Research article

Dynamics on semi-discrete Mackey-Glass model

  • Received: 18 December 2024 Revised: 17 January 2025 Accepted: 06 February 2025 Published: 14 February 2025
  • MSC : 39A28, 39A30, 39A60

  • Red blood cells play an extremely important role in human metabolism, and the study of hematopoietic models is of great significance in biology and medicine. A kind of semi-discrete hetmatopoietic model named Mackey-Glass Model was proposed and analyzed in this paper. The existences, stabilities, and local dynamics of the fixed points were discussed. By using bifurcation theory, we studied the Neimark-Sacker bifurcation, saddle-node bifurcation, and strong resonance of 1:4. The numerical simulations were presented to illustrate the results of theoretical analysis obtained in this paper, and complex dynamical behaviors were found such as invariant cycles, heteroclinic cycles and Li-Yorke chaos. In addition, a new periodic bubbling phenomenon was discovered in numerical simulations. These not only reflect the richer dynamical behaviors of the semi-discrete models, but also some reflect the complex metabolic characteristics of the hematopoietic system under environmental intervention.

    Citation: Yulong Li, Long Zhou, Fengjie Geng. Dynamics on semi-discrete Mackey-Glass model[J]. AIMS Mathematics, 2025, 10(2): 2771-2807. doi: 10.3934/math.2025130

    Related Papers:

    [1] A. Q. Khan, Ibraheem M. Alsulami . Discrete Leslie's model with bifurcations and control. AIMS Mathematics, 2023, 8(10): 22483-22506. doi: 10.3934/math.20231146
    [2] Ali Al Khabyah, Rizwan Ahmed, Muhammad Saeed Akram, Shehraz Akhtar . Stability, bifurcation, and chaos control in a discrete predator-prey model with strong Allee effect. AIMS Mathematics, 2023, 8(4): 8060-8081. doi: 10.3934/math.2023408
    [3] Abdul Qadeer Khan, Ayesha Yaqoob, Ateq Alsaadi . Discrete Hepatitis C virus model with local dynamics, chaos and bifurcations. AIMS Mathematics, 2024, 9(10): 28643-28670. doi: 10.3934/math.20241390
    [4] Figen Kangalgil, Seval Isșık . Effect of immigration in a predator-prey system: Stability, bifurcation and chaos. AIMS Mathematics, 2022, 7(8): 14354-14375. doi: 10.3934/math.2022791
    [5] Xiongxiong Du, Xiaoling Han, Ceyu Lei . Dynamics of a nonlinear discrete predator-prey system with fear effect. AIMS Mathematics, 2023, 8(10): 23953-23973. doi: 10.3934/math.20231221
    [6] Xijuan Liu, Yun Liu . Stability and bifurcation analysis of a discrete-time host-parasitoid model with Holling III functional response. AIMS Mathematics, 2023, 8(10): 22675-22692. doi: 10.3934/math.20231154
    [7] Xijuan Liu, Peng Liu, Yun Liu . The existence of codimension-two bifurcations in a discrete-time SIR epidemic model. AIMS Mathematics, 2022, 7(3): 3360-3378. doi: 10.3934/math.2022187
    [8] Abdul Qadeer Khan, Ayesha Yaqoob, Ateq Alsaadi . Neimark-Sacker bifurcation, chaos, and local stability of a discrete Hepatitis C virus model. AIMS Mathematics, 2024, 9(11): 31985-32013. doi: 10.3934/math.20241537
    [9] Yao Shi, Zhenyu Wang . Bifurcation analysis and chaos control of a discrete fractional-order Leslie-Gower model with fear factor. AIMS Mathematics, 2024, 9(11): 30298-30319. doi: 10.3934/math.20241462
    [10] Ahmad Suleman, Rizwan Ahmed, Fehaid Salem Alshammari, Nehad Ali Shah . Dynamic complexity of a slow-fast predator-prey model with herd behavior. AIMS Mathematics, 2023, 8(10): 24446-24472. doi: 10.3934/math.20231247
  • Red blood cells play an extremely important role in human metabolism, and the study of hematopoietic models is of great significance in biology and medicine. A kind of semi-discrete hetmatopoietic model named Mackey-Glass Model was proposed and analyzed in this paper. The existences, stabilities, and local dynamics of the fixed points were discussed. By using bifurcation theory, we studied the Neimark-Sacker bifurcation, saddle-node bifurcation, and strong resonance of 1:4. The numerical simulations were presented to illustrate the results of theoretical analysis obtained in this paper, and complex dynamical behaviors were found such as invariant cycles, heteroclinic cycles and Li-Yorke chaos. In addition, a new periodic bubbling phenomenon was discovered in numerical simulations. These not only reflect the richer dynamical behaviors of the semi-discrete models, but also some reflect the complex metabolic characteristics of the hematopoietic system under environmental intervention.



    Red blood cells (RBCs) are derived from stem cells in the bone marrow, and their main function is to distribute oxygen in the human body. In a healthy human body, RBCs tend to a stable state, and if their numbers suddenly decrease, it will stimulate the stem cells in the bone marrow to produce significant excitement, thereby generating new cells that return to a stable state. There must be a period of time between the moment RBCs decline and the moment they recover. Under normal circumstances, the process of forming new cells takes approximately 5 to 7 days [1].

    Currently, some diseases are known to cause the number of RBCs to gradually oscillate, which is represented by a periodic function. If RBCs are insufficient for a period of time, the human body will be unable to recover and die. To check the health status of the body, it can be achieved by detecting RBCs. In 1976, Lasota et al. [24] proposed the Lasota hematopoietic model, which perfectly described the changes in RBCs in the human body. Later, Mackey and Glass [1] combined the respiratory system to further characterize RBCs in the blood.

    In 1976, Mackey and Glass introduced the following hetmatopoietic system

    ξ(t)=δξ(t)+pξk(tτ)1+ξm(tτ), (1.1)

    where δ,p(0,+), k,mN. ξ(t) denotes the population of mature cells at time t, and the cells disappear from cycle with the velocity δ, and the delay τ>0 denotes the time of generating new RBCs.

    EI Sheikh et al. [4] provided some sufficient and necessary conditions for the oscillation of all positive solutions when k=0, while Gopalsamy [17] provided some sufficient and necessary conditions for the oscillation of all positive solutions and global attractiveness when k=0 or 1. Through numerical simulations of the system, Hela and Sternberg [18] found that the degree of chaos increases with the increase of time delay τ. System (1.1) with k=1 attracted many authors' attentions [3,12,28,29,36]. While for the case k=0 and k>1, there were less researchers concerned. Therefore, the paper deals with system (1.1) with k=0 and k>1.

    Hematopoietic models are usually divided into two categories: Continuous models described by differential equations, and discrete models obtained by discretizing continuous systems. These mathematical methods include the Euler difference method, Adams Bashforth method [49], piecewise-constant approximation method [22], the conformable derivatives [21], non-standard finite difference scheme (NSFDS) of Mickens-type [32], and semi discrete method [17,38,39]. The main methods are analyzed by using the Euler difference method and semi discrete method.

    In recent years, the advantages of discrete models have become increasingly apparent in various fields. To begin in scientific research, data is collected over a period of time, such as year, month, day, etc. Therefore, considering a discrete hematopoietic model is more in line with practical significance [38,39].

    Moreover, due to the fact that medical tests using RBC counting are often conducted at regular intervals, and changes in the number of RBCs are difficult to observe continuously over a long period of time, the discrete case is considered to be more appropriate to describle the real world sometimes, and it is obtained in the literatures that discrete systems have richer dynamical behaviors. Various of discrete-time models have been studied in detail recently [9,19,25,31,41,48,50]. There are a series of results on the qualitativeness and bifurcations of discrete-time models like flip bifurcation and Neimark-Sacker bifurcation [8,10,26,31,34,41,47].

    In 2002, Tamas and Gabor [13] proposed a semi-discretization method to study the delayed system, which is a simple and effective way to deal with the delayed term. Then, by using this method some new semi-discrete models have been proposed and analyzed. For example, in [39], the authors proposed a semi-discrete hematopoiesis model and discussed its dynamical behavior which included the stabilities of the fixed points and the Neimark-Sacker bifurcation. For other studies for semi-discrete models, one may see [9,20] and the references therein. Moreover, Yao and Li [40] provided bifurcation difference induced by different discrete methods in a discrete predator-prey model, which shows that it is meaningful to investigate the model with different discretization methods.

    When analyzing the semi-discrete Mackey-Glass model, we need to consider the situation where multiple parameters change simultaneously. Complicated bifurcations likely occur when more than one systemic parameter is varied at the same time [6]. Codimension-2 bifurcations in the discrete-time systems will occur when the dimension of the center manifold is changed by the approximation of extra multipliers to the unit circle, or some of the nondegeneracy conditions for the one-parameter bifurcations are violated [44,45]. Kuznetzov derived explicit formulas for the normal form coefficients to verify the nondegeneracy of codim-2 bifurcations of fixed points with 1 or 2 critical multipliers [45]. Li and He considered 1:2 and 1:4 resonances in a discrete-time Hindmarsh-Rose model [6], and Wu and Zhao derived codimension-two bifurcations with 1:2 resonance in a discrete-time predator-prey model [11]. Rana and Uddin analytically showed a flip-Neimark Sacker bifurcation of the discretized L¨u system [35]. More types of complicated bifurcations of discrete-time or continuous-time systems have been studied in [2,5,6,7,11,14,15,27,30,33,35,42,43,44,45,46]. Moreover, from the perspective of numerical convergence, discrete systems can reflect the dynamical behaviors of the original continuous system and generate richer dynamical phenomena [9,25,38,39,41,48,50].

    In this paper, the semi-discrete hetmatopoietic Mackey-Glass model is analyzed. We discuss the dynamical behaviors of system (1.1), including the existence of fixed points, local stability, and codim 1 and 2 bifurcations. We have found that with the parameters varying, there exist invariant curves when the system is undergoing Neimark-Sacker bifurcation, period-4 saddle points, and heteroclinic cycle composed by the separatrices of them when undergoing 1:4 resonance. The existence of Li-Yorke chaos is proved in this case. In the case of 2km, the system undergoes saddle-node bifurcation. In order to illustrate the correctness of theoretical analysis, numerical simulations are provided and some other complex dynamical behaviors are reviewed.

    The rest of paper is organized as follows: In Section 2, we discussed the existences and stabilities of fixed points of the M-G model with k=0 and theoretically verified the existence of Neimark-Sacker bifurcations and 1:4 resonances. In Section 3, these theoretical results are supported by the numerical simulations. Then, we consider the case of 1<km in Section 4. We analyzed the existences and stabilities of fixed points and verified the existence of saddle-node bifurcations, Neimark-Sacker bifurcations, and 1:4 resonances, which are more complex than the case in Section 2. In Section 5, the numerical simulations are presented to support the theoretical results and show the impact of negative feedback, mixed feedback, and positive feedback depending on the parameter k. Finally, a brief conclusion is organized in Section 6 which provides more directions for future works.

    In this section we discuss system (1.1) with k=0 as follows:

    ξ(t)=δξ(t)+p1+ξm(tτ), (2.1)

    where δ,τ,p(0,+) and mN. Next, we establish the semi-discrete model for system (2.1). For the sake of simplicity, intoducing the transformations s=th and N(t)=N(sτ)=η(s), (2.1) is then changed to

    dηds=δτη(s)+pτ1+ηm(s1),s0. (2.2)

    For simplicity, resetting δ=δτ, p=pτ in the above equation, one gets

    dηds=δη(s)+p1+ηm(s1),s0. (2.3)

    Through this transformation, the reduction of the time-delay parameter τ in the system is completed, which makes the problem simple.

    Let [s] denote the integer that is not bigger than s, and consider the semi-discrete model of (2.3)

    dηds=δη([s])+p1+ηm([s1]),s0. (2.4)

    It is easy to see that the right side of the Eq (2.4) is constant on the interval [n,n+1). Obviously, the following conclusions about Eq (2.4) are true.

    Lemma 2.1. The solution η(s) of Eq (2.4) satisfies

    1. η(s) is continuous on [0,+),

    2. dηds exists on +s=0(s,s+1),

    3. Equation (2.4) is true on every interval [m,m+1) for m=0,1,2,.

    For any s[n,n+1), we have [s]=n. Substituting it into Eq (2.4) we get

    dηds=δη([n])+p1+ηm([n1]),s0, (2.5)

    and then integrate the Eq (2.5) from n to s for any s[n,n+1) where n{0,1,2,3,}. The following difference equation can be found for s[n,n+1)

    η(s)η(n)=(δη(n)+p1+ηm([n1]))(sn).

    Let s(n+1) in the above equation, and then the semi-discrete system with no delay for system (2.1) will be obtained as follows:

    η(n+1)=(1δ)η(n)+p1+ηm(n1),

    which is a second-order difference equation. Take the normal transformation

    {xn=η(n1),yn=η(n),

    and we may achieve the discrete planar dynmaical system of (2.1) as follows:

    {xn+1=yn,yn+1=(1δ)yn+p1+xmn, (2.6)

    where δ>0, p>0, x0,y0(0,+). We shall discuss the dynamics on system (2.6) in detail.

    At the beginning of the discussion, we consider the existence of the fixed point of (2.1).

    Theorem 2.1. The conclusions of the fixed point for system (2.6) are true:

    (i) there is a unique fixed point E(x,y) of system (2.6) satisfying

    {x=y,δx=p1+xm;

    (ii) x is strictly increasing at p.

    Proof. It is easy to see that the fixed point of system (2.6) satisfies the following equations

    {x=y,δx=p1+xm.

    Next, we show the existence and uniqueness of the fixed point. Set

    F(x,p)=δxm+1+δxp,

    so we may study its positive zero points.

    Notice that

    Fx(x,p)=δ(m+1)xm+δ>0,xDx={x:x0},

    and

    F(0+,p)=p<0,F(pδ,p)=pm+1δm>0.

    Hence, there is a unique zero point of F(x,p) in Dx, and 0<x<pδ.

    (ii) Since Fx(x,p)>0, the implicit function theorem shows that there exists a function x=f(p) in the field Dxp={(x,p)|0<x<pδ,p>0} and

    dxdp=Fp(x,p)Fx(x,p)=1δ(m+1)xm+δ>0,

    which means x is strictly increasing in p. Next, we deal with the stability of the fixed point of system (2.6). The Jacobian matrix of (2.6) at E is

    JE=(01pmxm1(1+xm)21δ)=(01δmxm1+xm1δ),

    then Tr(JE)=1δ, Det(JE)=δmxm1+xm, and the corresponding characteristic equation is

    F(λ)=λ2Tr(JE)λ+Det(JE),

    hence,

    F(1)=δ+Det(JE)>0,F(1)=22δ+F(1)>0.

    Lemma 2.2. [39] Consider F(λ)=λ2+Bλ+C with two constant real parameters B and C. Suppose λ1 and λ2 are two zero points of F(λ). Then, the zero points of the equation are entirely determined by both F(1) and the parameters.

    Case I. If F(1)>0, then

    (1) |λ1|<1 and |λ2|<1 if, and only if, F(1)>0 and C<1.

    (2) |λ1|>1 and |λ2|>1 if, and only if, F(1)>0 and C>1.

    (3) |λ1|<1 and |λ2|>1 if, and only if, F(1)<0.

    (4) There exist λ1=1 and λ21 if, and only if, F(1)=0 and B2.

    (5) λ1=λ2=1 if, and only if, F(1)=0 and B=2.

    (6) There exist a pair of conjugate complex roots with |λ1|=|λ2|=1 if, and only if, 2<B<2 and C=1.

    Case II. If F(1)=0, i.e., λ1=1, then |λ2|>1(resp.,<1) if, and only if, |C|>1 (resp.,<1).

    Case III. If F(1)<0, then F(λ)=0 has one root λ1(1,+). Then, the following statements about the other root λ2 hold.

    (1) λ2<(=)1 if, and only if, F(1)<(=)0;

    (2) 1<λ2<1 if, and only if, F(1)>0.

    Therefore, the following conclusions can be obtained.

    Theorem 2.2. Let pNS=δ2mδm1(1δm1)1m, then

    (i)the fixed point E of system (2.6) is a sink for δm1 or δm>1 and p<pNS;

    (ii) the fixed point E of system (2.6) is a source for δm>1 and p>pNS;

    (iii) system (2.6) may undergo Neimark-Sacker bifurcation at the fixed point E for δm>1 and p=pNS.

    Proof. Notice that Det(JE)=δmxm1+xm; it is easy to derive that Det(JE)<1 holds for δm1.

    Meanwhile, for δm>1 and p=pNS, we have x=(1δm1)1m, which leads to Det(JE)=1. Since F(1)>0 and F(1)>0, by Lemma 1, one may know that λ1,2 are a pair of conjugate complex roots; moreover, |λ1,2|=1. Hence, system (2.6) probably undergoes Neimark-Sacker bifurcation at E (we will give the theoretical proof in the next subsection).

    Furthermore, we may see that Det(JE) is strictly increasing in x in the field Dx. Combining the results in Theorem 2.1 that x is strictly increasing in p, which in turn leads to the fact Det(JE) is strictly increasing in p, so for δm>1 and p<pNS, one gets Det(JE)<1 and |λ1,2|<1, which implies the fixed point E of system (2.6) is a sink.

    Whereas for δm>1 and p>pNS, we have Det(JE)>1 and |λ1,2|>1, which shows that the fixed point E of system (2.6) is a source.

    The proof is then completed.

    In this sbusection, we shall show the existence of Neimark-Sacker bifurcation and the stability of the bifurcated invariant curve.

    To start, we choose p as a bifurcation parameter, giving a perturbation p of parameter pNS, then we obtain the following pertubed system:

    {xy,y(1δ)y+pNS+p1+xm, (2.7)

    where |p|1.

    Next, taking u=xx,v=yy, then the fixed point E can be moved to the origin O(0,0) and system (2.7) becomes

    {uv,v(1δ)v+pNS+p1+(u+x)mδy. (2.8)

    The characteristic function of the linearization of system (2.8) at its fixed point O is:

    F(λ)=λ2a(p)λ+b(p),

    where a(p)=1δ, b(p)=m(pNS+p)xm1(1+xm)2=δmxm(p)1+xm(p). It is easy to see

    λ1,2|p=0=12[(1δ)±i3+2δδ2]

    is a pair of conjugate complex roots of F(λ)=0. As we know, to guarantee the existence of Neimark-Sacker bifurcation, the following conditions must be fufilled:

    (C.1)d|λ1,2|dp|p=00;(C.2)λk1,21,k=1,2,3,4.

    Notice that a(p)|p=0=1δ and b(p)=m(pNS+p)xm1(1+xm)2=δmxm(p)1+xm(p), so

    λ1,2|p=0=12[(1δ)±i3+2δδ2],

    which means the condition (C.2) is true. Furthermore, |λ1,2|=b(p)=Det(JE),

    d|λ1,2|dp|p=0=12Det(JE)dDet(JE)dxdxdpdpdp|p=0>0,

    which guarantees that the condition (C.1) holds. Therefore, system (2.6) undergoes Neimark-Sacker Bifrucation as p varies in the neighborhood of pNS.

    In the following, we shall discuss the stability of the invariant curve by three steps.

    The first step: Expand the right side of system (2.8) as Taylor series at (u,v)=(0,0)

    {ua10u+a01v+a20u2+a11uv+a02v2+a30u3+a12uv2+a21u2v+a03v3+O(ρ4),vb10u+b01v+b20u2+b11uv+b02v2+b30u3+b12uv2+b21u2v+b03v3+O(ρ4), (2.9)

    where ρ=u2+v2,

    a01=1, a10=a20=a11=a02=a30=a21=a12=a03=0,b10=1, b01=1δ, b11=b02=b12=b21=b03=0,b20=2+δδm2δ(1δm1)1m,b30=δ2(m2+3m2)+6δ(m1)66δ2(1δm1)2m.

    The second step: Take the matrix

    T=(013+2δδ221δ2),

    and it is not difficult to see

    T1=(δ13+2δδ223+2δδ210),

    introduce the invertable transformation (u,v)T=T(X,Y)T, and the normal form of system (2.9) is arrived as:

    {X1δ2X3+2δδ22Y+F(X,Y)+O(ρ4),Y3+2δδ22X+1δ2Y+G(X,Y)+O(ρ4), (2.10)

    where F(X,Y)=2b203+2δδ2Y2+2b303+2δδ2Y3, G(X,Y)=0,ρ=|X|2+|Y|2. By computation, we get

    FXXX|(0,0)=FXXY|(0,0)=FXYY|(0,0)=0,FYY|(0,0)=2γ03+2δδ2,FYYY|(0,0)=2γ203+2δδ2,FXX|(0,0)=FXY|(0,0)=GXX|(0,0)=GXY|(0,0)=GYY|(0,0)=0,GXXX|(0,0)=GXXY|(0,0)=GXYY|0,0=GYYY|(0,0)=0. (2.11)

    The last step is to compute the first Lyapunuov coefficient by

    a=Re[(12λ)¯λ21λL11L20]12|L11|2|L02|2+Re(¯λL21), (2.12)

    where

    L20=18[(FXXFYY+2GXY)+i(GXXGYY2FXY)]=b2023+2δδ2,L11=14[(FXX+FYY)+i(GXX+GXY)]=b203+2δδ2,L02=18[(FXXFYY2GXY)+i(GXXGYY+2FXY)]=b2023+2δδ2,L21=116[(FXXX+FXYY+GXXY+GYYY)+i(GXXX+GXYYFXXYFYYY)]=3b3043+2δδ2i.

    Therefore, by computation and simplification, one may achieve

    a=[2b220(δ+2)+3b30(δ+1)]8(δ+1)=(δ3mδ3+δ2m22δ22δm+2)16δ2(δ+1)(1δm1)2m.

    Next, we discuss the sign of a. Denote h(m)=(δ3mδ3+δ2m22δ22δm+2),

    h(m)=δ(δ2+2δm2).

    Obviously, h(m)<0 for δm>1. Therefore, h(m)h(2)<0 for δm>1 (in fact, δm>1 and mN+ means m2). This leads to a<0 directly. Hence, the following theorem can be obtained.

    Theorem 2.3. For δm>1, the first Lyapunov coefficient a<0 is always true, which means when p is varying in the sufficient small neighborhood of PNS, system (2.6) undergoes Neimark-Sacker bifurcation at the fixed point E; moreover, the bifurcated invariant curve is stable.

    Now we choose p and δ as the bifurcation parameters. Since λ1,2=±i when p=pSN and δ=δ4 =1, give a perturbation p and δ of parameter pNS and δ4=1. In order to ensure that the value of pNS is meaningful, m is not equal to 1 in this case. Then, we obtain the following pertubed system:

    {xy,y(1δ4δ)y+pNS+p1+xm, (2.13)

    where p=pNS+p, δ=δ4+δ, |p|,|δ|1.

    Next, taking u=xx,v=yy, then the fixed point E can be moved to the origin O(0,0) and system (2.13) becomes:

    {uv,v(1δ4δ)v+pNS+p1+(u+x)mδy. (2.14)

    The Jacobian of system (2.14) at its fixed point O is:

    J(0,0)=(01m(pNS+p)xm1(1+xm)21δ4δ).

    System (2.14) is expanded as the Taylor series at (u,v)=(0,0)

    {uv,vb10u+b01v+b20u2+b11uv+b02v2+b30u3+b12uv2+b21u2v+b03v3+O(ρ4), (2.15)

    where ρ=u2+v2,

    b10=m(pNS+p)xm1(1+xm)2, b01=1δ4δ,b11=b02=b12=b21=b03=0,b20=12m(pNS+p)xm2(1+xm)3(1m+xm+mxm),b30=16m(pNS+p)xm3(1+xm)4(23m+m2+4xm4m2xm+2x2m+3mx2m+m2x2m).

    If p=δ=0, the Jacobian will be equivalent to

    Λ0=J(0,0)|p=δ=0=(0110).

    The characteristic value of the linearization of system (2.14) at its fixed point O is

    λ1,2|p=δ=0=±i,

    and the eigenvectors are

    q1=(i,1)T,q2=(i,1)T.

    The map (2.15) can be written as

    (uv)(0110)(uv)+(0b20u2+b30v2+O(ρ4)), (2.16)

    where

    b20=3m2(1m1)1m,b30=(m2+3m2)+6(m1)66(1m1)2m.

    To analyze the bifurcation, we take any vector (u,v)TR2 in the form (u,v)T=zq1+ˉzq2, where z is a complex variable which transforms the map (2.16) to the complex form

    ziz+G(z,ˉz), (2.17)

    where

    G(z,ˉz)=b20(iz+iˉz)2+b30(iz+iˉz)32+O(ρ4)=k+l21k!l!gklzkˉzl,

    with

    g20=g02=b20,g11=b20,g30=g12=3ib30,g03=g21=3ib30. (2.18)

    Take an invertible parameter-dependent change of complex coordinate

    z=w+h202w2+h11wˉw+h022ˉw2,

    with

    h20=g20λ21λ1=12g20(i1),h11=g11|λ1|2λ1=12g11(i+1),h02=g02¯λ12λ1=12g02(i1).

    Then, eliminate all the quadratic terms in (2.17), and transform it to

    wiw+k+l31k!l!ρklzkˉzl, (2.19)

    with

    ρ30=3(1i)h20g203(iˉg02+(1+i)ˉh02)h11+3(1i)h220+3g11ˉh02+g30,ρ21=(iˉg02+(1+i)ˉh02)h02+(12i)g11h20+2g11ˉh11+g02ˉh02+g21+((1+3i)h202iˉg112(1i)ˉh11+(2+i)g20)h11,ρ12=(2(1+i)h11+ˉh20)g11+g20h02+((1i)h20+2iˉg11+2(1i)ˉh11)h022(1+i)h211(iˉg20+(1+i)ˉh20)h11+g12ih20g02+2g02ˉh11,ρ03=g03+3g11h02+3((i1)h11+iˉg20+(1+i)ˉh20)h02+3g02ˉh20+3ih11g02.

    Then, by setting

    h30=g30λ31λ1=g30i3i,h21=g21λ1λ1|λ1|2=g21ii|i|2,
    h12=g12¯λ1|λ1|2¯λ1=g12i|i|2+i,h03=g03¯λ13λ1=g03ˉi3i,

    we take another invertible parameter-dependent change of complex coordinate

    w=ζ+16h30ζ3+12h21ζ2ˉζ+12h12ζˉζ2+16h03ˉζ3,

    which changes the map (2.19) into

    ζΓ(ζ)=iζ+Cζ|ζ|2+Dˉζ3+O(|ζ|4), (2.20)

    where gkl are given by (2.18) and

    C=1+3i4g20g11+1i2|g11|21+i4|g02|2+12g21=32b2203i2b30=18[3(3+m)2+2i(7+m)(2+m)](11+m)2/m,D=i14g11g021+i4g02ˉg20+16g03=i2b220+i2b30=i24(5+m)(11+5m)(11+m)2/m.

    When m5 (D0), the fourth iterate of Γ allows for approximation by a complex flow, and the bifurcation of 1:4 resonance will be determined by

    ˜A=C|D|=9(m3)26i(m7)(m2)(m5)(5m11),

    and

    [9(m3)2]2+[6(m7)(m2)]2[(m5)(5m11)]2>0.

    We have

    |˜A|>1.

    By mN+, we have

    Re˜A=9(m3)2(m5)(5m11)0,form1,2,3,5,7,
    Im˜A1+(Re˜A)21(Re˜A)2,form1,2,3,5,7.

    When dealing with strong resonances, we will repeatedly use the approximation of maps near their fixed points by shifts along the orbits of certain systems of autonomous ordinary differential equations [6,44].

    Lemma 2.3. [44] The fourth iterate of the map (2.20) can be represented in the form

    Γ4(ζ)=φ1ζ+O(|ζ|4), (2.21)

    where φt is the flow of a planar system

    ˙ζ=4iCζ|ζ|24iDˉζ3. (2.22)

    Theorem 2.4. If m1,2,3,5,7 (means that pNS is meaningful, Re˜A0 and Im˜A0), and the parameters (p,δ) vary in a small neighborhood of (pSN,1), the system (2.6) undergoes a bifurcation of 1:2 resonance around the fixed point E, and admits the following bifurcation curves:

    There is a Neimark-Sacker bifurcation curve at the fixed point E of the system (2.6).

    There exist four saddle fixed points Sk when |˜A|>1, which are the corresponding period-4 points of (2.21). As the parameters (p,δ) in (2.6) varying in the neighborhood of (pNS,1), these nontrivial periodic points appear or disappear in the neighborhood of E.

    There is a square heteroclinic cycle around E composed by the separatrices of Sk, which is stable from the inside.

    Select δ=0.2, m=10, and the initial point E0(1.5,1.5). By the theoretical analysis in Section 2.1, one has pNS=0.4, and at this case the fixed point of system (2.6) is E(1,1), a=0.5687. See Figures 13 by numerical simulation.

    Figure 1.  The Neimark-Sacker bifurcation as p is varying when m=10, δ=0.2, E0 =(1.5,1.5).
    Figure 2.  The Neimark-Sacker bifurcation as p is varying when m=10, δ=0.2, E0 =(1.5,1.5).
    Figure 3.  The MLE as γ varying for p=2, δ=0.5.

    Figures 1(a)(c) show that the fixed point E is a sink when p<0.4; meanwhile for p=0.41, a stable invariant curve occurs, see the Figures 1(d), (f); it can be also shown in Figure 2 which is consistent with the results in Theorem 2.3.

    The Figure 3 shows the maximum Lyapunov exponent (MLE) varying with the parameter p, which shows the MLE is always less than zero when p<0.4; meanwhile for γ>0.4, the MLE will fluctuate below 0, which also means the oribit from the initial (1.5,1.5) will be periodic. Also the MLE is not greater than 0, which implies the chaos does not occour for system (2.6) at this time.

    Taking m=4 as an example, the same applies to other situations.

    Let δ vary form the neighborhood of δ4=1, while pNS=1.0131, x(pNS)=111/4 and let p vary in [pNS0.02,pNS+0.2]. Thus, it follows from Theorem 4.6 that the fixed point E of system (2.6) is a 1:4 resonance point.

    Figures 4(a), (b) show that there exists the invariant cycle around the fixed point when δ varies near δ=1. A nondegenerate Neimark-Sacker bifurcation curve when δ=0.8 is illustrated in Figure 4(c), and four period points when δ=0.9 and δ=1.01 are given in Figures 4(d), (e). Figures 4(b), (d) show that a complex heteroclinic curve will occur near the Neimark-Sacker bifurcation curve or period points when δ varies in the neighborhood of δ=0.9 or 1.01. A possible chaotic phenomenon is shown in Figure 4f, and in fact, we have verified it through numerical simulations (see Figures 6(a) and 7(b)).

    Figure 4.  The 1:2 resonance as (p,δ) varying when k=0,m=4.

    Figure 5 shows the phase portraits of system (2.6) corresponding to Figures 4(a)(e). Since the system undergoes Neimark-Sacker bifurcation, the fixed point E changes from local asymptotically stable to unstable and there occurs a stable closed invariant curve around it, which is shown in Figures 5(a), (b)

    Figure 5.  Phase portraits corresponding to Figure 4(a)(e): Supercritical Neimark-Sacker bifurcation curve (5(b), (c)) and 1:4 resonance (5(d)–(f)).

    From Figures 5(d)(f), we observe that a heteroclinic curve consists of the separatrices of four saddles; specially, Figure 5f shows the collision between the saddle points and the invariant circle. By selecting δ=1.685, p=2.01311 (see Figures 6(a), (b)), we calculate that the max Lyapunov exponent is L=0.0715457>0, and the system has 3-period points, which implies chaos [37].

    Figure 6.  Phase portraits corresponding to Li-Yorke chaos (6(a), (b)).

    The 3-dimensional bifurcation diagrams of map (2.6) in (p,δ,y) and the maximium Lyapunov exponents corresponding to δ and p are shown in Figure 7(a). The maximum Lyapunov exponents corresponding to p and δ are calculated and plotted in Figure 7(b) that confirms the existence of the period orbits and chaotic regions near the 1:4 resonance point E in the parametric space.

    Figure 7.  The bifurcation of 1:4 resonance of system (2.6) when k=0.

    In this section, we study the Makey-Glass model with 1<km.

    We shall see in this section, that the model with 1<km exhibits more complex dynamics (saddle-node bifurcation and chaos) than the model with k=0 and k=1.

    ξ(t)=δξ(t)+pξk(tτ)1+ξm(tτ), (4.1)

    where δ,τ,p(0,+), m,kN+,1<km. With the same way one can derive the semi-discrete system of system (4.1):

    η(n+1)=(1δ)η(n)+pηk(n1)1+ηm(n1). (4.2)

    Taking the transfomations {xn=η(n1),yn=η(n), we can obtain the planar dynamical system

    {xn+1=yn,yn+1=(1δ)yn+pxkn1+xmn. (4.3)

    In the following, we discuss the dynamics of system (4.3). It is shown that the origin O(0,0) is always the fixed point of system (4.3), which is a trivial fixed point. With the increasing of parameter p, the number and the stability of the fixed point will change.

    When the parameter p meets the first critical value p0, there will exist a positive fixed point E, and the system may undergo saddle-node bifurcation at E for p=p0. Meanwhile, for p>p0, there are two fixed points E1,2 bifurcated from E.

    Suppose E(x,y) is the fixed point of system (4.3), and it must satisfy the following equations

    {x=y,δy=pxk1+xm.

    Denote f(x)=δxm+1pxk+δx. Obvilusly if ˉx is the root of f(x)=0, then (ˉx,ˉx) is also the fixed point of system (4.3). It is easy to see that x=0 is always a root of f(x)=0, then E0(0,0) is the trivial fixed point of (4.3).

    To find the nonzero roots of f(x)=0, take g(x)=f(x)x=δxmpxk1+δ, and we may discuss the positive solutions of g(x)=0. Notice that

    g(x)=mδxm1p(k1)xk2=xk2[mδxmk+1p(k1)].

    Since xmk+1 is strictly increasing at x, g(x) will get to its minimum value at x=mk+1p(k1)mδx0. Let ρ1=pδ,ρ2=k1m, then we have ρ1>1,ρ2<1; hence, x0=(ρ1ρ2)1mk+1, and the minimum value of g(x) is

    g(x0)=δxm0+δpxk10=δ[(ρ1ρ2)mmk+1+1ρ1(ρ1ρ2)k1mk+1]=δ[(ρ1ρ2)11ρ2+1ρ1(ρ1ρ2)ρ21ρ2].

    Theorem 4.1. Let pSN=δρρ22(1ρ2)ρ21, x0=(ρ1ρ2)1mk+1, then the following conclusions about the fixed points of system (4.3) are true:

    (i) when p<pSN, there is only one trivial fixed point E0(0,0) of system (4.3);

    (ii) when p=pSN, there are two fixed points of system (4.3), which are the trivial fixed point E0(0,0) and the positive fixed point E(x,y), x=y=x0;

    (iii) when p>pSN, there are three fixed points of system (4.3), which are E0(0,0), E1(x1,y1),E2(x2,y2), where xi=yi statisfies δxm+1ipxki+δxi=0, i=1,2.

    Proof. It is not difficult to see that when p=pSN, ρ1=ρρ22(1ρ2)ρ21, the minimum value of g(x) satisfies

    g(x0)=δ[(ρ1ρ2)11ρ2+1ρ1(ρ1ρ2)ρ21ρ2]=δ[ρ21ρ2+1ρρ22(1ρ2)ρ21(ρ21ρ2)ρ2]=δ[11ρ211ρ2]=0,

    which shows there is only one solution x0 of g(x)=0.

    Whereas for p<pSN, we have

    g(x0)p=11ρ2ρρ21ρ21(ρ11ρ22ρρ21ρ22)<0,

    that is to say, g(x0) is strictly decreasing at p for ρ1>1, 0<ρ2<1. Hence, g(x0)>0 for p<pSN, which means there is no positive solution of g(x)=0; as to p>pSN, we have g(x0)<0, combined with the fact

    limx0+g(x)=δ>0,limx+g(x)=+,

    which means there are only two solutions x1,x2 for g(x)=0 and 0<x1<x0<x2.

    Notice that E0(0,0) is always the fixed point of system (4.3), hence, we can finish the proof.

    It is easy to see from the above theorem that system (4.3) probably undergoes saddle-node bifurcation. In the next subsection we shall verify the existence of saddle-node bifurcation theoretically.

    Theorem 4.2. Let p>pSN, E1(x1,y1), and E2(x2,y2) be the fixed points of system (4.3), then x1,x2, and the parameter p satisfy that xi is strictly increasing with respect to p in (pSN,+), i=1,2.

    Proof. Denote D2={(xi,p) | xi(x0,+),p(pSN,+)}, and xi satisfies g(xi,p)=0. Notice that gp(xi,p)=xk1i, gxi(xi,p)=δmxm1ip(k1)xk2i are continuous in D2, and

    gxi(xi,p)=δmxm1ip(k1)xk2i=δmxk21(xmk+1iρ1ρ2)>0.

    The implicit theorem shows that there exists a function xi=xi(p) satisfying g(xi(p),p)=0 when p>pSN, and

    dxi(p)dp=xiδm(xmk+1iρ1ρ2)>0,

    then the proof is finished.

    Next, we study the stabilities of the fixed points.

    To begin, we discuss the stability of the fixed point E0(0,0).

    The Jocabian Matrix of system (4.3) at E0(0,0) is

    JO=(01p(km)xm+k1+pkxk1(1+xm)21δ)|O=(0101δ),

    and the corresponding characteristic equation is

    F(λ)=λ2(1δ)λ=0,

    hence, λ1=0,λ2=1δ<1, which means the fixed point is a sink.

    Next, we study the stability of E(x,y) for p=pSN. The Jacobian matrix of (4.3) at E(x,y) is

    JE=(01p(km)xm+k1+pkxk1(1+xm)21δ)|E=(01δ(km)xm+δk1+xm1δ),

    so Tr(JE)=1δ, Det(JE)=δ(km)xm+δk1+xm, and the corresponding characteristic equation is

    F(λ)=λ2Tr(JE)λ+Det(JE).

    It is not difficult to see

    F(λ)=λ2Tr(JE)λ+Det(JE).
    F(1)=δ+Det(JE)=δδ(km)xm+δk1+xm=δ(k1)1+xm[(1ρ21)xm1]=δ(k1)1+xm[1ρ2ρ2ρ21ρ21]=0,

    and C=Det(JE)=δ, which means |C|<1; Lemma 2.2 implies that λ1=1 and |λ2|<1, that is to say, E is a nonhyperbolic fixed point.

    Last, we discuss the stabilities of the fixed points E1(x1,y1) and E2(x2,y2) for p>pSN.

    JE1=(01δ(km)xm1+δk1+xm11δ),

    and the corresponding characteristic equation is

    F(λ)=λ2+Bλ+C,

    where B=(δ1), C=δ(km)xm2+δk1+xm2,

    F(1)=δδ(km)xm1+δk1+xm1G(x1).

    It is easy to see that G(x1) is increasing at x1, while x1 is decreasing at p; hence,

    F(1)<G(x)=0.

    Combined with Theorem 4.2, we know that there is an eigenvalue λ1>1.

    Notice F(1)=22δ+G(x1), which is decreasing at p, and

    limx10+F(1)=2δδk,limxxF(1)=22δ.

    So when 2δδk<0, we have

    p=pchange=δ2mδ(mk)+(2δ)(δk+δ2δ(mk)+(2δ))1km.

    such that F(1)=0. At this case, x1=(δk+δ2δ(mk)+(2δ))1m, λ2=1. Moreover, when pSN<p<pchange, F(1)>0, which implies |λ2|<1; while for p>pchange, one has F(1)<0, and then |λ2|>1.

    Theorem 4.3. The conclusions of the fixed point E1 for system (4.3) are true:

    (i) If δk+δ20, then the fixed point E1 is a saddle;

    (ii) if δk+δ2>0, then

    (a) when pSN<p<pchange, E1 is a saddle;

    (b) when p>pchange, E1 is a source, where pchange is given below.

    The theorem shows that pchange is a critical value. When the parameter p varies in the neighborhood, the stability of the fixed point will be changed. So, the system may undergo bifurcation at E1, which will be discussed in the next section.

    With similar arguments, we may know that the characteristic equation of the Jacobian matrix for system (4.3) at E2(x2,y2) is:

    F(λ)=λ2+Bλ+C,

    where B=(δ1), C=δ(km)xm2+δk1+xm2, denote

    F(1)=δδ(km)xm2+δk1+xm2G(x2).

    It is easy to verify that G(x2) is increasing at x2, and by Theorem 4.2, x2 is also increasing at p for p>pSN. As a result, G(x2(p)) is increasing at p for p>pSN which leads to

    F(1)=G(x2)>G(x0)=0.

    Notice that F(1)=22δ+F(1), then F(1)>0 holds. As we know, C(p)=G(x2(p))δ is increasing at p, moreover,

    limpp+SNC(p)=δ,limp+C(p)=δ(mk),

    so there exists a point

    p=pNS:=δ2mδmδk1(δk+1δmδk1)1km,

    for δ(mk)>1, where x2=(δk+1δmδk1)1m, such that C(pNS)=1. According to Lemma 2.2, we know that the Jacobian matrix of system (4.3) at E1 has a pair of conjugate complex roots λ1,2 with |λ1,2|=1. When pSN<p<pNS, C<1, |λ1,2|<1; while for p>pNS, C>1, |λ1,2|>1. Hence, we get the following results.

    Theorem 4.4. The fixed point E2 of system (4.3) has the following properties:

    (ⅰ) in the case of δ(mk)1, the fixed point E2 is a sink;

    (ⅱ) in the case of δ(mk)>1,

    (a) if pSN<p<pNS, then the fixed point E2 is a sink;

    (b) if p=pNS, then system (4.3) perhaps undergoes Neimark-Sacker bifurcation at E2;

    (c) if p>pNS, then the fixed point E2 is a source.

    Theorem 4.2 shows that system (4.3) may undergo saddle-node bifurcation for p=pSN, and Theorem 4.4 shows system (4.3) may undergo Neimark-Sacker bifurcation at E2 for p=pNS. In this section, we shall verify the truth theoretically.

    To start, according to the analysis in the above section we know that for p=pSN, the eigenvalues of Jacobian matrix for system (4.3) at E(x,y) (where x=y=(ρ1ρ2)1mk+1) are λ1=1, λ2=δ. To verify the existence of saddle-node bifurcation, the following conditions must be fulfilled:

    (SN.1)WTGp(x,y;p)0;(SN.2)12WT(D2G(x,y;p)(V,V))0.

    where W is the eigenvector of JTE (namely, the transverse of JE) corresponding to the eigenvalue λ=1, V is the eigencector of JE corresponding to the eigenvalue λ=1, and

    G(x,y;p)=(y(1δ)y+pxk1+xm).

    It is easy to compute

    W=(δ1),V=(11),

    therefore

    WTGp(x,y;p)=(δ1)T(0ρkm2(1ρ2)k+1m)=ρkm2(1ρ2)k+1m>0,

    which means (SN.1) is true.

    D2G(x,y;p) is a double linear function, and

    D2G(x,y;p)=(B1(x,y;p)B2(x,y;p)B3(x,y;p)B4(x,y;p)))T,

    where

    Bj(x,y;p)=2k,l=12Gj(ξ,0)ξkξl|ξ=0ξkξl,j=1,2,3,4,ξ1=x,ξ2=y.

    Hence, we have

    12WT(D2G(x,y;p)(V,V))=12(δ1)T(0δ2xkb12pSN(mk+1))=δ2xkb12pSN(mk+1)<0,

    where

    b1=(mk)(k1)(2k22m)k(k1)m,

    namely, (SN.2) is fulfilled. So, system (4.3) undergoes saddle-node bifurcation when p varies in the neiborhood of pSN, and the semi-stable fixed point E is bifurcated to a sink E2 and a saddle E1.

    Next, we shall illustrate that system (4.3) undergoes Neimark-Sacker bifurcation. Similar to the proof in Section 2.2, to guarantee the existence of Neimark-Sacker bifurcation, we need to verify the transversal condition (C.1) and the nondegenerate condition (C.2). By virtue of arguements in Section 4.2, we know that Det(JE2)=1 when

    pNS=δ2mδmδk1(δk+1δmδk1)1km,

    and the Jacobian matrix of the perturbed system to system (4.3) at E2 is

    JE2=(01(km)(pNS+p)xm+k12+k(pNS+p)xk12(1+xm2)21δ),

    where x2=(δk+1δmδk1)1m. It is easy to compute the eigenvalues

    λ1,2|p=0=(1δ)±(3δ)(1+δ)i2,

    which satisfy the nondegereate condition (C.2).

    Notice that

    b(p)=Det(JE2)=(mk)(pNS+p)xm+k12k(pNS+p)xk12(1+xm2)2=δ(km)xm2+δk1+xm2,

    and |λ1,2|=b(p). Combining with Theorem 4.2, we have

    d|λ1,2|dp|p=0=δmxm2(xm2+1)2(δmxmk+12pNS(k1))>0,

    which verifies the transversal condition (C.1). As a result, there will be an invariant curve bifurcated from the fixed point E2 of system (4.3).

    What we need to do is to determine the stability of the invariant curve bifurcated from the fixed point E2 of system (4.3). With the similar arguments in Section 3.2, we need to compute the first Lyapunov coefficient

    a=[2b220(δ+2)+3b30(δ+1)]8(δ+1),

    where

    b20=δ2k2δ2km+2δkδm+δ+22δ(δk+1δmδk1)1m,b30=δ3k(k23k+2)+6δ(δk+1)2(k+m1)6(δk+1)36δ3(δk+1δmδk1)2mδ2(δk+1)(3k2+3km6k+m23m+2)6δ2(δk+1δmδk1)2m.

    As a result, we obtain the following theorem.

    Theorem 4.5. There are no k and m satisfying a=0, so a0, moreover,

    (i) if a<0, then the fixed point E2 of system (4.3) is a sink for p<pNS, while for p>pNS, there exists a stable invariant curve for system (4.3) and the fixed point E2 is a source;

    (ii) if a>0, then the fixed point E2 of system (4.3) is a source for p>pNS, while for p<pNS, there exists an unstable invariant curve for system (4.3) and the fixed point E2 is a sink.

    We still choose p and δ as the bifurcation parameters of (4.3), Now, consider the case of a perturbation p and δ of parameter pNS and δ4=1. In order to ensure that the value of pNS is meaningful, take mk1 in this case. Then, we obtain the following pertubed system:

    {xy,y(1δ4δ)y+pNS+p1+xmxk, (4.4)

    where p=pNS+p, δ=δ4+δ, |p|,|δ|1.

    Next, taking u=xx2,v=yy2, then the fixed point E2 can be moved to the origin O(0,0) and system (4.4) becomes

    {uv,v(1δ4δ)v+pNS+p1+(u+x2)m(u+x2)kδy2. (4.5)

    The Jacobian of system (4.5) at its fixed point O is:

    J(0,0)=(01(km)(pNS+p)xm+k12+k(pNS+p)xk12(1+xm2)21δ4δ).

    Expand the right of system (4.5) as Taylor series at (u,v)=(0,0)

    {uv,vb10u+b01v+b20u2+b11uv+b02v2+b30u3+b12uv2+b21u2v+b03v3+O(ρ4), (4.6)

    where ρ=u2+v2,

    b10=(km)(pNS+p)xm+k12+k(pNS+p)xk12(1+xm2)2, b01=1δ4δ,b11=b02=b12=b21=b03=0,b20=(pNS+p)[mxm+k222(1+xm2)3(1m+xm2+mxm2)kmx2+k+m2(1+xm2)2+(1+k)kx2+k22(1+xm2)],b30=(pNS+p)[mxm+k326(1+xm2)4(23m+m2+4xm24m2xm2+2x2m2+3mx2m2+m2x2m2)+kmx3+k+m2(1m+xm2+mxm2)2(1+xm2)3(1+k)kmx3+k+m22(1+xm2)2+(2+k)(1+k)kx3+k26(1+xm2).

    Substituting p=δ=0 and δ4=1 into (4.6), we have

    (uv)A0(uv)+(0b20u2+b30v2+O(ρ4)), (4.7)

    with

    b10=1,b01=0,b20=k2km+2km+32(k+1mk1)1m,b30=k(k23k+2)+6(k+1)2(k+m1)6(k+1)36(k+1mk1)2m(k+1)(3k2+3km6k+m23m+2)6(k+1mk1)2m,

    and

    A0=J(0,0)|p=δ=0=(0110).

    The characteristic value of the linearization of system (4.3) at its fixed point O is

    λ1,2|p=δ=0=±i,

    and the corresponding eigenvectors are

    q1=(i,1)T,q2=(i,1)T.

    Use the same discussion in Section 2.2.2. We take any vector (u,v)TR2 in the form (u,v)T=zq1+ˉzq2, where z is a complex variable which transforms the map (4.7) to the complex form

    ziz+G(z,ˉz), (4.8)

    where

    G(z,ˉz)=b20(iz+iˉz)2+b30(iz+iˉz)32+O(ρ4)=k+l21k!l!gklzkˉzl,

    with

    g20=g02=b20,g11=b20,g30=g12=3ib30,g03=g21=3ib30. (4.9)

    Take the invertible parameter-dependent change of complex coordinate the same as Section 2.2.3:

    z=w+h202w2+h11wˉw+h022ˉw2,

    by setting

    h20=g20λ21λ1,h11=g11|λ1|2λ1,h02=g02¯λ12λ1,

    and

    w=ζ+16h30ζ3+12h21ζ2ˉζ+12h12ζˉζ2+16h03ˉζ3,

    by setting

    h30=g30λ31λ1,h21=g21λ1¯λ1|λ1|2,h12=g12¯λ1|λ1|2λ1,h03=g03¯λ13λ1,

    with λ1=i. The map (4.8) is changed into

    ζΓ(ζ)=iζ+Cζ|ζ|2+Dˉζ3+O(|ζ|4), (4.10)

    where gkl are given by (4.9) and

    C=1+3i4g20g11+1i2|g11|21+i4|g02|2+12g21=32b2203i2b30,D=i14g11g021+i4g02ˉg20+16g03=i2b220+i2b30,

    which is determined by the coefficients in (4.7).

    Then, we have

    C=18(1+k1+m)2m×[3(3+2k+k2(1+k)m)2+2i(2(7+k(11+k(6+k)))3(1+k)(3+k)m+(1+k)m2)],
    D=i24(1+k1+m)2m×[(1+k)(5+3k)m2+6(1+k)(6+k(3+k))mk(80+k(54+k(16+3k)))55].

    When 1<km, there is no kN+ and mN+ satisfying C=0 or D=0, so we have that C0 and D0 holds. Then, the fourth iterate of Γ allows for approximation by a complex flow, and the bifurcation of 1:4 resonance will be determined by

    ˜A=C|D|,

    and by 1<km, k,mN+, and mk1, we have

    |˜A|<1,
    Re˜A1,
    Im˜A1+(Re˜A)21(Re˜A)2.

    Lemma 4.1. [44] The fourth iterate of the map (4.10) can be represented in the form

    Γ4(ζ)=φ1ζ+O(|ζ|4), (4.11)

    where φt is the flow of a planar system

    ˙ζ=4iCζ|ζ|24iDˉζ3. (4.12)

    Theorem 4.6. Let mk1. When the parameters (p,δ) vary in a small neighborhood of (pNS,1), the system (4.3) undergoes a bifurcation of 1:4 resonance around the fixed point E2 and admits the following bifurcation curves:

    There is a Neimark-Sacker bifurcation curve at the fixed point E of system (4.3).

    Besides E1, there exist four saddle fixed points Sk of (4.12) around E2 when |˜A|<1, which are the corresponding period-4 points of (4.3). As the parameters (p,δ) vary in the neighborhood of (pNS,1), these nontrivial periodic points appear or disappear in the neighborhood of E2.

    There is a square heteroclinic cycle around E2 composed by the separatrices of Sk which is stable from the inside.

    To illustrate the theoretical results, we present some numerical simulations by virtue of the Python program.

    Choose parameters δ=0.2, k=5, m=10. At this case, pSN=0.39. Let p vary in a neighborhood of pSN=0.39, and we obtain the following figure.

    Figure 8 shows that when p<pSN=0.39, there is only one fixed point O(0,0) of (4.3) (see Figure 8(a)); when p=pSN=0.39, there are two fixed points O(0,0) and E(see Figure 8 (b)); meanwhile for p>pSN=0.39, there are three fixed points O, E1, and E2 (see Figures 8(c), (d)). It is in accordance with Theorem 4.1.

    Figure 8.  The fixed point of system (4.3) for δ=0.2, k=5, m=10.

    Choose the parameters δ=0.433, m=10, k=4. At this case pNS=0.9989 and pchange=1.18. Take the initial value E0=(1.063,1.063), and p varies in the neighborhood of pNS. We get the following figures (see Figure 9). Figure 9 shows that when p<pNS=0.9989, the fixed point E2 of system (4.3) is a sink, the orbits from E0(1.063,1.063) are all convergent to E2 (see the first four sub-fiures of Figure 9). Meanwhile for p>pNS=0.9989, there will occur a stable invariant curve (see the last four sub-figures of Figure 9), which is in accordance with Theorem 4.2.

    Figure 9.  Neimark-Sacker bifurcation.

    To illustrate the existence chaos more clearly, we present the following figures, which show the change of maximum Lypunuov exponent with p. The figure exhibits that when p<1.32, the MLE is not greater than 0, while for p>1.32, the MLE is greater than 0, which means that there exists chaos for the system when p>1.32. The results are in accordance with Figure 10.

    Figure 10.  Neimark-Sacker bifurcation and the maximium Lyapunuov exponent.

    Select m=10,k=4, and let δ vary from the neighborhood of δ4=1, while pNS=2, and let p vary in [pNS0.02,pNS+0.2]. By Theorem 4.6, the fixed point x2 is a 1:4 resonance point. The phenomenon is similar to the case of k=0.

    Figures 11 and 12 show the complex bifurcation phenomena when k=4, m=10, δ=1.02 and p vary. The invariant cycle induced by a Neimark-Sacker bifurcation is shown in the Figures 11(a)(c).

    Figure 11.  Phase portraits corresponding to Figure 14(c): Supercritical Neimark-Sacker bifurcation curve (Figures 11(a)–(c)).
    Figure 12.  Phase portraits corresponding to Figure 14(c): Heteroclinic loop and 4-period saddles induced by 1:4 resonance (Figures 12(a), (b)) period-doubling phenomena (Figure 12(c)).

    As the parameter p changes, further 4-period saddle points and heteroclinic cycles are generated by undergoing 1:4 resonance, showed in the Figures 12(a)(c). A similar period-doubling phenomena occurs, and then the system enters chaos. See the Figures 13(a), (b).

    Figure 13.  Phase portraits corresponding to Figure 14(c): Chaos induced by the "period bubbling" phenomena (Figures 13(a), (b)).

    It is worth noting that from Figure 14(b) and Figures 15(a)(d), we can observe that there are period-doubling phenomena from the period-4 orbits occuring by undergoing the bifurcation of 1:4 resonance to chaos, which is different from the case of k=0. From Figure 14(a), there are "period bubbling" phenomena when δ varies less than the critical value (the combination of period-doubling and inverse period-doubling [6,16,23]).

    Figure 14.  The 1:4 resonance as (p,δ) varies when k=4,m=10.
    Figure 15.  The "period bubbling" phenomena when δ=1 and p varying when k=4,m=10.

    It should be noted that although there is a phenomenon of period doubling, there is no conventional flip bifurcation here, since multipliers of the fixed point are not equal to –1. This does not comply with the conventional flip bifurcation conditions. The phenomenon of doubling the period is a complex bifurcation phenomenon caused by a strong resonance of 1:4, similar phenomena have been mentioned in [6].

    To describe the bifurcation of 1:4 resonance more clearly, Figure 14(d) is given to show the three-dimensional bifurcation diagrams of map (4.3) in (p,δ,y) space, respectively. Moreover, by calculating the maximium Lyapunov exponent corresponding to the system (for example, δ, similarly p), we see that there exist both positive and negative maximium Lyapunov exponents via the changes of parameters from Figures 16(a), (b), so there exist stable fixed point or stable period windows in the chaotic region [6]. In general, the existences of the positive maximium Lyapunov exponents are always considered as characteristics of chaos.

    Figure 16.  The 1:4 resonance as (p,δ) is varying when k=4,m=10.

    In this paper, the dynamical behaviors of the semi-discrete M-G model with k=0 and 1<km are analyzed. The local dynamical behaviors of fixed points are theoretically discussd by using qualitative theory, and the possible bifurcations of the system are theoretically revealed, which are verified by numerical simulations, too. We find that when the value of k is different, the system exhibits different dynamic behaviors. When 1<k<m, and comparing with k=0, the system produces more complex phenomena.

    When k=0 and certain parameter conditions are satisfied, the system will produce Neimark-Sacker bifurcation and 1:4 resonance. By calculating the maximum Lyapunov exponent, we find that the system will generate chaos during the process of Neimark-Sacker bifurcation. The existence of invariant cycles ending heteroclinic loops means that the metabolic level of RBCs will be in a stable state, which is beneficial for human health. The bifurcation of 1:4 resonance proves that, in certain regions, invariant circles may bifurcate from the period-2 orbits, which means that the number of RBCs exhibit periodic oscillations, but overall remain in a stable metabolic state.

    Compared to the case of k=0, we find that the Neimark-Saker bifurcation and the 1:4 resonance still exist in the case of 1<km. However, more complex dynamical behaviors will occur. The system with k0 undergoes saddle-node bifurcation before the Neimark-Sacker bifurcation, and there are periodic bubbling phenomena induced by the bifurcation of 1:4 resonance.

    The Neimark-Sacker bifurcation and the 1:4 strong resonance of the system induce the existence of chaos, but we are still unclear about the types of chaotic attractors that exist in the system. Theoretical verification and explanation of different types of chaotic attractors will be one of our main works in the next stage. Some symmetric phenomena from the presented phase portraits and the coressponding bifurcation diagrams can be observed, which is a topic worth further exploration in the future.

    In addition, we have not yet demonstrated the impact of memory effect and hereditary properties. Whether the memory effect and hereditary properties can be used to improve the adequacy of the model is also a question worth considering.

    Yulong Li: Formal analysis, conceptualization, methodology, writing-review and editing, writing-original draft, resources, software; Long Zhou: Formal analysis, writing-review and editing, writing-original draft, software; Fengjie Geng: Supervision, writing-review and editing, funding acquisition, validation, project administration. All the authors have agreed and given their consent for the publication of this research paper.

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

    This work is partly supported by the Fundamental Research Funds for the Central Universities (#2652023058).

    The author declares no conflict of interest.



    [1] A. Beuter, L. Glass, M. C. Mackey, M. S. Titcombe, Nonlinear dynamics in physiology and medicine, Springer, 2003. http://dx.doi.org/10.1007/978-0-387-21640-9
    [2] A. Fahsi, M. Belhaq, Analytical approximation of heteroclinic bifurcation in a 1:4 resonance, Int. J. Bifurc. Chaos, 22 (2012), 1250294. http://dx.doi.org/10.1142/S021812741250294X doi: 10.1142/S021812741250294X
    [3] A. Wan, J. Wei, Bifurcation analysis of Mackey-Glass electronic circuits model with delayed feedback, Nonlinear Dynam., 57 (2009), 85–96. http://dx.doi.org/10.1007/s11071-008-9422-7 doi: 10.1007/s11071-008-9422-7
    [4] A. Zaghrout, A. Mmar, A. EI-Sheikh, Oscillations and global attractivity in delay differential equations of population dynamics, Appl. Math. Comput., 77 (1996), 195–204. http://dx.doi.org/10.1016/S0096-3003(95)00213-8 doi: 10.1016/S0096-3003(95)00213-8
    [5] B. Krauskopf, Bifurcations at in a model for 1:4 resonance, Ergod. Theory Dyn. Syst., 17 (1997), 899–931. http://dx.doi.org/10.1017/s0143385797085039 doi: 10.1017/s0143385797085039
    [6] B. Li, Z. He, 1:2 and 1:4 resonances in a two-dimensional discrete Hindmarsh-Rose model, Nonlinear Dyn., 79 (2015), 705–720. http://dx.doi.org/10.1007/s11071-014-1696-3 doi: 10.1007/s11071-014-1696-3
    [7] C. Bonatto, J. A. C. Gallas, Periodicity hub and nested spirals in the phase diagram of a simple resistive circuit, Phys. Rev. Lett., 101 (2008), 054101. http://dx.doi.org/10.1103/physrevlett.101.054101 doi: 10.1103/physrevlett.101.054101
    [8] C. Wang, X. Li, Further investigations into the stability and bifurcation of a discrete predator-prey model, J. Math. Anal. Appl., 422 (2015), 920–939. http://dx.doi.org/10.1016/j.jmaa.2014.08.058 doi: 10.1016/j.jmaa.2014.08.058
    [9] C. Wang, X. Li, Stability and Neimark-Sacker bifurcation of a semi-discrete population model, J Appl. Anal. Comput., 4 (2014), 419–435. http://dx.doi.org/10.11948/2014024 doi: 10.11948/2014024
    [10] D. Mukherjee, Dynamics of A discrete-time ecogenetic predator-prey model, Commun. Biomath. Sci., 5 (2023), 161–169. http://dx.doi.org/10.5614/cbms.2022.5.2.5 doi: 10.5614/cbms.2022.5.2.5
    [11] D. Wu, H. Zhao, Complex dynamics of a discrete predator-prey model with the prey subject to the Allee effect, J. Differ. Equ. Appl., 23 (2017), 1765–1806. http://dx.doi.org/10.1080/10236198.2017.1367389 doi: 10.1080/10236198.2017.1367389
    [12] E. Shahverdiev, R. A. Nuriev, L. H. Hashimova, E. M. Huseynova, R. H. Hashimov, Chaos synchronization in the multifeedback Mackey-Glass model, Int. J. Mod. Phys. B, 19 (2005), 3613–3618. http://dx.doi.org/10.1142/S0217979205032346 doi: 10.1142/S0217979205032346
    [13] I. Tamas, S. Gabor, Semi-discretization method for delayed systems, Int. J. Numer. Meth. Eng., 55 (2002), 503–518. http://dx.doi.org/10.1002/nme.505 doi: 10.1002/nme.505
    [14] J. G. Freire, J. A. C. Gallas, Cyclic organization of stable periodic and chaotic pulsations in Hartleys oscillator, Chaos Soliton. Fract., 59 (2014), 129–134. https://doi.org/10.1016/j.chaos.2013.12.007 doi: 10.1016/j.chaos.2013.12.007
    [15] J. Guckenheimer, Multiple bifurcation problems of codimension two, SIAM J. Math. Anal., 15 (1984), 1–49. http://dx.doi.org/10.1137/0515001 doi: 10.1137/0515001
    [16] J. Vandermeer, Period bubbling in simple ecological models: Pattern and chaos formation in a quartic model, Ecol. Model., 95 (1997), 311–317. https://doi.org/10.1016/S0304-3800(96)00046-4 doi: 10.1016/S0304-3800(96)00046-4
    [17] K. Gopalsamy, M. Kulenović, G. Ladas, Oscillations and global attractivity in models of hematopoiesis, J. Dyn. Differ. Equ., 2 (1990), 117–132. http://dx.doi.org/10.1007/BF01057415 doi: 10.1007/BF01057415
    [18] K. J. Hale, N. Sternberg, Onset of chaos in differential delay equations, J. Comput. Phys., 77 (1988), 221–239. https://doi.org/10.1016/0021-9991(88)90164-7 doi: 10.1016/0021-9991(88)90164-7
    [19] L. Cheng, H. Cao, Bifurcation analysis of a discrete-time ratio-dependent predator-prey model with Allee Effect, Commun. Nonlinear Sci., 38 (2016), 288–302. https://doi.org/10.1016/j.cnsns.2016.02.038 doi: 10.1016/j.cnsns.2016.02.038
    [20] L. Lv, X. Li, Stability and bifurcation analysis in a discrete predator-prey system of Leslie type with radio-dependent simplified Holling Type Ⅳ functionalresponse, Mathematics, 12 (2024), 1803. http://dx.doi.org/10.3390/math12121803 doi: 10.3390/math12121803
    [21] M. B. Almatrafia, M. Berkal, Stability and bifurcation analysis of predator-prey model with Allee effect using conformable derivatives, J. Math. Comput. Sci., 36 (2025), 299–316. http://dx.doi.org/10.22436/jmcs.036.03.05 doi: 10.22436/jmcs.036.03.05
    [22] M. Berkal, M. B. Almatrafi, Bifurcation and stability of two-dimensional activator-inhibitor model with fractional-order derivative, Fractal Fract., 7 (2023), 1–18. http://dx.doi.org/10.3390/fractalfract7050344 doi: 10.3390/fractalfract7050344
    [23] M. S. Peng, Multiple bifurcations and periodic "bubbling" in a delay population model, Chaos Soliton. Fract., 25(2005), 1123–1130. https://doi.org/10.1016/j.chaos.2004.11.087 doi: 10.1016/j.chaos.2004.11.087
    [24] M. Wazewsa, A. Lasota, Mathematical problems of the dynamics of a system of red blood cells, Math. Stosowana, 6 (1976), 23–40. http://dx.doi.org/10.14708/ma.v4i6.1173 doi: 10.14708/ma.v4i6.1173
    [25] M. Zhao, Y. Du, Stability and bifurcation analysis of an amensalism system with allee effect, Adv. Differ. Equ., 1 (2020), 1–13. http://dx.doi.org/10.1186/s13662-020-02804-9 doi: 10.1186/s13662-020-02804-9
    [26] M. Zhao, C. Li, J. Wang, Complex dynamic behaviors of a discrete-time predator-prey system, J. Appl. Anal. Comput., 7 (2017), 478–500. http://dx.doi.org/10.11948/2017030 doi: 10.11948/2017030
    [27] N. Yi, Q. L. Zhang, P. Liu, Y. P. Lin, Codimension-two bifurcations analysis and tracking control on a discrete epidemic model, J. Syst. Sci. Complex., 24 (2011), 1033–1056. http://dx.doi.org/10.1007/s11424-011-9041-0 doi: 10.1007/s11424-011-9041-0
    [28] P. Amil, C. Cabeza, C. Masoller, A. C. Martí, Organization and identification of solutions in the time-delayed Mackey-Glass model, Chaos Interd. J. Nonlinear Sci., 25 (2015), 035202–035204. http://dx.doi.org/10.1063/1.4918593 doi: 10.1063/1.4918593
    [29] P. Amil, C. Cabeza, A. C. Marti, Exact discrete-time implementation of the Mackey-Glass delayed model, IEEE T. Circuits-II, 62 (2015), 681–685. http://dx.doi.org/10.1109/TCSII.2015.2415651 doi: 10.1109/TCSII.2015.2415651
    [30] Q. L. Chen, Z. D. Teng, L. Wang, H. Jiang, The existence of codimension-two bifurcation in a discrete SIS epidemic model with standard incidence, Nonlinear Dyn. 71 (2013), 55–73. http://dx.doi.org/10.1007/s11071-012-0641-6 doi: 10.1007/s11071-012-0641-6
    [31] R. Ma, Y. Bai, F. Wang, Dynamical behavior of a two-dimensional discrete predator-prey model with prey refuse and fear factor, J. Appl. Anal. Comput., 10 (2020), 1683–1697. http://dx.doi.org/10.11948/20190426 doi: 10.11948/20190426
    [32] S. Akhtar, R. Ahmed, M. Batool, N. A. Shah, J. D. Chung, Stability, bifurcation and chaos control of a discretized Leslie prey-predator model, Chaos Soliton. Fract., 152 (2021), 1–10. https://doi.org/10.1016/j.chaos.2021.111345 doi: 10.1016/j.chaos.2021.111345
    [33] S. G. Ruan, D. M. Xiao, Global analysis in a predator-prey system with nonmonotonic functional response, SIAM J. Appl. Math., 61 (2001), 1445–1472. http://dx.doi.org/10.1137/S0036139999361896 doi: 10.1137/S0036139999361896
    [34] S. S. Rana, Bifurcations and chaos control in a discrete-time predator-prey system of Leslie type, J. Appl. Anal. Comput., 9 (2019), 31–44. http://dx.doi.org/10.11948/2019.31 doi: 10.11948/2019.31
    [35] S. S. Rana, M. Uddin, Dynamics of a discrete-time chaotic lü system, Pan-Am. J. Math., 1 (2022), 1–7. http://dx.doi.org/10.28919/cpr-pajm/1-7 doi: 10.28919/cpr-pajm/1-7
    [36] S. L. Badjate, S. V. Dudul, Prediction of Mackey-Glass chaotic time series with effect of superimposed noise using FTLRNN model, New York, 2008. Available from: https://api.semanticscholar.org/CorpusID: 59747728.
    [37] T. Y. Li, J. A. Yorke, Period three implies chaos, Am. Math. Mon., 82 (1975), 985–992. Available from: https://www.jstor.org/stable/2318254.
    [38] W. Cheng, X. Li, Stability and Neimark-Sacker bifurcation of a semi-discrete population model, J. Appl. Anal. Comput., 4 (2014), 419–435. http://dx.doi.org/10.11948/2014024 doi: 10.11948/2014024
    [39] W. Li, X. Li, Neimark-Sacker bifurcation of a semi-discrete hematopoiesis model, J. Appl. Anal. Comput., 8 (2018), 1679–1693. http://dx.doi.org/10.11948/2018.1679 doi: 10.11948/2018.1679
    [40] W. Yao, X. Li, Bifurcation difference induced by different discrete methods in a discrete predator-prey model, J. Nonlinear Model. Anal., 4 (2022), 64–79. http://dx.doi.org/10.12150/jnma.2022.64 doi: 10.12150/jnma.2022.64
    [41] X. Jin, X. Li, Dynamics of a discrete two-species competitive model with Michaelis-Menten type harvesting in the first species, J. Nonlinear Model. Anal., 5 (2023), 494–523. http://dx.doi.org/10.12150/jnma.2023.494 doi: 10.12150/jnma.2023.494
    [42] X. L. Liu, D. M. Xiao, Complex dynamic behaviors of a discrete-time predator-prey system, Chaos Soliton. Fract., 32 (2007), 80–94. https://doi.org/10.1016/j.chaos.2005.10.081 doi: 10.1016/j.chaos.2005.10.081
    [43] X. L. Liu, S. Q. Liu, Codimension-two bifurcations analysis in two-dimensional Hindmarsh-Rose model, Nonlinear Dyn., 67 (2012), 847–857. http://dx.doi.org/10.1007/s11071-011-0030-6 doi: 10.1007/s11071-011-0030-6
    [44] Y. A. Kuznetsov, Elements of applied bifurcation theory, 4 Eds., Springer, 2023. http://dx.doi.org/10.1007/978-3-031-22007-4
    [45] Y. A. Kuznetsov, H. Meijer, Numerical normal forms for codim 2 bifurcations of fixed points with at most two critical eigenvalues, SIAM J. Sci. Comput., 26 (2005), 1932–1954. http://dx.doi.org/10.1137/030601508 doi: 10.1137/030601508
    [46] Y. L. Li, D. M. Xiao, Bifurcations of a predator-prey system of Holling and Leslie types, Chaos Soliton. Fract., 34 (2007), 606–620. https://doi.org/10.1016/j.chaos.2006.03.068 doi: 10.1016/j.chaos.2006.03.068
    [47] Y. Hong, Global dynamics of a diffusive phytoplankton-zooplankton model with toxic substances effect and delay, Math. Biosci. Eng., 19 (2022), 6712–6730. https://doi.org/10.3934/mbe.2022316 doi: 10.3934/mbe.2022316
    [48] Y. Li, H. Cao, Bifurcation and comparison of a discrete-time Hindmarsh-Rose model, J. Appl. Anal. Comput., 13 (2023), 34–56. http://dx.doi.org/10.11948/20210204 doi: 10.11948/20210204
    [49] Z. Jing, J. Yang, Bifurcation and chaos in discrete-time predator–prey system, Chaos Soliton. Fract., 27 (2006), 259–277. https://doi.org/10.1016/j.chaos.2005.03.040 doi: 10.1016/j.chaos.2005.03.040
    [50] Z. Wei, Y. Xia, T. Zhang, Stability and bifurcation analysis of an amensalism model with weak allee effect, Qual. Theor. Dyn. Syst., 19 (2020), 1–15. https://doi.org/10.1007/s12346-020-00341-0 doi: 10.1007/s12346-020-00341-0
  • Reader Comments
  • © 2025 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(326) PDF downloads(42) Cited by(0)

Figures and Tables

Figures(16)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog