
Citation: Wilfredo Lopez, Alexis M. Page, Darby J. Carlson, Brad L. Ericson, Matyas F. Cserhati, Chittibabu Guda, Kimberly A. Carlson. Analysis of immune-related genes during Nora virus infection of Drosophila melanogaster using next generation sequencing[J]. AIMS Microbiology, 2018, 4(1): 123-139. doi: 10.3934/microbiol.2018.1.123
[1] | Shishi Wang, Yuting Ding, Hongfan Lu, Silin Gong . Stability and bifurcation analysis of $ SIQR $ for the COVID-19 epidemic model with time delay. Mathematical Biosciences and Engineering, 2021, 18(5): 5505-5524. doi: 10.3934/mbe.2021278 |
[2] | Hongfan Lu, Yuting Ding, Silin Gong, Shishi Wang . Mathematical modeling and dynamic analysis of SIQR model with delay for pandemic COVID-19. Mathematical Biosciences and Engineering, 2021, 18(4): 3197-3214. doi: 10.3934/mbe.2021159 |
[3] | Sarita Bugalia, Jai Prakash Tripathi, Hao Wang . Mathematical modeling of intervention and low medical resource availability with delays: Applications to COVID-19 outbreaks in Spain and Italy. Mathematical Biosciences and Engineering, 2021, 18(5): 5865-5920. doi: 10.3934/mbe.2021295 |
[4] | Honghua Bin, Daifeng Duan, Junjie Wei . Bifurcation analysis of a reaction-diffusion-advection predator-prey system with delay. Mathematical Biosciences and Engineering, 2023, 20(7): 12194-12210. doi: 10.3934/mbe.2023543 |
[5] | Yuting Ding, Gaoyang Liu, Yong An . Stability and bifurcation analysis of a tumor-immune system with two delays and diffusion. Mathematical Biosciences and Engineering, 2022, 19(2): 1154-1173. doi: 10.3934/mbe.2022053 |
[6] | Fang Liu, Yanfei Du . Spatiotemporal dynamics of a diffusive predator-prey model with delay and Allee effect in predator. Mathematical Biosciences and Engineering, 2023, 20(11): 19372-19400. doi: 10.3934/mbe.2023857 |
[7] | Xin-You Meng, Yu-Qian Wu . Bifurcation analysis in a singular Beddington-DeAngelis predator-prey model with two delays and nonlinear predator harvesting. Mathematical Biosciences and Engineering, 2019, 16(4): 2668-2696. doi: 10.3934/mbe.2019133 |
[8] | Jinhu Xu, Yicang Zhou . Bifurcation analysis of HIV-1 infection model with cell-to-cell transmission and immune response delay. Mathematical Biosciences and Engineering, 2016, 13(2): 343-367. doi: 10.3934/mbe.2015006 |
[9] | Xinyu Liu, Zimeng Lv, Yuting Ding . Mathematical modeling and stability analysis of the time-delayed $ SAIM $ model for COVID-19 vaccination and media coverage. Mathematical Biosciences and Engineering, 2022, 19(6): 6296-6316. doi: 10.3934/mbe.2022294 |
[10] | Zuolin Shen, Junjie Wei . Hopf bifurcation analysis in a diffusive predator-prey system with delay and surplus killing effect. Mathematical Biosciences and Engineering, 2018, 15(3): 693-715. doi: 10.3934/mbe.2018031 |
Epidemic refers to the disease that can spread in a certain scale within a relatively short period of time. It caused a major hazard to human health. For example, the H1N1 influenza virus, which erupted in Mexico in 2009 and soon spread around the globe, is a dangerous virus and can develop to pneumonia. Some infectives may even appear respiratory failure, multiple organ damage, and finally be kicked to death. The H1N1 has killed about 16 thousand people throughout the world then and over 1.3 million people had been infected. Studying the transmission mechanism of epidemics can help reduce the spread of the diseases. Many researchers use the mathematical models to describe the transmission of epidemics [1], and various methods are derived to investigate the dynamics of the diseases [2,3].
Researchers have proposed mathematical models of infectious diseases and studied their dynamical behaviours extensively [4,5]. The dynamical behaviours of ordinary differential equations have been widely studied, and abundant dynamical behaviours have been found, such as periodic oscillations, bifurcations, stable limit cycles and time-delay effects [6,7,8,9,10]. To investigate the dynamics of cholera which spread in the European Mediterranean regions, Capasso and Paveri-fontana proposed an epidemic model in 1979 as follows [6]:
$
{˙u(t)=−α1u(t)+av(t),˙v(t)=−α2v(t)+g(u(t)),
$
|
(1.1) |
where $ u $ and $ v $ represent the concentration of the infectious agent and the infected population, respectively. $ \frac{1}{\alpha_1} $ denotes the average lifetime of the infectious agent, $ \frac{1}{\alpha_2} $ denotes the average infectious period of infected people, thus $ \alpha_1 $, $ \alpha_2 > 0 $. $ a $ denotes the multiplicative factor of the virus caused by the increasing number of infected people. The function $ g(u) $ represents the infectivity of the virus to people, and can be affected by the concentration of the virus. They speculated that the infection process follow the nonlinear saturation pattern. They draw the conclusion of epidemic evolution by analyzing the equations in phase space, which provided a reference for public health policies. The validity of the model is verified by comparing it with available data from Bari town in Italy.
In fact, system (1.1) can be modified to simulate other epidemics with oral-faecal transmission [6], such as typhoid fever, infectious hepatitis, poliomyelitis, etc. Moreover, system (1.1) can also be used to investigate the spreading of man-environment epidemics [6,11,12]. The man-environment epidemics refer that, the infected people may increase the concentration of the virus in the environment, thus people maybe infected by eating the contaminated food [11].
Scientific researches show that the virus is highly sensitive to its environment. The concentration of the virus is not evenly distributed in space thus the effect of diffusion cannot be ignored. In 1997, Capasso and Wilson [11] studied the following diffusive epidemic model:
$ {ut(x,t)=d1Δu(x,t)−α1u(x,t)+av(x,t),x∈(0,l),t>0,vt(x,t)=−α2v(x,t)+g(u),x∈(0,l),t>0,u(0,t)=u(l,t)=0,t>0,u(x,0)=u0(x),v(x,0)=v0(x),x∈(0,l), $
|
(1.2) |
where $ u(x, t), \; v(x, t) $ represent the spatial concentration of the bacteria and the infected people, respectively. $ \alpha_1 $, $ \alpha_2 $ are the same as they are in (1.1). For system (1.2), they performed a detailed analysis of the steady-state bifurcation at the endpoints of a one-dimensional interval based on the monotone techniques. They speculated that the system (1.2) has saddle point structure in the natural function space, similar to the ODE case in which the diffusivity of bacteria is also set to zero.
In the transmission of the epidemics, an important factor that can not be ignored is the delay [13,14,15]. For example, there is time delay between the release of the virus from the infected person and its successful entry into the susceptible people. For those infected people, it also takes time from being infected to being able to release the virus. All of these can lead to time-delay feedback in the epidemics [16,17,18,19]. Time delay usually destroys the stability of the solutions, devotes to the complex dynamics of the system. Recently, systems with two delays have been used in characterizing the phenomena in various fields, such as power systems and predator-prey systems [20,21,22,23,24], etc.. Among them, Wu and Hsu [24] proposed a monostable epidemic model with double delays as follows:
$ {ut(x,t)=d1Δu(x,t)−α1u(x,t)+h(v(x,t−τ1)),vt(x,t)=d2Δv(x,t)−α2v(x,t)+g(u(x,t−τ2)). $
|
(1.3) |
They studied the existence of entire solutions in the system (1.3) with and without the quasi-monotone condition. Some special cases of system (1.3) have been extensively investigated. For example, when $ \tau_1 = d_2 = 0 $, $ \tau_2 $, $ d_1 > 0 $, Zhao and Wang [25] studied the properties of the existence and non-existence of the traveling wave fronts; when $ \tau_1 = \tau_2 = d_2 = 0 $, $ d_1 > 0 $, Xu and Zhao [26] proved the existence, uniqueness and global exponential stability of the bistable traveling wave fronts; Hsu et al. [27] studied the existence and exponential stability of traveling wave solutions in an extended form of system (1.3).
Motivated by the previous works, we study an epidemic model with two delays subjecting to the Neumann boundary conditions in the following form:
$ {∂u(x,t)∂t=d1Δu(x,t)−α1u(x,t)+S(v(x,t−τ1)),x∈(0,lπ),t>0,∂v(x,t)∂t=d2Δv(x,t)−α2v(x,t)+T(u(x,t−τ2)),x∈(0,lπ),t>0,∂u(x,t)∂x=0,∂v(x,t)∂x=0,x=0,lπ;t>0,u(x,t)=u0(x,t)≥0,v(x,t)=v0(x,t)≥0,x∈(0,lπ),t∈[−¯τ,0], $
|
(1.4) |
where $ u(x, t) $ and $ v(x, t) $ denote the concentration of the virus and the number of the infected people, respectively. $ \frac{1}{\alpha_1} $ denotes the average lifetime of the infectious agent, $ \frac{1}{\alpha_2} $ denotes the average infectious period of infected people, $ S(v) $ represents the contribution of infected people to the growth of the virus, $ T(u) $ represents the infection rate of the virus to people. $ d_1, d_2 > 0 $ are the diffusion coefficients, $ \tau_1 $ represents the average time taken by the infected people from being infected to being able to release the virus, $ \tau_2 $ represents the average time taken by the virus from being released to entrying the susceptible people and $ \overline{\tau} = \max\{\tau_1, \tau_2\} $.
In the spread of an epidemic, the concentration of the virus would increase with the number of the infected people, and the susceptible people would more likely to be infected. Therefore, $ S(v) $ and $ T(u) $ are increasing functions of $ v $ and $ u $ respectively in general. However, the increasing number of the virus poses a threat to public health, people will take effective measures to avoid them from being infected by the virus, including isolating the infected people, regular disinfection, reminding people to pay attention to hygiene through media reports and so on [28,29]. Scientific studies have shown that these measures are effective in reducing infection rates [30]. Once the virus exceeds a certain concentration $ u_0 $, even if the concentration continues to increase, people begin to take certain measures to reduce the infectivity of the virus. So in this paper, we assume that once if $ u > u_0 $, $ T(u) $ is a decreasing function with respect to $ u $.
In this paper, we study the existence of the positive constant steady state. By analyzing the zeros of characteristic equations of (1.4), we investigate the local stability of constant steady state. We found that when the sum of two delays reaches a critical value, the system undergoes Hopf bifurcation. By taking one of two delays as a bifurcation parameter, we calculate the normal form on the center manifold near the Hopf singularity.
This paper is organized as follows. In section 2, the stability of the positive constant steady state of the system (1.4) and the existence of Hopf bifurcation are established by analyzing the distribution of the eigenvalues. In section 3, the normal form on the center manifold near the Hopf bifurcation point is derived. In section 4, we carry out some numerical simulations to illustrate the theoretic results.
In this section, we will study the local stability of the positive constant steady state and the existence of Hopf bifurcation by analyzing the distribution of the eigenvalues.
For $ S(v) $ and $ T(u) $ in (1.4), we make the following assumption:
$ (H1){S,T∈C3(R,R),S(0)=T(0)=0, there exists a unique positive constantv∗∈R,such thatT(S(v∗)/α1)=α2v∗. $
|
Normally speaking, the contribution function $ S(v) $ is sufficiently smooth with respect to the infected people $ v $, the infection rate $ T(u) $ is sufficiently smooth with respect to $ u $. For the research purposes, we assume $ S, T\in C^3(\mathbb{R}, \mathbb{R}) $. Particularly, if no one is infected ($ v = 0 $), then no infected people contribute to the growth of the virus, i.e., $ S(0) = 0 $. Similarly, if the concentration of the virus is zero($ u = 0 $), then no one can be infected by the virus, i.e., $ T(0) = 0 $. In the early stage of a transmission, the virus concentration is low but the number of the susceptible people is large, thus the epidemic would spread at a certain speed; as the epidemic continues to spread, although the virus concentration is high, a reduction in the number of the susceptible people will limit the spread of the epidemic. Thus the virus and the population may reach a coexistence state. In $ {\bf (H_1)} $, we denote the number of infected people in this coexistence state as $ v^* $.
Define $ u^* = S(v^*)/\alpha_1 $, under the assumption $ (\bf H_1) $ system (4) has two constant steady states $ E_0(0, 0) $ and $ E_*(u^*, v^*) $. Obviously, $ E_* $ is a positive constant steady state. To investigate the dynamics of the epidemics, we focus on the properties of $ E_* $.
$ S_{v^*} = \frac{\partial S}{\partial v}\bigg|_{v^*}, \;T_{u^*} = \frac{\partial T}{\partial u}\bigg|_{u^*}, $ |
the linearized equation of system (1.4) at $ E_* $ is
$ ∂∂t(u(x,t)v(x,t))=(DΔ+A)(u(x,t)v(x,t))+B(u(x,t−τ1)v(x,t−τ1))+C(u(x,t−τ2)v(x,t−τ2)), $
|
(2.1) |
where
$ {D=(d100d2),A=(−α100−α2),B=(0Sv∗00),C=(00Tu∗0) }, $
|
and $ u(x, t) $, $ v(x, t) $ still satisfy the homogeneous Neumann boundary conditions.
The characteristic equations of the system (2.1) are given by
$ det(λI2−Mn−A−Be−λτ1−Ce−λτ2)=0,n∈N0, $
|
(2.2) |
where $ I_2 $ is a 2 $ \times $ 2 identity matrix, $ M_n = -\frac{n^2}{l^2}{D} $. Since
$ det(λI2−Mn−A−Be−λτ1−Ce−λτ2)=|λ+d1n2l2+α1Sv∗e−λτ1Tu∗e−λτ2λ+d2n2l2+α2|, $
|
denote $ m = S_{v^*}\times T_{u^*} $, Eq (2.2) becomes
$ (λ+d1n2l2+α1)(λ+d2n2l2+α2)−me−λ(τ1+τ2)=0,n∈N0. $
|
(2.3) |
Denote $ \tau = \tau_1+\tau_2 $, then Eq (2.3) becomes
$ (λ+d1n2l2+α1)(λ+d2n2l2+α2)−me−λτ=0,n∈N0. $
|
(2.4) |
For convenience, denote $ k_1 = d_1\frac{n^2}{l^2}+\alpha_1, \; k_2 = d_2\frac{n^2}{l^2}+\alpha_2 $, then we have $ k_1, k_2 > 0 $. Now the characteristic Eq (2.4) with $ \tau = 0 $ is in the following form:
$ λ2+(k1+k2)λ+k1k2−m=0. $
|
(2.5) |
To ensure all the roots of Eq (2.5) have negative real parts, we further make the following assumption
$ {\bf (H_2)}\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; m < \alpha_1\alpha_2.\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; $ |
Then we can conclude the following lemma.
Lemma 2.1. When $ \tau = 0 $, all the roots of Eq (2.4) have negative real parts under the assumption $ {\bf(H_2)} $.
Since the proof of Lemma 2.1 is intuitive, here we omit it.
Now we will study the stability of $ E_* $ and investigate the existence of Hopf bifurcation of system (1.4). Assume that $ \lambda = {\rm i}\omega\; (\omega > 0) $ is a root of (2.4). Substituting $ {\rm i}\omega $ into (2.4) and separating the real and imaginary parts leads to
$ {−ω2+k1k2=mcosωτ,(k1+k2)ω=−msinωτ. $
|
(2.6) |
This yields
$ \omega^4+(k_1^2+k_2^2)\omega^2+k_1^2k_2^2-m^2 = 0. $ |
Denote $ z = \omega^2 $, then the above equation becomes
$ z2+(k21+k22)z+k21k22−m2=0, $
|
(2.7) |
the roots of (2.7) can be given by
$ z_{1, 2} = \frac{1}{2}[-(k_1^2+k_2^2)\pm\sqrt{(k_1^2+k_2^2)^2-4(k_1^2k_2^2-m^2)}]. $ |
Clearly, $ z_2 < 0 $. To maintain the existence of a positive $ \omega $ that satisfies (2.6), $ z_1 > 0 $ should hold. Obviously, $ z_1 > 0 $ provided that $ m^2 > k_1^2k_2^2 $.
If $ m^2\leq \alpha_1^2 \alpha_2^2 $, then for all $ n\in\mathbb N_0 $, we have $ m^2\leq (d_1\frac{n^2}{l^2}+\alpha_1)^2(d_2\frac{n^2}{l^2}+\alpha_2)^2 $, and thus $ z_1\leq 0 $, all roots of (2.4) have negative real parts. Moreover, if $ m < \alpha_1\alpha_2 $ and $ m^2\leq \alpha_1^2 \alpha_2^2 $ are satisfied, i.e., $ -\alpha_1\alpha_2 \leq m < \alpha_1\alpha_2 $, then $ E_* $ is locally asymptotically stable. We can draw the following proposition.
Proposition 1. If $ -\alpha_1\alpha_2 \leq m < \alpha_1\alpha_2 $, all roots of (2.4) have negative real parts, then $ E_* $ is locally asymptotically stable.
Now we further make the following assumption:
$ (H3){∃N∈N0,s.t.m2>(d1n2l2+α1)2(d2n2l2+α2)2for0≤n<Nandm2≤(d1N2l2+α1)2(d2N2l2+α2)2. $
|
Then we study the existence of purely imaginary roots of (2.4) in the following lemma.
Lemma 2.2. If $ {\bf(H_3)} $ is satisfied, then Eq (2.4) with $ n\in\{0, 1, \cdots, N-1\} $ has a pair of purely imaginary roots $ \pm {\rm i}\omega_n $ when $ \tau = \tau_{j, n}, $ where
$ ωn=1√2[−(d1n2l2+α1)2−(d2n2l2+α2)2+√[(d1n2l2+α1)2−(d2n2l2+α2)2]2+4m2]12,τj,n=1ωn[arccos−ωn2+(d1n2l2+α1)(d2n2l2+α2)m+2jπ],j∈N0;n=0,1,⋯,N. $
|
(2.8) |
Proof. From $ {\bf(H_3)} $ we know that Eq (2.7) with $ n\in\{0, 1, \cdots, N-1\} $ has a positive root denoted by $ z_{n} $. Hence, $ \omega_n = \sqrt{z_{n}} > 0 $ makes sense. Define
$ \tau_{j, n} = \frac{1}{\omega_n}[\arccos\frac{{-\omega_n}^2+(d_1\frac{n^2}{l^2}+\alpha_1)(d_2\frac{n^2}{l^2}+\alpha_2)}{m}+2j\pi], \; \; j = 0, 1, 2, \cdots. $ |
From the second equation of (2.6) one obtains that $ \tau_{0, n}\omega_n\in(0, \pi] $. Then $ (\omega_n, \tau_{j.n}) $ solves (2.6), which implies that $ \pm{\rm i}\omega_n $ is a pair of purely imaginary roots of Eq (2.4) with $ \tau = \tau_{j, n} $.
Lemma 2.3. If assumptions $ {\bf(H_2)} $ and $ {\bf (H_3)} $ are satisfied, then $ \omega_n $ is monotone decreasing with respect to $ n $.
Proof. By Lemma 2.2, we know that $ \omega_n $ exists for $ n = 0, \; 1, \cdot\cdot\cdot\; N-1 $. Assumptions $ {\bf(H_2)} $ and $ {\bf(H_3)} $ imply that $ m < 0 $. We know that
$ zn=ωn2=12[−(d1n2l2+α1)2−(d2n2l2+α2)2+√[(d1n2l2+α1)2+(d2n2l2+α2)2]2+4[m2−(d1n2l2+α1)2(d2n2l2+α2)2]], $
|
$ dzndn=12{−2(d1n2l2+α1)(2d1nl2)−2(d2n2l2+α2)(2d2nl2)+122[(d1n2l2+α1)2+(d2n2l2+α2)2][2(d1n2l2+α1)(2d1nl2)+2(d2n2l2+α2)(2d2nl2)]√[(d1n2l2+α1)2+(d2n2l2+α2)2]2+4[m2−(d1n2l2+α1)2(d2n2l2+α2)2]−128[(d1n2l2+α1)(2d1nl2)(d2n2l2+α2)2+(d2n2l2+α2)(2d2nl2)(d1n2l2+α1)2]√[(d1n2l2+α1)2+(d2n2l2+α2)2]2+4[m2−(d1n2l2+α1)2(d2n2l2+α2)2]}<12{−2(d1n2l2+α1)(2d1nl2)−2(d2n2l2+α2)(2d2nl2)+[(d1n2l2+α1)2+(d2n2l2+α2)2][2(d1n2l2+α1)(2d1nl2)+2(d2n2l2+α2)(2d2nl2)]√[(d1n2l2+α1)2+(d2n2l2+α2)2]2−4[(d1n2l2+α1)(2d1nl2)(d2n2l2+α2)2+(d2n2l2+α2)(2d2nl2)(d1n2l2+α1)2]√[(d1n2l2+α1)2+(d2n2l2+α2)2]2+4[m2−(d1n2l2+α1)2(d2n2l2+α2)2]}=−2[(d1n2l2+α1)(2d1nl2)(d2n2l2+α2)2+(d2n2l2+α2)(2d2nl2)(d1n2l2+α1)2]√[(d1n2l2+α1)2+(d2n2l2+α2)2]2+4[m2−(d1n2l2+α1)2(d2n2l2+α2)2]<0. $
|
Hence, $ z_n $ is monotone decreasing with respect to $ n $. For $ i = 0, 1, \cdots, N-1 $ we have $ z_{i} > z_{i+1} $. Thus $ \omega_n $ is monotone decreasing with respect to $ n $.
Lemma 2.4. If assumption $ {\bf(H_2)} $ and $ {\bf(H_3)} $ are satisfied, then $ \tau_{0, n}\omega_n $ is monotone increasing with respect to $ n $. Moreover, $ \tau_{0, 0} = \min\{\tau_{0, n}\}_{0\leq n\leq N-1} $.
Proof. From Lemma 2.3, for $ n = 0, 1, \cdots, N-2 $, we have $ \omega_n > \omega_{n+1} $ and
$ \tau_{0, n}\omega_n = \arccos\frac{-{\omega_n}^2+(d_1\frac{n^2}{l^2}+\alpha_1)(d_2\frac{n^2}{l^2}+\alpha_2)}{m}. $ |
By $ -{\omega_n}^2 < -{\omega^2_{n+1}} $ and $ m < 0 $, we have $ \dfrac{-{\omega_n}^2+(d_1\frac{n^2}{l^2}+\alpha_1)(d_2\frac{n^2}{l^2}+\alpha_2)}{m} $ is monotone decreasing with respect to $ n $. Since the anti-trigonometric function $ y = \arccos x $ is monotone decreasing with respect to $ x $, from the properties of composite functions, we can obtain that $ \tau_{0, n}\omega_n $ is monotone increasing with respect to $ n $. The last conclusion follows from $ \omega_n > \omega_{n+1} $.
Under the assumption $ {\bf (H_2)} $ and $ {\bf (H_3)} $, let $ \lambda(\tau) = \alpha(\tau)+{\rm i}\beta(\tau) $ be the root of Eq (2.4) with $ n\in\{0, 1, \cdots, N-1\} $ satisfying $ \alpha(\tau_{j, n}) = 0 $ and $ \beta(\tau_{j, n}) = \omega_n $.
Lemma 2.5. Suppose that $ {\bf (H_2)} $ and $ {\bf (H_3)} $ are satisfied. Then
$ \alpha'(\tau)|_{\tau = \tau_{j, n}} > 0. $ |
Proof. Since Eq (2.4) can be rewritten in the following form:
$ λ2+(k1+k2)λ+k1k2+me−λτ=0. $
|
(2.9) |
Substituting $ \lambda(\tau) $ into (2.9) and taking the derivative of both sides of Eq (2.9) with respect to $ \tau $ gives
$ (dλdτ)−1=−τλ+2λ+(k1+k2)λ[−λ2−(k1+k2)λ−k1k2]. $
|
(2.10) |
Then we have
$ (dλdτ)−1|τ=τj,n=τωi+i[2ω3(k1+k2)−ω(k1+k2)(ω2−k1k2)](k1+k2)2ω4+ω2(ω2−k1k2)2+2ω2(ω2−k1k2)+(k1+k2)2ω2(k1+k2)2ω4+ω2(ω2−k1k2)2. $
|
Hence,
$ sign[α′(τj,n)]=sign[(dReλdτ)−1|τ=τj,n]=sign(2ω2(ω2−k1k2)+(k1+k2)2ω2(k1+k2)2ω4+ω2(ω2−k1k2)2)=sign(2ω4+(k21+k22)ω2(k1+k2)2ω4+ω2(ω2−k1k2)2)>0. $
|
This completes the proof. By Lemmas 2.1, 2.4, 2.5 and applying Corollary 2.4 in [31], we have the following conclusions on the distribution of the roots of the characteristic Eq (2.4).
Lemma 2.6. Suppose that $ {\bf(H_2)} $ and $ {\bf(H_3)} $ are satisfied. Then there exists a sequence values of $ \left\{\tau_{j, n}\right\}(j = 0, 1, 2, \cdots; n = 0, 1, \cdots, N-1) $, such that all the roots of Eq (2.4) have negative real parts when $ \tau_1+\tau_2\in[0, \tau_{0, 0}) $; When $ \tau_1+\tau_2 = \tau_{0, 0} $, all the other roots of Eq (2.4) have negative real parts, except for a pair of pure imaginary roots $ \pm{\rm i}\omega_0 $; When $ \tau_1+\tau_2 > \tau_{0, 0} $, Eq (2.4) has at least a couple of roots with positive real parts.
From Lemma 2.6, we have the following the conclusions on the stability of the positive equilibrium of the system (1.4) and the existence of Hopf bifurcations directly.
Theorem 2.7. Suppose that $ {\bf(H_1), (H_2)} $ and $ {\bf(H_3)} $ are satisfied. Then the positive constant steady state $ E^* $ of system (1.4) is asymptotically stable when $ \tau_1+\tau_2\in[0, \tau_{0, 0}) $, unstable when $ \tau_1+\tau_2 > \tau_{0, 0} $, and system (1.4) undergoes a Hopf bifurcation at $ E_* $ when $ \tau_1+\tau_2 = \tau_{j, n} $, $ j = 0, 1, 2, \cdots; \; \; n = 0, 1, \cdots, N-1. $
In this section, we will calculate the normal form of Hopf bifurcation on the center manifold. To investigate the dynamical behaviours of system (1.4) near the Hopf bifurcation point, we will study the normal form of Hopf bifurcation by making use of the normal form method of partial functional differential equations [32,33].
Without loss of generality, we assume that $ \tau_1 > \tau_2 $, let $ \overline u(x, t) = u(x, \tau_1t)-u^* $, $ \overline v(x, t) = v(x, \tau_1t)-v^* $ and take out of the bars, the system (1.4) becomes
$ ∂∂t(u(x,t)v(x,t))=τ1(DΔ+A)(u(x,t)v(x,t))+τ1B(u(x,t−1)v(x,t−1))+τ1C(u(x,t−τ2τ1)v(x,t−τ2τ1))+τ1(f1f2),x∈(0,lπ),t>0 $
|
(3.1) |
with the boundary and initial conditions:
$ {∂u(x,t)∂x=0,∂v(x,t)∂x=0,x=0,lπ,t>0,u(x,t)=u0(x,t)≥0,v(x,t)=v0(x,t)≥0,x∈(0,lπ),t∈[−1,0], $
|
(3.2) |
where
$ f1=S(v(x,t−1))−(Sv∗)v(x,t−1),f2=T(u(x,t−τ2τ1))−(Tu∗)u(x,t−τ2τ1). $
|
We define the real-valued Hilbert space
$ Y = \left\{(u, v)^T\in H^2(0, l\pi)\times H^2(0, l\pi):\frac{\partial u}{\partial x} = \frac{\partial v}{\partial x} = 0\;at\;x = 0, l\pi\right\} $ |
and the complexification of $ Y $
$ Y_\mathbb{C} = Y\oplus {\rm i}Y = \left\{U_1+{\rm i}U_2: U_1, U_2\in Y\right\} $ |
with the general complex-value $ L^2 $ inner product
$ \left( U, V\right) = \int_{0}^{l\pi} (\overline u_1v_1+\overline u_2v_2)\, dx $ |
for $ U = (u_1, u_2)^T, V = (v_1, v_2)^T \in Y_\mathbb{C} $.
Let $ \mathcal{C}: = C([-1, 0], Y_\mathbb{C}) $ represents the phase space with the sup norm. We write $ u_t\in\mathcal{C} $ for $ u_t(\theta) = u(t+\theta) $, $ -1\leq\theta\leq0 $.
In the previous section, we have obtained that when $ \tau = \tau_{0, 0} $, all the roots of Eq (2.4) have negative real parts except $ \pm{\rm i}\omega_0 $. We denote the Hopf bifurcation point by $ (\tau_1^*, \tau_2^*) $ with $ \tau_1^*+\tau_2^* = \tau_{0, 0} $, and introduce parameter $ \mu $ by setting $ \tau_1 = \tau_1^*+\mu $ with fixing $ \tau_2 = \tau_2^* $ and $ |\mu| $ sufficiently small. Then we have $ \tau = \tau_1+\tau_2 = \tau_1^*+\tau_2^*+\mu = \tau_{0, 0}+\mu $, and $ \mu = 0 $ is the Hopf bifurcation value of system (3.1).
Denote $ U(t) = (u(x, t), v(x, t))^T $ and substitute $ \tau_1 = \tau_1^*+\mu, \tau_2 = \tau_2^* $ into (3.1). We have
$ dU(t)dt=DμΔU(t)+Lμ(Ut)+f(μ,Ut), $
|
(3.3) |
where
$ Dμ=(τ∗1+μ)D,Lμ(Ut)=(τ∗1+μ)(AUt(0)+BUt(−1)+CUt(−τ∗2τ∗1+μ)) $
|
and
$ f(\mu, U_t) = (\tau_1^*+\mu)(f1f2) = (\tau_1^*+\mu) (S(v(x,t−1))−(Sv∗)v(x,t−1)T(u(x,t−τ∗2τ∗1+μ))−(Tu∗)u(x,t−τ∗2τ∗1+μ)) . $
|
Consider the linearized system of (3.3)
$ dU(t)dt=DμΔU(t)+Lμ(Ut). $
|
(3.4) |
In $ Y $, the eigenvalues of $ D\Delta $ are $ -d_1\frac{n^2}{l^2} $ and $ -d_2\frac{n^2}{l^2} $, $ n\in\mathbb N_0 $, with the corresponding normalized eigenvectors: $ \beta_n^{(1)}(x) = \gamma_n(x)e_1 $, $ \beta_n^{(2)}(x) = \gamma_n(x)e_2 $, where
$ γn(x)=cosnlx‖cosnlx‖L2={√1lπ,n=0,√2lπcosnlx,n≥1, $
|
and $ e_1 = (1, 0)^T, e_2 = (0, 1)^T $ are the unit coordinate vectors of $ \mathbb R^2 $.
Define the subspace of $ \mathcal{C} $ as $ \mathcal{B}_n $, by
$ \mathcal{B}_n: = \text{span}\left\{\left( v(\cdot), \beta_n^{(1)}\right)\beta_n^{(1)}, \left( v(\cdot), \beta_n^{(2)}\right)\beta_n^{(2)}\big|v\in\mathcal{C}\right\}, $ |
we write $ \left(v(\cdot), \beta_n\right) = \left(\left(v(\cdot), \beta_n^{(1)}\right), \left(v(\cdot), \beta_n^{(2)}\right)\right)^T $ for simplification. Then on $ \mathcal{B}_n $, the linear Eq (3.4) is equivalent to retarded functional differential equation on $ \mathbb{R}^2 $:
$ ˙x(t)=−n2l2Dμx(t)+Lμxt,n∈N0. $
|
(3.5) |
By Riese representation theorem, there exists a $ 2\times 2 $ matrix function $ \eta_n(\mu, \theta) $, $ \theta\in [-1, 0] $, whose entries are of bounded variation such that
$ -\frac{n^2}{l^2}D_{\mu}\phi(0)+L_{\mu}(\phi) = \int_{-1}^{0} \, d{\eta_n(\mu, \theta)}\phi(\theta), \;\phi\in\mathcal{C}. $ |
In fact, we can choose
$ ηn(μ,θ)={−(τ∗1+μ)B,θ=−1,0,θ∈(−1,−τ∗2τ∗1+μ],(τ∗1+μ)C,θ∈(−τ∗2τ∗1+μ,0),(τ∗1+μ)(A+C−n2l2D),θ=0. $
|
For simplicity, we write $ \eta_n(\theta) $ for $ \eta_n(0, \theta) $, define $ \mathcal{A} $ as the infinitesimal generator of the semigroup generated by Eq (3.5) with $ \mu = 0 $ and $ n = 0 $. Define the adjoint operator $ \mathcal{A}^* $ of $ \mathcal{A} $ on $ \mathcal{C}^*: = C([0, 1], (\mathbb{R}^2)^*) $ as
$ \mathcal{A}^*\psi(s) = -\dot{\psi}(s)+X_0(-s)\Big[ \int_{-1}^{0}\psi(\theta) d\eta_0(\theta)+\dot{\psi}(0)\Big], $ |
with the bilinear form
$ ⟨ψ,φ⟩=¯ψ(0)φ(0)−∫0−1∫θξ=0¯ψ(ξ−θ)dη0(θ)φ(ξ)dξ, $
|
(3.6) |
Let
$ q(\theta) = q(0)e^{{\rm i}\omega_0\tau_1^*\theta}(\theta\in[-1, 0]), \;q^*(s) = q^*(0)e^{{\rm i}\omega_0\tau_1^* s}(s\in[0, 1]) $ |
be the eigenvectors of $ \mathcal{A} $ and $ \mathcal{A}^* $ corresponding to $ {\rm i}\omega_0\tau_1^* $ and $ -{\rm i}\omega_0\tau_1^* $ respectively. Thus
$ q(0) = (1, q_1)^T, \;q^*(0) = \overline M(q_2, 1), $ |
where
$ q1=iω0+α1(Sv∗)e−iω0τ∗1,q2=−iω0+α2(Sv∗)eiω0τ∗1. $
|
We choose $ M $ as
$ M=[q1+¯q2+τ∗1q1¯q2(Sv∗)e−iω0τ∗1+τ∗2(Tu∗)e−iω0τ∗2]−1 $
|
which assures that $ \langle q^*(s), q(\theta)\rangle = 1 $.
Let
$ \mathcal{BC} = \left\{\phi:[-1, 0]\to Y_\mathbb{C}, \;\phi\; is\; continous\; on\; [-1, 0), \;\exists\lim\limits_{\theta\to0^-}\phi(\theta)\in Y_\mathbb{C}\right\}, $ |
and define $ \mathcal{A}_\mu $ as the infinitesimal generator of the $ C_0 $ semi-group of solution maps of the linear equation of (3.3):
$ \mathcal{A}_\mu:\mathcal{C}_0^1\cap\mathcal{BC}\to\mathcal{BC}, \;\mathcal{A_\mu}\phi = \dot{\phi}+X_0[D_\mu\Delta\phi(0)+L_\mu(\phi)-\dot{\phi}(0)] $ |
with
$ \rm dom(\mathcal{A}_\mu) = \left\{\phi\in\mathcal{C}:\dot{\phi}\in\mathcal{C}, \phi(0)\in \rm dom(\Delta)\right\}, $ |
where
$ X_0(\theta) = {0,−1≤θ<0,I,θ=0. $
|
On $ \mathcal{BC} $, Eq (3.3) can be rewritten as an abstract ordinary differential equation
$ dUtdt=AμUt+X0f(μ,Ut). $
|
(3.7) |
Using the same notations in Wu [34], we compute the coordinates to describe the center manifold $ \mathcal C_0 $ at $ \mu = 0 $. Let $ U_t $ be the solutions of Eq (3.3) with $ \mu = 0 $. Define
$ z(t)=⟨q∗,(Ut,β0)⟩,W(t,θ)=Ut(θ)−2Re{z(t)q(θ)}. $
|
On the center manifold $ \mathcal C_0 $, we have
$ W(t,θ)=(z(t),¯z(t),θ), $
|
where
$ z=(z1,z2)T,W(z,¯z,θ)=W20(θ)z22+W11(θ)z¯z+W02(θ)¯z22+W30(θ)z36+⋯, $
|
$ z $ and $ \bar z $ are local coordinates for center manifold $ \mathcal C_0 $ in the direction of $ q^* $ and $ \overline q^* $ respectively. Notice that $ W $ is real if $ U_t $ is real, we consider real solutions.
For the solution $ U_t\in\mathcal C_0 $ of the system (3.3) with $ \mu = 0 $,
$ ˙z(t)=iω0τ∗1z+⟨q∗(θ),f(0,W+2Re{z(t)q(θ)})⟩=iω0τ∗1z+¯q∗(0)f(0,W(z,¯z,0)+2Re{z(t)q(0)})△=iω0τ∗1z+¯q∗(0)F0(z,¯z). $
|
It can be denoted as
$ ˙z(t)=iω0τ∗1z(t)+g(z,¯z), $
|
(3.8) |
where
$ g(z,¯z)=¯q∗(0)f(0,W(z,¯z,0)+2Re{z(t)q(0)})=g20z22+g11z¯z+g02¯z22+g21z2¯z2+⋯. $
|
Base on (3.7) and (3.8), similar to the method used in Zhao and Wei [35], we obtain
$ ˙W=˙Ut−˙zq−˙¯z¯q={A0W−2Re{¯q∗(0)F0q(θ)},θ∈[−1,0),A0W−2Re{¯q∗(0)F0q(θ)}+F0,θ=0,△=A0W+H(z,¯z,θ), $
|
where
$ H(z,¯z,θ)=H20(θ)z22+H11(θ)z¯z+H02(θ)¯z22+⋯. $
|
(3.9) |
By expanding the above series and comparing the coefficients, we have
$ {(A0−2iω0τ∗1)W20(θ)=−H20(θ),A0W11(θ)=−H11(θ),(A0+2iω0τ∗1)W02(θ)=−H02(θ),⋯. $
|
(3.10) |
Notice that
$ q∗(0)=¯M(q2,1),v(t−1)=q1e−iω0τ∗1z+¯q1eiω0τ∗1¯z+W(1)(t,−1),u(t−τ∗2τ∗1)=e−iω0τ∗2z+eiω0τ∗2¯z+W(2)(t,−τ∗2τ∗1), $
|
where
$ W(1)(t,−1)=W(1)20(−1)z22+W(1)11(−1)z¯z+W(1)02(−1)¯z22+⋯,W(2)(t,−τ∗2τ∗1)=W(2)20(−τ∗2τ∗1)z22+W(2)11(−τ∗2τ∗1)z¯z+W(2)02(−τ∗2τ∗1)¯z22+⋯, $
|
and
$ F0=12τ∗1(S″(v∗)v2(t−1)T″(u∗)u2(t−τ∗2τ∗1))+16τ∗1(S‴(v∗)v3(t−1)T‴(u∗)u3(t−τ∗2τ∗1))+⋯. $
|
We have
$ g(z,¯z)=¯q∗(0)F0=M2τ∗1[¯q2(S″(v∗)v2(t−1)+S‴(v∗)3v3(t−1))+T″(u∗)u2(t−τ∗2τ∗1)+T‴(u∗)3u3(t−τ∗2τ∗1)]=M2τ∗1{¯q2[S″(v∗)(q21e−2iω0τ∗1z2+¯q21e2iω0τ∗1¯z2+2q1¯q1z¯z)+(S″(v∗)[¯q1eiω0τ∗1W(1)20(−1)+2q1e−iω0τ∗1W(1)11(−1)]+S‴(v∗)q21¯q1e−iω0τ∗1)z2¯z]+T″(u∗)(e−2iω0τ∗2z2+e2iω0τ∗2¯z2+2z¯z)+(T″(u∗)[eiω0τ∗2W(2)20(−τ∗2τ∗1)+2e−iω0τ∗2W(2)11(−τ∗2τ∗1)]+T‴(u∗)e−iω0τ∗2)z2¯z}. $
|
Comparing the coefficients with (3.9), we obtain that
$ g20=τ∗1M¯q2S″(v∗)q21e−2iω0τ∗1+τ∗1MT″(u∗)e−2iω0τ∗2,g11=2τ∗1M¯q2S″(v∗)q1¯q1+τ∗1MT″(u∗),g02=τ∗1M¯q2S″(v∗)¯q21e2iω0τ∗1+τ∗1MT″(u∗)e2iω0τ∗2,g21=τ∗1M{S″(v∗)[¯q1eiω0τ∗1W(1)20(−1)+2q1e−iω0τ∗1W(1)11(−1)]+S‴(v∗)q21¯q1e−iω0τ∗1+T″(u∗)[eiω0τ∗1W(2)20(−τ∗2τ∗1)+2e−iω0τ∗2W(2)11(−τ∗2τ∗1)]+T‴(u∗)e−iω0τ∗2}. $
|
(3.11) |
We need to compute $ W_{20}(\theta) $ and $ W_{11}(\theta) $ for $ \theta\in[-1, 0) $, since
$ H(z,¯z,θ)=2Re{¯z∗(0)F0q(θ)}=−gq(θ)−¯g¯q(θ)=−(g20z22+g11z¯z+g02¯z22+⋯)q(θ)−(¯g20¯z22+¯g11z¯z+¯g02z22+⋯)¯q(θ). $
|
Comparing the coefficients with (3.9), we obtain
$ H20(θ)=−g20q(θ)−¯g02¯q(θ), $
|
$ H11(θ)=−g11q(θ)−¯g11¯q(θ). $
|
From (3.10), it can be given that
$ ˙W20(θ)=2iω0τ∗1W20(θ)+g20q(0)eiω0τ∗1θ+¯g02¯q(0)e−iω0τ∗1θ. $
|
Solving for $ W_{20}(\theta) $, we have
$ W20(θ)=g20iω0τ∗1q(0)eiω0τ∗1θ−¯g023iω0τ∗1¯q(0)e−iω0τ∗1θ+E1e2iω0τ∗1θ, $
|
(3.12) |
and similarly
$ W11(θ)=g11iω0τ∗1q(0)eiω0τ∗1θ−¯g11iω0τ∗1¯q(0)e−iω0τ∗1θ+E2, $
|
(3.13) |
where $ E_1 $ and $ E_2 $ can be determined by setting $ \theta = 0 $ in $ H $. Since
$ H(z,¯z,0)=−2Re{¯q∗(0)F0q(0)}+F0, $
|
we have
$ H20(0)=−g20q(0)−¯g02¯q(0)+(S″(v∗)q21e−2iω0τ∗1T″(u∗)e−2iω0τ∗2), $
|
(3.14) |
$ H11(0)=−g11q(0)−¯g11¯q(0)+(S″(v∗)q1¯q1T″(u∗)). $
|
(3.15) |
From (3.10) and the definition of $ \mathcal A_0 $, we have
$ (−α100−α2)W20(0)+(0Sv∗00)W20(−1)+(00Tu∗0)W20(−τ∗2τ∗1)=2iω0τ∗1W20(0)−H20(0), $
|
(3.16) |
$ (−α100−α2)W11(0)+(0Sv∗00)W11(−1)+(00Tu∗0)W11(−τ∗2τ∗1)=−H11(0). $
|
Noting that
$ (−1−iω0Sv∗e−iω0τ∗1Tu∗e−iω0τ∗2−1−iω0)q(0)=0, $
|
substituting (3.12) into (3.16), we obtain
$ (−1−2iω0τ∗1Sv∗e−2iω0τ∗1Tu∗e−2iω0τ∗2−1−2iω0τ∗1)E1=−g20q(0)−¯g02¯q(0)−H20(0). $
|
(3.17) |
Substituting (3.14) into this, we have
$ (−(1+2iω0τ∗1)Sv∗e−2iω0τ∗1Tu∗e−2iω0τ∗2−(1+2iω0τ∗1))E1=(−S″(v∗)q21e−2iω0τ∗1−T″(u∗)e−2iω0τ∗2). $
|
Solving Eq (3.17) for $ (E_1^{(1)}, E_2^{(2)})^T = E_1 $, we get
$ E(1)1=(1+2iω0τ∗1)S″(v∗)q21e−2iω0τ∗1+Sv∗T″(u∗)e−2iω0(τ∗1+τ∗2)(1+2iω0τ∗1)2−Sv∗Tu∗e−2iω0(τ∗1+τ∗2), $
|
$ E(2)1=e−2iω0τ∗2[Tu∗S″(v∗)q21e−2iω0τ∗1+T″(u∗)(1+2iω0τ∗1)](1+2iω0τ∗1)2−Sv∗Tu∗e−2iω0(τ∗1+τ∗2). $
|
Similarly, we can get
$ (−1Sv∗Tu∗−1)E2=(−S″(v∗)q1¯q1−T″(u∗)), $
|
and hence,
$ E(1)2=S″(v∗)q1¯q1+T″(u∗)Sv∗1−Sv∗Tu∗,E(2)2=S″(v∗)Tu∗q1¯q1+T″(u∗)1−Sv∗Tu∗. $
|
Then $ g_{21} $ can be confirmed. Therefore, we can calculate the following quantities
$ c1(0)=i2ω0τ∗1(g20g11−2|g11|2−13|g02|2)+12g21,μ2=−Re(c1(0))Reλ′(τ∗1),β2=2Re(c1(0)),T2=−1ω0τ∗1(Im(c1(0))+μ2Im(λ′(τ∗1))). $
|
(3.18) |
By the general theory in Hassard [36], we obtain that $ \mu_2 $ determines the direction of the Hopf bifurcation: When $ \mu_2 > 0 (<0) $, the direction of the Hopf bifurcation is forward (backward), in other words, for $ \tau_1 > \tau_1^* (\tau_1 < \tau_1^*) $, there exist the bifurcating periodic solutions; $ \beta_2 $ determines that whether the bifurcating periodic solutions are stable: When $ \beta_2 < 0 (>0) $, the bifurcating periodic solutions are orbitally asymptotically stable (unstable); $ T_2 $ determines the period of the bifurcating periodic solutions: When $ T_2 > 0 (<0) $, the period of the bifurcating periodic solutions increases (decreases).
In this section, we choose
$ S(vτ1)=K1v(x,t−τ1),T(uτ2)=K3u(x,t−τ2)(1−u(x,t−τ2)K2), $
|
(4.1) |
and take the following data:
$ {\bf (A)}\; \; l = 3, \alpha_1 = \alpha_2 = 0.8, d_1 = d_2 = 1, \; K_1 = 40, \; \; K_2 = 20, \; \; K_3 = 0.1. $ |
Then system (1.4) with (4.1) and (A) has a unique positive constant steady state $ E_* = (16.8, 0.336) $. By direct calculation, we have $ S_{v^*} = 40 $, $ T_{u^*} = -0.068 $ and $ m = -2.72 $. Thus the assumptions $ (\bf{H_1}) $ and $ (\bf{H_2}) $ are satisfied, and $ E_* $ is asymptotically stable when $ \tau_1 = \tau_2 = 0 $.
By calculation, we get that when $ N = 3, \; m^2 > (d_1\frac{n^2}{l^2}+\alpha_1)^2(d_2\frac{n^2}{l^2}+\alpha_2)^2 $ for $ 0\leq n < N\; and\; m^2\leq (d_1\frac{N^2}{l^2}+\alpha_1)^2(d_2\frac{N^2}{l^2}+\alpha_2)^2, $ thus assumption $ (\bf{H_3}) $ is satisfied. From Eq (2.8), we get that
$ ω0≈1.4422,τj,0≈0.7023+π0.7211j,ω1≈1.3747,τj,1≈0.8515+π0.6874j,ω2≈0.5411,τj,2≈1.5799+π0.2706j. $
|
And Eq (2.7) has no positive roots for $ n\geq3 $. Clearly, $ \tau_{0, 0}\approx 0.7023. $
From Theorem 2.7, we know that the positive constant steady state $ E_* $ is asymptotically stable when $ \tau_1+\tau_2\in [0, \tau_{0, 0}) $, and unstable when $ \tau_1+\tau_2 > \tau_{0, 0} $. Meanwhile, (1.4) undergoes a Hopf bifurcation at $ E_* $ when $ \tau_1+\tau_2 = \tau_{j, n} $, $ n = 0, 1, 2; j\in\mathbb N_0 $. The bifurcation set on $ (\tau_1, \tau_2) $ plane is shown by Figure 1.
When $ (\tau_1, \tau_2) $ are chosen as $ P_1(0.4, 0.2) $, we have $ \tau_1+\tau_2 < 0.7023 $, thus $ E_* $ is asymptotically stable, this is illustrated in Figure 2.
We choose the point $ P_3(0.55115, 0.15115) $ on the curve $ \tau = \tau_{0, 0} $, that is $ \tau_1^* = 0.55115 $ and $ \tau_2^* = 0.15115 $. According to the algorithm given in the previous section, we can obtain that $ c_1(0) = -0.00155-0.00290{\rm i} $, $ \lambda'(\tau_1^*) = 0.92063-0.94716{\rm i} $, $ \mu_2 = 0.00168 > 0 $, $ \beta_2 = -0.0031 < 0 $ and $ T_2 = 0.00565 > 0 $. This implies that the direction of the Hopf bifurcation is forward, the bifurcating periodic solutions are stable and the period of the bifurcating periodic solutions increases, which are illustrated by Figure 3.
In fact, with the help of the norm form calculated in section 2, we get that when $ \tau_1^* $ and $ \tau_2^* $ satisfying $ \tau_1^*+\tau_2^* = \tau_{0, 0} $ and $ \tau_1^* > \tau_2^* $, the corresponding $ \text{Re}c_1(0) $ would be negative, we list some values of $ c_1(0) $ in Table 1.
$ \text{Point} $ | $ (\tau_1^*, \tau_2^*) $ | $ \text{The value of}\; c_1(0) $ |
$ 1 $ | (0.7023, 0) | -0.00210-0.00389$ \rm{i} $ |
$ 2 $ | (0.6023, 0.1) | -0.00175-0.00323$ \rm{i} $ |
$ 3 $ | (0.5023, 0.2) | -0.00134-0.00261$ \rm{i} $ |
$ 4 $ | (0.4023, 0.3) | -0.00089-0.00211$ \rm{i} $ |
In this paper, we study the dynamics of epidemics by using a diffusive model with two delays. First we perform the stability analysis at the constant steady state in the system. Then we present the conditions for the existence of the purely imaginary roots of the characteristic equations. Moreover, we investigate the existence of Hopf bifurcation by taking two delays as parameters. Theoretical result shows that the sum of two delays affect the stability of the steady state and the existence of Hopf bifurcation. Considering the sum of two delays as one parameter, we verify that $ \tau_1^*+\tau_2^* = \tau_{0, 0} $ is the critical Hopf bifurcation curve and the stability switching curve of this system. To calculate the normal form on the center manifold near the Hopf singularity, we fix $ \tau_2 $ as a constant and take $ \tau_1 $ as a parameter. Finally, numerical simulations are carried out to verify the theoretical results. The simulations show that when $ \tau_1+\tau_2 < \tau_{0, 0} $, $ E_* $ is locally asymptotically stable; when $ \tau_1+\tau_2 > \tau_{0, 0} $, the bifurcating periodic solutions are stable. Both theoretical verifications and numerical simulations reveal that when $ \tau_1+\tau_2 < \tau_{0, 0} $, the system is locally stable at $ E_* $; when $ \tau_1+\tau_2 > \tau_{0, 0} $ and close enough to $ \tau_{0, 0} $, $ E_* $ is unstable, the periodic solutions generated by the Hopf bifurcation is stable. All the analysis shows that the sum of two delays plays an important role in the dynamics of the system.
The authors are grateful to the anonymous referees for their helpful comments and valuable suggestions which have improved the presentation of the paper. This research is supported by National Natural Science Foundation of China (No.11771109).
The authors declared that they have no conflicts of interest to this work. We declare that we do not have any commercial or associative interest that represents a conflict of interest in connection with the work submitted.
[1] |
Habayeb MS, Ekengren SK, Hultmark D (2006) Nora virus, a persistent virus in Drosophila, defines a new picorna-like virus family. J Gen Virol 87: 3045–3051. doi: 10.1099/vir.0.81997-0
![]() |
[2] |
Webster CL, Waldron FM, Robertson S, et al. (2015) The discovery, distribution, and evolution of viruses associated with Drosophila melanogaster. PLoS Biol 13: e1002210. doi: 10.1371/journal.pbio.1002210
![]() |
[3] |
Johnson KN, Christian PD (1998) The novel genome organization of the insect picorna-like virus Drosophila C virus suggests this virus belongs to a previously undescribed virus family. J Gen Virol 79: 191–203. doi: 10.1099/0022-1317-79-1-191
![]() |
[4] |
Habayeb MS, Jens-Ola E, Dan H (2009) Nora virus persistent infections are not affected by the RNAi machinery. PLoS One 4: e5731. doi: 10.1371/journal.pone.0005731
![]() |
[5] | Ericson BL, Carlson DJ, Carlson KA (2016) Characterization of Nora virus structural proteins via Western blot analysis. Scientifica 2016: 9067848. |
[6] |
Sadanandan SA, Ekström JO, Jonna VR, et al. (2016) VP3 is crucial for the stability of Nora virus virions. Virus Res 223: 20–27. doi: 10.1016/j.virusres.2016.06.011
![]() |
[7] |
Ekström JO, Habayeb MS, Srivastava V, et al. (2011) Drosophila Nora virus capsid proteins differ from those of other picorna-like viruses. Virus Res 160: 51–58. doi: 10.1016/j.virusres.2011.05.006
![]() |
[8] |
Habayeb MS, Cantera R, Casanova G, et al. (2009) The Drosophila Nora virus is an enteric virus, transmitted via feces. J Invertebr Pathol 101: 29–33. doi: 10.1016/j.jip.2009.02.003
![]() |
[9] |
De Gregorio E, Spellman PT, Tzou P, et al. (2002) The Toll and Imd pathways are the major regulators of the immune response in Drosophila. EMBO J 21: 2568–2579. doi: 10.1093/emboj/21.11.2568
![]() |
[10] |
Ferreira AG, Naylor H, Esteves SS, et al. (2014) The Toll-dorsal pathway is required for resistance to viral oral infection in Drosophila. PLoS Pathog 10: e1004507. doi: 10.1371/journal.ppat.1004507
![]() |
[11] |
Baeg GH, Zhou R, Perrimon N (2005) Genome-wide RNAi analysis of JAK/STAT signaling components in Drosophila. Gene Dev 19: 1861–1870. doi: 10.1101/gad.1320705
![]() |
[12] |
Zambon RA, Vakharia VN, Wu LP (2006) RNAi is an antiviral immune response against a dsRNA virus in Drosophila melanogaster. Cell Microbiol 8: 880–889. doi: 10.1111/j.1462-5822.2006.00688.x
![]() |
[13] |
Cordes EJ, Licking-Murray KD, Carlson KA (2013) Differential gene expression related to Nora virus infection of Drosophila melanogaster. Virus Res 175: 95–100. doi: 10.1016/j.virusres.2013.03.021
![]() |
[14] | Corney DC (2012) RNA-seq using next generation sequencing. Mater Method 3: 203. |
[15] |
Livak KJ, Schmittgen TD (2001) Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) method. Methods 25: 402–408. doi: 10.1006/meth.2001.1262
![]() |
[16] |
Mi H, Huang X, Muruganujan A, et al. (2017) PANTHER version 11: Expanded annotation data from Gene Ontology and Reactome pathways, and data analysis tool enhancements. Nucleic Acids Res 45: D183–D189. doi: 10.1093/nar/gkw1138
![]() |
[17] |
Thomas PD, Kejariwal A, Guo N, et al. (2006) Applications for protein sequence-function evolution data: mRNA/protein expression analysis and coding SNP scoring tools. Nucleic Acids Res 34: W645–W650. doi: 10.1093/nar/gkl229
![]() |
[18] |
Huang DW, Sherman BT, Lempicki RA (2009) Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc 4: 44–57. doi: 10.1038/nprot.2008.211
![]() |
[19] | Huang DW, Sherman BT, Lempicki RA (2009) Bioinformatics enrichment tools: Paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Res 37: 1–13. |
[20] |
Gramates LS, Marygold SJ, Santos GD (2017) FlyBase at 25: Looking to the future. Nucleic Acids Res 45: D663–D671. doi: 10.1093/nar/gkw1016
![]() |
[21] | Lemaitre Lab Uplem, List of Drosophila gene potentially involved in the immune response, 2007. Available from: https://lemaitrelab.epfl.ch/page-7767-en.html. |
[22] | Society for Developmental Biology, the Interactive Fly: RNAi and posttranscriptional gene silencing, 2017. Available from: http://www.sdbonline.org/sites/fly/aignfam/rnaistuf.htm. |
[23] | Smith AM, Perelson AS (2011) Influenza A virus infection kinetics: Quantitative data and models. Wires Syst Biol Med 3: 429–445. |
[24] |
Read EL, Tovodwyer AA, Chakraborty AK (2012) Stochastic effects are important in intrahost HIV evolution even when viral loads are high. Proc Natl Acad Sci USA 109: 19727–19732. doi: 10.1073/pnas.1206940109
![]() |
[25] |
Pearson A, Lux A, Krieger M (1995) Expression cloning of dSR-CI, a class C macrophage-specific scavenger receptor from Drosophila melanogaster. Proc Natl Acad Sci USA 92: 4056–4060. doi: 10.1073/pnas.92.9.4056
![]() |
[26] |
Ligoxygakis P, Pelte N, Ji C, et al. (2002) A serpin mutant links Toll activation to melanization in the host defence of Drosophila. EMBO J 21: 6330–6337. doi: 10.1093/emboj/cdf661
![]() |
[27] |
Gendrin M, Zaidman-Rémy A, Broderick NA, et al. (2013) Functional analysis of PGRP-LA in Drosophila immunity. PLoS One 8: e69742. doi: 10.1371/journal.pone.0069742
![]() |
[28] |
Rynes J, Donohoe CD, Frommolt P, et al. (2012) Activating transcription factor 3 regulates immune and metabolic homeostasis. Mol Cell Biol 32: 3949–3962. doi: 10.1128/MCB.00429-12
![]() |
[29] |
Wang H, Cai Y, Chia W, et al. (2006) Drosophila homologs of mammalian TNF/TNFR-related molecules regulate segregation of Miranda/Prospero in neuroblasts. EMBO J 25: 5783–5793. doi: 10.1038/sj.emboj.7601461
![]() |
[30] |
Werner T, Liu G, Kang D, et al. (2000) A family of peptidoglycan recognition proteins in the fruit fly Drosophila melanogaster. Proc Natl Acad Sci USA 97: 13772–13777. doi: 10.1073/pnas.97.25.13772
![]() |
[31] |
Tsail CW, Mcgraw EA, Ammar ED, et al. (2008) Drosophila melanogaster mounts a unique immune response to the Rhabdovirus sigma virus. Appl Environ Microb 74: 3251–3256. doi: 10.1128/AEM.02248-07
![]() |
[32] |
Kim YS, Ryu JH, Han SJ, et al. (2000) Gram-negative bacteria-binding protein, a pattern recognition receptor for lipopolysaccharide and β-1,3-glucan that mediates the signaling for the induction of innate immune genes in Drosophila melanogaster cells. J Biol Chem 275: 32721–32727. doi: 10.1074/jbc.M003934200
![]() |
[33] |
Chamy LE, Leclerc V, Caldelari I, et al. (2008) Danger signal and PAMP sensing define binary signaling pathways upstream of Toll. Nat Immunol 9: 1165–1170. doi: 10.1038/ni.1643
![]() |
[34] |
Watson FL, Püttmannholgado R, Thomas F, et al. (2005) Extensive diversity of Ig-superfamily proteins in the immune system of insects. Science 309: 1874–1878. doi: 10.1126/science.1116887
![]() |
[35] |
Zambon RA, Nandakumar M, Vakharia VN, et al. (2005) The Toll pathway is important for an antiviral response in Drosophila. Proc Natl Acad Sci USA 102: 7257–7262. doi: 10.1073/pnas.0409181102
![]() |
[36] |
Grech A, Quinn R, Srinivasan D, et al. (2000) Complete structural characterisation of the mammalian and Drosophila TRAF genes: Implications for TRAF evolution and the role of RING finger splice variants. Mol Immunol 37: 721–734. doi: 10.1016/S0161-5890(00)00098-5
![]() |
[37] |
Cha GH, Cho KS, Lee JH, et al. (2003) Discrete functions of TRAF1 and TRAF2 in Drosophila melanogaster mediated by c-Jun N-terminal kinase and NF-κB-dependent signaling pathways. Mol Cell Biol 23: 7982–7991. doi: 10.1128/MCB.23.22.7982-7991.2003
![]() |
[38] |
Berkey CD, Blow N, Watnick PI (2009) Genetic analysis of Drosophila melanogaster susceptibility to intestinal Vibrio cholerae infection. Cell Microbiol 11: 461–474. doi: 10.1111/j.1462-5822.2008.01267.x
![]() |
[39] |
Moy RH, Gold B, Molleston JM, et al. (2014) Antiviral autophagy restricts Rift Valley fever virus infection and is conserved from flies to mammals. Immunity 40: 51–65. doi: 10.1016/j.immuni.2013.10.020
![]() |
[40] |
Xu YC, Wu RF, Gu Y, et al. (2002) Involvement of TRAF4 in oxidative activation of c-Jun N-terminal kinase. J Biol Chem 277: 28051–28057. doi: 10.1074/jbc.M202665200
![]() |
[41] |
Ha EM, Oh CT, Bae YS, et al. (2005) A direct role for dual oxidase in Drosophila gut immunity. Science 310: 847–850. doi: 10.1126/science.1117311
![]() |
[42] |
Tian C, Gao B, Rodriguez MC, et al. (2008) Gene expression, antiparasitic activity, and functional evolution of the drosomycin family. Mol Immunol 45: 3909–3916. doi: 10.1016/j.molimm.2008.06.025
![]() |
[43] |
Lematire B, Nicolas E, Michaut L, et al. (1996) The dorsoventral regulatory gene cassette spätzle/Toll/cactus controls the potent antifungal response in Drosophila adults. Cell 86: 973–983. doi: 10.1016/S0092-8674(00)80172-5
![]() |
[44] |
Kurucz E, Markus R, Zsamboki J, et al. (2007) Nimrod, a putative phagocytosis receptor with EGF repeats in Drosophila plasmatocytes. Curr Biol 17: 649–654. doi: 10.1016/j.cub.2007.02.041
![]() |
[45] |
Van Mierlo JT, Bronkhorst AW, Overheul GJ, et al. (2012) Convergent evolution of Argonaute-2 slicer antagonism in two distinct insect RNA viruses. PLoS Pathog 8: e1002872. doi: 10.1371/journal.ppat.1002872
![]() |
[46] |
Kemp C, Imler JL (2009) Antiviral immunity in Drosophila. Curr Opin Immunol 21: 3–9. doi: 10.1016/j.coi.2009.01.007
![]() |
[47] |
Baker DA, Russell S (2009) Gene expression during Drosophila melanogaster egg development before and after reproductive diapause. BMC Genomics 10: 1–16. doi: 10.1186/1471-2164-10-1
![]() |
[48] | Potempa J, Korzus E, Travis J (1994) The serpin superfamily of proteinase inhibitors: Structure, function, and regulation. J Biol Chem 269: 15957–15960. |
[49] |
Irving P, Trozler L, Heuer TS, et al. (2001) A genome-wide analysis of immune responses in Drosophila. Proc Natl Acad Sci USA 98: 15119–15124. doi: 10.1073/pnas.261573998
![]() |
[50] |
Tanj T, Yun EY, Ip YT (2010) Heterodimers of NF-κB transcription factors DIF and Relish regulate antimicrobial peptide genes in Drosophila. Proc Natl Acad Sci USA 107: 14715–14720. doi: 10.1073/pnas.1009473107
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
1. | Yongli Song, Yahong Peng, Tonghua Zhang, The spatially inhomogeneous Hopf bifurcation induced by memory delay in a memory-based diffusion system, 2021, 300, 00220396, 597, 10.1016/j.jde.2021.08.010 |
$ \text{Point} $ | $ (\tau_1^*, \tau_2^*) $ | $ \text{The value of}\; c_1(0) $ |
$ 1 $ | (0.7023, 0) | -0.00210-0.00389$ \rm{i} $ |
$ 2 $ | (0.6023, 0.1) | -0.00175-0.00323$ \rm{i} $ |
$ 3 $ | (0.5023, 0.2) | -0.00134-0.00261$ \rm{i} $ |
$ 4 $ | (0.4023, 0.3) | -0.00089-0.00211$ \rm{i} $ |