to converge. Thus,

    limnxn=x[0,˜x),

    and

    limnxn+1=limnxαner(1xβn)x=xαer(1xβ).

    As the only solution to x=xαer(1xβ) in [0,˜x) is x=0, {xn} converges to 0 if x0[0,˜x).

    For ˆx<x0<, as f(x) is decreasing on (ˆx,), we have limxf(x)=0<f(x0)<f(ˆx)=˜x. This implies 0<x1<˜x, as previously shown to converge to 0.

    Theorem 4. Consider the difference Eq (1.2), where α>1 and α1β<r<α+1β. Let ˆx˜x satisfy f(ˆx)=˜x. Then the basin of attraction of the positive equilibrium ˉx is the interval (˜x,ˆx).

    Proof. Let α>rβ. In this scenario, according to Lemma (1), the global maximum of the function f(x) exceeds 1. We proceed by dissecting the interval (˜x,ˆx) into four sub-intervals to scrutinize the behavior of the sequence {xn} within each subdivision.

    Case 1. ˜x<x01: We aim to demonstrate that if x0(˜x,1], then {xn} converges to 1. We employ an inductive argument akin to the proof of Theorem (2.3). Suppose x0(˜x,1]. We establish, by induction, that {xn} remains within the interval (˜x,1]. The case n=0 is trivial since x0(˜x,1]. Assume the statement holds for n=k, where k is a positive integer, i.e., ˜x<xk1. Given that f(x) is increasing on (0,βαrβ) by Lemma (1), we have f(˜x)<f(xk)f(1), implying ˜x<xk+11. This proves {xn}(˜x,1] for all integers n. Additionally, since f(x)>x on (˜x,1], we deduce xn+1=f(xn)>xn, rendering {xn} an increasing sequence. As {xn} is both bounded and increasing, it converges. Hence,

    limnxn=x(˜x,1]

    and

    limnxn+1=limnxαner(1xβn)x=xαer(1xβ).

    Given that the sole solution to x=xαer(1xβ) in (˜x,1] is x=1, {xn} converges to 1 if x0(˜x,1].

    Case 2. 1x0<βαrβ: The proof parallels Case 1.

    Case 3. βαrβ<x0x with f(x)=1:

    βαrβ<x0xf(βαrβ)>f(x0)f(x)βαrβ>f(βαrβ)>x11.

    This aligns with Case 2.

    Case 4. xx0<ˆx:

    xx0<ˆxf(x)f(x0)>f(ˆx)1x1>˜x.

    Thus, we revert to Case 1.

    When αrβ, the proof can be obtained in a similar approach.

    Theorem 5. Consider the difference Eq (1.2) with parameters satisfying 0<α1, β>0, and r>0. This equation possesses two equilibria: one at 0 and another at the positive value of 1.

    Proof. The equilibrium points of the difference Eq (1.2) correspond to the solutions of the following equation:

    x=f(x)x=xαer(1xβ)0=xxαer(1xβ)0=x(1x(α1)er(1xβ)).

    Hence, the solutions to f(x)=x are x=0, and the solutions to 1x(α1)er(1xβ)=0. Since x0,

    1x(α1)er(1xβ)=0x(1α)er(1xβ)=0.

    Define g(x)=x(1α)er(1xβ). We find that g(0)=er<0, g(1)=0, and limxg(x)=. According to the intermediate value theorem, since g(x) is continuous and real-valued, it must intersect the line x=0 at least once on the interval (0,). Therefore, the Eq (1.2) possesses at least one positive equilibrium. Notice that g(x)=(1α)xα+rβx(β1)er(1xβ). Since 0<α1, g(x)>0 for all x(0,), and g(x) is strictly increasing on (0,). Consequently, the Eq (1.2) has exactly one positive equilibrium at x=1.

    Theorem 6. Consider the difference Eq (1.2), where 0<α1, β>0, and r>0. Then the following statements are true:

    (ⅰ) The equilibrium at 0 is unstable.

    (ⅱ) If 0rβ1<α<rβ+1, then the positive equilibrium at 1 is locally asymptotically stable.

    Proof. Part (ⅰ): We compute

    f(x)=(αrβxβ)x(α1)er(1xβ)=αrβxβx(1α)er(1xβ). (2.5)

    For 0<α<1, f(0) is undefined. However, limx0+|f(x)|=>1, indicating that the equilibrium at 0 is unstable.

    When α=1, f(0)=er>1 for r>0, hence the equilibrium at 0 is again unstable.

    Part (ⅱ):

    |f(1)|<11<αrβ<1rβ1<α<rβ+1.

    The following theorem sheds light on the dynamics of populations by examining the difference Eq (1.2). This equation serves as a representation of a population model under specific conditions: 0<α1, β>0, and r>0. Capturing the intricate interplay between growth parameters and environmental factors, it reflects dynamic behaviors pertinent to population dynamics. Through a thorough analysis of its properties, we can effectively assess its suitability for modeling population phenomena.

    Theorem 7. The difference Eq (1.2) represents a population model when 0<α1, β>0, and r>0.

    Proof. According to the definition of a population model presented in [1], the model xt+1=f(xt) represents a population model when it satisfies the following conditions:

    (1) The function f is a continuous function.

    (2) f(0)=0.

    (3) The model has a unique positive equilibrium point f(ˉx)=ˉx.

    (4) f(x)>x for 0<x<ˉx and f(x)<x for x>ˉx.

    (5) If f(x) has a maximum xM in (0,ˉx), then f(x) is monotonically decreasing for x<xM.

    It is evident that the model xn+1=f(xn)=xαner(1xβn) satisfies all the above conditions.

    Theorem 8. The equilibrium point 1 of the difference Eq (1.2), when 0<α1, β>0, and r>0, under the condition 0rβ1<α<rβ+1, is globally stable.

    Proof. The global stability of the equilibrium point is established by invoking Theorem 1 part (a) from Cull's study [1]. Initially, it is noted that the equilibrium point remains globally stable when the difference equation (1.2) does not exhibit a maximum point within the interval (0,1), a condition satisfied when α>rβ.

    Assuming the presence of a maximum point xM=βαrβ within (0,1), Theorem 2 part (a1) in Cull's work [1] is invoked. This theorem implies that establishing global stability necessitates demonstrating that f(x)=xαer(1xβ) maintains consistent concavity within the interval (xM,x). Concavity is investigated by ascertaining the second derivative and establishing its sign within (xM,x).

    The second derivative of f(x) is expressed as:

    f(x)=x(α2)er(1xβ)((β2r2xβ)2rβ(β+2α1)xβ+α2α).

    The roots of f(x) coincide with those of the following equation:

    (xβ)2β2r2rβ(β+2α1)xβ+α2α=0. (2.6)

    These roots are expressed as:

    x=β1rβ(α+β212±4βα+(β1)22). (2.7)

    It is demonstrated that the roots (2.7) lie outside the interval (xM,1) through proof by contradiction.

    Assuming:

    αrβ<β1rβ(α+β212+4βα+(β1)22)<1,α<α+β212+4βα+(β1)22<rβ,0<β212+4βα+(β1)22<rβα<1,β212+4βα+(β1)22<1,4βα+(β1)2<1β.

    This last inequality is invalid for positive parameter values. Similarly, the same approach reveals that the other root cannot reside in the interval (xM,1). Thus, the function maintains its concavity within the specified interval, establishing the theorem.

    In this section, we embark on a comprehensive exploration of potential bifurcations inherent in our discrete difference Eq (1.2). Initially, guided by Theorem (2.1), we observe the occurrence of a fold bifurcation, where a pair of equilibrium points collide and subsequently vanish as the parameter undergoes variation. This collision event reduces the equilibrium count to two, but as the parameter evolves further, a new pair of equilibrium points emerges, restoring the equilibrium count to three. This phenomenon is visually represented in Figure 1, where the diagonal y=x tangentially intersects the function f(x)=xαer(1xβ), precisely occurring when α=rβ+1.

    From Theorem 2, we further analyze the behavior of the equilibrium point x=1 under different parameter regimes. When α<rβ+1, the slope of the tangent at x=1 can reach 1, leading to a period-doubling bifurcation. Figure 2 presents a bifurcation diagram, illustrating how parameter variations r, α, and β induce transitions from stable equilibria to periodic orbits and chaotic dynamics. The diagram clearly shows how the equilibrium point x=1 transitions from stability to a period-2 orbit, followed by higher-order periodic orbits, ultimately culminating in chaos, a characteristic feature of nonlinear dynamical systems.

    Figure 2.  Bifurcation diagram of equilibrium point 1.

    Additionally, a period-doubling bifurcation may also occur at the other positive equilibrium point, ˜x. A similar approach can be applied to analyze the system's behavior near this equilibrium point, analogous to the analysis we will conduct for equilibrium point 1.

    Figure 2 offers a detailed view of the system's bifurcation structure as r increases. Initially, the system remains stable at the equilibrium point, but as r approaches a critical value, a period-doubling bifurcation occurs. This leads to a cascade of period-doubling bifurcations, progressing through period-4, period-8, and eventually chaotic behavior. The bifurcation diagram is a crucial tool that highlights how the system's qualitative dynamics shift as parameters vary, confirming theoretical predictions of period-doubling and chaos.

    Figure 3.  Cobweb diagram for various values of parameter r from 4 to 5.55.

    The cobweb diagrams in Figure 3 provide further insight into the iterative behavior of the system, particularly the transition from stability to chaos as r increases. As shown, the stable equilibrium at r=4 transitions through a series of period-doubling bifurcations as r increases, eventually leading to chaotic behavior beyond r=5.07. These diagrams offer a step-by-step visual demonstration of how the system's dynamics unfold, with particular emphasis on the parameter sensitivity and cascading bifurcations.

    Figure 4.  Time series diagram for various values of parameter r from 4 to 5.55.

    Figure 4 presents time series diagrams for the same parameter values as the bifurcation and cobweb diagrams, providing a temporal perspective on the system's behavior. The time series reinforce the bifurcation and cobweb findings by showing how the state variable xn evolves over time. For smaller r, the system remains stable, but as r increases, periodic and chaotic oscillations emerge, clearly linking the time-domain behavior to the bifurcation phenomena discussed earlier.

    The numerical simulations across Figures 24, combined with the theoretical insights, offer a robust understanding of the system's dynamics. The fold bifurcation observed early in the analysis is followed by a series of period-doubling bifurcations, eventually leading to chaos, a result that aligns well with the predictions from Theorem (2.1). By using these diagrams as analytical tools, rather than mere illustrations, we highlight the critical role of parameter variation in controlling system stability and behavior. This comprehensive exploration of bifurcations not only confirms the theoretical framework but also deepens our understanding of the system's complex dynamics.

    In conclusion, the generalized discrete Ricker map introduced in this paper extends the conventional Ricker map by incorporating additional parameters, enhancing its versatility and applicability across various fields. Through a comprehensive analysis of equilibria, stability conditions, and bifurcation diagrams, we have illuminated the intricate dynamics of the generalized Ricker map, showcasing its behavior across different parameter regimes.

    Our investigation revealed that the map's behavior is highly contingent upon the values of parameters α, β, and r, with distinct stability properties emerging under varying parameter combinations. The existence of multiple equilibria and their respective stability conditions underscore the rich dynamical landscape of the generalized Ricker map, aligning with findings in the literature while also introducing novel dynamics not previously explored in similar models.

    Furthermore, our exploration of basins of attraction elucidated the regions of phase space where trajectories converge to specific equilibria, providing crucial insights into the long-term behavior of the system. This analysis not only enhances understanding of transient and asymptotic dynamics but also resonates with prior studies on basin behavior, establishing a connection between our findings and established theories in dynamical systems.

    By comparing our results with previous studies that examine bifurcation structures and chaotic dynamics in related models, we highlight the unique contributions of the generalized Ricker map. Our results reveal new bifurcation scenarios and parameter interactions, enriching the existing body of knowledge and suggesting implications for modeling complex systems.

    Moving forward, future research endeavors may investigate extensions and refinements of the generalized Ricker map, delving deeper into its chaotic properties, exploring higher-dimensional analogues, and examining its applications in emerging fields such as network dynamics and machine learning. Additionally, comparing our findings with those involving the Allee effect and other related phenomena could further enhance the understanding of its dynamics.

    In summary, the generalized discrete Ricker map stands as a testament to the enduring relevance and utility of iterated maps in modeling complex systems, offering a rich tapestry of dynamical phenomena ripe for exploration and discovery. By situating our findings within the broader context of existing literature, we not only highlight the contributions made by this study but also lay the groundwork for future research.

    M. Y. Hamada: Investigation, Writing-original draft, Validation, Writing-review and editing; H. El-Metwally: Conceptualization, Methodology, Investigation, Supervision; Ibraheem M. Alsulami: Writing-original draft, Writing-review and editing. All authors have read and approved the final version of the manuscript for publication.

    We would like to thank the editor and the anonymous referees for their valuable comments and helpful suggestions, which have led to a great improvement of the initial version.

    The authors declare that they have no conflict of interests regarding the publication of this paper.



    [1] P. Cull, Global stability of population models, Bull. Math. Biol., 43 (1981), 47–58. https://doi.org/10.1016/S0092-8240(81)80005-5 doi: 10.1016/S0092-8240(81)80005-5
    [2] P. Cull, K. Walsh, J. Wherry, Stability and instability in one dimensional pop ulation models, Sci. Math. Japon., 67 (2008), 105–124.
    [3] L. Dai, Nonlinear dynamics of piecewise constant systems and implementation of piece wise constant arguments, Singapore: World Scientific, 2008.
    [4] R. Devaney, An introduction to chaotic dynamical systems, London: CRC press, 2018.
    [5] H. El-Metwally, A. Q. Khan, M. Y. Hamada, Allee effect in a ricker type discrete time predator–prey model with holling type-ii functional response, J. Biol. Syst., 31 (2023), 591–610. https://doi.org/10.1142/S0218339023500201 doi: 10.1142/S0218339023500201
    [6] S. N. Elaydi, An introduction to difference equations, New York: Springer, 2005.
    [7] S. N. Elaydi, Discrete chaos: with applications in science and engineering, New York: Chapman and Hall/CRC, 2007. https://doi.org/10.1201/9781420011043
    [8] M. Y. Hamada, T. El-Azab, H. El-Metwally, Allee effect in a ricker type predator-prey model, J. Math. Comput. Sci., 29 (2022), 239–251. http://dx.doi.org/10.22436/jmcs.029.03.03 doi: 10.22436/jmcs.029.03.03
    [9] M. Y. Hamada, T. El-Azab, H. El-Metwally. Bifurcations and dynamics of a discrete predator-prey model of ricker type, J. Appl. Math. Comput., 69 (2022), 113–135. https://doi.org/10.1007/s12190-022-01737-8 doi: 10.1007/s12190-022-01737-8
    [10] M. Y. Hamada, T. El-Azab, H. El-Metwally, Bifurcation analysis of a two-dimensional discrete-time predator-prey model, Math. Meth. Appl. Sci., 46 (2023), 4815–4833. https://doi.org/10.1002/mma.8807 doi: 10.1002/mma.8807
    [11] R. M. May, Simple mathematical models with very complicated dynamics, Nature, 261 (1976), 459–467. https://doi.org/10.1038/261459a0 doi: 10.1038/261459a0
    [12] J. R. Pounder, T. D. Rogers, Global stability in population models, Math. Model., 3 (1982), 207–214. https://doi.org/10.1016/0270-0255(82)90025-2 doi: 10.1016/0270-0255(82)90025-2
    [13] A. G. Radwan, On some generalized discrete logistic maps, J. Adv. Res., 4 (2013), 163–171. https://doi.org/10.1016/j.jare.2012.05.003 doi: 10.1016/j.jare.2012.05.003
    [14] W. E. Ricker, Stock and recruitment, J. Fisher. Board Can., 11 (1954), 559–623. https://doi.org/10.1139/f54-039 doi: 10.1139/f54-039
    [15] A. Sarkovskii, Coexistence of cycles of a continuous map of a line to itself, Ukr. Mat. Z., 16 (1964), 61–71.
    [16] A. Stewart, S. Dean, J. Boffenmyer, B. Willie, The existence and stability of Equilibria of the generalized Ricker's model, LSU Summer Math Integrated Learning Experience, 2010.
    [17] S. H. Strogatz, Nonlinear dynamics and chaos: with applications to physics, biology, chemistry, and engineering, Lendon: CRC press, 2018.
    [18] S. Wiggins, M. Golubitsky, Introduction to applied nonlinear dynamical systems and chaos, New York: Springer, 1990. http://dx.doi.org/10.1007/978-1-4757-4067-7
    [19] S. Willson, Explorations into global stability of population models, 1995,144–153.
  • 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(769) PDF downloads(52) Cited by(0)

Figures and Tables

Figures(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog