Loading [MathJax]/jax/output/SVG/jax.js
Research article

The uniqueness of limit cycles in a predator-prey system with Ivlev-type group defense

  • Received: 01 September 2024 Revised: 21 October 2024 Accepted: 04 November 2024 Published: 27 November 2024
  • MSC : 34C07, 34C23, 92D25

  • This paper discusses the uniqueness of limit cycles in a two-dimensional autonomous Gause predator-prey model with an Ivlev-type group defense introduced by D. M. Xiao, S. G. Ruan, Codimension two bifurcations in a predator-prey system with group defense, Int. J. Bifurcat. Chaos, 11 (2001). We proved their conjecture that the system can exhibit at most one limit cycle. Furthermore, we compared the qualitative differences between this system and two similar systems with group defense: One system with the same Ivlev-type functional response function but with Leslie-Gower predator dynamics and another system with a comparable functional response function. For both systems, we show that two limit cycles can occur.

    Citation: Jin Liao, André Zegeling, Wentao Huang. The uniqueness of limit cycles in a predator-prey system with Ivlev-type group defense[J]. AIMS Mathematics, 2024, 9(12): 33610-33631. doi: 10.3934/math.20241604

    Related Papers:

    [1] Jawdat Alebraheem . Asymptotic stability of deterministic and stochastic prey-predator models with prey herd immigration. AIMS Mathematics, 2025, 10(3): 4620-4640. doi: 10.3934/math.2025214
    [2] Shanshan Yu, Jiang Liu, Xiaojie Lin . Multiple positive periodic solutions of a Gause-type predator-prey model with Allee effect and functional responses. AIMS Mathematics, 2020, 5(6): 6135-6148. doi: 10.3934/math.2020394
    [3] Binfeng Xie, Na Zhang . Influence of fear effect on a Holling type III prey-predator system with the prey refuge. AIMS Mathematics, 2022, 7(2): 1811-1830. doi: 10.3934/math.2022104
    [4] Yudan Ma, Ming Zhao, Yunfei Du . Impact of the strong Allee effect in a predator-prey model. AIMS Mathematics, 2022, 7(9): 16296-16314. doi: 10.3934/math.2022890
    [5] Qinghui Liu, Xin Zhang . Chaos detection in predator-prey dynamics with delayed interactions and Ivlev-type functional response. AIMS Mathematics, 2024, 9(9): 24555-24575. doi: 10.3934/math.20241196
    [6] Xinjing Wang, Guangwei Du . Strongly-coupled and predator-prey subelliptic system on the Heisenberg group. AIMS Mathematics, 2024, 9(10): 29529-29555. doi: 10.3934/math.20241430
    [7] Aeshah A. Raezah, Jahangir Chowdhury, Fahad Al Basir . Global stability of the interior equilibrium and the stability of Hopf bifurcating limit cycle in a model for crop pest control. AIMS Mathematics, 2024, 9(9): 24229-24246. doi: 10.3934/math.20241179
    [8] 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
    [9] Chuanfu Chai, Yuanfu Shao, Yaping Wang . Analysis of a Holling-type IV stochastic prey-predator system with anti-predatory behavior and Lévy noise. AIMS Mathematics, 2023, 8(9): 21033-21054. doi: 10.3934/math.20231071
    [10] Teekam Singh, Ramu Dubey, Vishnu Narayan Mishra . Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response. AIMS Mathematics, 2020, 5(1): 673-684. doi: 10.3934/math.2020045
  • This paper discusses the uniqueness of limit cycles in a two-dimensional autonomous Gause predator-prey model with an Ivlev-type group defense introduced by D. M. Xiao, S. G. Ruan, Codimension two bifurcations in a predator-prey system with group defense, Int. J. Bifurcat. Chaos, 11 (2001). We proved their conjecture that the system can exhibit at most one limit cycle. Furthermore, we compared the qualitative differences between this system and two similar systems with group defense: One system with the same Ivlev-type functional response function but with Leslie-Gower predator dynamics and another system with a comparable functional response function. For both systems, we show that two limit cycles can occur.



    In this article, we study the limit cycle problem for the following Gause predator-prey system introduced in [26] by Xiao and Ruan:

    dx(t)dt=rx(t)(1x(t)k)x(t)eβx(t)y(t),dy(t)dt=y(t)(D+μx(t)eβx(t)), (1.1)

    where all parameters are positive: r>0, β>0, k>0, D>0, μ>0. For convenience, we will write x and y for x(t) and y(t), respectively. The real variables x and y represent the nonnegative density of the prey and predator population, respectively.

    The main result of this paper is the proof of the conjecture posed in [26] concerning the number of limit cycles in (1.1).

    Theorem 1.1. System (1.1) has at most one limit cycle. If it exists, it is stable and hyperbolic.

    In a recent paper [15], system (1.1) was discussed for the special case of a singular perturbation, essentially studying the occurrence of so-called canard cycles for small D and μ.

    System (1.1) is a special case of the more general Gause predator-prey system [5]:

    dxdt=h(x)p(x)y,dydt=y(D+q(x)). (1.2)

    In system (1.1), the natural prey growth h(x) was chosen to be logistic. The function p(x) is often referred to as the functional response function, describing the effectiveness with which the predator attacks the prey population. Typically, in Gause systems, the numerical response function q(x) is taken to be proportional to the functional response function p(x).

    With the choice of p(x) in system (1.1), the system exhibits group defense by the prey population. This means that if the prey population is sufficiently large, they can form a more effective defense against attacks by predators. The particular form of the functional response function xeβx is similar to the well-known Ivlev functional response in the absence of group defense p(x)=1eβx; see, for example, [11]. The essential difference is that the traditional Ivlev functional response function is monotonically increasing in x, whereas the group defense functional response in (1.1) has a global maximum for some positive x.

    From a biological point of view, one of the first studies of this kind of group defense appeared in the book by Tener [24]. There, a single muskox is easily eaten by wolves, but when the number of muskoxen increases, it becomes less likely to be attacked by wolves. Another fundamental paper in this area is by Holmes and Bethel [8], where it was discussed that parasites can influence host behavior by changing evolutionary strategies to achieve group defense. The relationship with mathematical modeling was made in the fundamental paper by Freedman and Wolkowicz [3]. Recent research on the topic of group defense can be found in [2,9,21].

    From a mathematical modeling point of view, one of the first papers, to our knowledge, mentioning this effect was Wolkowicz's paper [25]. Her model is similar to (1.1) but with a different functional response function, p(x)=mxax2+bx+1, where m, a, and b are positive constants satisfying p(0)=0, p(0)>0, p(0)<0, and p(x)>0 for x>0. For small x, this function approximates the Holling Type II functional response.

    Mathematically, group defense implies that the predators will be less effective in their attack when there are more prey. This leads to some consequences on the behavior of the functional response function p(x). The following definition was introduced by [3].

    Definition 1.2. In Gause systems (1.2), the prey population exhibits group defense if p(x)>0, p(x)>0 (<0) for 0<x<M (x>M).

    Figure 1.1 shows the function p(x) in the case of group defense.

    Figure 1.1.  The functional response function p(x) with group defense as in Definition 1.2. The function is initially increasing as a function of x, i.e., more prey is killed by the predators if more prey is available. However, when x is sufficiently large, i.e., x>M, the prey population becomes strong enough to create a group defense against the predators and the effective killing per predator will go down.

    The main problem in the study of these Gause systems is the determination of the number and stability of limit cycles. This is related to a more general problem for planar autonomous ordinary differential equations, which asks for the number of limit cycles for systems of the following form:

    dxdt=P(x,y,μ),dydt=Q(x,y,μ), (1.3)

    where xR, yR, and tR. The system typically depends on real parameters, indicated by μ.

    In particular, the unsolved second part of the 16th Hilbert problem asks for an upper bound on the number of limit cycles for systems, where P(x,y) and Q(x,y) are polynomial functions of x and y; see [1,29].

    There are few results with the current technology for the study of systems that do not involve small parameters. Specifically, obtaining global results on the qualitative behavior of solutions for all parameters is a notoriously difficult problem. The most successful results have been obtained in the study of the non-existence, existence, and uniqueness of limit cycles, in particular the study of systems where at most one limit cycle exists. There are few studies involving systems with at most two limit cycles. In those few cases, there are restrictive conditions on the functions P(x,y) and Q(x,y), such as satisfying certain symmetry conditions; see the books [29,33]. For the solution of the 16th Hilbert problem, the development of tools to determine upper bounds for the number of limit cycles for systems with at most two limit cycles would be a first natural but difficult step.

    For systems with at most one limit cycle, the strongest methods have been developed for a special case of (1.3), so-called generalized Liénard systems.

    dxdt=F(x)ϕ(y),dydt=g(x). (1.4)

    More on such systems can be found in the books [29,33].

    To apply one of the many theorems about limit cycles available for Liénard systems to (1.3), a nonlinear transformation of the variables needs to be found, changing the system into the form (1.4). This is not always possible, but for the system discussed in this paper, (1.1), it is relatively straightforward. For a more general discussion, see [17].

    Additionally, we will compare system (1.1) with similar group defense mechanisms in other predator-prey models. One instance is when the predator growth dynamics is of Leslie-Gower type, while keeping the same dynamics for the prey. A second instance is to take a similar Gause model, but with a slightly different functional response function p(x), in particular the well-known case of Wolkowicz [25] and Rothe-Shafer [22]. In both systems, we will show that typically situations occur with two limit cycles and that the uniqueness of limit cycles in system (1.1) is a rather rare occurrence in systems with group defense. Although it is difficult to formulate a general statement, we will indicate the mathematical reasons for why such differences occur.

    In Section 2, we will outline the use of Liénard systems and introduce the uniqueness theorem that will be the main tool in the study of (1.1). In Section 3, we will transform system (1.1) into Liénard form and apply the uniqueness theorem to prove Theorem 1.1. In Section 4, we show that a similar system with group defense and Leslie-Gower predator dynamics can have a weak focus of order two and that two limit cycles can be created in a degenerate Hopf bifurcation. Finally, we discuss the mathematical reasons why (1.1) only has at most one limit cycle, while an apparently similar Gause model introduced by [25] and further studied by [22,27] can have two limit cycles.

    The use of Liénard systems (1.4) in the theory of limit cycles has proved to be advantageous. They contain some of the strongest available theorems about the existence or uniqueness of limit cycles.

    In practice, we can try to convert a planar system of nonlinear autonomous ordinary differential equations (1.3) into a Liénard system (1.4) and then analyze it.

    To prove that a certain Liénard system has at most one limit cycle, we will adapt two theorems for our purposes. The first is a simple theorem introduced in [12] (Theorem 1.1) to show that for some cases, no limit cycles can occur:

    Theorem 2.1. ([12]) Suppose system (1.4) with F(x)=xxgf(ˉx)dˉxC3, g(x)C2, f(x)C2, satisfies the following conditions on the interval r1<x<r2:

    (i) dϕ(y)dy>0,

    (ii) xg(r1,r2) is the unique value such that (xxg)g(x)>0, for xxg, and g(xg)=0,

    (iii) xf(r1,r2) is the unique value such that (xxf)f(x)<0, for xxf, and f(xf)=0,

    (iv) cR, such that f(x)cg(x) has no zeroes in r1<x<r2.

    Then, (1.4) does not have limit cycles in the strip r1<x<r2.

    There is a graphical interpretation of condition (iv) of Theorem (2.1). It implies that the function f(x)g(x) has a gap. It means that there is a horizontal line L: y=c such that part of the graph of y=f(x)g(x) lies above L and part of it below L, but without any intersections. See Figure 2.1.

    Figure 2.1.  A typical example of the function f(x)g(x) with a gap according to Condition (iv) Theorem 2.1. The function f(x)cg(x) has no zeroes for the value of c corresponding to the gap in the figure, separating two branches of the graph of the function.

    Condition (i) of Theorem 2.1 is a condition typically satisfied in applications. In Gause predator-prey systems, ϕ(y)=ey and the condition is obviously satisfied.

    Condition (ii) of Theorem 2.1 means that only limit cycles surrounding one singularity are considered. In Gause systems, this is typically true, but for exceptions (for example, in cases where an Allee effect is introduced in the natural growth of the prey), see [6,16].

    Condition (iii) of Theorem 2.1 implies that the divergence of system (1.4) changes sign for exactly one value of x. This condition can be relaxed, but for this paper it is more convenient to use this form. The differentiability conditions on the three functions F(x), g(x), f(x) and the sign of the condition (xxf)f(x)<0 can be relaxed as well, but were chosen to be consistent with the system studied in this paper.

    The non-existence Theorem (2.1), in practice, can be combined with a uniqueness theorem for limit cycles. For our purposes, the following uniqueness theorem will be sufficient. It is a special case of a more general theorem presented in [34] (Theorem 3, page 485).

    Theorem 2.2. ([34]) Suppose system (1.4) with F(x)=xxgf(ˉx)dˉxC3, g(x)C2, f(x)C2, satisfies the following conditions on the interval r1<x<r2:

    (i) dϕ(y)dy>0,

    (ii) xg(r1,r2) is the unique value such that (xxg)g(x)>0, for xxg, and g(xg)=0,

    (iii) xf(r1,r2) is the unique value such that (xxf)f(x)<0, for xxf, and f(xf)=0, and xg<xf,

    (iv) df(x)g(x)dx<0 in r1<x<xg and xf<x<r2.

    Then, (1.4) has at most one limit cycle in the strip r1<x<r2, which is stable and hyperbolic if it exists.

    Theorems 2.1 and 2.2 are in fact complimentary and can be combined in one theorem, which is more natural for practical applications. Conditions (i)–(iii) of Theorem 2.2 are the same as the conditions in Theorem 2.1. The difference lies in Condition (iv). As mentioned above, in Theorem 2.1, there is a geometrical gap interpretation of the condition. A similar approach is possible for Condition (iv) in Theorem 2.2.

    Lemma 2.3. A sufficient condition for Condition (iv) of Theorem (2.2) to hold true is: cR, f(x)cg(x) has no multiple zeroes on the interval r1<x<xg and xf<x<r2.

    Proof. Suppose the function (f(ˉx)g(ˉx)) has a zero ˉx, then we can write:

    f(ˉx)g(ˉx)=f(ˉx)g(ˉx).

    Denote the function value of f(ˉx)g(ˉx) at x=ˉx by c, i.e.,

    f(ˉx)cg(ˉx)=0,f(ˉx)g(ˉx)=f(ˉx)g(ˉx).

    These equations imply that

    f(ˉx)cg(ˉx)=0.

    That means (f(ˉx)g(ˉx)) has a multiple zero, which is not possible. Therefore, (f(ˉx)g(ˉx)) does not have a zero and, from Conditions (ii) and, (iii) of Theorem 2.2, it follows that it has to be negative.

    This simple lemma has a geometrical interpretation, as shown in Figure 2.2: If a differentiable function has a local minimum or maximum at x=ˉx, then necessarily the horizontal line y=f(ˉx)g(ˉx) corresponds to a value c where f(x)cg(x) has a multiple zero, because the line is tangent to the graph of the function.

    Figure 2.2.  An example of a non-monotonic function f(x)g(x) as described in Lemma 2.3. At the local minimum, the tangent line y=c corresponds to a multiple zero of f(x)cg(x). It implies that a sufficient condition for the function f(x)g(x) to be monotonic is that f(x)cg(x) does not have a multiple zero for any cR.

    Conditions (iv) of Theorems 2.1 and 2.2 are complimentary, because of the geometrical interpretation in Lemma 2.3. We combine both theorems into one convenient theorem. This leads to a similar theorem as Lemma 1.4 in [12]. The only difference with that lemma is the interval for x in Condition (iv). In Lemma 1.4 [12], the condition needs to hold true for the whole interval r1<x<r2, while in Theorem (2.2), the interval is restricted. This subtle difference will simplify the proof for the uniqueness of limit cycles in (1.1).

    Theorem 2.4. Suppose system (1.4) with F(x)=xxgf(ˉx)dˉxC3, g(x)C2, f(x)C2 satisfies the following conditions on the interval r1<x<r2, yR:

    (i) dϕ(y)dy>0,

    (ii) xg(r1,r2) is the unique value such that (xxg)g(x)>0, for xxg, and g(xg)=0,

    (iii) xf(r1,r2) is the unique value such that (xxf)f(x)<0, for xxf, and f(xf)=0, and xg<xf.

    (iv) Either (iv): cR, such that f(x)cg(x) has no zeroes in r1<x<r2,

    or (iv): cR, f(x)cg(x) has no multiple zeroes on the interval r1<x<xg and xf<x<r2.

    Then, (1.4) has at most one limit cycle in the strip r1<x<r2, and if it exists, it is hyperbolic. In case (iv), no limit cycles occur.

    Proof. The theorem is a combination of Theorems 2.1 and 2.2, which have the same Conditions (i)–(iii). Condition (iv) is Condition (iv) from Theorem 2.1 and Condition (iv) is Condition (iv) from Theorem 2.2 rewritten in terms of Lemma 2.3.

    Conditions (i)–(iii) can easily be verified in our case. The critical conditions are (iv) and (iv). These are, in principle, two independent conditions with different implications, but in typical families (the system in this paper is an example of this), their combination is sufficient to cover all parameter values in a system.

    Compared to the general system (1.3), Liénard systems have another advantage. It is straightforward to determine the position of the singularities, because they satisfy an equation with one variable only.

    Lemma 2.5. In system (1.4) with dϕ(y)dy>0, a necessary and sufficient condition for the existence of a singularity at (xg,yg) is g(xg)=0, and ϕ(1)[F(xg)] exists. If g(xg)>0 (<0), then the singularity is an anti-saddle (saddle).

    Proof. Singularities of the system satisfy F(xg)ϕ(yg)=0, g(xg)=0. From this, the necessary and sufficient conditions for the coordinates of a singularity follow.

    The Jacobian matrix at a singularity, where g(xg)=0, F(xg)ϕ(yg)=0, is given by [F(xg)dϕ(yg)dyg(xg)0]. The product of the eigenvalues is dϕ(yg)dyg(xg), from which the result of the lemma follows.

    The strategy for the analysis of system (1.1) is to convert it into (1.4) first, then to study the singularities using Lemma 2.5, and finally, to use Theorem 2.4 to establish the uniqueness of limit cycles.

    To apply the results of the previous section to system (1.1), the system first needs to be converted into Liénard form. This is a well-known procedure for Gause predator-prey systems [4]. In general, different transformations into Liénard form exist, a concept not widely understood yet. The choice of transformation is not unique, and different choices will lead to different results. For examples of how to convert predator-prey systems to a Liénard system, see [4,17,28,30]. In this paper, we choose the simplest transformation with the justification that it is sufficient for getting proof of the uniqueness of limit cycles.

    Lemma 3.1. System (1.1) can be transformed into a Liénard system (1.4) with F(x)=reβx(1xk), g(x)=Deβxx+μ, ϕ(y)=ey.

    Proof. Let ttp(x), with p(x)=xeβx, and y=ev. After the transformation, we label for convenience v again by y and get:

    dxdt=reβx(1xk)ey=F(x)ey,dydt=Deβxx+μ=g(x). (3.1)

    For future reference, we define the divergence of the vector field by:

    f(x)=F(x)=reβx[β(1xk)1k]. (3.2)

    To find the position of the singularities in the first quadrant of the phase plane (since x(t)0 and y(t)0), we apply Lemma 2.5 to the Liénard system (3.1) in Lemma 3.1.

    Lemma 3.2. In system (1.1) define μ=Dβe and ˜μDeβkk, with 0<μ<˜μ.

    If μ<μ, then the equivalent Liénard system (3.1) has an anti-saddle with x-coordinate xAS and if μ<μ<˜μ, it has a saddle at x=xS. For the x-coordinates of the singularities, the inequality 0<xAS<1β<xS<k holds, if the saddle exists.

    If μ<μ, then system (3.1) has no singularities.

    Proof. The necessary and sufficient conditions for the existence of a singularity in Liénard system (3.1), according to Lemma 2.5, translate into the following conditions for the system in Lemma 3.1.

    g(xg)=Deβxgxg+μ=0, 0<xg<k. (3.3)

    The number of solutions to the equation g(xg)=0 follows from looking at the derivative g(x)=Deβx(1βx)x2. A double zero can only occur when xg=1β. Substituting this into g(xg), we find the condition for a double zero to be μ=μ=Dβe. If μ>μ, then g(xg)=0 has two solutions xg>0, and for μ<μ, there are no solutions. The Jacobian matrix of system (3.1) is:

    J=[f(x)eyg(x)0]=[reβx[(1xk)1k]eyDeβx(1βx)x20].

    Under the condition μ>μ, we find that for the two zeroes xAS and xS of g(x), with xAS<1β<xS the determinant is positive and negative, respectively, i.e., xAS is an anti-saddle and xS is a saddle. Moreover, 0<xAS<1β<xS<k.

    The last inequality in (3.3) is the condition that yg=ln(F(xg)=ln(reβxg(1xgk)) exists. Since xg satisfies g(xg)=0, we get that 0<xg<k implies that μ<˜μDeβkk. For the two critical values of μ, the inequality 0<μ<˜μ holds true, because kβ>1.

    The condition μ<˜μ reflects the fact that the saddle in the original system (1.1) is located in the first quadrant. For μ>˜μ, the saddle moves into the region y<0, which is irrelevant from a biological point of view. Since the Liénard system was obtained through a transformation y=ev, only singularities in the first quadrant of the original system will appear in the Liénard form.

    Since limit cycles cannot appear in a system without singularities, we will impose in the following the condition μ>μ of Lemma 3.2. It means that implicitly we will assume that μ>Dβe. The conclusions of Lemma 3.1 are in line with the results of [26]. In principle, there are two cases to be considered according to the lemma: μ<˜μ (saddle exists) and μ>˜μ (saddle does not exist). However, for the proof of the uniqueness of limit cycles, this distinction will not be relevant. In both cases, the condition μ>μ ensures that the equation g(xg)=0 has two solutions denoted by xAS and xS. We will prove the uniqueness of limit cycle in the extended interval 0<x<xS even if that zero does not represent a saddle.

    The stability of the anti-saddle is determined by the sign of f(x) for x=xAS. If it is negative (positive), then the singularity is stable (unstable). In the case when f(xAS)=0, the sign of the first focal value will determine the stability. For a Liénard system (1.4), there exists a simple method to determine it. The unique zero of f(x) is given by xf=k1β<k.

    Lemma 3.3. Under the condition xf=xAS, system (1.1) has a weak focus, which is stable and of order one. In terms of the parameters of the original system, this occurs when μ=μwkDβekβ1kβ1.

    Proof. According to system (3.1) and Eq (3.2), we have f(x)=reβx[β(1xk)1k], g(x)=μxDeβxx. Then f(xf)=0, with xf=k1β. Therefore, a weak focus occurs for g(xf)=0, which is equivalent to μ=μwkDβekβ1kβ1.

    Under this condition, we have xf=xAS and the derivatives of the functions are given by f(x)=rβeβx[β(1xk)2k], f(x)=rβeβx[β(1xk)3k], which evaluated at xf leads to f(xf)=rβeβxf[β(1xfk)2k]=rβeβxf(1k)<0, f(xf)=rβeβxf[β(1xfk)3k]=rβeβxf(2k)<0.

    Similarly, we get g(xf)=eβxf(1βxf)x2f, g(xf)=eβxfx3f(β2x2f+2βxf2).

    From Lemma 3.2, we know that 0<xAS=xf<1β, so g(xf)=g(xAS)=eβxAS(1βxAS)x2AS>0, and g(xf)=g(xAS)=eβxASx3AS(β2x2AS+2βxAS2)<g(1β)=eβ3(1)<0.

    Under the condition xf=xAS, we get according to [31] the following expression for the first focal value for a generalized Liénard system.

    V1f(xf)g(xf)f(xf)g(xf)<0. (3.4)

    This implies that Liénard system (3.1) has a weak focus, which is stable and of order one.

    System (1.1) is equivalent to the Liénard system (3.1). Therefore, if the original system (1.1) has a weak focus, it is stable and has order one.

    To apply Theorem 2.4, we need to verify that all conditions are satisfied on a relevant interval for x, i.e., the strip in the phase plane where limit cycles are residing. First, we determine the interval r1<x<r2. We can determine the interval by using the properties of Liénard systems.

    Lemma 3.4. Limit cycles in (1.1) are restricted to the strip 0<x<xS in the phase plane. Here, xS is one of the two solutions xAS<xS of g(x)=0. If μ<μ<˜μ, according to Lemma 3.2, xS represents the x-coordinate of the saddle.

    Proof. Limit cycles surround the anti-saddle with x-coordinate xAS with 0<xAS<xS and cannot cross the vertical line x=xS in the phase plane. This is true, because if the saddle at x=xS exists, then on the line dxdt=F(xS)ϕ(yS) changes sign exactly at the saddle. Any limit cycle crossing the line would have to contain the saddle, which is impossible, because the system only has two singularities, one anti-saddle and one saddle.

    If there is no saddle at x=xS, then it implies that k<xS. In the original system (1.1), the line x=k is a line without contact and cannot be crossed by limit cycles. Therefore, limit cycles cannot cross any other vertical line x=c>k.

    It means that for the application of Theorem 2.4, we will effectively use r1=0, r2=xS, even if xS does not correspond to a saddle.

    Lemma 3.5. Under the Condition μ>Dβe, Liénard system (3.1) satisfies the first three conditions (i)–(iii) of Theorem 2.4 on the interval r1<x<r2, with r1=0, r2=xS.

    Proof. Since ϕ(y)=ey, g(x)=μxDeβxx, and F(x)=reβx(1xk) in system (3.1), and r1=0, r2=xS according to Lemma 3.4, we get dϕ(y)dy>0 and f(x)=F(x)=reβx[β(1xk)1k]. Moreover, the zeroes of g(x) satisfy 0<xAS<1β<xS<k according to Lemma 3.2 and the unique zero of f(x) is given by xf=k1β(0,xS). If xf>xS, then the divergence of the system has fixed sign on 0<x<xS and no limit cycles can exist. Therefore, we can safely exclude this case in the following.

    If we denote xg=xAS, then xg(0,xS) is the unique value such that g(xg)=0. We know g(1β)=0 according to Lemma 3.2. As a consequence, (xxAS)g(x)>0 where xxAS, i.e., (xxg)g(x)>0 where xxg.

    To apply Theorem 2.4, we need to investigate the zeroes of the function f(x)cg(x) and check whether one of the two Conditions (iv) or (iv) holds true. To simplify the analysis, we rewrite the expression in the following way:

    f(x)cg(x)=s(x)[ˉf(x)ˉcˉg(x)], (3.5)

    where

    s(x)=reβxβxk,
    ˉf(x)=βx(βxβxf),
    ˉc=cDβkr,
    ˉg(x)=μβDβxeβx1.

    Evidently, s(x)<0 for 0<x<xS, and to check Condition (iv) we can restrict ourselves to analyzing the zeroes of ˉf(x)ˉcˉg(x). The change of variables t=βx does not change the nature of the zeroes and simplifies the expression further. We get:

    ˉz=ˉfLˉg=t(ttf)CLtetL, (3.6)

    where ˉf=t(ttf), ˉg=CtetL, C=μDβ, L>0. We only need to consider the case L>0, because ˉf(x)ˉcˉg(x) can only have a zero on the intervals r1<t<tAS,tf<t<tS, where we used the notation tAS=βxAS and tS=βxS. In short, we will study in the following the zeroes of the function ˉz=ˉfLˉg for L>0.

    Next, we will check Condition (iv) of Theorem 2.4. As we will see, the two subcases (iv) and (iv) correspond to the relative positions of tAS and tf.

    In this case, we expect no limit cycles and construct a constant c such that Condition (iv) in Theorem 2.4 is satisfied.

    Proposition 3.6. Under the condition μ>Dβe, there exists a constant c such f(x)cg(x) in Liénard system (3.1) has fixed sign on the interval 0<x<xS.

    Proof. Following the discussion of the previous section, we will consider the equivalent expressions for the variable t=βx. Under the condition μ>Dβe, we know tAS>tf where tAS=βxAS, tf=βxf. We want to choose L in such a way that the tangent lines of ˉf(t) at tf and Lˉg(t) at tAS have the same slope, i.e., ˉf(tf)=Lˉg(tAS). Denote the tangent lines of ˉf and Lˉg by lf and lAS, respectively. We get the situation of Figure 3.1.

    Figure 3.1.  For the case xAS>xf: The images of ˉf(t) and Lˉg(t) with the tangent lines at the zeroes tf and tAS, respectively, as described in the proof of Proposition 3.6. By choosing the scaling parameter L in such a way that the tangent lines lf and lAS are parallel, the graphs of y=ˉf(t) and y=Lˉg(t) will not intersect, establishing that ˉf(t)Lˉg(t) does not have a zero. This shows that Condition (iv) is satisfied in Theorem 2.4.

    From Figure 3.1, we see that the graph of ˉf(t) lies above lf, and the graph of Lˉg(t) lies below lAS. This would imply that ˉf(t)=Lˉg(t) has no solution for this value of L.

    To prove analytically that the function f(t)Lg(t) has fixed sign for this choice of L, we consider the expression for lf and lAS:

    lf:yf=ˉf(tf)(ttf),
    lAS:yAS=Lˉg(tAS)(ttAS),

    where L=ˉf(tf)ˉg(tAS).

    Next, let H1(t)=ˉf(t)yf=ˉf(t)ˉf(tf)(ttf). Then, the condition that the graph of ˉf(t) lies above lf is equivalent to H1(t)0. Taking derivatives we get: H1(t)=ˉf(t)>0, H1(tf)=0, and H1(tf)=0. From this, it follows that H1(t)=ttft2tfH1(t1)dt1dt20.

    Similarly, let H2(t)=yASLˉg(t)=L(ˉg(tAS)(ttAS)ˉg(t)). The condition that the graph of Lˉg(t) lies below lAS is equivalent to H2(t)0. For this function, we have H2(tAS)=H2(tAS)=0 and H2(t)>0 for 0<t<2. Taking derivatives leads to H2(t)0 in t(0,2), and H2(2)>0. It follows that H2(t)=ttASt2tASH2(t1)dt1dt20 for 0<t<2.

    For t2, H2(t)=L[ˉg(tAS)ˉg(t)]>0, and H2(t) increases monotonically for t2. Therefore, H2(t)H2(t)min=H2(2)>0. This implies that H2(t)0 for t>0 as we intended to prove. It follows that the graph of Lˉg(t) lies below lAS.

    We have proved that ˉf(t)ˉf(tf)(ttf)>Lˉg(tAS)(ttAS)Lˉg(t), which leads to the result we are looking for: There exists an L=L such that f(x)Lg(x)0.

    In this case, we expect at most one limit cycle and will show that Condition (iv) in Theorem 2.4 is satisfied.

    Proposition 3.7. Under the condition μ>Dβe, f(x)cg(x), cR, in Liénard system (3.1) does not have a multiple zero on the intervals 0<x<xAS and xf<x<xS.

    Proof. Again, we will consider the equivalent expressions for the variable t=βx. First, we prove t(0,tAS) such that ˉz(t)=ˉz(t)=0.

    According to (3.6), we know ˉf(t)=2, ˉg(t)(t2)<0 for t(0,2) and tAS<2, so ˉz(t)=ˉf(t)Lˉg(t)>0, for t(0,2).

    Suppose t, such that ˉz(t)=ˉz(t)=0. Since ˉz(t)>0 for t(0,2), we know that ˉz(t)(tt)0, i.e., ˉz(t)<0(>0), for 0<t<t(t<t<tAS<2). It follows that

    ˉz(0)=0tˉz(s)ds=t0ˉz(s)ds, (3.7)
    ˉz(tAS)=tAStˉz(s)ds>0. (3.8)

    However, ˉz(tAS)=ˉf(tAS)Lˉg(tAS)=ˉf(tAS)<0. This contradicts the assumption that ˉz(t)=ˉz(t)=0. Therefore, t(0,tAS) such that ˉz(t)=ˉz(t)=0. The above calculation shows that the result is true for the extended interval (0,2), i.e., t(0,2) such that ˉz(t)=ˉz(t)=0.

    Finally, we prove that under the condition μ>Dβe, ˉz=ˉfLˉg does not have a multiple zero on the interval 2<t<tS.

    According to (3.6), ˉz(t)=ˉf(t)Lˉg(t) and ˉf(t)>0, ˉg(t)<0 in (2,tS). It follows that ˉz(t)>0 in (2,tS) and there cannot be a t such that ˉz(t)=ˉz(t)=0, for t(2,tS). Therefore, y=ˉf(t) and y=Lˉg(t) have exactly one transversal intersection in (2,tS).

    Combining the results, we find that under the condition μ>Dβe, f(x)cg(x) in Liénard system (3.1) does not have a multiple zero on the intervals 0<x<xAS and xf<x<xS. The images of ˉf and Lˉg are shown in Figure 3.2.

    Figure 3.2.  For the case xAS<xf: the figure depicts an example of y=ˉf(t) and y=Lˉg(t) with only transversal intersections, implying that no multiple zeroes can occur for the function ˉf(t)Lˉg(t). This shows that Condition (iv) is satisfied in Theorem 2.4.

    With the results of the previous sections, we have sufficient information to prove the main result of this paper.

    Theorem 3.8. System (1.1) has at most one limit cycle. If it exists, it is stable and hyperbolic.

    Proof. According to Lemma 3.5, Propositions 3.6 and 3.7, we know that Liénard system (3.1) satisfies all conditions of Theorem (2.4). Therefore, Liénard system (3.1) has at most one limit cycle.

    In particular, Liénard system (3.1) has no limit cycles (at most one limit cycle) when μμWF (μ>μWF) if the Condition (iv) (Condition (iv)) in (2.4) is satisfied according to Proposition (3.6) (Proposition (3.7)). Here, μWF is defined in Lemma 3.3.

    System (1.1) is equivalent to the Liénard system (3.1) and the uniqueness of limit cycle in the Liénard system (3.1) implies the uniqueness of the limit cycle in the original system (1.1), i.e., system (1.1) has at most one limit cycle.

    In the case when μ>μWF, the divergence of the vector field f(x) at x=xg is positive in the case where a limit cycle can exist (xAS<xf). For system (1.1), this means that the limit cycle, if it exists, must be stable, because the singularity inside the limit cycle is unstable.

    Using Lemmas 3.2 and 3.3, we establish the existence of a limit cycle by a suitable perturbation of the weak focus leading to a small-amplitude limit cycle created in a Hopf bifurcation. A further change of the parameters will lead to the disappearance of the stable limit cycle in a stable saddle loop. Figure 3.3 shows a numerical scenario where the parameter k is changed while the other parameters are fixed.

    Figure 3.3.  A numerical scenario of a limit cycle born in a Hopf-bifurcation and disappearing in a saddle loop under the variation of the parameter k. The other parameters are fixed and are equal to r=0.8;β=1;D=1.7;xg=0.5;μ=Deβxgxg.
    (a) k=1.35: A stable anti-saddle without limit cycles;
    (b) k=1.51: An unstable anti-saddle with a small-amplitude stable limit cycle created in a Hopf-bifurcation;
    (c) k=1.6: An unstable anti-saddle with a growing stable limit cycle;
    (d) k=1.93: An unstable anti-saddle with a stable limit cycle close to disappearing in a saddle loop;
    (e) k=2.0: An unstable anti-saddle without limit cycles.

    Even though it is difficult to show a clear pattern in the asymptotic behaviour of the solutions for all values of the parameters, the numerical example of Section 3.6 shows what can typically be expected.

    In the case when xAS>xf, system (1.1) does not have limit cycles, as indicated in the proof of Theorem 3.8. A typical situation is depicted in Figure 3.3(a). The anti-saddle is stable and locally attracts solutions.

    It means that in the positive quadrant of the phase plane, for some initial values, the solution [x(t),y(t)] will tend to a positive equilibrium, while for some initial values it will tend to a stable node situated at x=k,y=0, i.e., leading to extinction of the predators. The regions in the phase plane leading to extinction of the predators or a positive equilibrium are determined by the relative position of the separatrix of the saddle lying in the first quadrant. The exact positions can only be found numerically.

    In the case when xAS<xf, system (1.1) has an unstable positive singularity. For increasing k, see Figure 3.3(b)(d), a growing stable limit cycle surrounds the equilibrium. In those situations, depending on the position in the phase plane of the separatrices of the saddle and the position of the limit cycle, the solution will either tend to the stable limit cycle or to the stable node situated at x=k,y=0, i.e., the predator and prey populations will approach a periodic limit set or the predator population will die out.

    When k is large enough, the limit cycle has disappeared into a saddle loop and the ensuing system has a positive unstable anti-saddle, no limit cycles, a saddle, and a stable node situated at x=k,y=0; see Figure 3.3(e). In that case, regardless of the initial position, the predator population will die out. The biological interpretation is that if the carrying capacity is large, the prey population can grow big enough to contain the predation effect through group defense, and the predator population will die.

    A natural question is how typical the behavior of system (1.1) is, compared to other systems with group defense.

    The first system with group defense that was studied systematically in the literature [22,25,27] contained cases with two limit cycles. A first question therefore would be: Why does (1.1) only have one limit cycle? In [13], the group defense property of the functional response function p(x) in a Gause system was discussed. It was observed that actually a weak focus of second order can only be created for those values of x where p(x) is increasing, implying that the group defense property is not the essential reason to why more than one limit cycle can occur.

    To study this difference in the number of limit cycles between the various predator-prey models with group defense, we look at two systems that have a similar structure to (1.1).

    (modified Leslie-Gower with group defense)

    dxdt=rx(1xk)xeβxy,dydt=y(Dμyλ+x). (4.1)

    (Wolkowicz-Rothe-Shafer with group defense)

    dxdt=rx(1xk)xx2+ax+by,dydt=y(D+μxx2+ax+b). (4.2)

    Both systems bear a similarity with (1.1) in the following sense:

    System (4.1) has the same dynamics for the prey, i.e., the same logistic growth and same functional response function, but the predator dynamics contains a version of the carrying capacity term introduced by Leslie-Gower [14]. It has not occurred in the literature as far as we know. Its purpose here is purely mathematical: If the predator dynamics in system (1.1) are changed into a Leslie-Gower form while keeping the same dynamics for the prey population including group defense of Ivlev-type, what will be the consequences, especially in terms of the number of limit cycles?

    System (4.2) is the system studied by Wolkowicz-Rothe-Shafer [22,25], while [27] showed that a second-order weak focus can occur in the system. The only difference with (1.1) is the denominator in the functional response function p(x)=xM(x): MXR(x)=eβx in [26] versus MWRS(x)=x2+ax+b in [22,25]. These two functions have similar behavior for x>0: M(x)>0 (the parameters in (4.2) satisfy conditions to ensure this; see [22]), and second- and higher-order derivatives cannot be negative. One difference is that the first derivative of MXR(x) is positive, while the first derivative of MWRS(x) can be negative (if a<0). We note that for system (4.2), the conjecture is that for all parameters, at most two nested limit cycles can occur [22].

    Proposition 4.1. System (4.1) can have two small-amplitude limit cycles created in an Andronov-Hopf bifurcation.

    Proof. First, we transform system (4.1) into a Liénard system. This is more complicated than for a standard Gause system because of the y2-term in dydt. Following [30], we apply the transformation y=ω(x)u, with ω(x)=exx0μ(λ+ˉx)ˉxeβˉxdˉx, for some constant x0>0, and a rescaling of time ttxeβxω(x), which leads to the Liénard system

    dxdt=F(x)ey,dydt=g(x), (4.3)

    with

    f(x)=F(x)=1x2e2βxω(x)˜f(x),
    g(x)=1x2e2βxω(x)˜g(x),

    with

    ˜f(x)=rx[x(λ+x)(βkβx1)eβx+μ(xk)]k(λ+x),˜g(x)=x[δkeβx(λ+x)+rμ(xk)]k(λ+x). (4.4)

    The expressions are quite involved and their full analysis is beyond the scope of this article. To prove the existence of a second-order weak focus, we first impose the condition of a weak focus, i.e., the parameter values for which f(x) and g(x) have a zero in common. To construct such a situation, we introduce a new parameter xf corresponding to a zero of f(x). Solving f(xf)=0 for the parameter r, we get:

    r=kδxf(kββxf1).

    In the same spirit, we can introduce a new parameter xg corresponding to a zero of g(x) by removing the parameter μ:

    μ=eβxgxf(kβλ+kβxgβλxfβxfxgλxg)kxg.

    Next, we create a weak focus by setting xf=xg and λ=0 to simplify the expressions. The quotient f(x)g(x) becomes:

    f(x)g(x)=(xxg)xeβx(kxg)(kββxβxg1)xg(kββxg1)[xeβx(kxg)xgeβxg(kx)]. (4.5)

    This function is useful, because it determines the focal values of the weak focus at x=xg=xf; see the proof of Lemma 3.3, Eq (3.4). We get, according to (3.4), the following expression determining the first focal value:

    V1β3k2x2g3β3kx3g+2β3x4g2β2xgk2+5β2x2gkβ2x3g+2βk24βxgk2k. (4.6)

    To create an example of a second-order weak focus, we take β=14, xg=1, k=175+1513750. With this choice, the first focal value V1 vanishes. Using Maple, it can be shown that the second focal value is negative and therefore the original system has a weak focus of order two. Since the system can have a weak focus of order one, which is either stable or unstable (V1 can be made positive or negative by choosing appropriate values of the parameters in the system), we can create a stable small-amplitude limit cycle surrounding an unstable weak focus of order one, by changing the sign of V1. Then, changing the parameters in such a way that the focus becomes a stable strong focus, a second, unstable limit cycle is created inside the already existing stable limit cycle. We note that this particular system has a weak focus of at most second order and the conjecture is that for all parameters at most two nested limit cycles can occur.

    In Figure 4.1, a numerical example is shown of system (4.1) with two nested limit cycles.

    Figure 4.1.  A numerical example of system (4.1) with two nested limit cycles. The system has the same prey-dynamics as system (1.1) studied in this paper. The predator dynamics of (4.1) are different and of Leslie-Gower type. The parameters were chosen as k=9.711;D=1;β=0.25;λ=0;μ=0.12025;r=7.2197. The two nested limit cycles LC1LC2 establish a two-layer limit for different initial conditions of the prey and predator population: for initial conditions outside LC1, solutions tend to the periodic solution LC2, while solutions starting inside LC1 will approach the stable equilibrium. This example shows that a Gause predator-prey system (1.2) with a group defense mechanism similar to (1.1) can have two limit cycles, in contrast to system (1.1) which can have only one limit cycle.

    To compare the two systems (1.1) and (4.2), we transform to a Liénard system and compare the functions F(x) and g(x).

    In Lemma 3.1, we obtained the Liénard system (3.1) of system (1.1). Next, we transform system (4.2) to a Liénard system in a similar way.

    Let t=t(x2+ax+b)x, y=ey. Then, system (4.1) becomes a Liénard system (1.4).

    dxdt=r(x2+ax+b)(1xk)ey,dydt=D(x2+ax+b)+μxx. (4.7)

    In Liénard system (3.1), we obtained F(x)=reβx(1xk). Then, f(x)=F(x)=reβx[β(1xk)1k]. Let f(xf)=0, which leads to xf=k1β. It follows that δ1>0, such that F(x) has one local maximum F(xf) in (xfδ1,xf+δ1).

    Similarly, in Liénard system (4.7), we have F(x)=r(x2+ax+b)(1xk). Then, f(x)=F(x)=rk[3x2+2(ka)x+akb]. Setting f(x)=0, we get two positive solutions x1, x2. It means that in the case of the Wolkowicz-Rothe-Shafer system, for the function F(x), the situation of simultaneously one local minimum and one local maximum can occur.

    This paper showed that the conjecture in [26] regarding the uniqueness of limit cycles in a Gause predator-prey model with a specific functional response function for group defense is true. The method to achieve this was an adaptation of a standard uniqueness theorem in Liénard systems.

    By comparing the system to similar systems, it was shown that it is difficult in general to relate the number of limit cycles to the specific form of the group defense mechanism. In relation to these observations, we make some suggestions for future research.

    The functional response functions to model group defense in systems (1.1) and (4.2) had very similar properties. Since their form does not seem to have a specific justification from a biological point of view, the choice of the function in a practical situation would be based on parameter fitting to field data. However, it has been a notoriously difficult problem to fit functional response functions to field data. For example, the four Holling functional response functions, originally introduced in [7], were a poor fit to field data, but nevertheless have become a standard in predator-prey modeling. Similarly, in more recent biological studies [10,35], it was shown that obtaining good fits with popular functional response functions is not easy.

    The functions p(x)=xeβx and p(x)=xx2+ax+b used in (1.1) and (4.2), respectively, have similar mathematical properties and are virtually indistinguishable when used for data fitting. This leads to the following issues, that need to be resolved in future research.

    ● (Mathematical) What causes the difference between systems with almost similar functional response functions, while modelling group defense? The functional response functions p(x)=xeβx and p(x)=xx2+ax+b have similar properties, but in the Liénard form, the functions F(x) have one local maximum and one local maximum + one local minimum, respectively, as shown in Section 4.2. It is a worthwhile future research topic to investigate what the real mathematical factor is determining the maximum number of limit cycles.

    ● (Biological) It is difficult to obtain a good fit with field data for the functional response function. This could either be circumvented by finding a biological interpretation of the parameters in a suitable functional response function or a detailed statistical analysis of field data. This seems to be a general problem that needs to be resolved in many predator-prey models introduced in the literature.

    ● (Mathematical) It is not clear if some systems can have more than two limit cycles. Several recent papers contain examples with more than two limit cycles [20,23,31]. Many other systems seem to have at most two limit cycles, see [18,19] amongst many others. In general, no methods exist to get global upper bounds for the number of limit cycles if the number is two or more. Any progress in this direction for predator-prey systems would be of importance for the still open second part of the 16th Hilbert problem [1,29,33], which asks for upper bounds for the number of limit cycles in polynomial systems.

    ● (Mathematical) The parameters in a predator-prey system typically do not correspond to a rotated vector field (see Chapter IV, section 3 in [33] or Chapter 3 in [29]). This means it is difficult to trace limit cycles and saddle loops as a function of the parameter. From a biological point of view, it makes it difficult to indicate in the positive quadrant x0,y0 which solutions stay positive and which solutions lead to extinction of either the predator or prey population. An understanding of the role of the parameters in the Gause system (1.2) or specifically in (1.1) is needed. For example, in Figure 3.3, a clear numerical pattern of the phase portraits as a function of the parameter k, i.e., the carrying capacity of the growth rate of the prey population in the absence of predators is observable. However, a rigorous mathematical proof of this pattern is still needed.

    Jin Liao: Formal analysis, writing-original draft; André Zegeling: Validation, writing-review and editing; Wentao Huang: Validation, review. All authors have read and approved the final version of the manuscript for publication.

    This work is partially supported by Innovation Project of Guangxi Graduate Education(YCSW2024152) and by the National Natural Science Foundation of China (No.12461099).

    The authors declare that there is no conflict of interest.



    [1] J. C. Artés, J. Llibre, D. Schlomiuk, N. Vulpe, Geometric configurations of singularities of planar polynomial differential systems, Birkhäuser, 2021. https://doi.org/10.1007/978-3-030-50570-7
    [2] L. Depalo, C. Gallego, R. Ortells-Fabra, C. Salas, R. Montal, A. Urbaneja, et al., Advancing tomato crop protection: Green leaf volatile-mediated defense mechanisms against Nesidiocoris tenuis plant damage, Biol. Control, 192 (2024), 105517. https://doi.org/10.1016/j.biocontrol.2024.105517 doi: 10.1016/j.biocontrol.2024.105517
    [3] H. I. Freedman, G. S. K. Wolkowicz, Predator-prey systems with group defence: The paradox of enrichment revisited, Bull. Math. Biol., 48 (1986), 493–508. https://doi.org/10.1016/S0092-8240(86)90004-2 doi: 10.1016/S0092-8240(86)90004-2
    [4] A. Gasull, Differential equations that can be transformed into equations of Liénard type, Actas del XVLL colloquio Brasileiro de matematica, 1989.
    [5] G. F. Gause, The struggle for existence, Baltimore: Williams and Wilkins, 1934.
    [6] E. González-Olivares, B. González-Yañez, J. M. Lorca, A. Rojas-Palma, J. D. Flores, Consequences of double Allee effect on the number of limit cycles in a predator-prey model, Comput. Math. Appl., 62 (2011), 3449–3463. https://doi.org/10.1016/j.camwa.2011.08.061 doi: 10.1016/j.camwa.2011.08.061
    [7] C. S. Holling, The components of predation as revealed by a study of small-mammal predation of the European pine sawfly, Can. Entomol., 91 (1959), 293–320. https://doi.org/10.4039/Ent91293-5 doi: 10.4039/Ent91293-5
    [8] J. C. Holmes, W. M. Bethel, Modification of intermediate host behavior by parasites, Zool. J. Linn. Soc., 51 (1972), 123–149.
    [9] X. Y. Hou, Y. Li, X. Y. Zhang, S. J. Ge, Y. Mu, J. Y. Shen, Unraveling the intracellular and extracellular self-defense of Chlorella sorokiniana toward highly toxic pyridine stress, Bioresour. Technol., 385 (2023), 129366. https://doi.org/10.1016/j.biortech.2023.129366 doi: 10.1016/j.biortech.2023.129366
    [10] C. Jost, S. P. Ellner, Testing for predator dependence in predator-prey dynamics: A non-parametric approach, Proc. Roy. Soc. B, 267 (2000), 1611–1620. https://doi.org/10.1098/rspb.2000.1186 doi: 10.1098/rspb.2000.1186
    [11] R. E. Kooij, A. Zegeling, A predator-prey model with Ivlev's functional response, J. Math. Anal. Appl., 198 (1996), 473–489. https://doi.org/10.1006/jmaa.1996.0093 doi: 10.1006/jmaa.1996.0093
    [12] R. E. Kooij, A. Zegeling, Qualitative properties of two-dimensional predator-prey systems, Nonlinear Anal., 29 (1997), 693–715. https://doi.org/10.1016/S0362-546X(96)00068-5 doi: 10.1016/S0362-546X(96)00068-5
    [13] R. E. Kooij, A. Zegeling, Predator-prey models with non-analytical functional response, Chaos Solit. Fractals, 123 (2019), 163–172. https://doi.org/10.1016/j.chaos.2019.03.036 doi: 10.1016/j.chaos.2019.03.036
    [14] P. H. Leslie, J. C. Gower, The properties of a stochastic model for a predatorprey type of interaction between two species, Biometrika, 47 (1960), 219–234. https://doi.org/10.2307/2333294 doi: 10.2307/2333294
    [15] Q. Li, Y. Y. Zhang, Y. N. Xiao, Canard phenomena for a slow-fast predator-prey system with group defense of the prey, J. Math. Anal. Appl., 527 (2023), 127418. https://doi.org/10.1016/j.jmaa.2023.127418 doi: 10.1016/j.jmaa.2023.127418
    [16] Y. L. Li, D. M. Xiao, Bifurcations of a predator-prey system of Holling and Leslie types, Chaos Solit. Fractals, 34 (2007), 606–620. https://doi.org/10.1016/j.chaos.2006.03.068 doi: 10.1016/j.chaos.2006.03.068
    [17] Y. Liu, A. Zegeling, W. T. Huang, The application of Liénard transformations to predator-prey systems, Qual. Theory Dyn. Syst., 23 (2024), 91. https://doi.org/10.1007/s12346-023-00947-0 doi: 10.1007/s12346-023-00947-0
    [18] M. Lu, D. Z. Gao, J. C. Huang, H. Wang, Relative prevalence-based dispersal in an epidemic patch model, J. Math. Biol., 86 (2023), 52. https://doi.org/10.1007/s00285-023-01887-8 doi: 10.1007/s00285-023-01887-8
    [19] M. Lu, J. C. Huang, S. G. Ruan, P. Yu, Bifurcation analysis of an SIRS epidemic model with a generalized nonmonotone and saturated incidence rate, J. Differ. Equ., 267 (2019), 1859–1898. https://doi.org/10.1016/j.jde.2019.03.005 doi: 10.1016/j.jde.2019.03.005
    [20] M. Lu, J. C. Huang, S. G. Ruan, P. Yu, Global dynamics of a susceptible-infectious-recovered epidemic model with a generalized nonmonotone incidence rate, J. Dyn. Differ. Equ., 33 (2021), 1625–1661. https://doi.org/10.1007/s10884-020-09862-3 doi: 10.1007/s10884-020-09862-3
    [21] J. Morcuende, J. Martín-García, P. Velasco, T. Sánchez-Gómez, Ó. Santamaría, V. M. Rodríguez, J. Poveda, Effective biological control of chickpea rabies (Ascochyta rabiei) through systemic phytochemical defenses activation by Trichoderma roots colonization: From strain characterization to seed coating, Biol. Control, 193 (2024), 105530. https://doi.org/10.1016/j.biocontrol.2024.105530 doi: 10.1016/j.biocontrol.2024.105530
    [22] F. Rothe, D. S. Shafer, Multiple bifurcation in a predator-prey system with nonmonotonic predator response, Proc. R. Soc. Edinb. A: Math., 120 (1992), 313–347. https://doi.org/10.1017/S0308210500032169 doi: 10.1017/S0308210500032169
    [23] Y. L. Tang, F. Li, Multiple stable states for a class of predator-prey systems with two harvesting rates, J. Appl. Anal. Comput., 14 (2024), 506–514. https://doi.org/10.11948/20230295 doi: 10.11948/20230295
    [24] J. S. Tener, Muskoxen in Canada, a biological and taxonomic review, Canadian Government, Ottawa, 1965.
    [25] G. S. K. Wolkowicz, Bifurcation analysis of a predator-prey system involving group defence, SIAM J. Appl. Math., 48 (1988), 592–606. https://doi.org/10.1137/0148033 doi: 10.1137/0148033
    [26] D. M. Xiao, S. G. Ruan, Codimension two bifurcations in a predator-prey system with group defense, Int. J. Bifurc. Chaos, 11 (2001), 2123–2131. https://doi.org/10.1142/S021812740100336X doi: 10.1142/S021812740100336X
    [27] D. M. Xiao, H. P. Zhu, Multiple focus and Hopf bifurcations in a predator-prey system with nonmonotonic functional response, SIAM J. Appl. Math., 66 (2006), 802–819. https://doi.org/10.1137/050623449 doi: 10.1137/050623449
    [28] R. C. Yang, L. Q. Yang, Y. L. Tang, Some research on limit cycles of Liénard system, Math. Theory Appl., 41 (2021), 59–95.
    [29] Y. Q. Ye, Theory of limit cycles, American Mathematical Society, 1986.
    [30] A. Zegeling, R. E. Kooij, Uniqueness of limit cycles in polynomial systems with algebraic invariants, Bull. Aust. Math. Soc., 49 (1994), 7–20. https://doi.org/10.1017/S0004972700016026 doi: 10.1017/S0004972700016026
    [31] A. Zegeling, R. E. Kooij, Several bifurcation mechanisms for limit cycles in a predator-prey system, Qual. Theory Dyn. Sys., 20 (2021), 65. https://doi.org/10.1007/s12346-021-00501-w doi: 10.1007/s12346-021-00501-w
    [32] A. Zegeling, H. L. Wang, G. Z Zhu, Uniqueness of limit cycles in a predator-prey model with sigmoid functional response, J. Nonlinear Model. Anal., 5 (2023), 790–802. https://doi.org/10.12150/jnma.2023.790 doi: 10.12150/jnma.2023.790
    [33] Z. F. Zhang, T. R. Ding, W. Z. Huang, Z. X. Dong, Qualitative theory of differential equations, American Mathematical Society, 1992.
    [34] Y. R. Zhou, C. W. Wang, D. Blackmore, The uniqueness of limit cycles for Liénard system, J. Math. Anal. Appl., 304 (2005), 473–489. https://doi.org/10.1016/j.jmaa.2004.09.037 doi: 10.1016/j.jmaa.2004.09.037
    [35] B. Zimmerman, H. Sand, P. Wabakken, O. Liberg, H. P. Andreassen, Predator-dependent functional response in wolves: From food limitation to surplus killing, J. Anim. Ecol., 84 (2015), 102–112. https://doi.org/10.1111/1365-2656.12280 doi: 10.1111/1365-2656.12280
  • 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(823) PDF downloads(103) Cited by(0)

Figures and Tables

Figures(7)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog