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

The dynamics of a delayed predator-prey model with square root functional response and stage structure

  • Received: 10 March 2024 Revised: 30 April 2024 Accepted: 11 May 2024 Published: 15 May 2024
  • In recent years, one of the most prevalent matters in population ecology has been the study of predator-prey relationships. In this context, this paper investigated the dynamic behavior of a delayed predator-prey model considering square root type functional response and stage structure for predators. First, we obtained positivity and boundedness of the solutions and existence of equilibrium points. Second, by applying the stability theory of delay differential equations and the Hopf bifurcation theorem, we discussed the system's local stability and the existence of a Hopf bifurcation at the positive equilibrium point. Moreover, the properties of the Hopf bifurcation were deduced by using the central manifold theorem and normal form method. Analytical results showed that when the time delay was less than the critical value, the two populations will coexist, otherwise the ecological balance will be disrupted. Finally, some numerical simulations were also included to verify the theoretical results.

    Citation: Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang. The dynamics of a delayed predator-prey model with square root functional response and stage structure[J]. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150

    Related Papers:

    [1] Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215
    [2] Mengting Sui, Yanfei Du . Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay. Electronic Research Archive, 2023, 31(9): 5124-5150. doi: 10.3934/era.2023262
    [3] Wenbin Zhong, Yuting Ding . Spatiotemporal dynamics of a predator-prey model with a gestation delay and nonlocal competition. Electronic Research Archive, 2025, 33(4): 2601-2617. doi: 10.3934/era.2025116
    [4] Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109
    [5] San-Xing Wu, Xin-You Meng . Hopf bifurcation analysis of a multiple delays stage-structure predator-prey model with refuge and cooperation. Electronic Research Archive, 2025, 33(2): 995-1036. doi: 10.3934/era.2025045
    [6] Meng Wang, Naiwei Liu . Qualitative analysis and traveling wave solutions of a predator-prey model with time delay and stage structure. Electronic Research Archive, 2024, 32(4): 2665-2698. doi: 10.3934/era.2024121
    [7] Ailing Xiang, Liangchen Wang . Boundedness of a predator-prey model with density-dependent motilities and stage structure for the predator. Electronic Research Archive, 2022, 30(5): 1954-1972. doi: 10.3934/era.2022099
    [8] Ruizhi Yang, Dan Jin . Dynamics in a predator-prey model with memory effect in predator and fear effect in prey. Electronic Research Archive, 2022, 30(4): 1322-1339. doi: 10.3934/era.2022069
    [9] Xiaowen Zhang, Wufei Huang, Jiaxin Ma, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and schooling behavior. Electronic Research Archive, 2022, 30(7): 2510-2523. doi: 10.3934/era.2022128
    [10] Maurıicio F. S. Lima, Jaume Llibre . Hopf bifurcation for a class of predator-prey system with small immigration. Electronic Research Archive, 2024, 32(7): 4604-4613. doi: 10.3934/era.2024209
  • In recent years, one of the most prevalent matters in population ecology has been the study of predator-prey relationships. In this context, this paper investigated the dynamic behavior of a delayed predator-prey model considering square root type functional response and stage structure for predators. First, we obtained positivity and boundedness of the solutions and existence of equilibrium points. Second, by applying the stability theory of delay differential equations and the Hopf bifurcation theorem, we discussed the system's local stability and the existence of a Hopf bifurcation at the positive equilibrium point. Moreover, the properties of the Hopf bifurcation were deduced by using the central manifold theorem and normal form method. Analytical results showed that when the time delay was less than the critical value, the two populations will coexist, otherwise the ecological balance will be disrupted. Finally, some numerical simulations were also included to verify the theoretical results.



    There are various biological species in nature, in which many interactions exist. Since the pioneering progress of the work of Lotka [1] and Volterra [2] in the 1920s on the interaction between predators and prey, the dynamics of predator-prey models have been widely studied by scholars [3,4,5,6].

    The functional responses are the functions of the density of prey due to the predation rate. In the famous Lokta-Volterra model, the number of prey captured by predators per unit of time increases with the increase of the prey quantity, without considering the impact of digestion saturation on the system, which has some shortcomings. In 1965, Holling [7] developed three types of functional responses based on different species. The Holling-type functional responses can reflect the phenomenon of predator density gradually saturating as the number of prey increases, which sparked extensive research [8,9,10]. With further research, scholars have put forward some functional responses that are more in line with reality [11,12,13,14]. When exploring the herd behavior between populations, Ajraldi et al. [15] found that cattle herds gather together for defensive purposes. Usually, weaker individuals occupy the interior of the herd, while stronger individuals are outside. Due to the discovery that the number of cattle is directly proportional to the square root of density, they considered that using the square root of the prey density to simulate the functional response of prey exhibiting this herd behavior was appropriate. Braza [16] considered a predator-prey model with square root functional response, in which prey exhibit strong herd structure. Salman et al. [17] presented a discrete predator-prey model with square root functional response and studied the stability and bifurcation of the system. Suleman et al. [18] structured a slow-fast predator-prey model with herd behavior, where the presence and stability of fixed points were investigated. He and Li [19] introduced the square root response function into a Leslie-Gower predator-prey model and discussed the impact of square root functional response on the dynamic behaviors of the system. Furthermore, researchers also studied other systems with square root response functions [20,21,22,23,24,25].

    In [22], Mortuja et al. proposed a predator-prey model with nonlinear prey harvesting and square root functional response:

    ˙X(t)=rX(t)(1X(t)p)αX(t)Y(t)1+thαX(t)qEX(t)m1E+m2X(t),˙Y(t)=βY(t)+mαX(t)Y(t)1+thαX(t), (1.1)

    where X(t) and Y(t) represent the densities of the prey and predator population at the time t. r denotes the growth rate of the prey and p is the environmental carrying capacity. α represents the search efficiency of the predator for prey and β represents the natural mortality rate of the predator in the absence of prey. th is the average processing time of the predator for the prey, and m is the consumption rate. In the nonlinear harvesting section, q is the capture capability coefficient, and E is the harvest effort. m1 and m2 are appropriate constants. They explored the stability of the system and discussed Hopf bifurcation and saddle-node bifurcation.

    According to the characteristics and growth laws of organisms, young biological individuals cannot compete and reproduce for a long time. Therefore, it is more practical to divide the population into different stage structures. The predator-prey models with stage structure have been paid much attention by scholars. The predators or prey are divided into immature and mature stages in these models. Dai et al. [26] considered a Holling type-III predator-prey model with stage structure for prey and researched the asymptotic stability of the positive equilibrium point. Meng et al. [27] obtained global Hopf bifurcation in a predator-prey model with a stage structure for prey. A predator-prey model with a time delay and stage structure for predators was analyzed by Zhou [28]. He found that the combination of the time delay and stage structure can affect the dynamic behavior of the system. Zhao and Zeng [29] dealt with a predator-prey model with stage structure for predator and ratio-dependent functional response. They extended the existence of stationary distribution and obtained sufficient conditions for the extinction of predator populations.

    In addition, interactions between populations inevitably cause time delays. For example, the reproductive cycle of a population, the time it takes for predators to absorb and convert their prey into their energy, and so on, are phenomena that exist in a time delay. From a dynamic perspective, differential equations with a time delay typically exhibit more complex dynamic behavior than ordinary differential equations. Introducing a time delay into predator-prey models has practical significance. Zhang et al. [30] came up with a predator-prey model with a discrete time delay, and the results showed that predators would become extinct with increasing the time delay. Yang and Jin [31] proposed and analyzed a diffusive model with a time delay. They studied the stability of the system and Hopf bifurcation. Some scholars have also studied predator-prey models with time delays [32,33,34,35,36].

    Based on the above discussions and inspiration from the work of Mortuja et al. [22], in this paper, we consider introducing a stage structure into the system (1.1) and dividing the predator into two stages: an immature predator and a mature predator, and the prey is only preyed upon by mature predators. In addition, the time delay of the predator is also considered. The predator-prey model with stage structure and a time delay is as follows:

    ˙X(t)=rX(t)(1X(t)p)αX(t)Y2(t)1+thαX(t)qEX(t)m1E+m2X(t),˙Y1(t)=mαX(tτ)Y2(tτ)1+thαX(tτ)βY1(t)DY1(t),˙Y2(t)=DY1(t)dY2(t), (1.2)

    where X(t) represents the population density of the prey, and Y1(t) and Y2(t) denote the population density of immature and mature predators, respectively. D is the rate at which immature predators become mature predators. β and d respectively express the natural mortality rates of immature and mature predators. τ is taken as a gestation delay of the mature predator. The other parameters r, p, α, th, m, q, E, m1, and m2 represent the same meaning as in system (1.1).

    For convenience, we have scaled the variables and the parameters of the system (1.2). The variables are scaled as

    x=Xp,y1=αY1rp,y2=αY2rp,ˉt=rt,

    and the other parameters are made dimensionless as follows:

    a=thαp,b=mαpr,d1=βr,d2=dr,ˉD=Dr,h=qErm2p,s=m1Em2p,ˉτ=rτ.

    In addition, to avoid heavy notation, we still use t, τ, D to represent ˉt, ˉτ, ˉD. Through the above transformation, the dimensionless form of the system (1.2) is as follows:

    ˙x(t)=x(t)(1x(t))x(t)y2(t)1+ax(t)hx(t)x(t)+s,˙y1(t)=bx(tτ)y2(tτ)1+ax(tτ)d1y1(t)Dy1(t),˙y2(t)=Dy1(t)d2y2(t), (1.3)

    with initial conditions

    x(θ)=ϕ1(θ)0,y1(θ)=ϕ2(θ)0,y2(θ)=ϕ3(θ)0,θ[τ,0),
    ϕ1(0)>0,ϕ2(0)>0,ϕ3(0)>0,(ϕ1(θ),ϕ2(θ),ϕ3(θ))([τ,0],R3+0),

    where R3+0={(x1,x2,x3):xi0,i=1,2,3}.

    According to the above definition, all parameters here are positive.

    The rest of this article is organized as follows: In Section 2, we discuss positivity, boundedness, the stability of the equilibrium points, and the existence of Hopf bifurcation. Section 3 is devoted to the nature of Hopf bifurcation. In Section 4, some numerical simulations are conducted to verify our analytical results. The article ends with a simple summary.

    In this section, we will study the positivity and boundedness of the solution, the local stability of the equilibria of the system (1.3), and prove the existence of Hopf bifurcation.

    Considering the biological background, positivity means the survival of the population, so it is necessary to prove the positivity of the solution. The system (1.3) can be rewritten into the following matrix form:

    dXdt=J(X),

    where X=(x(t),y1(t),y2(t))TR3, and J(X) can be written as:

    J(X)=(J1(X)J2(X)J3(X))=(x(t)(1x(t))x(t)y2(t)1+ax(t)hx(t)x(t)+sbx(tτ)y2(tτ)1+ax(tτ)d1y1(t)Dy1(t)Dy1(t)d2y2(t)).

    Because J(X) and JX are continuous on R3+, then J:R3+R3 is locally Lipschitz. Through the standard theory of the ordinary differential equation system, the unique solution of the system (1.3) under any initial condition J(0)=J0=(x(0),y1(0),y2(0))R3+ can be obtained.

    Then, the system (1.3) can be rewritten as:

    dxdt=xϕ1(x,y1,y2),dy1dt=y1ϕ2(x,y1,y2),dy2dt=y2ϕ3(x,y1,y2),

    where

    ϕ1(x,y1,y2)=1x(t)y2(t)x(t)+ax(t)hx(t)+s,ϕ2(x,y1,y2)=bx(tτ)y2(tτ)(1+ax(tτ))y1(t)d1D,ϕ3(x,y1,y2)=Dy1(t)y2(t)d2.

    Taking the second equation as an example, we can obtain

    1y1dy1dt=bx(tτ)y2(tτ)(1+ax(tτ))y1(t)d1D,

    that is

    d(lny1)dt=bx(tτ)y2(tτ)(1+ax(tτ))y1(t)d1D.

    Integrating both sides of the above equation from 0 to t, we obtain

    lny1(t)lny1(0)=t0[bx(tτ)y2(tτ)(1+ax(tτ))y1(t)d1D]dt,

    then

    lny1(t)y1(0)=t0[bx(tτ)y2(tτ)(1+ax(tτ))y1(t)d1D]dt,

    and we have

    y1(t)y1(0)=et0[bx(tτ)y2(tτ)(1+ax(tτ))y1(t)d1D]dt.

    Therefore, we have

    y1(t)=y1(0)et0[bx(tτ)y2(tτ)(1+ax(tτ))y1(t)d1D]dt0.

    Similarly, it can be concluded that x(t)0, y2(t)0. Therefore, for any t>0, the solution X(t)=(x(t),y1(t),y2(t)) satisfying the initial condition X(0)=(x(0),y1(0),y2(0))R3+ remains nonnegative in the region R3+.

    Next, we show that all solutions of system (1.3) under the initial condition are bounded.

    Boundedness is indicated as a result of limited resources. We know that all solutions of system (1.3) with the initial condition are positive. From the first equation of system (1.3), we obtain ˙x(t)x(t)(1x(t)). Considering that ˙s(t)=s(t)(1s(t)), the solution that satisfies the critical condition is s(t)=s(0)1s(0)1+et. By comparing x(t) and s(t), it is obtained that x(t)s(t) for all t0. Then, all solutions of system (1.3) satisfy 0<x(t)1 for t0.

    Next, we define that V(t)=x(tτ)+y1(t)b+y2(t)b. The derivative of V(t) along the solution of system (1.3) is

    ˙V(t)=x(tτ)(1x(tτ))hx(tτ)x(tτ)+sd1by1(t)d2by2(t),x(tτ)(1x(tτ))d1by1(t)d2by2(t),

    and we choose μ=min{d1,d2} as a constant. Then,

    ˙V(t)μV(t)+(1+μ)x(tτ)x2(tτ),μV(t)+(1+μ)24.

    Let (1+μ)24=M, and we obtain that

    0V(t)eμt(V(0)+t0Meμsds)=eμtV(0)+Mμ(1eμt).

    Then, for a sufficiently large t, there is 0<V(t)Mμ. Therefore, the solutions x(t), y1(t), and y2(t) under the initial conditions are bounded.

    The system (1.3) has nonnegative equilibrium points.

    (i) System (1.3) always has a trivial equilibrium point E0(0,0,0).

    (ii) (a) If (H1) 1s>0, (H2) (1s)24(hs)>0, and (H3) h>s hold, system (1.3) has two predator-free equilibrium points Ei(xi,0,0)(i=1,2), where

    x1=12[(1s)+(1s)24(hs)],x2=12[(1s)(1s)24(hs)].

    (b) If (H1), (H2), and (H4) h<s hold, only a boundary equilibrium E1(x1,0,0) exists.

    (c) When (H1) and (H5) (1s)2=4(hs) hold, system (1.3) has a predator-free equilibrium point E3(x3,0,0), where x3=1s2.

    (iii) The unique positive equilibrium point E(x,y1,y2) exists in the system (1.3) if condition (H6) c0>c1>0, where c0=1x, c1=hx+s>0, and (H7) bDad2(d1+D)>d2(d1+D) are met, where x=[d2(d1+D)bDad2(d1+D)]2, y1=b(c0c1)d1+Dx, y2=Dd2y1.

    In this part, we mainly discuss the stability of the trivial equilibrium point E0 and the positive equilibrium point E.

    Theorem 2.1. If (H4) h<s holds, the equilibrium point E0(0,0,0) is unstable. The equilibrium point E0(0,0,0) is locally asymptotically stable, when (H3) h>s is satisfied.

    Proof. The characteristic equation at the equilibrium point E0 is

    (λ+d2)(λ+d1+D)(1hsλ)=0. (2.1)

    All of the three characteristic roots of the equation can be obtained: λ1=d2<0, λ2=(d1+D)<0, λ3=1hs. If condition (H4) h<s holds, the eigenvalue λ3 is a positive real root, and therefore, E0 is unstable. If condition (H3) h>s holds, λ3 is a negative real root. All eigenvalues are negative, which leads to E0 being locally asymptotically stable. The proof is complete.

    Next, we analyze the stability of the unique positive equilibrium point E of the system (1.3).

    Let ˉx(t)=x(t)x(t), ˉy1(t)=y1(t)y1(t), and ˉy2(t)=y2(t)y2(t). For convenience, we still use x(t), y1(t), y2(t) to represent ˉx(t), ˉy1(t), ˉy2(t). We get

    ˙x(t)=a11x(t)+a13y2(t)+i+k2f(ik)1x(i)(t)y(k)2(t),˙y1(t)=a22y1(t)+b21x(tτ)+b23y2(tτ)+j+m+n2f(jmn)2y(j)1(t)x(m)(tτ)y(n)2(tτ),˙y2(t)=a32y1(t)+a33y2(t)+j+k2f(jk)3y(j)1(t)y(k)2(t), (2.2)

    where the coefficients and functions of the system (2.2) are listed in Appendix A.

    Then, the system (1.3) is linearized to

    ˙x(t)=a11x(t)+a13y2(t),˙y1(t)=a22y1(t)+b21x(tτ)+b23y2(tτ),˙y2(t)=a32y1(t)+a33y2(t). (2.3)

    Therefore, the characteristic equation of system (2.3) is written as

    λ3+l12λ2+l11λ+l10+eτλ(h11+h10λ)=0, (2.4)

    where

    l10=a11a22a33,l11=a11a33+a11a22+a22a33,l12=(a11+a22+a33),h10=a32b23,
    h11=a11a32b23a13a32b21.

    Next, we will discuss two different situations.

    When τ=0, Eq (2.4) becomes

    λ3+l12λ2+l21λ+l20=0, (2.5)

    where l20=l10+h11, l21=l11+h10.

    Therefore, by the Routh-Hurwitz discriminant method, we obtain that all of the roots of Eq (2.5) have the negative real part, if (H8) l12>0, l20>0, l12l21l20>0 holds. This means that the positive equilibrium point E is locally asymptotically stable.

    When τ0, let iω(ω>0) be a pure imaginary root of Eq (2.4). By separating the real and imaginary parts, it can be obtained that

    h10ωcosωτh11sinωτ=ω3l11ω,h10ωsinωτ+h11cosωτ=l12ω2l10. (2.6)

    By squaring and adding, we have

    ω6+e12ω4+e11ω2+e10=0, (2.7)

    where

    e10=l210h211,e11=l2112l10l12h210,e12=l2122l11.

    Let ω2=υ, and Eq (2.7) can be written as

    υ3+e12υ2+e11υ+e10=0. (2.8)

    Let

    g(υ)=υ3+e12υ2+e11υ+e10. (2.9)

    We can get g(0)=e10, limυ+g(υ)=+, and the derivative of Eq (2.9) is g(υ)=3υ2+2e12υ+e11.

    Therefore, the roots of Eq (2.8) can be studied by using the method of reference [37], along with the following lemma.

    Lemma 2.1. For Eq (2.8),

    (i) if (H9) e100, Δ=e2123e110 holds, then Eq (2.8) does not have positive roots.

    (ii) if (H10) e100, Δ=e2123e11>0, υ=e11+Δ3>0, g(υ)0 or (H11) e10<0 holds, then Eq (2.8) has positive roots.

    Therefore, we suppose that Eq (2.8) has positive roots. Without loss of generality, we assume that it has three positive roots, denoted by υ1, υ2 and υ3, respectively. Then Eq (2.7) also has three positive roots, defined as ωk=υk, k=1,2,3. Furthermore, the critical value corresponding to the time delay τ(j)k can be expressed as

    τ(j)k=1ωkarccos{h10ω4k+(l12h11l11h10)ω2kl10h11h210ω2k+h211}+2πjωk, (2.10)
    k=1,2,3;j=0,1,2,.

    Let λ(τ)=ς(τ)+iω(τ) be the root of Eq (2.4) and ς(τjk)=0, thus ±ωk is a pair of purely imaginary roots of Eq (2.4) with τ=τ(j)k. We define τ=mink=1,2,3;j=0,1,2,{τ(j)k} and ω=ωk.

    Lemma 2.2. If (H12) g(ω2)0 holds, then {d(Reλ)dτ}λ=iω0.

    Proof. Taking the derivative of Eq (2.4) on both sides with respect to τ, we obtain

    (dλdτ)1=3λ2+2l12λ+l11λ(λ3+l12λ2+l11λ+l10)+h10λ(h11+h10λ)τλ. (2.11)

    Further calculation reveals that

    Re(dλdτ)1λ=iω=Re(3λ2+2l12λ+l11λ(λ3+l12λ2+l11λ+l10))λ=iω+Re(h10λ(h11+h10λ))λ=iω
    =3ω4+(2l2124l11)ω2+l2112l10l12(ω3l11ω)2+(l12ω2l10)2h210h210ω2+h211.

    According to Eq (2.6), it can be calculated that

    (ω3l11ω)2+(l12ω2l10)2=h211+h210ω2. (2.12)

    Since {d(Reλ)dτ}λ=iω and {Re(dλdτ)1}λ=iω have the same symbol, then

    sign{d(Reλ)dτ}λ=iω=sign{Re(dλdτ)1}λ=iω=sign3(ω2)2+2e12ω2+e11(ω3l11ω)2+(l12ω2l10)2=signg(ω2)(ω3l11ω)2+(l12ω2l10)2. (2.13)

    So, if (H12) g(ω2)0 is met, {d(Reλ)dτ}λ=iω0 can be obtained.

    The proof of the transversality condition is complete.

    In summary, the following conclusions can be drawn.

    Theorem 2.2. For system (1.3), suppose (H6), (H7), and (H8) hold, then the following results satisfy:

    (i) If (H9) holds, the positive equilibrium point E is locally asymptotically stable when τ0.

    (ii) If (H10) or (H11) and (H12) are satisfied, the positive equilibrium point E is locally asymptotically stable when τ[0,τ) and unstable when τ>τ. In addition, when τ=τ, a Hopf bifurcation occurs at the positive equilibrium point E.

    In this section, we will apply the central manifold theorem and the normal form method in [38] to analyze the properties of Hopf bifurcation at the positive equilibrium point E.

    Let τ=τ+ξ, ξR, t=sτ, x(sτ)=˜x(s), y1(sτ)=˜y1(s), and y2(sτ)=˜y2(s), where x=˜x, y1=˜y1, y2=˜y2, and t=s. Then the system (1.3) can be described as the following functional differential equation in C=C([1,0],R3)

    u(t)=Lξ(ut)+F(ξ,ut), (3.1)

    where u(t)=(x(t),y1(t),y2(t))TC, ut(θ)=u(t+θ)=(x(t+θ),y1(t+θ),y2(t+θ))TC, and Lξ:CR3, F:R×CR3 are written as

    Lξ(ϕ)=(τ+ξ)Eϕ(0)+(τ+ξ)Lϕ(1), (3.2)

    and

    F(ξ,ϕ)=(τ+ξ)(F1,F2,F3)T, (3.3)

    where the specific form of Lξ(ϕ) and F(ξ,ϕ) are given in Appendix B.

    Therefore, according to the Riesz representation theorem, there exists a bounded variation function η(θ,ξ) of 3×3 where θ[1,0], such that

    Lξϕ=01dη(θ,ξ)ϕ(θ), for ϕC. (3.4)

    We can choose

    η(θ,ξ)=(τ+ξ)Eδ(θ)+(τ+ξ)Lδ(θ+1), (3.5)

    where δ(θ)={0,θ0,1,θ=0.

    For ϕC([1,0],R3), we define

    A(ξ)ϕ={dϕ(θ)dθ,1θ<0,01dη(θ,ξ)ϕ(θ),θ=0, (3.6)

    and

    Rξ(ϕ)={0,1θ<0,F(ξ,ϕ),θ=0. (3.7)

    Then, Eq (3.1) can be rewritten in the form of the following operator equation

    ut=A(ξ)ut+R(ξ)ut. (3.8)

    For φC([1,0],(R3)), where (R3) is a three-dimensional row vector space, define the adjoint operator A of A(0)

    Aφ(s)={dφ(s)ds,s(0,1],01dηT(t,0)φ(t),s=0. (3.9)

    For ϕC([1,0],R3) and φC([1,0],(R3)), we define bilinear inner product maps

    φ(s),ϕ(s)=ˉφ(0)ϕ(0)01θμ=0ˉφ(μθ)dη(θ)ϕ(μ)dμ, (3.10)

    where η(θ)=η(θ,0), A=A(0), and A are adjoint operators.

    According to the analysis in the previous section, there is a pair of pure imaginary roots ±iωτ for A(0) and A. We assume that q(θ)=(1,q1,q2)Teiωτθ is the eigenvector of the feature root A(0) corresponding to iωτ, and q(s)=Ω(1,q1,q2)eiωτs is the eigenvector of the feature root A corresponding to iωτ. After some calculations, it can be concluded that

    q1=ω2(a11+a33)iω+a11a33a13a32,q2=iωa11a13,
    q1=iω+a11b21eiωτ,q2=ω2+(a11+a22)iω+a11a22a32b21eiωτ.

    Thus, combined with Eq (3.10), there is

    q(s),q(θ)=ˉΩ(1,ˉq1,ˉq2)(1,q1,q2)T01θμ=0ˉq(μθ)dη(θ)q(μ)dμ=ˉΩ[1+q1ˉq1+q2ˉq201(1,ˉq1,ˉq2)θeiωτθdη(θ)(1,q1,q2)T]=ˉΩ[1+q1ˉq1+q2ˉq2+(b21+q2b23)ˉq1τeiωτ]. (3.11)

    Hence, it is obtained that

    ˉΩ=11+q1ˉq1+q2ˉq2+(b21+q2b23)ˉq1τeiωτ, (3.12)

    so that q(s),q(θ)=1, q(s),ˉq(θ)=0.

    Next, adopting the approach in [38] and the calculation process similar to [37], we obtain the coefficients that can be applied to determine the nature of the Hopf bifurcation.

    The details of the derivations of coefficients g20, g11, g02 and g21 are given in Appendix C.

    Then, the following formulas can be calculated to obtain:

    c1(0)=i2ωτ(g20g112|g11|2|g02|23)+g212,ρ2=Re{c1(0)}Re{λ(τ)},β2=2Re{c1(0)},T2=Im{c1(0)}+ρ2Im{λ(τ)}ωτ. (3.13)

    Through the above analysis, the following conclusions can be drawn.

    Theorem 3.1. For system (1.3),

    (i) the sign of ρ2 can determine the direction of the Hopf bifurcation, if ρ2>0(ρ2<0), then the Hopf bifurcation is supercritical (subcritical).

    (ii) the sign of β2 can determine the stability of the bifurcation periodic solutions, if β2<0(β2>0), then the bifurcation periodic solutions are stable (unstable).

    (iii) the sign of T2 determines the periodic change of the bifurcating periodic solutions, if T2>0(T2<0), it represents an increase (decrease) in the period of the bifurcating periodic solutions.

    In this section, we use MATLAB software to perform some numerical simulations to demonstrate the results of the analysis in the above sections.

    In the system (1.3), we select a set of parameters a=0.4, b=0.4, h=0.2, s=0.4, d1=0.1, d2=0.1, and D=0.1. We found the trivial equilibrium point E0(0,0,0) is always unstable.

    When the conditions (H6) and (H7) hold, system (1.3) has a unique positive equilibrium E(0.3906,0.2784,0.2784). For τ=0, (H8) is also met under this set of parameters. Then the positive equilibrium E is locally asymptotically stable, which is presented in Figure 1.

    Figure 1.  Numerical simulation shows that E is locally asymptotically stable when τ=0.

    For τ0, we have ω0.0710 and τ8.9781, and the transversality condition is satisfied. According to Theorem 2.2, when τ=5.75<τ, the positive equilibrium E is locally asymptotically stable, and two populations will be stable eventually, which is shown in Figure 2. For τ=9.73>τ, E becomes unstable. Thus, the Hopf bifurcation occurs when τ8.9781 and a family of periodic solutions bifurcate from the positive equilibrium E. From Figure 3, it can be seen that the two populations are unstable. It can further prove the instability of the system caused by a time delay, inducing species to fluctuate.

    Figure 2.  Numerical simulation shows that E is locally asymptotically stable when τ=5.75<τ=8.9781.
    Figure 3.  Numerical simulation shows that a Hopf bifurcation occurs when τ=9.73>τ=8.9781.

    Furthermore, after the calculation of Eq (3.13), we can gain c1(0)=43.2389125.4193i, β2=86.4778, ρ2=7.6070×104, and T2=428.9949. From Theorem 3.1, it can be obtained that the Hopf bifurcation is supercritical, the bifurcating periodic solutions are stable, and its period increases. From a biological perspective, the stable bifurcating periodic solutions indicate the coexistence of prey and predator in an oscillatory mode.

    A delayed predator-prey model with square root functional response and stage structure was studied in this paper inspired by the work of Mortuja et al. [22], where predators were divided into two stages: immature predators and mature predators, and τ represented the gestation delay of the mature predator in prey and predator interaction.

    First, we analyzed the positivity and boundedness of the solution, and studied the existence of three types of different equilibrium points, as well as the relevant stability conditions for the trivial equilibrium and positive equilibrium of the system (1.3). We found that the equilibrium E0 was always unstable, and if (H8) l12>0, l20>0, l12l21l20>0 were satisfied, then the positive equilibrium E was locally asymptotically stable when τ=0(see Figure 1). On this basis, the sufficient conditions for the Hopf bifurcation of the system (1.3) at the positive equilibrium point E when τ>0 were also studied, and the critical value of the time delay was calculated. When the time delay was less than the critical value τ, the positive equilibrium E was locally asymptotically stable (see Figure 2). When the time delay increased and passed the critical value τ, the stability of positive equilibrium E switched from locally asymptotically stable to unstable and generated a Hopf bifurcation, as well as a family of periodic solutions bifurcated from the positive equilibrium E (see Figure 3). The research results indicate that the time delay τ plays an important role in the system (1.3) which can affect the stability of this system and the existence of the Hopf bifurcation. In addition, by using the central manifold theorem and normal form theory, explicit formulas were deduced to determine the direction of the Hopf bifurcation and the properties of the periodic solutions. In the end, numerical simulations validated our theoretical analysis. Mortuja et al. [22] observed that the harvesting rate can influence population coexistence and ecological balance. Compared with reference [22], we introduced time delay and stage structure into the system and have shown that gestation delay is a very important factor to understand the dynamics of the system. The stable bifurcating periodic solutions mean that the prey, immature predator, and mature predator can oscillate periodically to survive together. This is valuable from the view of ecology.

    This paper investigated a predator-prey model with a stage structure for predators. What kind of dynamic behavior would occur if the prey had a stage structure or if both populations had a stage structure? We will leave this for future research.

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

    The authors would like to thank the editors and referees for their careful reading of the manuscript and valuable suggestions. This work was supported by the National Natural Science Foundation of China (Grant Nos. 12102148 and 12372012).

    The authors declare that there are no conflicts of interest.

    Appendix A. The coefficients and functions of the system (2.2)

    a11=12xhs(x+s)2y22(1+ax)2x,a13=x1+ax,
    a22=d1D,b21=by22(1+ax)2x,b23=bx1+ax,
    a32=D,a33=d2,
    f1=x(t)(1x(t))x(t)y2(t)1+ax(t)hx(t)x(t)+s,
    f2=bx(tτ)y2(tτ)1+ax(tτ)d1y1(t)Dy1(t),
    f3=Dy1(t)d2y2(t),
    f(ik)1=1i!k!i+kf1xi(t)yk2(t)|(x,y1,y2),
    f(jmn)2=1j!m!n!j+m+nf2yj1(t)xm(tτ)yn2(xτ)|(x,y1,y2),
    f(jk)3=1j!k!j+kf3yj1(t)yk2(t)|(x,y1,y2).

    Appendix B. The specific form of Lξ(ϕ) and F(ξ,ϕ)

    Lξ:CR3, F:R×CR3 are written as

    Lξ(ϕ)=(τ+ξ)Eϕ(0)+(τ+ξ)Lϕ(1), (B1)

    and

    F(ξ,ϕ)=(τ+ξ)(F1,F2,F3)T, (B2)

    where

    ϕ(θ)=(ϕ1(θ),ϕ2(θ),ϕ3(θ))TC,
    E=(a110a130a2200a32a33),L=(000b210b23000),
    F1=k11ϕ21(0)+k12ϕ1(0)ϕ3(0)+k13ϕ31(0)+k14ϕ21(0)ϕ3(0)+,
    F2=k21ϕ21(1)+k22ϕ1(1)ϕ3(1)+k23ϕ31(1)+k24ϕ21(1)ϕ3(1)+,
    F3=0,
    k11=2+y24(x)3+3ay24x(1+ax)3+2hs(x+s)3,k12=12x(1+ax)2,
    k13=3y28(x)5+3ay22x2+15a2y28(x)3(1+ax)46hs(x+s)4,k14=14(x)3+3a4x(1+ax)3,
    k21=by24(x)3+3aby24x(1+ax)3,k22=b2x(1+ax)2,
    k23=3by28(x)5+3aby22x2+15a2by28(x)3(1+ax)4,k24=b4(x)3+3ab4x(1+ax)3.

    Appendix C. The details of the derivations of coefficients g20, g11, g02, and g21

    We need to calculate the coordinate representation of the center manifold C0 at the origin. We define

    {z(t)=q,ut,W(t,θ)=ut(θ)2Re{z(t)q(θ)}=W(z(t),ˉz(t),θ) (C1)

    on the central manifold C0, and there is

    W(z(t),ˉz(t),θ)=12W20(θ)z2+W11(θ)zˉz+12W02ˉzˉz+. (C2)

    We obtain

    ˙z(t)=iωτz(t)+ˉq(0)F0(z,ˉz)=iωτz(t)+g(z,ˉz), (C3)

    where

    g(z,ˉz)=ˉq(0)F0(z,ˉz)=ˉq(0)F(0,ut)=τΩ(1,ˉq1,ˉq2)×(F1(0,ut),F2(0,ut),F3(0,ut))T=12g20(θ)z2+g11(θ)zˉz+12g02(θ)zˉz+12g21(θ)z2ˉz+, (C4)

    with

    F1(0,ut)=k11x2t(0)+k12xt(0)y2t(0)+k13x3t(0)+k14x2t(0)y2t(0)+,
    F2(0,ut)=k21x2t(1)+k22xt(1)y2t(1)+k23x3t(1)+k24x2t(1)y2t(1)+,
    F3(0,ut)=0.

    From Eq (C1), we get

    ut(θ)=(xt(θ),y1t(θ),y2t(θ))=W(t,θ)+zq(θ)+ˉzˉq(θ),

    therefore, it is obtained that

    xt(0)=z+ˉz+12W(1)20(0)z2+W(1)11(0)zˉz+12W(1)02(0)ˉzˉz+,
    y2t(0)=q2z+ˉq2ˉz+12W(3)20(0)z2+W(3)11(0)zˉz+12W(3)02(0)ˉzˉz+,
    xt(1)=zeiωτ+ˉzeiωτ+12W(1)20(1)z2+W(1)11(1)zˉz+12W(1)02(1)ˉzˉz+,
    y2t(1)=q2zeiωτ+ˉq2ˉzeiωτ+12W(3)20(1)z2+W(3)11(1)zˉz+12W(3)02(1)ˉzˉz+.

    From the above formula, it can be calculated that

    g(z,ˉz)=τˉΩ(F1+ˉq1F2)=τˉΩ[k11x2t(0)+k12xt(0)y2t(0)+k13x3t(0)+k14x2t(0)y2t(0)+ˉq1(k21x2t(1)+k22xt(1)y2t(1)+k23x3t(1)+k24x2t(1)y2t(1))]. (C5)

    By comparing the coefficients of Eqs (C4) and (C5), we get

    g20=2τˉΩ[k11+k12q2+ˉq1(k21e2iωτ+k22q2e2iωτ)],g11=τˉΩ[2k11+k12(q2+ˉq2)+ˉq1(2k21+k22(q2+ˉq2))],g02=2τˉΩ[k11+k12ˉq2+ˉq1(k21e2iωτ+k22ˉq2e2iωτ)],g21=2τˉΩ{k11[2W(1)11(0)+W(1)20(0)]+k12[W(3)11(0)+12W(3)20(0)+12ˉq2W(1)20(0)+q2W(1)11(0)]+3k13+k14(2q2+ˉq2)+ˉq1[k21(2eiωτW(1)11(1)+eiωτW(1)20(1))+k22(eiωτW(3)11(1)+12eiωτW(3)20(1)+12ˉq2eiωτW(1)20(1)+q2eiωτW(1)11(1))+3k23eiωτ+k24(2q2+ˉq2)eiωτ]}. (C6)

    Through further calculation, we obtain the unknown coefficients W20(θ) and W11(θ) of g21

    W20(θ)=ig20ωτq(0)eiωτθ+iˉg023ωτˉq(0)eiωτθ+M1e2iωτθ,W11(θ)=ig11ωτq(0)eiωτθ+iˉg11ωτˉq(0)eiωτθ+M2 (C7)

    where M1=(M(1)1,M(2)1,M(3)1)TR3 and M2=(M(1)2,M(2)2,M(3)2)TR3 are constant vectors, and given

    (2iωa110a13b21e2iωτ2iωa22b23e2iωτ0a322iωa33)M1=2(H1H2H3),(a110a13b21a22b230a32a33)M2=(P1P2P3), (C8)

    with

    H1=k11+k12q2,P1=2k11+k12(q2+ˉq2),
    H2=k21e2iωτ+k22q2e2iωτ,P2=2k21+k22(q2+ˉq2),
    H3=0,P3=0.


    [1] Lotka, Elements of Physical Biology, Williams and Wilkins Company, Baltimore, 1925.
    [2] V. Volterra, Fluctuations in the abundance of a species considered mathematically, Nature, 119 (1927), 12–13. https://doi.org/10.1038/119012a0 doi: 10.1038/119012a0
    [3] Y. Y. Huang, F. Y. Li, J. P. Shi, Stability of synchronized steady state solution of diffusive Lotka-Volterra predator-prey model, Appl. Math. Lett., 105 (2020). https://doi.org/10.1016/j.aml.2020.106331
    [4] B. Ghanbari, S. Djilali, Mathematical and numerical analysis of a three-species predator-prey model with herd behavior and time fractional-order derivative, Math. Meth. Appl. Sci., 43 (2020), 1736–1752. https://doi.org/10.1002/mma.5999 doi: 10.1002/mma.5999
    [5] Y. Z. Liu, Y. P. Yang, Dynamics and bifurcation analysis of a delay non-smooth Filippov Leslie-Gower prey-predator model, Nonlinear Dyn., 111 (2023), 18541–18557. https://doi.org/10.1007/s11071-023-08789-w doi: 10.1007/s11071-023-08789-w
    [6] M. L. Deng, Y. B. Fan, Invariant measure of a stochastic hybrid predator-prey model with infected prey, Appl. Math. Lett., 124 (2022). https://doi.org/10.1016/j.aml.2021.107670
    [7] C. S. Holling, Some characteristics of simple types of predation and parasitism, Can. Entomol., 91 (1959), 385–395. https://doi.org/10.4039/Ent91385-7 doi: 10.4039/Ent91385-7
    [8] H. A. A. El-Saka, S. Lee, B. Jang, Dynamic analysis of fractional-order predator-prey biological economic system with Holling type Ⅱ functional response, Nonlinear Dyn., 96 (2019), 407–416. https://doi.org/10.1007/s11071-019-04796-y doi: 10.1007/s11071-019-04796-y
    [9] C. L. Qin, J. J. Du, Y. X. Hui, Dynamical behavior of a stochastic predator-prey model with Holling-type Ⅲ functional response and infectious predator, AIMS Math., 7 (2022), 7403–7418. https://doi.org/10.3934/math.2022413 doi: 10.3934/math.2022413
    [10] S. M. Li, X. Wang, X. Li, K. Wu, Relaxation oscillations for Leslie-type predator-prey model with Holling Type I response functional function, Appl. Math. Lett., 120 (2021). https://doi.org/10.1016/j.aml.2021.107328
    [11] M. J. Ruan, C. Li, X. Y. Li, Codimension two 1: 1 strong resonance bifurcation in a discrete predator-prey model with Holling IV functional response, AIMS Math., 7 (2021), 3150–3168. https://doi.org/10.3934/math.2022174 doi: 10.3934/math.2022174
    [12] J. R. Beddington, Mutual interference between parasites or predators and its effect on searching efficiency, J. Anim. Ecol., (1975), 331–340. https://doi.org/10.2307/3866
    [13] D. DeAngelis, R. A. Goldstein, R. V. O'Neill, A model for tropic interaction, Ecology, 56 (1975) 881–892. https://doi.org/10.2037/1936298
    [14] P. H. Crowley, E. K. Martin, Functional responses and interference within and between year classes of a dragonfly population, J. N. Am. Benthol. Soc., 8 (1989), 211–221. https://doi.org/10.2307/1467324 doi: 10.2307/1467324
    [15] V. Ajraldi, M. Pittavino, E. Venturino, Modeling herd behavior in population systems, Nonlinear Anal. Real World Appl., 12 (2011), 2319–2338. https://doi.org/10.1016/j.nonrwa.2011.02.002 doi: 10.1016/j.nonrwa.2011.02.002
    [16] P. A Braza, Predator-prey dynamics with square root functional responses, Nonlinear Anal. Real World Appl., 13 (2012), 1837–1843. https://doi.org/10.1016/j.nonrwa.2011.12.014 doi: 10.1016/j.nonrwa.2011.12.014
    [17] S. M. Salman, A. M. Yousef, A. A. Elsadany, Stability, bifurcation analysis and chaos control of a discrete predator-prey system with square root functional response, Chaos Solit. Fractals, 93 (2016), 20–31. https://doi.org/10.1016/j.chaos.2016.09.020 doi: 10.1016/j.chaos.2016.09.020
    [18] A. Suleman, R. Ahmed, F. S. Alshammari, N. A Shah, Dynamic complexity of a slow-fast predator-prey model with herd behavior, AIMS Math., 8 (2023), 24446–24472. https://doi.org/10.3934/math.20231247 doi: 10.3934/math.20231247
    [19] M. X. He, Z. Li, Global dynamics of a Leslie-Gower predator-prey model with square root response function, Appl. Math. Lett., 140 (2023). https://doi.org/10.1016/j.aml.2022.108561
    [20] M. Lin, Y. Chai, X. Yang, Y. Wang, Spatiotemporal patterns induced by Hopf bifurcations in a homogeneous diffusive predator-prey system, Math. Probl. Eng., 2019 (2019). https://doi.org/10.1155/2019/3907453 doi: 10.1155/2019/3907453
    [21] P. Chakraborty, U. Ghosh, S. Sarkar, Stability and bifurcation analysis of a discrete prey-predator model with square root functional response and optimal harvesting, J. Biol. Syst., 28 (2020), 91–110. https://doi.org/10.1142/S0218339020500047 doi: 10.1142/S0218339020500047
    [22] M. G. Mortuja, M. K. Chaube, S. Kumar, Dynamic analysis of a predator-prey system with nonlinear prey harvesting and square root functional response, Chaos Solit. Fractals, 148 (2021). https://doi.org/10.1016/j.chaos.2021.111071
    [23] J. G. Tan, W. J. Wang, J. F. Feng, Transient dynamics analysis of a predator-prey system with square root functional responses and random perturbation, Mathematics, 10 (2022) 1–12. https://doi.org/10.3390/math10214087
    [24] X. Y. Meng, F. L Meng, Bifurcation analysis of a special delayed predator-prey model with herd behavior and prey harvesting, AIMS Math., 6 (2021), 5695–5719. https://doi.org/10.3934/math.2021336 doi: 10.3934/math.2021336
    [25] M. S. Rahman, S. Pramanik, E. Venturino, An ecoepidemic model with healthy prey herding and infected prey drifting away, Nonlinear Anal.-Model Control, 28 (2023), 326–364. https://doi.org/10.15388/namc.2023.28.31549 doi: 10.15388/namc.2023.28.31549
    [26] L. H. Dai, J. J. Wang, Y. G. Ni, B. Xu, Dynamical analysis of a new fractional-order predator-prey system with Holling type-Ⅲ functional, Adv. Differ. Equations, 2021 (2021), 1–13. https://doi.org/10.1186/s13662-020-03169-9 doi: 10.1186/s13662-020-03169-9
    [27] X. Y. Meng, H. F. Huo, X. B. Zhang, Stability and global Hopf bifurcation in a Leslie-Gower predator-prey model with stage structure for prey, J. Appl. Math. Comput., 60 (2019), 1–25. https://doi.org/10.1007/s12190-018-1201-0 doi: 10.1007/s12190-018-1201-0
    [28] X. Y. Zhou, Stability and Hopf bifurcation analysis of a stage-structured predator-prey model with delay, Axioms, 11 (2022). https://doi.org/10.3390/axioms11100575
    [29] X. Zhao, Z. J. Zeng, Stationary distribution and extinction of a stochastic ratio-dependent predator-prey system with stage structure for the predator, Physica A, 545 (2020). https://doi.org/10.1016/j.physa.2019.123310
    [30] X. Zhang, R. X. Shi, R. Z. Yang, Z. Z. Wei, Dynamical behaviors of a delayed prey-predator model with Beddington-DeAngelis functional response: stability and periodicity, Int. J. Bifurcation Chaos, 30 (2020). https://doi.org/10.1142/S0218127420502442
    [31] R. Z. Yang, D. Jin, W. L. Wang, A diffusive predator-prey model with generalist predator and time delay, AIMS Math., 7 (2022), 4574–4591. https://doi.org/10.3934/math.2022255 doi: 10.3934/math.2022255
    [32] X. W. Zhang, W. F. Huang, J. X. Ma, R. Z. Yang, Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and schooling behavior, Electron. Res. Arch., 30 (2022), 2510–2523. https://doi.org/10.3934/era.2022128 doi: 10.3934/era.2022128
    [33] M. Peng, R. Lin, Y. Chen, Z. D. Zhang, M. M. Khater, Qualitative analysis in a Beddington-DeAngelis type predator-prey model with two time delays, Symmetry-Basel, 14 (2022). https://doi.org/10.3390/sym14122535
    [34] Q. M. Zhang, D. Q. Jiang, Dynamics of stochastic predator-prey systems with continuous time delay, Chaos Solit. Fractals, 152 (2021). https://doi.org/10.1016/j.chaos.2021.111431
    [35] C. J. Xu, D. Mu, Y. L. Pan, C. Aouiti, L.Y. Yao, Exploring bifurcation in a fractional-order predator-prey system with mixed delays, J. Appl. Math. Comput., 13 (2023), 1119–1136. https://doi.org/10.11948/20210313 doi: 10.11948/20210313
    [36] Y. J. Xiang, Y. Q. Jiao, X. Wang, R. Z. Yang, Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator, Electron. Res. Arch., 31 (2023), 2120–2138. https://doi.org/10.3934/era.2023109 doi: 10.3934/era.2023109
    [37] Y. L. Song, J. J. Wei, Bifurcation analysis for Chen's system with delayed feedback and its application to control of chaos, Chaos Solit. Fractals, 22 (2004), 75–91. https://doi.org/10.1016/j.chaos.2003.12.075 doi: 10.1016/j.chaos.2003.12.075
    [38] B. D. Hassard, N. D. Kazarinoff, Y. H. Wan, Theory and Application of Hopf Bifurcation, Cambridge University Press, Cambridge, 1981.
  • This article has been cited by:

    1. Miao Peng, Rui Lin, Lei Huang, Zhengdi Zhang, Ammar Alsinai, Bifurcation Analysis of a Delayed Predator–Prey Model With Square Root Response Functions, 2024, 2024, 2314-4629, 10.1155/2024/8120170
    2. San-Xing Wu, Xin-You Meng, Hopf bifurcation analysis of a multiple delays stage-structure predator-prey model with refuge and cooperation, 2025, 33, 2688-1594, 995, 10.3934/era.2025045
    3. Alyaa Hussain Naser, Dahlia Khaled Bahlool, Dynamics of an Intraguild Predation Food Web Cooperation Model Under the Influence of Fear and Hunting, 2025, 13, 2079-3197, 128, 10.3390/computation13060128
  • 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(1256) PDF downloads(84) Cited by(3)

Figures and Tables

Figures(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog