
In this paper, the stability and bifurcation of a two–dimensional p53 gene regulatory network without and with time delay are taken into account by rigorous theoretical analyses and numerical simulations. In the absence of time delay, the existence and local stability of the positive equilibrium are considered through the Descartes' rule of signs, the determinant and trace of the Jacobian matrix, respectively. Then, the conditions for the occurrence of codimension–1 saddle–node and Hopf bifurcation are obtained with the help of Sotomayor's theorem and the Hopf bifurcation theorem, respectively, and the stability of the limit cycle induced by hopf bifurcation is analyzed through the calculation of the first Lyapunov number. Furthermore, codimension-2 Bogdanov–Takens bifurcation is investigated by calculating a universal unfolding near the cusp. In the presence of time delay, we prove that time delay can destabilize a stable equilibrium. All theoretical analyses are supported by numerical simulations. These results will expand our understanding of the complex dynamics of p53 and provide several potential biological applications.
Citation: Xin Du, Quansheng Liu, Yuanhong Bi. Bifurcation analysis of a two–dimensional p53 gene regulatory network without and with time delay[J]. Electronic Research Archive, 2024, 32(1): 293-316. doi: 10.3934/era.2024014
[1] | Jianmin Hou, Quansheng Liu, Hongwei Yang, Lixin Wang, Yuanhong Bi . Stability and bifurcation analyses of p53 gene regulatory network with time delay. Electronic Research Archive, 2022, 30(3): 850-873. doi: 10.3934/era.2022045 |
[2] | 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 |
[3] | Qixiang Wen, Shenquan Liu, Bo Lu . Firing patterns and bifurcation analysis of neurons under electromagnetic induction. Electronic Research Archive, 2021, 29(5): 3205-3226. doi: 10.3934/era.2021034 |
[4] | 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 |
[5] | Rina Su . Dynamic analysis for a class of hydrological model with time delay under fire disturbance. Electronic Research Archive, 2022, 30(9): 3290-3319. doi: 10.3934/era.2022167 |
[6] | Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang . The dynamics of a delayed predator-prey model with square root functional response and stage structure. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150 |
[7] | Yichao Shao, Hengguo Yu, Chenglei Jin, Jingzhe Fang, Min Zhao . Dynamics analysis of a predator-prey model with Allee effect and harvesting effort. Electronic Research Archive, 2024, 32(10): 5682-5716. doi: 10.3934/era.2024263 |
[8] | Xianjun Wang, Huaguang Gu, Bo Lu . Big homoclinic orbit bifurcation underlying post-inhibitory rebound spike and a novel threshold curve of a neuron. Electronic Research Archive, 2021, 29(5): 2987-3015. doi: 10.3934/era.2021023 |
[9] | Yuan Xue, Jinli Xu, Yuting Ding . Dynamics analysis of a diffusional immunosuppressive infection model with Beddington-DeAngelis functional response. Electronic Research Archive, 2023, 31(10): 6071-6088. doi: 10.3934/era.2023309 |
[10] | Fang Yan, Changyong Dai, Haihong Liu . Oscillatory dynamics of p53 pathway in etoposide sensitive and resistant cell lines. Electronic Research Archive, 2022, 30(6): 2075-2108. doi: 10.3934/era.2022105 |
In this paper, the stability and bifurcation of a two–dimensional p53 gene regulatory network without and with time delay are taken into account by rigorous theoretical analyses and numerical simulations. In the absence of time delay, the existence and local stability of the positive equilibrium are considered through the Descartes' rule of signs, the determinant and trace of the Jacobian matrix, respectively. Then, the conditions for the occurrence of codimension–1 saddle–node and Hopf bifurcation are obtained with the help of Sotomayor's theorem and the Hopf bifurcation theorem, respectively, and the stability of the limit cycle induced by hopf bifurcation is analyzed through the calculation of the first Lyapunov number. Furthermore, codimension-2 Bogdanov–Takens bifurcation is investigated by calculating a universal unfolding near the cusp. In the presence of time delay, we prove that time delay can destabilize a stable equilibrium. All theoretical analyses are supported by numerical simulations. These results will expand our understanding of the complex dynamics of p53 and provide several potential biological applications.
Dynamical analysis of gene regulatory networks (GRNs) characterized by mathematical models plays an important role in understanding the underlying mechanism of the corresponding biological processes and predicting what will happen[1,2,3]. A GRN is composed of genes, RNAs, and proteins represented by nodes, which are connected through the edges that indicate the transformation, promotion, and inhibition between them[4]. The expression level of proteins in the GRN is closely related to the mechanisms of a number of biochemical processes, including cell differentiation[5,6,7], neural plasticity[8,9,10] and the development of cancer[11,12,13]. Typically, cell fates in response to stresses are closely related to the dynamics of the tumor suppressor protein p53 in p53 GRN with core p53–Mdm2 feedback loops[14,15]. p53 maintains a low level under a homeostatic condition while moderate and high stresses make p53 exhibit oscillation and high level, which lead to cell cycle arrest and apoptosis[16,17], respectively. Therefore, more and more research has been focusing on exploring p53 dynamics numerically and theoretically through the construction of mathematical models with the help of bifurcation diagrams[18,19].
Bifurcation diagrams of mathematical models is a useful tool for analyzing the dynamics of GRN [20]. Bifurcation diagrams of p53 GRN in [16,17,21] display various types of bifurcation, such as codimension–1 saddle–node, Hopf bifurcation and codimension–2 Bogdanov–Takens bifurcation. Saddle–node bifurcation can generate the bistability with a low and high expression levels of p53, which results in cell survival and cell apoptosis, respectively. Hopf bifurcation can induce the appearance of a stable limit cycle [22] corresponding to the oscillation expression of p53, which result in cell cycle arrest[23]. Codimension–2 Bogdanov–Takens bifurcation may give rise to the coexistence of a stable steady state and a stable limit cycle[24], which correspond to cell survival and cell cycle arrest, respectively. Therefore, the analysis of the conditions under which these bifurcations occur can allow for a deeper understanding of cell fate decision in response to different parameters. More research has focused on bifurcation analyses of high–dimensional p53 GRNs through numerical simulations[25,26,27]. However, theoretical analysis of low-dimensional p53 GRNs contribute to the understanding of cell fate decisions under different conditions[28,29].
Theoretical analyses of the bifurcation of dynamical systems described by ordinary differential equations play an important role in unveiling their complex dynamic properties. Previously, extensive effort has been devoted to bifurcation analyses of predator–prey models and SIR models of infectious diseases with various factors[30], while several recent studies have theoretically revealed the bifurcation of GRNs. These studies focus on investigating the existence and stability of possible equilibria of the system and deriving the rigorous mathematical proofs for the existence of bifurcations, such as saddle-node bifurcation, Hopf bifurcation of codimension–1 and Bogdanov–Takes bifurcation of codimension–2 or codimension–3[31,32,33,34]. Besides, Hopf bifurcation may be caused by time delay in GRNs, and it is analyzed by studying the associated characteristic equation of the corresponding linearized system[18,35]. Although there has been much research on bifurcation analyses for p53 GRNs through numerical simulations, there is scant theoretical analyses of bifurcations of low dimensional p53 GRNs.
In this paper, we investigated the bifurcation of a two–dimensional p53 GRN without and with time delay, as described in [14], by performing rigorous mathematical analysis. Firstly, the existence of all possible positive equilibria are investigated by applying Descartes' rule of signs and the local stability of the positive equilibria are analyzed. Then, in the absence of time delay, the conditions for the appearance of codimension–1 saddle–node, Hopf bifurcation and codimension–2 Bogdanov–Takens bifurcation are derived by using Sotomayor's theorem[36], Hopf bifurcation theorem[37] and the normal form method, respectively, and the first Lyapunov number is calculated to obtain the stability of the limit cycle. Furthermore, in the presence of time delay, the Hopf bifurcation induced by time delay is analyzed on the basis of the Hopf bifurcation theorem. These theoretical results are numerically supported by bifurcation diagrams and phase portraits, and they can be considered as a complement to existing literature on the dynamics of p53 GRN.
The organization of this paper is as follows. The mathematical model of the p53 GRN is given in Section 2. The existence and local stability of positive equilibria of the p53 GRN without time delay are analyzed in Section 3. Section 4 presents the conditions for the occurrence of codimension–1 saddle–node bifurcation, Hopf bifurcation and codimension–2 Bogdanov–Takens bifurcation of the p53 GRN without time delay through theoretical analyses, which are supported by bifurcation diagrams and phase portraits. Section 5 is devoted to the analysis of Hopf bifurcation induced by time delay. We end the paper with the conclusion in Section 6.
In the present work, we consider a core p53 GRN with p53 and its key regulator Mdm2 in [14,15], as shown in Figure 1. Figure 1 includes a p53 self–induction positive feedback loop and a negative feedback loop, where p53 elevates the expression level of the Mdm2 protein and Mdm2 promotes the degradation of p53. Here, the degradation of p53 is regulated by the concentration of Mdm2 at some previous time. The rate equations for the concentration of p53 (denoted by x) and that of Mdm2 (denoted by y) are given by the following delay different equations:
{dxdt=r1+v1x2k21+x2−v2y(t−τ1)xk2+x−d1x,dydt=r2+v3x(t−τ2)2k23+x(t−τ2)2−d2y. | (2.1) |
Here, r1 and r2 denote the basal production rates of p53 and Mdm2, respectively. Also the production of both p53 and Mdm2 activated by p53 are modeled by using Hill functions with the production rates v1 and v3, and Michaelis constants k1 and k3, respectively. d1 and d2 are the basal degradation rates of p53 and Mdm2, respectively. Besides, p53 is degraded by Mdm2 at a rate v2 in a Michaelis–Menten function with the Michaelis constant k2. Time delays τ1 and τ2 characterize the periods of time for gene expression to protein production of p53 and Mdm2.
In this section, the existence and stability of the positive equilibria of the system (2.1) are investigated and the qualitative behavior of the system (2.1) are given in the following subsections.
In the section, we focus on analyzing the conditions for the existence of a positive equilibrium in the system (2.1) for biological reasons. Assuming a positive equilibrium of the system (2.1) is E(x∗,y∗), which satisfies the following equations:
{f1(x∗,y∗)=r1+v1x2∗k21+x2∗−v2y∗x∗k2+x∗−d1x∗=0,f2(x∗,y∗)=r2+v3x2∗k23+x2∗−d2y∗=0. | (3.1) |
Obviously, the second equation of Eq (3.1) is equivalent to
y∗=r2d2+v3x2∗d2(k23+x2∗). | (3.2) |
Rearranging the first equation of Eq (3.1), we get
g(x∗)=F(x∗)S(x∗)=0, | (3.3) |
where
F(x∗)=C6x6∗+C5x5∗+C4x4∗+C3x3∗+C2x2∗+C1x∗+C0, | (3.4) |
S(x∗)=d2(k21+x2∗)(k2+x∗)(k23+x2∗), | (3.5) |
C0=d2k21k2k23r1,C1=k21k23(d2r1−r2v2−d1d2k2),C2=d2k21k2r1+k23(d2k2r1+d2k2v1−d1d2k21),C3=−k21v2v3+k23(d2v1+d2r1−r2v2−d1d2k2)+k21(d2r1−r2v2−d1d2k2),C4=−d1d2k23+(d2k2r1+d2k2v1−d1d2k21),C5=d2r1+d2v1−r2v2−v2v3−d1d2k2,C6=−d1d2. | (3.6) |
Obviously, x∗ is the root of the following equation
F(x)=0. | (3.7) |
If the root x∗ of Eq (3.4) is positive, y∗ is positive according to Eq (3.2) with positive rate constants. Therefore, the conditions for the existence of positive roots of Eq (3.4) are suitable for the one of the positive equilibria of the system (2.1).
Applying Descartes' rule of signs to Eq (3.4), we obtained the number of possible positive equilibria in the system (2.1) is concluded in Table 1. Obviously, the system (2.1) must have at least one positive equilibrium for C0C6<0. Besides, according to Eq (3.6), we conclude that if C2<0, then C4<0 and if C4>0, then C2>0. Hence, based on the cases 1–4 in Table 1, a unique positive equilibrium E(x∗,y∗) in the system (2.1) exists under the conditions in the following theorem.
Case | C0 | C1 | C2 | C3 | C4 | C5 | C6 | Numberofsignchanges | Numberofpossiblepositiveequilibria(E) |
1 | + | − | − | − | − | − | − | 1 | 1 |
2 | + | + | − | − | − | − | − | 1 | 1 |
3 | + | + | + | − | − | − | − | 1 | 1 |
4 | + | + | + | + | − | − | − | 1 | 1 |
5 | + | + | + | + | + | − | − | 1 | 1 |
6 | + | + | + | + | + | + | − | 1 | 1 |
7 | + | + | − | + | − | − | − | 3 | 1, 3 |
8 | + | − | − | + | − | − | − | 3 | 1, 3 |
9 | + | − | + | + | − | − | − | 3 | 1, 3 |
10 | + | − | − | − | − | + | − | 3 | 1, 3 |
11 | + | + | − | − | − | + | − | 3 | 1, 3 |
12 | + | + | + | − | + | + | − | 3 | 1, 3 |
13 | + | − | + | + | + | + | − | 3 | 1, 3 |
14 | + | + | + | − | + | − | − | 3 | 1, 3 |
15 | + | + | + | + | − | + | − | 3 | 1, 3 |
16 | + | − | + | + | + | − | − | 3 | 1, 3 |
17 | + | + | + | − | − | + | − | 3 | 1, 3 |
18 | + | − | + | − | − | − | − | 3 | 1, 3 |
19 | + | + | − | + | − | + | − | 5 | 1, 3, 5 |
20 | + | − | + | − | − | + | − | 5 | 1, 3, 5 |
21 | + | − | + | + | − | + | − | 5 | 1, 3, 5 |
22 | + | − | − | + | − | + | − | 5 | 1, 3, 5 |
23 | + | − | + | − | + | − | − | 5 | 1, 3, 5 |
24 | + | − | + | − | + | + | − | 5 | 1, 3, 5 |
Theorem 3.1. The system (2.1) has a unique positive equilibrium E(x∗,y∗) if one of the following conditions holds:
(i)C2>0,C3>0,C5>0;
(ii)C1>0,C3>0,C4>0;
(iii)C1>0,C2>0,C4>0,C5>0.
Theorem 3.1 is verified by the nullclines of x and y in Figure 2, where the system (2.1) has a unique positive equilibrium for the condition (i) C2=−0.1197<0, C3=−0.1384<0, C5=−0.000723<0 with v1=0.18, v2=0.003, v3=0.6, d1=0.034, d2=0.02, r1=0.01, r2=0.001, k1=6, k2=4, k3=4 in Figure 2 (a) and the condition (ii) C1=0.0057>0, C3=0.1080>0, C4=0.0024>0 with v1=0.63, v2=0.01, v3=0.985, d1=0.004, d2=0.0025, r1=0.021, r2=0.0001, k1=1.5, k2=2, k3=9 in Figure 2 (b). Besides, Figure 2 (c) shows that the system (2.1) has three positive equilibria for C1=−0.1585>0, C2=0.0932>0, C3=0.0173>0, C4=−0.0107>0, C5=−0.0018>0 in case 9 with v1=0.18, v2=0.01, v3=0.55, d1=0.034, d2=0.03, r1=0.011, r2=0.001, k1=2.4, k2=2, k3=4.
Here, the conditions for the existence of the positive equilibria in system (2.1) are given in Table 1. Furthermore, the stability of the positive equilibria is analyzed in the next section.
To investigate the stability of any positive equilibrium E(x∗,y∗) of the system (2.1), the corresponding Jacobian matrix J is given by
J(E)=(2k21v1x∗(k21+x2∗)2−k2v2y∗(k2+x∗)2−d1−v2x∗k2+x∗2k23v3x∗(k23+x2∗)2−d2). |
The characteristic equation of the system (2.1) is
λ2−tr(J)λ+det(J)=0. |
The trace tr(J) and the determinant det(J) of the Jacobian matrix J are given by
tr(J)=2k21v1x∗(k21+x2∗)2−k2v2y∗(k2+x∗)2−d1−d2,det(J)=k2v2d2y∗(k2+x∗)2−2k21v1d2x∗(k21+x2∗)2+2k23v2v3x2∗(k2+x∗)(k23+x2∗)2+d1d2. | (3.8) |
The local stability of the positive equilibrium E(x∗,y∗) is decided by the signs of tr(J) and det(J). Next, we will first study the sign of det(J). Since
det(J)=k2v2d2y∗(k2+x∗)2−2k21v1d2x∗(k21+x2∗)2+2k23v2v3x2∗(k2+x∗)(k23+x2∗)2+d1d2=−d2g′(x∗)=−d2F′(x∗)S(x∗)−F(x∗)S′(x∗)S2(x∗). | (3.9) |
According to F(x∗)=0 in Eq (3.4), we conclude that
det(J)=−F′(x∗)(k2+x)(k21+x2)(k23+x2), | (3.10) |
and the signs of det(J) and F′(x∗) are opposite. Therefore, the signs of F′(x∗) and tr(J) decide the stability of the positive equilibrium E(x∗,y∗), which are given in the following theorem.
Theorem 3.2. The stability of the positive equilibrium E(x∗,y∗) under different conditions is described in Table 2.
Case | Conditions | Eigenvalues | Properties | |
1 | F′(x∗)<0 | tr(J)<0 | Reλ1<0,Reλ2<0 | Asymptotically stable |
2 | tr(J)=0 | λ1=−i√det(J),λ2=i√det(J) | Linear center | |
3 | tr(J)>0 | Reλ1>0,Reλ2>0 | Unstable | |
4 | F′(x∗)=0 | tr(J)<0 | λ1=tr(J)<0,λ2=0 | Non-hyperbolic |
5 | tr(J)=0 | λ1=λ2=0 | Non-hyperbolic | |
6 | tr(J)>0 | λ1=0,λ2=tr(J)>0 | Unstable(Non-hyperbolic) | |
7 | F′(x∗)>0 | ∀tr(J) | λ1λ2<0 | Unstable(saddle) |
However, the signs of F′(x∗) and tr(J) are not decided explicitly due to their complex expression, so we will give some numerical examples to illustrate the stability of E(x∗,y∗) in the following section. The stability of E(x∗,y∗) can be changed by the bifurcation, which will be explored in the following bifurcation analysis.
Bifurcation changes the stability and the number of equilibria in the system as the parameter varies through a critical value. In this section, we investigate the conditions for the occurrence of codimension–1 saddle–node and Hopf bifurcation with respect to v3 and codimension–2 Bogdanov–Takens bifurcation with respect to v3 and d2 in the system (2.1) without time delay.
According to Sotomayor theorem[36], we shall establish the conditions under which the system (2.1) experiences saddle–node bifurcation at the equilibrium E(x∗,y∗) when the control parameter v3 crosses the critical value vSN3.
The first condition is that tr(J(E))|v3=vSN3≠0 and det(J(E))|v3=vSN3=0, which correspond to
(SN.1)2k21v1x∗(k21+x2∗)2−k2v2y∗(k2+x∗)2−d1−d2≠0,vSN3=2k21v1d2x∗(k2+x∗)(k23+x2∗)22k23v2x2∗(k21+x2∗)2−k2v2d2y∗(k23+x2∗)22k23v2x2∗(k2+x∗)−d1d2(k2+x∗)(k23+x2∗)22k23v2x2∗, |
based on Eqs (3.8) and (3.10).
Next, to obtain the transversality conditions, the eigenvectors of the matrices J(E,vSN3) and JT(E,vSN3) with the eigenvalue λ=0 are given as follows, respectively,
V=(V1V2)=(d2(k23+x2∗)22k23vSN3x∗1),W=(W1W2)=(−d2(k2+x∗)v2x∗1). |
Denoting f(x,y)=(f1(x,y),f2(x,y))T,
fv3(E;vSN3)=(0x2∗k23+x2∗),D2f(E;vSN3)(V,V)=(∂2f1∂x2V21+2∂2f1∂x∂yV1V2+∂2f1∂2yV22∂2f2∂x2V21+2∂2f2∂x∂yV1V2+∂2f2∂2yV22)(E;vSN3)=(d2(k23+x2∗)22k23vSN3x∗[−2k2v2(k2+x∗)2+d2(k23+x2∗)22k23vSN3x∗(k21v1(k21−3x2∗)(k21+x2∗)3+k2v2y∗(k2+x∗)3)]d22(k23−3x2∗)(k23+x2∗)4k23vSN3x2∗). |
Clearly, the transversality conditions are
WTfv3(E;vSN3)=x2∗k23+x2∗≠0,(SN.2)WT[D2f(E;vSN3)(V,V)]=d22(k23+x2∗)22k23vSN3x2∗[2k2(k2+x∗)+(k23−3x2∗)2(k23+x2∗)−d2(k23+x2∗)22k23vSN3x∗(k21v1(k21−3x2∗)(k2+x∗)v2(k21+x2∗)3+k2y∗(k2+x∗)2)]≠0. |
Therefore, based on the above analyses, we get the following theorem.
Theorem 4.1. The system (2.1) experiences saddle–node bifurcation at the positive equilibrium E(x∗,y∗) as the parameter v3 crosses the critical value vSN3 if the conditions (SN.1) and (SN.2) hold. Here vSN3=2k21v1d2x∗(k2+x∗)(k23+x2∗)22k23v2x2∗(k21+x2∗)2−k2v2d2y∗(k23+x2∗)22k23v2x2∗(k2+x∗)−d1d2(k2+x∗)(k23+x2∗)22k23v2x2∗.
Theorem 4.1 is verified by the bifurcation diagram of x with respect to v3 in Figure 3(a) and phase portraits of x and y for five typical values of v3 in Figure 3(b)–(f) with the same parameters v1=0.18, v2=0.01, d1=0.034, d2=0.03, r1=0.011, r2=0.001, k1=2.4, k2=2, k3=4. In Figure 3(a), black solid and dashed lines represent stable and unstable equilibria, respectively, which meet at two saddle–node bifurcation points SN1 and SN2 with vSN13=0.4606397 and vSN23=0.5991365, respectively. In Figure 3(b)–(f), the solid lines represent the trajectory running along the arrows and red solid and hollow dots denote the stable and unstable equilibria, respectively.
As shown in Figure 3(a), the system (2.1) undergoes saddle–node bifurcation at E1(x∗,y∗)=(0.7593,0.5674) and E2(x∗,y∗)=(1.3543,2.0872) as v3 passes through vSN13=0.4606397 and vSN23=0.5991365, respectively, with tr(J(E1,vSN13))=−0.0263≠0, WT[D2f(E1;vSN13)(V,V)]=0.0118≠0 and tr(J(E2,vSN23))=−0.0315≠0, WT[D2f(E2;vSN23)(V,V)]=0.0104≠0, which meet all conditions in Theorem {4.1}. vSN13 and vSN23 divide the region in Figure 3(a) into five parts, in which phase portraits of x and y are illustrated in Figure 3(b)–(f). Only a stable equilibrium appears for v3=0.45<vSN13 in Figure 3(b) and v3=0.61>vSN23 in Figure 3(f). Two equilibria coexist at v3=vSN13 in Figure 3(c) and v3=vSN23 in Figure 3(e). There are three equilibria for v3 that vary between vSN13 and vSN23 in Figure 3(d)).
Furthermore, the stability and property of these equilibria E(x∗,y∗) in Figure 3(b)–(f) are listed in Table 3 to verify three cases in Table 2. The properties of unstable non-hyperbolic equilibria and unstable saddle are consistent with the conditions of cases 4 and 7 in Table 2, respectively. Other equilibria are asymptotically stable nodes, which correspond to case 1 in Table 2.
v3 | Ei(x∗,y∗) | F′(x∗) | tr(J) | Stability | Phase portraits |
0.45 | (2.1967,3.5089) | −0.2792 | −0.0273 | Asymptotically stable | Figure 3(b) |
0.4606397 | (0.7593,0.5674) | 0 | −0.0263 | Non-hyperbolic(saddle-node) | Figure 3(c) |
(2.1538,3.4845) | −0.2573 | −0.0267 | Asymptotically stable | Figure 3(c) | |
0.55 | (0.6400,0.4910) | −0.03146 | −0.0305 | Asymptotically stable | Figure 3(d) |
(1.0352,1.1842) | 0.0246 | −0.0206 | Unstable(saddle) | Figure 3(d) | |
(1.7665,3.0254) | −0.0981 | −0.0218 | Asymptotically stable | Figure 3(d) | |
0.5991365 | (0.6186,0.4999) | −0.0391 | −0.0315 | Asymptotically stable | Figure 3(e) |
(1.3543,2.0872) | 0 | −0.0190 | Non-hyperbolic(saddle-node) | Figure 3(e) | |
0.61 | (0.6147,0.5025) | −0.0405 | −0.0316 | Asymptotically stable | Figure 3(f) |
In this section, we try to explore the conditions under which a positive equilibrium E(x∗,y∗) loses the stability through Hopf bifurcation under some parametric restriction. Here, considering v3 as the bifurcation parameter, we shall establish the conditions under which the system (2.1) experiences Hopf bifurcation at the positive equilibrium E(x∗,y∗) when v3 crosses the critical value vHB3.
The first condition is that the Jacobian matrix J(E,vHB3) has a pair of purely imaginary eigenvalues, that is tr(J(E,vHB3))=0 and det(J(E,vHB3))>0, which correspond to
(HB.1)vHB3=−d2(d1+d2)(k2+x∗)2(k23+x2∗)k2v2x2∗+2d2k21v1x∗(k2+x∗)2(k23+x2∗)k2v2x2∗(k21+x2∗)2−r2(k23+x2∗)x2∗,andF′(x∗,vHB3)<0. |
Besides, the transversality condition that ensures the changes of stability of the positive equilibrium through non–degenerate Hopf bifurcation is dRe(λi)dv3∣v3=vHB3≠0, i.e.,
(HB.2)dtr(J(E))dv3|v3=vHB3=x3∗(k21v2+v3x2∗)F′(x∗)[2k41v1−6k21v1x2∗(k21+x2∗)3+k2r2v2d2(k2+x∗)2+k2v2v3(k23+3x2∗+2x∗k2)d2(k23+x2∗)2(k2+x∗)2]−k2v2x2∗d2(k23+x2∗)(k2+x∗)2≠0. |
Lastly, the first Lyapunov number Γ at the equilibrium E(x∗,y∗) is computed to analyze the stability of the limit cycle. By the transformation X=x−x∗, Y=y−y∗, the system (2.1) becomes
{dXdt=a10X+a01Y+a20X2+a11XY+a02Y2+a30X3+a21X2Y+a12XY2+a03Y3+Q1(|X,Y|4),dYdt=b10X+b01Y+b20X2+b11XY+b02Y2+b30X3+b21X2Y+b12XY2+b03Y3+Q2(|X,Y|4), |
where
a10=2k21v1x∗(k21+x2∗)2−k2v2y∗(k2+x∗)2−d1,a01=−v2x∗k2+x∗,a20=k21v1(k21−3x2∗)(k21+x2∗)3+k2v2y∗(k2+x∗)3,a30=−4k21v1x∗(k21−x2∗)(k21+x2∗)4−k2v2(k2+x∗)4,a11=−k2v2(k2+x∗)2,a21=k2v2(k2+x∗)3,b10=2k23v3x∗(k23+x2∗)2,b01=−d2,b20=k23v3(k23−3x2∗)(k23+x2∗)3,b30=−4k23v3x∗(k23−x2∗)(k23+x2∗)4,a02=a12=a03=b02=b11=b03=b12=b21=b03=0. |
According to the formula in [21], the first Lyapunov number Γ is given as follows
(HB.3)Γ=−32a01Φ32[a10b10a211−2a10a01a220−2a201a20b20−(a01b10−2a210)a11a20+(3a01a30−2a10a21)(a210+a01b10)], |
where Φ=a10b01−a01b10.
Therefore, we get the following theorem.
Theorem 4.2. The system (2.1) experiences Hopf bifurcation at the positive equilibrium E(x∗,y∗) when v3 crosses the critical value vHB3 with the conditions (HB.1) and (HB.2). A supercritical (subcritical) Hopf bifurcation occurs for the first Lyapunov number Γ<0(>0) in (HB.3). Here vHB3=−d2(d1+d2)(k2+x∗)2(k23+x2∗)k2v2x2∗+2d2k21v1x∗(k2+x∗)2(k23+x2∗)k2v2x2∗(k21+x2∗)2−r2(k23+x2∗)x2∗.
Note that the sign of the first Lyapunov number Γ cannot be determined directly due to its complex expression; we illustrate Hopf bifurcation of the system (2.1) through the following numerical examples to verify the correctness of Theorem 4.2.
Two supercritical Hopf bifurcation points HBsup1 and HBsup2, and a subcritical Hopf bifurcation point HBsub are shown in bifurcation diagrams of x with respect to v3 in Figure 4 (a)–(b), where black solid and dashed lines respectively represent stable and unstable equilibria while green and purple lines denote stable and unstable limit cycles. For v1=0.8153, v2=0.2, d1=0.015, d2=0.04, r1=0.03, r2=0.002, k1=7, k2=3 and k3=7, the first critical value vHB3=0.2096 labeled by HBsup1, where the conditions, E(x∗,y∗)=(5.1628,1.8957), F′(x∗)=−34.7231<0, dtr(J(E))dv3=0.1612>0 and Γ=−0.1455<0, imply the appearance of a stable limit cycle. Then the stable limit cycle disappears at another critical value vHB3=0.2589 labeled by HBsup2 with E(x∗,y∗)=(3.3111,1.2332), F′(x∗)=−19.3163<0, dtr(J(E))dv3=−0.2941<0 and Γ=−0.9058<0. For v3=0.25 between HBsup1 and HBsup2, a stable limit cycle surrounding an unstable equilibrium is illustrated in the phase portrait of x and y in Figure 4 (c). Beside, Figure 4 (b) shows a subcritical Hopf bifurcation point HBsub at vHB3=0.1852 with the conditions E(x∗,y∗)=(4.4394,4.1289), F′(x∗)=−2.1014<0, dtr(J(E))dv3=2.7028>0 and Γ=0.4856>0. An unstable limit cycle surrounding a stable focus coexists with a saddle and a node for v3=0.1838 in the phase portrait in Figure 4 (d).
Apart from codimension–1 saddle–node and Hopf bifurcation, the system (2.1) may undergo codimension–2 Bogdanov–Takens bifurcation at the positive equilibria E(x∗,y∗), which will be analyzed by considering v3 and d2 as bifurcation parameters in the next section.
Let vBT3 and dBT2 be two critical values of v3 and d2, at which det(J(E))|(v3,d2)=(vBT3,dBT2)=0 and tr(J(E))|(v3,d2)=(vBT3,dBT2)=0. Perturbing v3 and d2 by v3=vBT3+μ1 and d2=dBT2+μ2 with μ1 and μ2 in a small neighborhood of (0,0), the system (2.1) becomes
{dxdt=r1+u2x2k21+x2−v2yxk2+x−d1x,dydt=r2+(vBT3+μ1)x2k23+x2−(dBT2+μ2)y. | (4.1) |
Transforming the equilibrium E(x∗,y∗) to the origin (0,0) by z1=x−x∗ and z2=y−y∗, the system (4.1) becomes
{˙z1=a10(μ)z1+a01(μ)z2+a20(μ)z21+a11(μ)z1z2+R1(z,μ),˙z2=b00(μ)+b10(μ)z1+b01(μ)z2+b20(μ)z21+R2(z,μ), | (4.2) |
where
a10(μ)=2k21v1x∗(k21+x2∗)2−k2v2y∗(k2+x∗)2−d1,a01(μ)=−v2x∗k2+x∗,a20(μ)=k21v1(k21−3x2∗)(k21+x2∗)3+k2v2y∗(k2+x∗)3,a11(μ)=−k2v2(k2+x∗)2,b00(μ)=r2+(v3+μ1)x2∗k23+x2∗−(d2+μ2)y∗,b10(μ)=2k23(vBT3+μ1)x∗(k23+x2∗)2,b01(μ)=−(dBT2+μ2),b20(μ)=k23(vBT3+μ1)(k23−3x2∗)(k23+x2∗)3, |
and z=(z1,z2)T, μ=(μ1,μ2)T. Ri(z,μ)=O(‖z‖3) (i=1,2) denotes the power series as to z1, z2 with the order 3 and more. The coefficients of Ri(z,μ)(i=1,2) and aij, bij smoothly depend on μ1 and μ2. According to b00(0)=0, we rewrite the system (4.2) at μ1=0 and μ2=0 in the following form
dzdt=J0z+F(z), |
where
(BT.1)J0=(a10(0)a01(0)b10(0)b01(0))≠0 ,F(z)=(a20(μ)z21+a11(μ)z1z2+R1(z,μ)b20(μ)z21+R2(z,μ)). |
Let a10(0)=a10, a01(0)=a01, b10(0)=b10 and b01(0)=b01. Since J0 has two zero eigenvalues, we get a10+b01=0 and a10b01=a01b10. Then, we choose α0=(1,−a10a01)T,α1=(0,1a01)T and β0=(1,0)T,β1=(a10,a01)T as the eigenvector and generalized eigenvector with zero eigenvalues for J0 and JT0, respectively, which satisfy ⟨α0,β0⟩= ⟨α1,β1⟩=1 and ⟨α1,β0⟩=⟨α0,β1⟩=0. The linearly independent vectors α0 and α1 form a basis of R2. Thus, we make the following transformation
z=γ1α0+γ2α1, |
i.e., γ1=z1, γ2=a10z1+a01z2. Then the system (4.2) becomes
{˙γ1=γ2+c20(μ)γ12+c11(μ)γ1γ2+R3(γ,μ),˙γ2=d00(μ)+d10(μ)γ1+d01(μ)γ2+d20(μ)γ12+d11(μ)γ1γ2+R4(γ,μ), | (4.3) |
where
c20(μ)=a20(μ)−a10a11(μ)a01,c11(μ)=a11(μ)a01,d00(μ)=a01b00(μ),d10(μ)=a01b10(μ)−a10b01(μ),d01(μ)=a10+b01(μ),d20(μ)=a10a20(μ)−a210a11(μ)a01+a01b20(μ),d11(μ)=a10a11(μ)a01, |
and R3,4(γ,μ)=O(‖γ‖3) (γ=(γ1,γ2)T). Due to a10+b01=0 and a10b01=a01b10, we obtain d00(0)=d10(0)=d01(0)=0.
Next, making the transformations η1=γ1 and η2=γ2+c20(μ)γ12+c11(μ)γ1γ2+R3(γ,μ), system (4.3) becomes
{˙η1=η2,˙η2=e00(μ)+e10(μ)η1+e01(μ)η2+e20(μ)η12+e11(μ)η1η2+e02(μ)η22+R5(η,μ), | (4.4) |
where
e00(μ)=d00(μ),e10(μ)=d10(μ)+c11(μ)d00(μ),e01(μ)=d01(μ),e02(μ)=c11(μ),e20(μ)=d20(μ)+c11(μ)d10(μ)−c20(μ)d01(μ),e11(μ)=d11(μ)+2c20(μ), |
and R5(η,μ)=O(‖η‖3), η=(η1,η2)T. Moreover, we have
e00(0)=e10(0)=e01(0)=0,e20(0)=d20(0),e11(0)=d11(0)+2c20(0),e02(0)=c11(0). |
And we assume that
(BT.2)e11(0)=d11(0)+2c20(0)≠0, |
Then making a coordinate shift η1=ω1+ξ(μ),η2=ω2, where ξ(μ)≈−e01(μ)e11(0), the system (4.4) reduces to
{˙ω1=ω2,ω2=f00(μ)+f10(μ)ω1+f20(μ)ω21+f11(μ)ω1ω2+f02(μ)ω22+R6(ω,μ), | (4.5) |
where
f00(μ)=e00(μ)+e10(μ)ξ(μ)+⋯,f10(μ)=e10(μ)+2e20(μ)ξ(μ)+⋯,f02(μ)=e02+ξ(μ)+⋯,f11(μ)=e11(μ)+2ξ(μ)+⋯,f20(μ)=e20(μ)+3ξ(μ)+⋯, | (4.6) |
and R6(ω,μ)=O(‖ω‖3) (ω=(ω1,ω2)T).
Next, introducing a new time variable τ1 by dt=(1+θ(μ)ω1)dτ1, θ(μ)=−f02(μ), system (4.5) becomes
{˙ω1=ω2,˙ω2=h00(μ)+h10(μ)ω1+h20(μ)ω21+h11(μ)ω1ω2+R7(ω,μ), | (4.7) |
in which
h00(μ)=f00(μ),h10(μ)=f10(μ)−2f00(μ)f02(μ),h20(μ)=f20(μ)−2f10(μ)f02(μ)+f00(μ)f02(μ)2,h11(μ)=f11(μ), |
and R7(ω,μ)=O(‖ω‖3).
If the following condition holds,
(BT.3)h20(0)=d20(0)≠0, |
then we introduce the new variables τ2=|h20(μ)h11(μ)|τ1, y1=h211(μ)h20(μ)ω1, y2=sign(h20(μ)h11(μ))h311(μ)h220(μ)ω2, the system (4.7) is changed to
{˙y1=y2,˙y2=m00(μ)+m10(μ)y1+y21+sy1y2+R8(y,μ), | (4.8) |
where
m00(μ)=h411(μ)h320(μ)h00(μ),m10(μ)=h211(μ)h220(μ)h10(μ), | (4.9) |
s=sign(h20(μ)h11(μ))=sign(h20(0)h11(0))=sign(e20(0)e11(0))=±1, |
and R8(y,μ)=O(‖y‖3), y=(y1,y2)T.
If the following transversality condition holds,
(BT.4)det(∂(m00,m10)∂(μ1,μ2))μ1=μ2=0≠0, |
the system (4.1) experiences Bogdanov–Takens bifurcation when μ=(μ1,μ2) is in a small neighborhood of (0,0) based on the results in [22]. According to the above discussion, the following theorem is obtained.
Theorem 4.3. The system (2.1) experiences codimension–2 Bogdanov–Takens bifurcation at the positive equilibrium E(x∗,y∗) as (v3,d2) varies near (vBT3,dBT2) and the conditions (BT.1)–(BT.4) are satisfied. The local representations of the bifurcation curves are given as follows:
(i) the saddle–node bifurcation curve SN={(μ1,μ2)∣4m00−m210=0};
(ii) the Hopf bifurcation curve H={(μ1,μ2)∣m00=0,m10<0};
(iii) the homoclinic bifurcation curve HL={(μ1,μ2)∣m00=−625m210+O(m210),m10<0}.
Theorem 4.3 is verified by codimension–2 bifurcation diagram of v3 and d2 in Figure 5 and phase portraits of x and y in Figure 6 with r1=0.011, r2=0.001, v1=0.63, v2=0.06, k1=4.5, k2=2, k3=4 and d1=0.034. Through numerical simulation, we find that system (2.1) undergoes Bogdanov–Takens bifurcation at (v3,d2)=(0.3419136,0.04331208), where the conditions (BT.1)–(BT.4) in Theorem 4.3 are as follows,
J0=(0.043312−0.0344130.054513−0.043312)≠0 |
|∂(m00,m10)∂(μ1,μ2)|μ1=μ2=0=|0.0094380.563671−0.1633265.187334|=0.141019≠0, |
e11(0)=−0.002882≠0, h20(0)=d20(0)=−0.000125≠0, and s=sign(0.042991)=+1. Moreover, the local representations of the bifurcation curves are given as follows,
(i) the saddle–node bifurcation curve SN
{(μ1,μ2)∣0.0000002763620326μ1−0.000002097660245μ2+0.000003702622789μ21+2.500675014μ1μ2−20.19183593μ22=0}; |
(ii) the Hopf bifurcation curve H
{(μ1,μ2)∣0.000003273285777μ1−0.00002645487898μ2+3.14829075μ1μ2−25.22103078μ22=0,m10<0}; |
(iii) the homoclinic bifurcation curve HL
{(μ1,μ2)∣0.000000690907144μ1−0.000006244213952μ2−0.000005734914096μ21+6.251472238μ1μ2−50.47629929μ22=0,m10<0}. |
Figure 5 illustrates the curves SN, H and HL in the (μ1, μ2) parameter plane. These curves divide the small neighborhood of the origin (0, 0) in the (μ1,μ2) plane into four parts, in which the phase portraits of x and y are given in Figure 6 with the trajectory in blue solid lines and the stable and unstable equilibria in red solid and hollow dots, respectively. In order to see the variation around Bogdanov–Takens bifurcation point, phase portraits of x and y are given in insets in Figure 6(a)–(f).
(a) The system (2.1) has a cusp of codimension–2 Bogdanov–Takens bifurcation point EBT2 with another stable node for (μ1,μ2)=(0,0) (see Figure 6(a)).
(b) The system (2.1) has a unique stable node when (μ1,μ2) lies in region I (see Figure 6(b)).
(c) A saddle and an unstable focus coexist with another stable node when (μ1,μ2) enters into region II from region I through the branch SN− of the curve SN (see Figure 6(c)).
(d) The unstable focus becomes stable and is surrounded by an unstable limit cycle when (μ1,μ2) crosses the subcritical Hopf bifurcation curve H into region III (see Figure 6(d)).
(e) A homoclinic loop occurs for (μ1,μ2) on the curve HL (see Figure 6(e)).
(f) The unstable limit cycle disappears and three stable equilibria are left when (μ1,μ2) crosses the HL curve into region IV (see Figure 6(f)).
In this section, we will investigate the effect of time delay on the stability of the positive equilibrium E(x∗,y∗) in system (2.1).
The system (2.1) is linearized at E(x∗,y∗) as follows,
(˙x(t)˙y(t))=(p1100p22)(x(t)y(t))+(0p1200)(x(t−τ1)y(t−τ1))+(00p210)(x(t−τ2)y(t−τ2)) | (5.1) |
where
p11=2k21v1x∗(k21+x2∗)2−d1,p12=−v2x∗k2+x∗,p21=2k23v3x∗(k23+x2∗)2,p22=−d2. | (5.2) |
The characteristic equation of the linearized system (5.1) is
λ2−(p11+p22)λ+p11p22−p12p21e−λ(τ1+τ2)=0. | (5.3) |
Assuming that λ=iω(ω>0) is the root of Eq (5.3) and τ=τ1+τ2, we get the following equation
ω2+(p11+p22)iω−p11p22+p12p21(cosωτ−isinωτ)=0. | (5.4) |
Separating the real and imaginary parts of Eq (5.4) results in
{p12p21cos(ωτ)+ω2−p11p22=0,p12p21sin(ωτ)−(p11+p22)ω=0. | (5.5) |
Thus, cos(ωτ) and sin(ωτ) are given by
{cos(ωτ)=p11p22−ω2p12p21,sin(ωτ)=(p11+p22)p12p21ω, | (5.6) |
which implies that
ω4+(p211+p222)ω2+(p22p11)2−(p12p21)2=0. | (5.7) |
The discriminant of Eq (5.7) is
Δω=(p211+p222)2−4((p22p11)2−(p12p21)2)=(p211−p222)2+4(p12p21)2>0. | (5.8) |
Therefore, Eq (5.7) has two different roots ω21 and ω22, and ω21+ω22=−(p211+p222)>0, ω21ω22=(p22p11)2−(p12p21)2. Therefore, if (p22p11)2−(p12p21)2>0, Eq (5.7) has a purely imaginary root iω0 and
ω0=√−(p211+p222)+√Δω2. | (5.9) |
Then, according to Eq (5.6), the critical value of τ is given as follows,
τ(j)0=1ω0arccos(p11p22−ω20p12p21)+2jπω0,j=0,1,2,⋯. | (5.10) |
Let
τ0=min{τ(j)0|j=0,1,2,⋯}. | (5.11) |
Next, we will verify the transversality condition sign{[dRe(λ(τ))dτ]|τ=τ0}≠0. By differentiating both sides of Eq (5.3) with respect to τ, we get
(dλ(τ)dτ)−1=−(2λ−(p11+p22))eλτp12p21λ−τλ. |
At τ=τ0, we have
Re(dλdτ)−1=Re{−(2λ−(p11+p22))eλτp12p21λ−τλ}=Re{(p11+p22)cosω0τ0+2ω0sinω0τ0+i((p11+p22)sinω0τ0−2ω0cosω0τ0)p12p21iω0}=1(p12p21)2{2ω02+[(p11+p22)2−2p11p22]}=1(p12p21)2√Δω>0. |
Hence,
sign{[dRe(λ)dτ]|τ=τ0}=sign{Re(dλdτ)−1|τ=τ0}>0. |
Finally, we have the following theorem based on the Hopf bifurcation theorem [37].
Theorem 5.1. Let τ0 and p11,p12,p21,p22 be defined by Eqs (5.11) and (5.2), If (p22p11)2−(p12p21)2<0, the positive equilibrium E(x∗,y∗) of system (2.1) is asymptotically stable for τ∈(0,τ0] and the system (2.1) undergoes Hopf bifurcations at τ=τ0.
Theorem 5.1 is verified by the bifurcation diagram of x with respect to τ2 in Figure 7 (a) and phase portraits of x and y in Figure 7 (b)–(d) with v1=0.86,v2=0.6,v3=0.89,d1=0.017,d2=0.49,r1=0.17,r2=0.004,k1=1.07,k2=0.03 and k3=0.42. For these parameters, the system (2.1) has a unique equilibrium E(x∗,y∗)=(0.2126,0.3785) and (p22p11)2−(p12p21)2=−0.4940<0, then ω0=0.7677, τ0=0.4677 in Theorem 5.1. Based on Theorem 5.1, the positive equilibrium E(x∗,y∗)=(0.2126,0.3785) is locally asymptotically stable for τ∈(0,τ0) and the system (2.1) undergoes supcritical Hopf bifurcation at τ=τ0, which is accord with bifurcation diagram in Figure 7 (a). In Figure 7 (a), black solid and dashed lines respectively denote stable and unstable equilibria while green dots represent the maxima and minima of stable limit cycle, supcritical Hopf bifurcation HBsup occurs at τ2=0.2677=τ0−τ1 with τ1=0.2. Besides, in the phase diagrams of x and y in Figure 7 (b)–(d), red solid and hollow dots denote stable and unstable equilibria, respectively, and blue and green lines respectively denote the trajectory and a stable limit cycle, the positive equilibrium E(x∗,y∗) is stable for τ2=0.1<τ0−τ1 in Figure 7 (b); and it lose stability at τ2=0.2677=τ0−τ1 in Figure 7 (c); then it become unstable one with the appearance of a stable limit cycle at τ2=0.35>τ0−τ1 in Figure 7 (d).
In this paper, we are mainly concerned with the stability and bifurcation of system (2.1) without and with time delay τ. For τ=0, the existence of positive equilibria of system (2.1) are analyzed through the Descartes' rule of signs to obtain the conditions under which the system (2.1) has a unique positive equilibrium in Theorem 3.1, which are verified by the nucllines of x and y in system (2.1) in Figure 2. The stability of positive equilibria of system (2.1) without time delay is presented in Table 2 due to the complex expression of the determinant and trace of the Jacobian matrix. For positive equilibria, selecting v3 as a bifurcation parameter, the conditions of saddle-node and Hopf bifurcation in system (2.1) are given in Theorems 4.1 and 4.2, which gives the first Lyapunov number that determines the stability of limit cycle. These two theorems are verified by codimension-1 bifurcation diagram of x with respect to v3 in Figures 3 and 4, which include saddle-node bifurcation and supercritical and subcritical Hopf bifurcation, respectively. Furthermore, by choosing two parameters v3 and d2 in system (2.1) as bifurcation parameters, we prove that the system exhibits codimension-2 Bogdanov-Takens bifurcation under the conditions in Theorem 4.3, which is obtained by calculating universal unfolding near the cusp. Theorem 4.3 is verified by codimension- 2 bifurcation diagram in Figure 5, which includes Bogdanov-Takens bifurcation point generating the curves of saddle-node, Hopf and homoclinic bifurcation. For τ≠0, we find time delay induce the superscribe Hopf bifurcation under the conditions in Theorem 5.1, which is verified by the bifurcation diagram of x with respect to τ2 and phase portraits of x and y.
Our results give rigorous mathematical proofs of the stability and bifurcation for a two-dimensional p53 GRN to expand the understanding of p53 GRN. Besides, the theoretical analyses of GRNs with noise and space will be further to explored[38,39,40]. In recent years, fractional-order differential equations (FODEs) are used to describe GRNs because they possess memory, after-affects and hereditary properties, which are more compatible with reality than the integer-order differential equations[41,42,43,44]. Therefore, it is worth to explore the stability and bifurcation of GRNs described by three or four dimensional FODEs.
The authors declare that they have not used Artificial Intelligence tools in the creation of this article.
This work is supported by the National Natural Science Foundation of China (Nos. 12062017 and 12262025), Natural Science Foundation of Inner Mongolia Autonomous Region of China (Grants 2021ZD01), and Program for Innovative Research Team in Universities of Inner Mongolia Autonomous Region of China (No. NMGIRT2008). The authors acknowledge the reviewers for their valuable reviews and kind suggestions.
The authors declare that there is no conflict of interest.
[1] |
Q. Zhu, J. Shen, F. Han, W. Lu, Bifurcation analysis and probabilistic energy landscapes of two-component genetic network, IEEE Access, 8 (2020), 150696–150708. https://doi.org/10.1109/ACCESS.2020.3013615 doi: 10.1109/ACCESS.2020.3013615
![]() |
[2] |
L. Fang, Y. Li, L. Ma, Q. Xu, F. Tan, G. Chen, GRNdb: decoding the gene regulatory networks in diverse human and mouse conditions, Nucleic Acids Res., 49 (2021), D97–D103. https://doi.org/10.1093/nar/gkaa995 doi: 10.1093/nar/gkaa995
![]() |
[3] |
X. Zhang, X. Zhao, K. He, L. Lu, Y. Cao, J. Liu, et al., Inferring gene regulatory networks from gene expression data by path consistency algorithm based on conditional mutual information, Bioinformatics, 28 (2012), 98–104. https://doi.org/10.1093/bioinformatics/btr626 doi: 10.1093/bioinformatics/btr626
![]() |
[4] |
T. Yu, X. Zhang, G. Zhang, B. Niu, Hopf bifurcation analysis for genetic regulatory networks with two delays, Neurocomputing, 164 (2015), 190–200. https://doi.org/10.1016/j.neucom.2015.02.070 doi: 10.1016/j.neucom.2015.02.070
![]() |
[5] |
B. S. Stikker, R. W. Hendriks, R. Stadhouders, Decoding the genetic and epigenetic basis of asthma, Allergy, 78 (2023), 940–956. https://doi.org/10.1111/all.15666 doi: 10.1111/all.15666
![]() |
[6] |
B. Huang, M. Lu, M. Galbraith, H. Levine, J. N. Onuchic, D. Jia, Decoding the mechanisms underlying cell-fate decision-making during stem cell differentiation by random circuit perturbation, J. R. Soc. Interface, 17 (2020), 20200500. https://doi.org/10.1098/rsif.2020.0500 doi: 10.1098/rsif.2020.0500
![]() |
[7] |
A. Ghaffarizadeh, G. J. Podgorski, N. S. Flann, Applying attractor dynamics to infer gene regulatory interactions involved in cellular differentiation, Biosystems, 155 (2017), 29–41. https://doi.org/10.1016/j.biosystems.2016.12.004 doi: 10.1016/j.biosystems.2016.12.004
![]() |
[8] |
S. Vyas, A. J. Rodrigues, J. M. Silva, F. Tronche, O. F. Almeida, N. Sousa, et al., Chronic stress and glucocorticoids: from neuronal plasticity to neurodegeneration, Neural Plast., 2016 (2016), 6391686. https://doi.org/10.1155/2016/6391686 doi: 10.1155/2016/6391686
![]() |
[9] |
J. Chrol-Cannon, Y. Jin, Computational modeling of neural plasticity for self-organization of neural networks, Biosystems, 125 (2014), 43–54. https://doi.org/10.1016/j.biosystems.2014.04.003 doi: 10.1016/j.biosystems.2014.04.003
![]() |
[10] |
Y. Meng, Y. Jin, J. Yin, Modeling activity-dependent plasticity in BCM spiking neural networks with application to human behavior recognition, IEEE Trans. Neural Networks, 22 (2011), 1952–1966. https://doi.org/10.1109/TNN.2011.2171044 doi: 10.1109/TNN.2011.2171044
![]() |
[11] |
X. Shi, M. Sun, H. Liu, Y. Yao, R. Kong, F. Chen, et al., A critical role for the long non-coding RNA GAS5 in proliferation and apoptosis in non-small-cell lung cancer, Mol. Carcinog., 54 (2015), E1–E12. https://doi.org/10.1002/mc.22120 doi: 10.1002/mc.22120
![]() |
[12] |
D. Chudasama, V. Bo, M. Hall, V. Anikin, J. Jeyaneethi, J. Gregory, et al., Identification of cancer biomarkers of prognostic value using specific gene regulatory networks (GRN): a novel role of RAD51AP1 for ovarian and lung cancers, Carcinogenesis, 39 (2018), 407–417. https://doi.org/10.1093/carcin/bgx122 doi: 10.1093/carcin/bgx122
![]() |
[13] |
H. C. Lo, J. H. Hsu, L. C. Lai, M. H. Tsai, E. Y. Chuang, MicroRNA-107 enhances radiosensitivity by suppressing granulin in PC-3 prostate cancer cells, Sci. Rep., 10 (2020), 14584. https://doi.org/10.1038/s41598-020-71128-1 doi: 10.1038/s41598-020-71128-1
![]() |
[14] |
J. Eliaš, C. K. Macnamara, Mathematical modelling of p53 signalling during DNA damage response: a survey, Int. J. Mol. Sci., 22 (2021), 10590. https://doi.org/10.3390/ijms221910590 doi: 10.3390/ijms221910590
![]() |
[15] |
Q. Zheng, J. Shen, Z. Wang, Pattern formation and oscillations in Reaction-Diffusion model with p53-Mdm2 feedback Loop, Int. J. Bifurcation Chaos, 29 (2019), 1930040. https://doi.org/10.1142/S0218127419300404 doi: 10.1142/S0218127419300404
![]() |
[16] |
Y. Bi, Q. Liu, L. Wang, W. Yang, X. Wu, Bifurcation and potential landscape of p53 dynamics depending on pdcd5 level and atm degradation rate, Int. J. Bifurcation Chaos, 30 (2020), 2050134. https://doi.org/10.1142/S0218127420501345 doi: 10.1142/S0218127420501345
![]() |
[17] |
Y. Bi, Z. Yang, C. Zhuge, J. Lei, Bifurcation analysis and potential landscapes of the p53-mdm2 module regulated by the co-activator programmed cell death 5, Chaos, 25 (2015), 113103. https://doi.org/10.1063/1.4934967 doi: 10.1063/1.4934967
![]() |
[18] |
J. Hou, Q. Liu, H. Yang, L. Wang, Y. Bi, Stability and bifurcation analyses of p53 gene regulatory network with time delay, Electron. Res. Arch., 30 (2022), 850–873. https://doi.org/10.3934/era.2022045 doi: 10.3934/era.2022045
![]() |
[19] |
C. Gao, F. Chen, Dynamics of p53 regulatory network in DNA damage response, Appl. Math. Modell., 88 (2020), 701–714. https://doi.org/10.1016/j.apm.2020.06.057 doi: 10.1016/j.apm.2020.06.057
![]() |
[20] |
Y. Song, X. Cao, T. Zhang, Bistability and delay-induced stability switches in a cancer network with the regulation of microRNA, Commun. Nonlinear Sci. Numer. Simul., 54 (2018), 302–319. https://doi.org/10.1016/j.cnsns.2017.06.008 doi: 10.1016/j.cnsns.2017.06.008
![]() |
[21] |
T. Sun, R. Yuan, W. Xu, F. Zhu, P. Shen, Exploring a minimal two-component p53 model, Phys. Biol., 7 (2010), 036008. https://doi.org/10.1088/1478-3975/7/3/036008 doi: 10.1088/1478-3975/7/3/036008
![]() |
[22] | L. Perko, Differential Equations and Dynamical Systems, 1991. https://doi.org/10.1007/978-1-4684-0392-3 |
[23] |
B. Hat, M. Kochanczyk, M. N. Bogdal, T. Lipniacki, Feedbacks, bifurcations, and cell fate decision-making in the p53 system, PLoS Comput. Biol., 12 (2016), e1004787. https://doi.org/10.1371/journal.pcbi.1004787 doi: 10.1371/journal.pcbi.1004787
![]() |
[24] | Y. A. Kuznetsov, Elements of Applied Bifurcation Theory, 1998. https://doi.org/10.1007/b98848 |
[25] |
D. Mu, C. Xu, Z. Liu, Y. Pang, Further insight into bifurcation and hybrid control tactics of a chlorine dioxide-iodine-malonic acid chemical reaction model incorporating delays, Match-Commun. Math. Comput. Chem., 89 (2023), 529–566. https://doi.org/10.46793/match.89-3.529M doi: 10.46793/match.89-3.529M
![]() |
[26] |
Y. Xiang, Y. Jiao, X. Wang, R. 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
![]() |
[27] |
M. Sui, Y. Du, Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay, Electron. Res. Arch., 31 (2023), 5124–5150. https://doi.org/10.3934/era.2023262 doi: 10.3934/era.2023262
![]() |
[28] |
C. Wang, F. Yan, H. Liu, Y. Zhang, Theoretical study on the oscillation mechanism of p53-Mdm2 network, Int. J. Biomath., 11 (2018), 1850112. https://doi.org/10.1142/S1793524518501127 doi: 10.1142/S1793524518501127
![]() |
[29] |
Y. Bi, Y. Li, J. Hou, Q. Liu, Multiple time delays induced dynamics of p53 gene regulatory network, Int. J. Bifurcation Chaos, 31 (2021), 2150234. https://doi.org/10.1142/S0218127421502345 doi: 10.1142/S0218127421502345
![]() |
[30] |
J. Xia, X. Li, Bifurcation analysis in a discrete predator-prey model with herd behaviour and group defense, Electron. Res. Arch., 31 (2023), 4484–4506. https://doi.org/10.3934/era.2023229 doi: 10.3934/era.2023229
![]() |
[31] |
D. Hu, H. Cao, Stability and bifurcation analysis in a predator-prey system with Michaelis-Menten type predator harvesting, Nonlinear Anal. Real World Appl., 33 (2017), 58–82. https://doi.org/10.1016/j.nonrwa.2016.05.010 doi: 10.1016/j.nonrwa.2016.05.010
![]() |
[32] |
M. Liu, F. Meng, D. Hu, Bogdanov-Takens and Hopf bifurcations analysis of a genetic regulatory network, Qual. Theory Dyn. Syst., 21 (2022), 45. https://doi.org/10.1007/s12346-022-00575-0 doi: 10.1007/s12346-022-00575-0
![]() |
[33] |
H. Zhou, B. Tang, H. Zhu, S. Tang, Bifurcation and dynamic analyses of non-monotonic predator-prey system with constant releasing rate of predators, Qual. Theory Dyn. Syst., 21 (2022), 10. https://doi.org/10.1007/s12346-021-00539-w doi: 10.1007/s12346-021-00539-w
![]() |
[34] |
C. Shan, Y. Yi, H. Zhu, Nilpotent singularities and dynamics in an SIR type of compartmental model with hospital resources, J. Differ. Equations, 260 (2016), 4339–4365. https://doi.org/10.1016/j.jde.2015.11.009 doi: 10.1016/j.jde.2015.11.009
![]() |
[35] |
C. Xu, Q. Cui, Z. Liu, Y. Pan, X. Cui, W. Ou, et al., Extended hybrid controller design of bifurcation in a delayed chemostat model, Match-Commun. Math. Comput. Chem., 90 (2023), 609–648. https://doi.org/10.46793/match.90-3.609X doi: 10.46793/match.90-3.609X
![]() |
[36] | J. Sotomayor, Generic bifurcations of dynamical systems, Dyn. Syst., (1973), 561–582. https://doi.org/10.1016/B978-0-12-550350-1.50047-3 |
[37] | B. D. Hassard, N. D. Kazarinoff, Y. H. Wan, Theory and Applications of Hopf Bifurcation, 1981. |
[38] |
P. Wan, Dynamic behavior of stochastic predator-prey system, Electron. Res. Arch., 31 (2023), 2925–2939. https://doi.org/10.3934/era.2023147 doi: 10.3934/era.2023147
![]() |
[39] |
Y. Hou, C. Wei, Y. Ding, Dynamic analysis of reaction-diffusion dual carbon model considering economic development in China, Electron. Res. Arch., 31 (2023), 2438–2500. https://doi.org/10.3934/era.2023126 doi: 10.3934/era.2023126
![]() |
[40] |
W. Li, H. Wang, Dynamics of a three-molecule autocatalytic Schnakenberg model with cross-diffusion: turing patterns of spatially homogeneous Hopf bifurcating periodic solutions, Electron. Res. Arch., 31 (2023), 4139–4154. https://doi.org/10.3934/era.2023211 doi: 10.3934/era.2023211
![]() |
[41] |
C. Xu, Z. Liu, P. Li, J. Yan, L. Yao, Bifurcation mechanism for fractional-order three-triangle multi-delayed neural networks, Neural Process. Lett., 55 (2023), 6125–6151. https://doi.org/10.1007/s11063-022-11130-y doi: 10.1007/s11063-022-11130-y
![]() |
[42] |
C. Xu, D. Mu, Y. Pan, C. Aouiti, L. Yao, Exploring bifurcation in a fractional-order predator-prey system with mixed delays, J. Appl. Anal. Comput., 13 (2023), 1119–1136. https://doi.org/10.11948/20210313 doi: 10.11948/20210313
![]() |
[43] |
P. Li, Y. Lu, C. Xu, J. Ren, Insight into Hopf bifurcation and control methods in fractional order BAM neural networks incorporating symmetric structure and delay, Cognit. Comput., 15 (2023), 1825–1867. https://doi.org/10.1007/s12559-023-10155-2 doi: 10.1007/s12559-023-10155-2
![]() |
[44] |
P. Li, X. Peng, C. Xu, L. Han, S. Shi, Novel extended mixed controller design for bifurcation control of fractional-order Myc/E2F/miR-17-92 network model concerning delay, Math. Methods Appl. Sci., 46 (2023), 18878–18898. https://doi.org/10.1002/mma.9597 doi: 10.1002/mma.9597
![]() |
1. | Yuanhong Bi, Xiaoqi Zhang, Quansheng Liu, Deterministic and stochastic dynamic analysis of a Parkinson’s disease model, 2025, 195, 09600779, 116207, 10.1016/j.chaos.2025.116207 |
Case | C0 | C1 | C2 | C3 | C4 | C5 | C6 | Numberofsignchanges | Numberofpossiblepositiveequilibria(E) |
1 | + | − | − | − | − | − | − | 1 | 1 |
2 | + | + | − | − | − | − | − | 1 | 1 |
3 | + | + | + | − | − | − | − | 1 | 1 |
4 | + | + | + | + | − | − | − | 1 | 1 |
5 | + | + | + | + | + | − | − | 1 | 1 |
6 | + | + | + | + | + | + | − | 1 | 1 |
7 | + | + | − | + | − | − | − | 3 | 1, 3 |
8 | + | − | − | + | − | − | − | 3 | 1, 3 |
9 | + | − | + | + | − | − | − | 3 | 1, 3 |
10 | + | − | − | − | − | + | − | 3 | 1, 3 |
11 | + | + | − | − | − | + | − | 3 | 1, 3 |
12 | + | + | + | − | + | + | − | 3 | 1, 3 |
13 | + | − | + | + | + | + | − | 3 | 1, 3 |
14 | + | + | + | − | + | − | − | 3 | 1, 3 |
15 | + | + | + | + | − | + | − | 3 | 1, 3 |
16 | + | − | + | + | + | − | − | 3 | 1, 3 |
17 | + | + | + | − | − | + | − | 3 | 1, 3 |
18 | + | − | + | − | − | − | − | 3 | 1, 3 |
19 | + | + | − | + | − | + | − | 5 | 1, 3, 5 |
20 | + | − | + | − | − | + | − | 5 | 1, 3, 5 |
21 | + | − | + | + | − | + | − | 5 | 1, 3, 5 |
22 | + | − | − | + | − | + | − | 5 | 1, 3, 5 |
23 | + | − | + | − | + | − | − | 5 | 1, 3, 5 |
24 | + | − | + | − | + | + | − | 5 | 1, 3, 5 |
Case | Conditions | Eigenvalues | Properties | |
1 | F′(x∗)<0 | tr(J)<0 | Reλ1<0,Reλ2<0 | Asymptotically stable |
2 | tr(J)=0 | λ1=−i√det(J),λ2=i√det(J) | Linear center | |
3 | tr(J)>0 | Reλ1>0,Reλ2>0 | Unstable | |
4 | F′(x∗)=0 | tr(J)<0 | λ1=tr(J)<0,λ2=0 | Non-hyperbolic |
5 | tr(J)=0 | λ1=λ2=0 | Non-hyperbolic | |
6 | tr(J)>0 | λ1=0,λ2=tr(J)>0 | Unstable(Non-hyperbolic) | |
7 | F′(x∗)>0 | ∀tr(J) | λ1λ2<0 | Unstable(saddle) |
v3 | Ei(x∗,y∗) | F′(x∗) | tr(J) | Stability | Phase portraits |
0.45 | (2.1967,3.5089) | −0.2792 | −0.0273 | Asymptotically stable | Figure 3(b) |
0.4606397 | (0.7593,0.5674) | 0 | −0.0263 | Non-hyperbolic(saddle-node) | Figure 3(c) |
(2.1538,3.4845) | −0.2573 | −0.0267 | Asymptotically stable | Figure 3(c) | |
0.55 | (0.6400,0.4910) | −0.03146 | −0.0305 | Asymptotically stable | Figure 3(d) |
(1.0352,1.1842) | 0.0246 | −0.0206 | Unstable(saddle) | Figure 3(d) | |
(1.7665,3.0254) | −0.0981 | −0.0218 | Asymptotically stable | Figure 3(d) | |
0.5991365 | (0.6186,0.4999) | −0.0391 | −0.0315 | Asymptotically stable | Figure 3(e) |
(1.3543,2.0872) | 0 | −0.0190 | Non-hyperbolic(saddle-node) | Figure 3(e) | |
0.61 | (0.6147,0.5025) | −0.0405 | −0.0316 | Asymptotically stable | Figure 3(f) |
Case | C0 | C1 | C2 | C3 | C4 | C5 | C6 | Numberofsignchanges | Numberofpossiblepositiveequilibria(E) |
1 | + | − | − | − | − | − | − | 1 | 1 |
2 | + | + | − | − | − | − | − | 1 | 1 |
3 | + | + | + | − | − | − | − | 1 | 1 |
4 | + | + | + | + | − | − | − | 1 | 1 |
5 | + | + | + | + | + | − | − | 1 | 1 |
6 | + | + | + | + | + | + | − | 1 | 1 |
7 | + | + | − | + | − | − | − | 3 | 1, 3 |
8 | + | − | − | + | − | − | − | 3 | 1, 3 |
9 | + | − | + | + | − | − | − | 3 | 1, 3 |
10 | + | − | − | − | − | + | − | 3 | 1, 3 |
11 | + | + | − | − | − | + | − | 3 | 1, 3 |
12 | + | + | + | − | + | + | − | 3 | 1, 3 |
13 | + | − | + | + | + | + | − | 3 | 1, 3 |
14 | + | + | + | − | + | − | − | 3 | 1, 3 |
15 | + | + | + | + | − | + | − | 3 | 1, 3 |
16 | + | − | + | + | + | − | − | 3 | 1, 3 |
17 | + | + | + | − | − | + | − | 3 | 1, 3 |
18 | + | − | + | − | − | − | − | 3 | 1, 3 |
19 | + | + | − | + | − | + | − | 5 | 1, 3, 5 |
20 | + | − | + | − | − | + | − | 5 | 1, 3, 5 |
21 | + | − | + | + | − | + | − | 5 | 1, 3, 5 |
22 | + | − | − | + | − | + | − | 5 | 1, 3, 5 |
23 | + | − | + | − | + | − | − | 5 | 1, 3, 5 |
24 | + | − | + | − | + | + | − | 5 | 1, 3, 5 |
Case | Conditions | Eigenvalues | Properties | |
1 | F′(x∗)<0 | tr(J)<0 | Reλ1<0,Reλ2<0 | Asymptotically stable |
2 | tr(J)=0 | λ1=−i√det(J),λ2=i√det(J) | Linear center | |
3 | tr(J)>0 | Reλ1>0,Reλ2>0 | Unstable | |
4 | F′(x∗)=0 | tr(J)<0 | λ1=tr(J)<0,λ2=0 | Non-hyperbolic |
5 | tr(J)=0 | λ1=λ2=0 | Non-hyperbolic | |
6 | tr(J)>0 | λ1=0,λ2=tr(J)>0 | Unstable(Non-hyperbolic) | |
7 | F′(x∗)>0 | ∀tr(J) | λ1λ2<0 | Unstable(saddle) |
v3 | Ei(x∗,y∗) | F′(x∗) | tr(J) | Stability | Phase portraits |
0.45 | (2.1967,3.5089) | −0.2792 | −0.0273 | Asymptotically stable | Figure 3(b) |
0.4606397 | (0.7593,0.5674) | 0 | −0.0263 | Non-hyperbolic(saddle-node) | Figure 3(c) |
(2.1538,3.4845) | −0.2573 | −0.0267 | Asymptotically stable | Figure 3(c) | |
0.55 | (0.6400,0.4910) | −0.03146 | −0.0305 | Asymptotically stable | Figure 3(d) |
(1.0352,1.1842) | 0.0246 | −0.0206 | Unstable(saddle) | Figure 3(d) | |
(1.7665,3.0254) | −0.0981 | −0.0218 | Asymptotically stable | Figure 3(d) | |
0.5991365 | (0.6186,0.4999) | −0.0391 | −0.0315 | Asymptotically stable | Figure 3(e) |
(1.3543,2.0872) | 0 | −0.0190 | Non-hyperbolic(saddle-node) | Figure 3(e) | |
0.61 | (0.6147,0.5025) | −0.0405 | −0.0316 | Asymptotically stable | Figure 3(f) |