Research article Special Issues

Global Hopf bifurcation of a cholera model with media coverage


  • We propose a model for cholera under the impact of delayed mass media, including human-to-human and environment-to-human transmission routes. First, we establish the extinction and uniform persistence of the disease with respect to the basic reproduction number. Then, we conduct a local and global Hopf bifurcation analysis by treating the delay as a bifurcation parameter. Finally, we carry out numerical simulations to demonstrate theoretical results. The impact of the media with the time delay is found to not influence the threshold dynamics of the model, but is a factor that induces periodic oscillations of the disease.

    Citation: Jie He, Zhenguo Bai. Global Hopf bifurcation of a cholera model with media coverage[J]. Mathematical Biosciences and Engineering, 2023, 20(10): 18468-18490. doi: 10.3934/mbe.2023820

    Related Papers:

    [1] Xuejuan Lu, Shaokai Wang, Shengqiang Liu, Jia Li . An SEI infection model incorporating media impact. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1317-1335. doi: 10.3934/mbe.2017068
    [2] Jingli Xie, Hongli Guo, Meiyang Zhang . Dynamics of an SEIR model with media coverage mediated nonlinear infectious force. Mathematical Biosciences and Engineering, 2023, 20(8): 14616-14633. doi: 10.3934/mbe.2023654
    [3] Cunjuan Dong, Changcheng Xiang, Wenjin Qin, Yi Yang . Global dynamics for a Filippov system with media effects. Mathematical Biosciences and Engineering, 2022, 19(3): 2835-2852. doi: 10.3934/mbe.2022130
    [4] Ya-Dong Zhang, Hai-Feng Huo, Hong Xiang . Dynamics of tuberculosis with fast and slow progression and media coverage. Mathematical Biosciences and Engineering, 2019, 16(3): 1150-1170. doi: 10.3934/mbe.2019055
    [5] Pengfei Liu, Yantao Luo, Zhidong Teng . Role of media coverage in a SVEIR-I epidemic model with nonlinear incidence and spatial heterogeneous environment. Mathematical Biosciences and Engineering, 2023, 20(9): 15641-15671. doi: 10.3934/mbe.2023698
    [6] Wenjie Qin, Jiamin Zhang, Zhengjun Dong . Media impact research: a discrete SIR epidemic model with threshold switching and nonlinear infection forces. Mathematical Biosciences and Engineering, 2023, 20(10): 17783-17802. doi: 10.3934/mbe.2023790
    [7] 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
    [8] Kazuo Yamazaki, Xueying Wang . Global stability and uniform persistence of the reaction-convection-diffusion cholera epidemic model. Mathematical Biosciences and Engineering, 2017, 14(2): 559-579. doi: 10.3934/mbe.2017033
    [9] Chenwei Song, Rui Xu, Ning Bai, Xiaohong Tian, Jiazhe Lin . Global dynamics and optimal control of a cholera transmission model with vaccination strategy and multiple pathways. Mathematical Biosciences and Engineering, 2020, 17(4): 4210-4224. doi: 10.3934/mbe.2020233
    [10] Fred Brauer, Zhisheng Shuai, P. van den Driessche . Dynamics of an age-of-infection cholera model. Mathematical Biosciences and Engineering, 2013, 10(5&6): 1335-1349. doi: 10.3934/mbe.2013.10.1335
  • We propose a model for cholera under the impact of delayed mass media, including human-to-human and environment-to-human transmission routes. First, we establish the extinction and uniform persistence of the disease with respect to the basic reproduction number. Then, we conduct a local and global Hopf bifurcation analysis by treating the delay as a bifurcation parameter. Finally, we carry out numerical simulations to demonstrate theoretical results. The impact of the media with the time delay is found to not influence the threshold dynamics of the model, but is a factor that induces periodic oscillations of the disease.



    In modern society, media coverage has become an important strategy for controlling and preventing disease transmission. It can alter an individual's behavior, such as wearing protective masks, vaccination, self-isolating or avoiding gathering activities, and hence reduce the possibility of contracting the infection [1]. Additionally, the media coverage also influences the implementation of a public health policy intervention and control polices [2]. Therefore, how to quantify this impact through mathematical modeling is an important issue for epidemics control.

    Many mathematical models have been developed to investigate the impact of media coverage on disease transmission and control. The most general approach to modeling the media impact is to change the transmission coefficient as a nonlinear decreasing function with respect to the number of infectious individuals (I). For example, Cui et al. [3] used a contact transmission rate μemI to describe the impact of media coverage, where μ>0 is the probability of baseline transmission and m>0 reflects the effect of the media. Li and Cui [4] employed β2Im+I to reflect the reduced amount of contact rate due to media coverage, producing the transmission rate of the form β1β2Im+I(β1>β2>0,m>0). By considering the correlation between media impact and the number of infected individuals at different disease stages, Liu et al. [5] devised the media impact function βea1Ea2Ia3H with E denoting exposed individuals, I infectives, H hospitalized individuals and nonnegative constants a1,a2,a3. However, it is noted that individuals can also change their behavior in response to the rate of change in case numbers. Xiao et al. [6] used an exponentially decreasing function eM(I,dIdt) with M(I,dIdt)=max{0,p1I+p2dIdt} to model such media impacts. For more studies on media impacts, we refer readers to [1,2,7,8,9,10] and references therein.

    However, as mentioned in [10], impact of media coverage on the population transmission dynamics has lag, describing both the lag time of individuals' response to the media and the lag time of media reports about an infectious disease outbreak. Some studies have been conducted to investigate the impact of delayed mass media, see [1,8,10,11]. In this paper, we will consider the time delay of the media impact in the mathematical modeling of cholera, an acute intestinal infectious disease caused by the bacterium Vibrio cholerae (or, V. cholerae). Several mathematical models have been proposed to study cholera dynamics, including, but not limited to, [12,13,14,15,16,17,18]. However, to our knowledge, few of these have specifically taken into account the response delay in media impacts.

    Our model is inspired by the disease transmission models in [10,13]. We divide the population into three subclasses: Susceptible S(t), infectives I(t) and recovered R(t). The concentration of bacteria V. cholerae in contaminated water is indicated by B(t). To study the lag effect of the impact of the media on cholera transmission, we incorporate a decreasing factor emI(tτ) into the direct and indirect incidence rate. That is,

    β1emI(tτ)S(t)I(t)andβ2emI(tτ)S(t)B(t).

    Here, β1 and β2 are, respectively, the baseline direct and indirect transmission rate, τ represents the report delay and response time of individuals to the current infection and m is the weight of media impact sensitive to the case number. The basic SIRB model under consideration is then

    {dS(t)dt=Λβ1emI(tτ)S(t)I(t)β2emI(tτ)S(t)B(t)μS(t)+σR(t),dI(t)dt=β1emI(tτ)S(t)I(t)+β2emI(tτ)S(t)B(t)(μ+γ)I(t),dR(t)dt=γI(t)(μ+σ)R(t),dB(t)dt=ξI(t)δB(t), (1.1)

    where Λ stands for the influx rate of susceptible humans, μ is the natural death rate, σ is the immunity waning rate, γ is the recovery rate, ξ denotes the rate of contribution to V. cholerae in the aquatic environment and δ=δ1δ2>0 is the net death rate of V. cholerae in the aquatic environment, where δ1,δ2 are the natural washout rate and the natural growth rate of V. cholerae in the aquatic environment, respectively. All the parameters of model (1.1) are assumed to be positive.

    Note that the total population size N(t)=S(t)+I(t)+R(t) satisfies

    dN(t)dt=ΛμN(t),

    and has a unique equilibrium N=Λμ that is globally stable in R+. We then consider the following limiting system:

    {dI(t)dt=β1emI(tτ)(NI(t)R(t))I(t)+β2emI(tτ)(NI(t)R(t))B(t)(μ+γ)I(t),dR(t)dt=γI(t)(μ+σ)R(t),dB(t)dt=ξI(t)δB(t). (1.2)

    The remainder of this work is organized as follows. In Section 2, we first give the well-posedness of system (1.2) and then establish its threshold dynamics concerning the basic reproduction number. In Section 3, we are devoted to the stability and Hopf bifurcation analysis of the positive equilibrium. In Section 4, we investigate the global continuation of a local branch. In Section 5, we perform numerical simulations to illustrate our analytical results. The last section is a brief discussion.

    Let C:=C([τ,0],R3) and C+:=C([τ,0],R3+). Then (C,C+) is an ordered Banach space equipped with the supremum norm ϕ=supτθ0|ϕ(θ)| for ϕC. For any given continuous function u=(u1,u2,u3):[τ,ρ)R3 with ρ>0, we define utC by

    ut(θ)=(u1(t+θ),u2(t+θ),u3(t+θ)),θ[τ,0],

    for any t[0,ρ). Set

    Γ={φ=(φ1,φ2,φ3)C+:φ1(θ)+φ2(θ)N,θ[τ,0]}.

    We then have the following result.

    Lemma 2.1. For any φΓ, system (1.2) admits a unique nonnegative bounded solution u(t,φ) on [0,) with u0=φ, and ut(φ)Γ,t0.

    Proof. For any φ=(φ1,φ2,φ3)Γ, we define

    f(φ)=(emφ1(τ)(Nφ1(0)φ2(0))(β1φ1(0)+β2φ3(0))(μ+γ)φ1(0)γφ1(0)(μ+σ)φ2(0)ξφ1(0)δφ3(0)).

    Obviously, f(φ) is continuous in φΓ and Lipschitz in φ on each compact subset of Γ. It follows from [19, Theorems 2.2.3] that system (1.2) has a unique solution u(t,φ) on its maximal existence interval [0,tφ) with u0=φ.

    For any given φΓ, one sees that if φi(0)=0 for some i{1,2,3}, then fi(φ)0. By [20, Theorem 5.2.1], we obtain u(t,φ)0 for all t[0,tφ). Let H(t)=I(t)+R(t). Then one has

    dH(t)dt|H(t)=N=[β1emI(tτ)(NH(t))I(t)+β2emI(tτ)(NH(t))B(t)μH(t)σR(t)]|H(t)=N=μNσR(t)0,t(0,tφ),

    which implies that It+RtN for t[0,tφ), and hence, ut(φ)Γ for all t[0,tφ). Moreover, by the third equation of (1.2), one easily sees that u3(t,φ) is bounded on [0,tφ). As a result, [19, Theorem 2.3.1] means that tφ=.

    Note that the disease-free equilibrium of (1.2) is E0=(0,0,0). Since the R class do not participate in the transmission of cholera, we only consider the infected equations of the linearization of (1.2) at E0:

    {dI(t)dt=β1NI(t)+β2NB(t)(μ+γ)I(t),dB(t)dt=ξI(t)δB(t). (2.1)

    Based on the next generation matrix method [21], we define

    F=(β1Nβ2N00)andV=(μ+γ0ξδ).

    Hence, the basic reproduction number of system (1.2) is given by

    R0=r(FV1)=Rd1+Ri2=β1Nμ+γ+β2ξNδ(μ+γ),

    where r(FV1) is the spectral radius of FV1, and Rd1 (resp. Ri2 is the basic reproduction number for the direct (resp. indirect) transmission in the absence of indirect (resp. direct) transmission.

    Now we begin to study the local stability of the disease-free equilibrium. The Jacobian matrix of system (1.2) at E0 has the form

    (β1Nμγ0β2Nγμσ0ξ0δ),

    and the corresponding characteristic equation is

    λ3+bλ2+cλ+d=0, (2.2)

    where

    b=δ+μ+σ+(1R0)(μ+γ)+β2ξNδ,c=(μ+σ+δ)(μ+γ)(1R0)+(μ+σ)(β2ξNδ+δ),d=δ(μ+σ)(μ+γ)(1R0).

    When R0<1, one easily sees

    b>0,c>0,d>0,bcd>0.

    Using the Routh-Hurwitz criterion, we know that all roots of (2.2) have negative real parts. Hence, E0 is locally stable if R0<1. In the case where R0>1, the Eq (2.2) has at least one positive root since d<0, and hence, E0 is unstable. Now we proceed to the global stability of E0.

    Theorem 2.1. If R0<1, then E0 is globally asymptotically stable in Γ.

    Proof. From the second and third equations of (1.2), we see

    {dI(t)dtβ1NI(t)+β2NB(t)(μ+γ)I(t),dB(t)dt=ξI(t)δB(t),

    which is equivalent to

    (dI(t)dtdB(t)dt)(FV)(I(t)B(t)). (2.3)

    Since the matrix V1F is nonnegative, irreducible and R0=r(FV1)=r(V1F), there exists a positive left eigenvector v such that

    vV1F=R0v.

    Define the functional L:ΓR as follows

    L(ϕ)=vV1(ϕ1(0),ϕ3(0))T,ϕ=(ϕ1,ϕ2,ϕ3)Γ.

    For any solution u(t,φ) of (1.2) with u0=φΓ, with the help of (2.3), direct calculation yields

    ddtL(ut(φ))=vV1(dI(t)dt,dB(t)dt)TvV1(FV)(I(t),B(t))T=(R01)v(I(t),B(t))T0. (2.4)

    Thus, L is a Lyapunov functional on Γ relative to system (1.2).

    Let M be the largest compact invariant subset in the set {ϕΓ:˙L(1.2)(ϕ)=0}, where ˙L(1.2)(ϕ) denotes the derivative of L along the solution of (1.2). It is easy to see from (2.4) that ˙L(1.2)(ϕ)=0 implies that I(t)=B(t)=0 for any t0, and hence, M={ϕΓ:ϕ1=ϕ3=0}. It then follows from the LaSalle invariance principle (see, e.g., [22, Theorem 1]) that limt(I(t),B(t))=(0,0) and further, limtR(t)=0. Therefore, E0 is globally attractive in Γ. This, together with the local stability of E0, gives the desired result.

    Next we are ready to show that the disease is uniformly persistent when R0>1. To this end, we need the following lemma.

    Lemma 2.2. For any given φΓ, if there exists some t00 such that I(t0,φ)>0 or B(t0,φ)>0, then I(t,φ)>0 and B(t,φ)>0 for all t>t0.

    Proof. If I(t0,φ)>0 for some t0, then

    I(t,φ)>I(t0,φ)e(μ+γ)(tt0)>0,tt0.

    By the third equation of (1.2), we have

    B(t,φ)=eδ(tt0)B(t0,φ)+tt0eδ(ts)ξI(s,φ)ds>0,t>t0.

    Similarly, if B(t0,φ)>0 for some t0, we obtain that B(t,φ)>0 for all tt0. We will show that H(t,φ):=I(t,φ)+R(t,φ)<N for all t>t0. If the assertion was false, then there exists t1(t0,) such that

    H(t,φ)<N,t(t0,t1)andH(t1,φ)=N.

    Clearly, dH(t,φ)dt|t=t10 must hold. However, in view of the first two equations of (1.2), we find

    dH(t,φ)dt|t=t1=μNσR(t1,φ)<0.

    This contradiction means that no such t1 can exist. Accordingly, by the I-equation in (1.2), we have

    I(t,φ)e(μ+γ)(tt0)I(t0,φ)+tt0e(μ+γ)(ts)β2emI(sτ,φ)(NI(s,φ)R(s,φ))B(s,φ)ds>0

    for all t>t0. Summarizing these two cases, we establish the desired result.

    Theorem 2.2. If R0>1, then system (1.2) admits a unique componentwise positive equilibrium. Moreover, there exists η>0 such that for any φΓ with φ1(0)0 or φ3(0)0, the solution u(t,φ)=(I(t,φ),R(t,φ),B(t,φ)) satisfies

    lim inftmin{I(t,φ),B(t,φ)}η. (2.5)

    Proof. Define the following two sets:

    Γ0={φΓ:φ1(0)>0 and φ3(0)>0},Γ0=ΓΓ0={φΓ:φ1(0)=0orφ3(0)=0}.

    By the form of (1.2), it can be verified that both Γ and Γ0 are positively invariant. Clearly, Γ0 is relatively closed in Γ. Let Φ(t) be the solution maps of system (1.2), namely

    Φ(t)φ=ut(φ),t0,φΓ.

    Note that for each t>τ, Φ(t) is continuous and compact (see [19, Theorem 3.6.1]). The ultimate boundedness of solutions implies that Φ(t) is point dissipative. It then follows from [23, Theorem 3.4.8] that Φ(t) has a global attractor K.

    Let ω(φ) be the omega limit set of the forward orbit through φ for Φ(t) and define

    M={φΓ0:Φ(t)φΓ0,t0}.

    Then the following claim holds true.

    Claim 1. E0 is globally stable for Φ(t) in M.

    For any given φM, we have Φ(t)φΓ0,t0. Hence, for each t0, either I(t,φ)=0 or B(t,φ)=0. We further show

    MM0:={φΓ0:φ1(0)=0andφ3(0)=0}. (2.6)

    Suppose to the contrary that φM0. Then either φ1(0)=I(0,φ)>0 or φ3(0)=B(0,φ)>0. With the help of Lemma 2.2, we have that I(t,φ)>0 and B(t,φ)>0 for all t>0, which contradicts the fact that φM. Thus, (2.6) holds. It then follows that I(t,φ)=0 and B(t,φ)=0 for all φM and t0, and further, limtR(t,φ)=0. Therefore, ω(φ)=E0 for any φM. In other words, E0 is globally attractive for Φ(t) in M. Note that system (1.2) restricted to M becomes dR(t)dt=(μ+σ)R(t). In view of [24, Lemma 2.2.1], E0 is locally Lyapunov stable for Φ(t) in M. So, the above claim is proved.

    Since R0>1, we can choose ϵ>0 small enough such that

    Rϵ0=emϵ(β1+β2ξδ)(N2ϵ)μ+γ>1.

    Claim 2. lim suptΦ(t)φE0ϵ,φΓ0. Suppose the claim is not true. Then there exists some ψΓ0 such that

    lim suptΦ(t)ψE0<ϵ.

    Thus, there is ˉt>0 such that

    0<I(t,ψ),R(t,ψ),B(t,ψ)<ϵ,tˉt.

    It follows that for all tˉt+τ, we have

    {dI(t)dtβ1emϵ(N2ϵ)I(t)+β2emϵ(N2ϵ)B(t)(μ+γ)I(t),dB(t)dt=ξI(t)δB(t).

    Let (x(t),y(t)) be the solution of the following system:

    {dx(t)dt=β1emϵ(N2ϵ)x(t)+β2emϵ(N2ϵ)y(t)(μ+γ)x(t),dy(t)dt=ξx(t)δy(t), (2.7)

    and its Jacobian matrix is

    Mϵ=(β1emϵ(N2ϵ)μγβ2emϵ(N2ϵ)ξδ).

    Recall that the stability modulus of Mϵ is defined by

    s(Mϵ):=max{Reλ:λis an eigenvalue ofMϵ}.

    Since Mϵ is quasi-positive and irreducible, it follows from [20, Corollary 3.2] that s(Mϵ) is a simple eigenvalue of Mϵ with a strongly positive eigenvector vϵ. In particular, es(Mϵ)tvϵ is a solution of (2.7). By using the proof of [21, Theorem 2], we know that s(Mϵ)>0 if and only if Rϵ0>1.

    Since (I(t,ψ),B(t,ψ))(0,0) for all t>0, there exists κ>0 such that

    (I(t,ψ),B(t,ψ))κes(Mϵ)tvϵ,t[ˉt,ˉt+τ].

    Hence by comparison,

    (I(t,ψ),B(t,ψ))κes(Mϵ)tvϵ,tˉt+τ.

    In view of s(Mϵ)>0, we see that I(t,ψ) and B(t,ψ) tend to infinity as t. This contradicts the boundedness of solutions, and thus the claim holds.

    Claim 1 shows that E0 cannot form a cycle in Γ0, and claim 2 implies that E0 is an isolated invariant set in Γ and Ws(E0)Γ0=, where Ws(E0) is the stable set of E0 for Φ(t). According to the acyclicity theorem of the uniform persistence of maps (see, e.g., [24, Theorem 1.3.1,Remark 1.3.1 and Remark 1.3.2]), we conclude that Φ(t):ΓΓ is uniformly persistent with respect to (Γ0,Γ0). Moreover, by [25, Theorem 2.4], there exists a global attractor A0 for Φ(t) in Γ0 and system (1.2) has an stationary coexistence state ˉφ=(ˉφ1,ˉφ2,ˉφ3)Γ0, and Φ(t)ˉφ=ˉφ for all t0. Noting that ˉφ is a constant function, we let ˉI=ˉφ1(0), ˉR=ˉφ2(0) and ˉB=ˉφ3(0). Then ˉRR+ and (ˉI,ˉB)int(R2+). We further claim that ˉRR+{0}. Suppose that ˉR=0. By the R-equation in (1.2), we then get 0=γˉI, and hence, ˉI=0, a contradiction. Therefore, E1=(ˉI,ˉR,ˉB) is a componentwise positive equilibrium of (1.2).

    Next, we show the uniqueness of the positive equilibrium. Suppose that (ˆI,ˆR,ˆB) is the equilibrium of (1.2). Then (ˆI,ˆR,ˆB) satisfies

    {β1emˆI(NˆIˆR)ˆI+β2emˆI(NˆIˆR)ˆB(μ+γ)ˆI=0,γˆI(μ+σ)ˆR=0,ξˆIδˆB=0. (2.8)

    Set ˆS=NˆIˆR. A direct verification yields

    Λβ1emˆIˆSˆIβ2emˆIˆSˆBμˆS+σˆR=0.

    This implies that (ˆS,ˆI,ˆR,ˆB) is the equilibrium of (1.1). In view of (2.8), we see that

    ˆS=F(ˆI):=NˆIγμ+σˆIandˆS=G(ˆI):=μ+γβ1emˆI+β2emˆIξ/δ.

    Therefore, F(ˆI) is strictly decreasing in ˆIR+, and G(ˆI) is strictly increasing in ˆIR+. If R0>1, then F(0)>G(0), which means that there is a unique intersection in R2+ between F(ˆI) and G(ˆI), and thus (ˆI,ˆR,ˆB) is the unique positive equilibrium of (1.2). Moreover, the uniqueness of positive equilibrium also implies that of E1.

    Finally, to derive the practical persistence, we define a continuous function p:ΓR+ by

    p(φ)=min{φ1(0),φ3(0)},φΓ.

    Clearly, Γ0=p1(0,) and Γ0=p1(0). Since A0 is a compact subset of Γ0, it follows that infφA0=minφA0p(φ)>0. Consequently, there exists η>0 such that

    lim inftmin{I(t,φ),B(t,φ)}=lim inftp(Φ(t)φ)η,φΓ0.

    Moreover, for any given φΓ with φ1(0)0 or φ3(0)0, Lemma 2.2 suggests that I(t,φ)>0 and B(t,φ)>0 for all t>0. Fix a t0>τ. By using the fact Φ(t)φ=Φ(tt0)(Φ(t0)φ),t>t0, we see that statement (2.5) holds true.

    Remark 2.1. The uniqueness of the positive equilibrium E1 can be obtained directly. In fact, from (1.2) we get

    em(μ+σ)μ+σ+γΛμmˉI(m(μ+σ)μ+σ+γΛμmˉI)=m(μ+σ)(μ+γ)em(μ+σ)μ+σ+γΛμ(β1+β2ξδ)(μ+σ+γ),ˉR=γμ+σˉI,ˉB=ξδˉI.

    Solving the above equation with respect to ˉI gives rise to

    ˉI=μ+σμ+σ+γΛμ1mLambert W(m(μ+σ)(μ+γ)em(μ+σ)μ+σ+γΛμ(β1+β2ξδ)(μ+σ+γ)),

    where the definition of Lambert W function is seen in [26]. It is positive (i.e. ˉI>0) provided

    m(μ+σ)μ+σ+γΛμ>Lambert W(m(μ+σ)(μ+γ)em(μ+σ)μ+σ+γΛμ(β1+β2ξδ)(μ+σ+γ)),

    which is equivalent to R0>1. In addition, this explicit expression makes the numerical approximation of (ˉI,ˉR,ˉB) more convenient.

    In this section, we will consider the global stability of the endemic equilibrium E1 and the existence of local Hopf bifurcations of system (1.2).

    The linearized form of system (1.2) at E1=(ˉI,ˉR,ˉB) is given by

    {dI(t)dt=β1AI(t)CI(t)(μ+γ)I(t)m(μ+γ)ˉII(tτ)CR(t)+β2AB(t),dR(t)dt=γI(t)(μ+σ)R(t),dB(t)dt=ξI(t)δB(t), (3.1)

    where A=emˉI(NˉIˉR) and C=β1emˉIˉI+β2emˉIˉB. The characteristic equation of system (3.1) is

    p(λ,τ):=λ3+a0λ2+a1λ+a2+eλτ(b0λ2+b1λ+b2)=0, (3.2)

    where

    a0=μ+σ+C+β2Aξδ+δ,a1=C(γ+δ)+(μ+σ)(β2Aξδ+δ+C),a2=Cδ(μ+σ+γ),b0=m(μ+γ)ˉI,b1=m(μ+γ)(μ+σ+δ)ˉI,b2=mδ(μ+γ)(μ+σ)ˉI.

    When τ=0, the equation (3.2) becomes

    λ3+(a0+b0)λ2+(a1+b1)λ+a2+b2=0. (3.3)

    In view of a0+b0>0 and

    (a0+b0)(a1+b1)(a2+b2)=X2(μ+σ)+(μ+σ)2X+Cγ(β2Aξδ+C+m(μ+γ)ˉI+μ+σ)+CXδ+mδ(μ+σ)ˉIX>0,

    where X=β2Aξδ+δ+C+m(μ+γ)ˉI, we deduce from the Routh-Hurwitz criterion that all roots of (3.3) have negative real parts. Hence, the following result holds true.

    Theorem 3.1. Suppose that R0>1. Then the endemic equilibrium E1 of system (1.2) is locally asymptotically stable when τ=0.

    Now we present the geometric approach developed in [27] to study the global stability of E1. Consider the following autonomous system

    dxdt=f(x),xD,fC1, (3.4)

    where DRn is a simply connected open set. Let x(t,x0) be the solution of (3.4) such that x(0,x0)=x0. The second compound equation with respect to x(t,x0)D is

    dzdt=fx[2](x(t,x0))z,

    where fx[2] is the additive compound matrix of the Jacobian matrix fx (see [28]).

    Let P(x) be a k×k matrix-valued C1 function, where k=12n(n1), and suppose P1(x) exists for xD. Set

    Q=PfP1+Pfx[2]P1,

    where Pf is the matrix obtained by replacing each entry pij in P with its directional derivative in the direction of f. The Lozinskiǐ measure of Q with respect to the matrix norm || in Rn×n is defined as

    μ(Q)=limh0+|I+hQ|1h.

    Define a quantity ˉq2 as

    ˉq2=lim suptsupx0K1tt0μ(Q(x(s,x0)))ds,

    where K denotes a compact absorbing set in D. The following result from [27, Theorem 3.5] is used to prove the global stability of E1.

    Lemma 3.1. Assume that system (3.4) has a unique equilibrium x in D. Then x is globally stable in D if ˉq2<0.

    Theorem 3.2. If R0>1, τ=0 and μ+γ>σ, then E1 is globally asymptotically stable in Ω={(I,R,B)R3+:I+RN,I>0,B>0}.

    Proof. Note that system (1.2) is equivalent to

    {dS(t)dt=Λβ1emI(t)S(t)I(t)β2emI(t)S(t)B(t)μS(t)+σ(NS(t)I(t)),dI(t)dt=β1emI(t)S(t)I(t)+β2emI(t)S(t)B(t)(μ+γ)I(t),dB(t)dt=ξI(t)δB(t). (3.5)

    Thus, under the condition R0>1, the global stability of E1 is equivalent to that of the unique positive equilibrium ˆE1 of system (3.5), where ˆE1=(NˉIˉR,ˉI,ˉB). Define

    ˜Ω={(S,I,B)R3+:S+IN,I>0,B>0}.

    Clearly, ˜Ω is a simply connected region in R3. Similar to the arguments in Theorem 2.2, one can prove that system (3.5) is uniformly persistent in ˜Ω. This, together with the boundedness of solutions to (3.5), implies the existence of a compact absorbing set K˜Ω. Therefore, by Lemma 3.1, it suffices to show that there exists a matrix-valued function P(x) such that ˉq2<0.

    Define the diagonal matrix P as

    P(S,I,B)=diag(1,IB,IB).

    Then P is C1 and nonsigular in ˜Ω. Let f denote the vector field of (3.5). Then

    PfP1=diag(0,IIBB,IIBB),

    where is the derivative with respect to time t. The Jacobian matrix J associated with a general solution (S(t),I(t),B(t)) to (3.5) is

    J=(g(I,B)μσβ1emIS+mSg(I,B)σβ2emISg(I,B)β1emISmSg(I,B)μγβ2emIS0ξδ),

    where g(I,B)=β1emII+β2emIB. The second additive compound matrix of J=(jik)3×3, denoted by J[2], is expressed by

    J[2]=(j11+j22j23j13j32j11+j33j12j31j21j22+j33)=(g(I,B)γ2μσ+β1emISmSg(I,B)β2emISβ2emISξg(I,B)σμδβ1emIS+mSg(I,B)σ0g(I,B)β1emISmSg(I,B)μγδ).

    As a result, the matrix Q=PfP1+PJ[2]P1 can be written in block form as follows:

    Q=(Q11Q12Q21Q22),

    where

    Q11=g(I,B)γ2μσ+β1emISmSg(I,B),Q12=(β2emISBI,β2emISBI),Q21=(IBξ,0)T,Q22=(g(I,B)σμδ+IIBBβ1emIS+mSg(I,B)σg(I,B)β1emISmSg(I,B)μγδ+IIBB).

    The vector norm || in R3 is taken as

    |(u,v,w)|=max{|u|,|v|+|w|}.

    As described in [27], we have the following estimate

    μ(Q)sup{g1,g2},

    with

    g1=μ1(Q11)+|Q12|=g(I,B)γ2μσ+β1emISmSg(I,B)+β2emISBI,g2=|Q21|+μ1(Q22)=IBξ+μ1(Q22).

    Here, |Q12| and |Q21| are matrix norms induced by the l1 norm, and μ1 denotes the Lozinskiǐ measure concerning the l1 norm. Moreover, using the method in [29], we calculate μ1(Q22) as

    μ1(Q22)=max{μσδ+IIBB,H(S,I,B)μγδσ+IIBB+|H(S,I,B)|},=μσδ+IIBB+sup{0,2H(S,I,B)γ},

    where H(S,I,B)=β1emISmSg(I,B)+σ. In view of

    g2g1=μ1(Q22)+IBξ+g(I,B)+μ+σ+mSg(I,B)IIμσδ+IIBB+IBξ+g(I,B)+μ+σ+mSg(I,B)II=μσδBB+δ+BB+g(I,B)+μ+σ+mSg(I,B)=g(I,B)+mSg(I,B)>0,

    we have

    μ(Q)sup{g1,g2}=g2=μσδ+IIBB+IBξ+sup{0,2H(S,I,B)γ}=μσ+II+sup{0,2β1emIS2mSg(I,B)+2σγ}=μσ+II+sup{0,2S(β1emII)2mSβ2emIB+2σγ}μσ+II+sup{0,2S(β1emII)+2σγ}.

    Define the function

    ˜f(S,I)={0,if2S(β1emII)+2σγ0,2S(β1emII)+2σγ,if2S(β1emII)+2σγ>0.

    Along each solution (S(t),I(t),B(t)) to (3.5) such that (S(0),I(0),B(0))K, we find that

    lim supt1tt0˜f(S(s),I(s))ds=0or2σγ,lim supt(1tlnI(t)I(0)μσ)=μσ.

    Here we have used the boundedness of solutions to (3.5). Observe that

    1tt0μ(Q)ds1tlnI(t)I(0)μσ+1tt0˜f(S(s),I(s))ds,

    then ˉq2 has two possibilities:

    (i)ˉq2μσ<0;(ii)ˉq2μγ+σ<0(duetoμ+γ>σ),

    for all (S(0),I(0),B(0))K. Therefore, the desired result holds.

    From Theorem 3.1, we know that if R0>1 and τ=0, E1 is locally asymptotically stable. Further, 0 cannot be an eigenvalue of (3.2) due to p(0,τ)=a2+b2>0 for any τ0. Therefore, the stability of E1 changes only when at least a pair of eigenvalues of (3.2) cross the imaginary axis to the right. We thus suppose that λ=iω (ω>0) is a purely imaginary solution of (3.2) for some τ>0, that is,

    (ω3+a1ω)ia0ω2+a2+eiωτ(b0ω2+ib1ω+b2)=0.

    Separating the real and imaginary parts, we obtain

    ω3+a1ω+b0ω2sin(ωτ)+b1ωcos(ωτ)b2sin(ωτ)=0,a0ω2+a2b0ω2cos(ωτ)+b1ωsin(ωτ)+b2cos(ωτ)=0.

    Equivalently,

    {cos(ωτ)=G(ω)=b1ω(a1ωω3)+(b2b0ω2)(a2a0ω2)(b2b0ω2)2+b21ω2,sin(ωτ)=N(ω)=(b2b0ω2)(a1ωω3)b1ω(a2a0ω2)(b2b0ω2)2+b21ω2. (3.6)

    Squaring and adding both equations of (3.6) yields

    ω6+pω4+qω2+r=0, (3.7)

    with

    p=a202a1b20,q=a212a0a2b21+2b0b2,r=a22b22.

    Let x=ω2. Then equation (3.7) becomes

    x3+px2+qx+r=0.

    Therefore, if iω is a purely imaginary root of (3.2), then the equation

    h(x):=x3+px2+qx+r=0

    has a positive root x=ω2. Define the set W as

    W={(p,q,r)R3|h(x)=0has only one positive real rootxandh(x)>0}.

    By using the properties for the general cubic equations given in [30, Lemma A.2], we have (p,q,r)W if and only if one of the following conditions holds:

    (C1) Δ=(pq9r)24(p23q)(q23pr)>0,r<0;

    (C2) r=0,q=0,p<0;

    (C3) Δ<0,q>0,p>0,r<0.

    In the case where (p,q,r)W, let ω0=x. Solving (3.6) for τ and ω=ω0, we obtain

    τn=τ0+2nπω0,τ0={arccosG(ω0)ω0,N(ω0)0,2πarccosG(ω0)ω0,N(ω0)<0,n=0,1,2,. (3.8)

    Then we have the following result.

    Theorem 3.3. Let R0>1. For (p,q,r)W, E1 is locally asymptotically stable for τ[0,τ0) and unstable for τ>τ0. Besides, system (1.2) undergoes Hopf bifurcation at E1 when τ=τn, n=0,1,2,.

    Proof. Differentiating both sides of equation (3.2) with respect to τ yields

    [dλ(τ)dτ]1=3λ2+2a0λ+a1λ4a0λ3a1λ2a2λτλ+2b0λ+b1b0λ3+b1λ2+b2λ.

    Substituting τ=τn into the above equality, we obtain

    [dReλ(τ)dτ]1|τ=τn=Re(a13ω2+2a0ωia1ω2ω4+a0ω3ia2ωi)Re(2b0ωi+b1(b2ωb0ω3i)b1ω2)=ω23ω2+2ω2(2a1+a20b20)+a212a0a2+2b0b2b21ω2[(b2b0ω2)2+b21ω2]=ω2ω2[(b2b0ω2)2+b21ω2]dh(x)dx|x=ω2.

    Since ω2[(b2b0ω2)2+b21ω2]>0, we have

    sign(dReλ(τ)dτ|τ=τn)=sign([dReλ(τ)dτ]1|τ=τn)=sign(dh(x)dx|x=ω2)=1.

    Therefore, the transversality condition holds, and so the desired result follows.

    Theorem 3.3 states that periodic solutions can bifurcate from E1 when τ is near the local Hopf bifurcation values τn, n=0,1,2,. In this section, we use the global Hopf bifurcation theorem (see [31, Therorem 3.2]) to study the global continuation of these locally bifurcating periodic solutions. In the remainder of this section, we always assume that R0>1, μ+γ>σ and (p,q,r)W.

    Our arguments are similar to those in [1,10,30,32,33,34,35]. Let z(t)=(I(τt),R(τt),B(τt))T. Rewrite system (1.2) as the following functional differential equation

    dz(t)dt=F(zt,τ,T),(t,τ,T)R+×(0,)×R+, (4.1)

    where ztY:=C([1,0],R3+) with zt(θ)=z(t+θ) for θ[1,0], parameter T is the period of the non-constant periodic solution of (4.1), and

    F(zt,τ,T)=τ(emz1t(1)(Nz1t(0)z2t(0))(β1z1t(0)+β2z3t(0))(μ+γ)z1t(0)γz1t(0)(μ+σ)z2t(0)ξz1t(0)δz3t(0)) (4.2)

    with zt=(z1t,z2t,z3t)Y. Restricting F to the subspace of Y, we get a restricted function

    ˜F(z,τ,T):=F|R3×(0,)×R+=τ(emz1(Nz1z2)(β1z1+β2z3)(μ+γ)z1γz1(μ+σ)z2ξz1δz3).

    Clearly, ˜F is twice continuously differentiable, that is, the assumption (A1) in [31] holds.

    By Theorems 2.1 and 2.2, the set of stationary solutions of system (4.1) is given by

    N(F)={(E0,τ,T):(τ,T)(0,)×R+}{(E1,τ,T):(τ,T)(0,)×R+}.

    For any stationary solution (˜z,τ,T)N(F), the characteristic matrix is

    Δ(˜z,τ,T)(λ)=λIdDF(˜z,τ,T)(eλId)=(τβ1˜A+τ˜C+τm(μ+γ)˜z1eλ+τγ+τμ+λτ˜Cτβ2˜Aτγτμ+τσ+λ0τξ0τδ+λ),

    where Id is the 3 × 3 identity matrix, ˜A=em˜z1(N˜z1˜z2) and ˜C=β1em˜z1˜z1+β2em˜z1˜z3. Thus, for any stationary solution (˜z,τ,T), the characteristic equation reads

    detΔ(˜z,τ,T)(λ)=λ3+˜a0τλ2+˜a1τ2λ+˜a2τ3+eλ(˜b0τλ2+˜b1τ2λ+˜b2τ3)=0,

    where

    ˜a0=μ+σ+˜C+β2˜Aξδ+δ,˜a1=˜C(γ+δ)+(μ+σ)(β2˜Aξδ+δ+˜C),˜a2=˜Cδ(μ+σ+γ),˜b0=m(μ+γ)˜z1,˜b1=m(μ+γ)(μ+σ+δ)˜z1,˜b2=mδ(μ+γ)(μ+σ)˜z1.

    Under the condition R0>1, 0 cannot be an eigenvalue of any stationary solution of (4.1). Therefore, the condition (A2) in [31] holds. From Eq (4.2), it can be easily verified that the smoothness condition (A3) in [31] is also valid.

    As defined in [36], the stationary solution (˜z,˜τ,˜T) of (4.1) is called a center if detΔ(˜z,˜τ,˜T)(ik2π˜T)=0 for some positive integer k. A center is isolated if no other center exists in some neighborhood of (˜z,˜τ,˜T) and there are only finite pure imaginary eigenvalues of the form ik2π˜T. Let J(˜z,˜τ,˜T) represent the set of all such positive integers k. Theorem 3.3 implies that if (p,q,r)W, for any integer n0, (E1,τn,2πω0τn) is an isolated center of (4.1), and there is only one pure imaginary root of the form ik2π˜T with k=1 and ˜T=2πω0τn. Therefore,

    J(˜z,˜τ,˜T)={1}. (4.3)

    Moreover, it follows from Theorem 3.3 that the crossing number at each of these center is

    γ1(E1,τn,2πω0τn)=sign(Reλ(τn))=sign(h(x))=1. (4.4)

    Thus the condition (A4) in [31] holds.

    Next, we define a closed subset Σ(F) of Y×(0,)×R+ by

    Σ(F)=Cl{(z,τ,T)Y×(0,)×R+: z is a T -periodic nontrivial solution of (4.1)}.

    For each integer n0, let C(E1,τn,2πω0τn) denote the connected branch of C(E1,τn,2πω0τn) in Σ(F). Theorem 3.3 guarantees that C(E1,τn,2πω0τn) is a nonempty subset of Σ(F). The global bifurcation theorem [31, Theorem 3.4] means that one of the following holds:

    (ⅰ) C(E1,τn,2πω0τn) is unbounded in Y×(0,)×R+;

    (ⅱ) C(E1,τn,2πω0τn) is bounded, C(E1,τn,2πω0τn)N(F) is finite and

    (˜z,τ,T)C(E1,τn,2πω0τn)N(F)γk(˜z,τ,T)=0,

    for all k=1,2,3,, where γk(˜z,τ,T) is the k-th crossing number of (˜z,τ,T) if kJ(˜z,τ,T), otherwise, γk(˜z,τ,T) is zero.

    For each n=1,2,, (˜z,τ,T)C(E1,τn,2πω0τn), based on (4.3) and (4.4), we see

    (˜z,τ,T)C(E1,τn,2πω0τn)N(F)γk(˜z,τ,T)=γ1(˜z,τ,T)=1.

    Hence, (ⅱ) fails and (ⅰ) holds. The following two lemmas help us confirm the boundedness of the projections of C(E1,τn,2πω0τn) on z-space and T-space.

    Lemma 4.1. For initial value ϕ=(ϕ1,ϕ2,ϕ3)Y with ϕ1(θ)+ϕ2(θ)Λμ,θ[1,0], all periodic solutions of system (4.1) are uniformly bounded.

    The proof is a direct result of Lemma 2.1, and hence is omitted.

    Lemma 4.2. If R0>1 and μ+γ>σ, then system (4.1) has no periodic solutions of period 1.

    Proof. Suppose that z(t)=(z1(t),z2(t),z3(t)) is the periodic solution of system (4.1) with period 1, then z(t) is a periodic solution of the following ordinary differential equations

    {dz1(t)dt=τemz1(t)(Nz1(t)z2(t))(β1z1(t)+β2z3(t))τ(μ+γ)z1(t),dz2(t)dt=τγz1(t)τ(μ+σ)z2(t),dz3(t)dt=τξz1(t)τδz3(t). (4.5)

    By Theorem 3, system (4.5) has a unique positive equilibrium that is globally stable, and thus no periodic solutions appear. This leads to a contradiction.

    Lemma 4.1 shows that for any integer n0, the projection of C(E1,τn,2πω0τn) onto z-space is bounded. Lemma 4.2 implies that system (4.1) also has no periodic solutions of period 1n+1 for any n0, Moreover, with the help of (3.8), we obtain

    1n+1<2πω0τn<1,n=1,2,.

    Therefore, the projection of C(E1,τn,2πω0τn) onto T-space is bounded. Accordingly, the projection of C(E1,τn,2πω0τn) onto τ-space is unbounded.

    Summarizing the above discussion, we arrive at the following result.

    Theorem 4.1. Assume that R0>1, μ+γ>σ and (p,q,r)W, then for any τ>τ1 system (1.2) has at least one nontrivial periodic solution.

    In this section, we carry out numerical simulations to demonstrate the theoretical results. Particularly, the global Hopf branches are computed by a Matlab package DDE-BIFTOOL developed by Engelborghs et al. [37,38].

    For illustrative purpose, we choose the parameters of system (1.1) as follows:

    Λ=31,β1=0.09,β2=0.08,μ=0.81,m=0.1,σ=0.3,γ=0.1,ξ=0.5,δ=0.2.

    Direct calculation gives R0=12.1964>1 and (p,q,r)W. By using (3.8), we further obtain τ0=5.5131, τ1=18.3517, τ2=31.1903,. Figure 1 shows that the endemic equilibrium E1=(17.8880,1.6115,44.7201) is asymptotically stable when τ=4<τ0, while Figure 2 displays that E1 loses stability and a Hopf bifurcation occurs when τ=8>τ0. These numerical results agree with the result of Theorem 3.3.

    Figure 1.  A solution converges to the stable equilibrium E1 when τ=4<τ0.
    Figure 2.  A solution converges to a stable periodic solution when τ=8>τ0.

    Moreover, we depict the global Hopf branches of periodic solutions emanating from the Hopf bifurcation points τ0, τ1 and τ2. As seen in Figure 3, when τ0<τ<τ1, system (1.2) has only one periodic solution originating from τ0. When τ lies between τ1 and τ2, the periodic solutions from τ0 and τ1 coexist. With the further increase of τ and τ>τ2, three periodic solutions originating from τ0, τ1 and τ2 coexist.

    Figure 3.  Global Hopf branches of system (1.2) at τ0=5.5131, τ1=18.3517, and τ2=31.1903, respectively.

    Finally, we use the delay as the bifurcation parameter to plot the bifurcation diagram. Figure 4 demonstrates the onset and global continuation of Hopf bifurcations as τ varies.

    Figure 4.  Bifurcation diagram of (1.2) with respect to τ, where red dashed line represents unstable equilibrium.

    Models related to the impact of media coverage on disease spread have shown great popularity in recent years. To study the effect of media coverage on cholera transmission, we considered a cholera model with delayed media impact. We showed that the basic reproduction number R0 is an epidemic threshold parameter that determines the extinction and uniform persistence of the disease. However, this threshold phenomenon is not influenced by the delayed media impact, since R0 is independent of m and τ. This observation motivates us to further explore the impact of media coverage. We proved that the positive equilibrium E1 is locally asymptotically stable when R0>1 and τ[0,τ0), and is unstable when τ>τ0 (see Theorem 3.3). Furthermore, system (1.2) undergoes a Hopf bifurcation at E1 along the sequence τn, n=0,1,2,. To examine the onset and termination of periodic solutions bifurcated from E1, we use delay as the bifurcation parameter and establish the existence of global bifurcation (see Theorem 4.1).

    The local stability and the Hopf bifurcation analysis of the positive equilibrium E1 are critically dependent on the existence and distribution of the roots of the cubic equation h(x)=0 [39]. In this paper, we considered only case where h(x)=0 has only one simple positive root. Therefore, it is exciting to study properties of the global Hopf branches that accompany the stability switch when h(x)=0 has exactly two or three simple positive roots. Another possible project is to consider a spatial version of model (1.1). For such a model with multiple compartments, the investigation of Hopf bifurcation is challenging. We leave these topics for future research.

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

    This research was supported by the National Natural Science Foundation of China (No. 11971369) and the Fundamental Research Funds for the Central Universities (No. QTZX23004).

    The authors declare that there is no conflict of interest.



    [1] N. Wang, L. Qi, M. Bessane, M. Hao, Global Hopf bifurcation of a two-delay epidemic model with media coverage and asymptomatic infection, J. Differ. Equations, 369 (2023), 1–40. https://doi.org/10.1016/j.jde.2023.05.036 doi: 10.1016/j.jde.2023.05.036
    [2] Y. Xiao, S. Tang, J. Wu, Media impact switching surface during an infectious disease outbreak, Sci. Rep., 5 (2015), 7838. http://doi.org/10.1038/srep07838 doi: 10.1038/srep07838
    [3] J. Cui, Y. Sun, H. Zhu, The impact of media on the control of infectious diseases, J. Dyn. Differ. Equations, 20 (2008), 31–53. http://doi.org/10.1007/s10884-007-9075-0 doi: 10.1007/s10884-007-9075-0
    [4] Y. Li, J. Cui, The effect of constant and pulse vaccination on SIS epidemic models incorporating media coverage, Commun. Nonlinear Sci. Numer. Simul., 14 (2009), 2353–2365. https://doi.org/10.1016/j.cnsns.2008.06.024 doi: 10.1016/j.cnsns.2008.06.024
    [5] R. Liu, J. Wu, H. Zhu, Media/psychological impact on multiple outbreaks of emerging infectious diseases, Comput. Math. Methods Med., 8 (2007), 153–164. http://doi.org/10.1080/17486700701425870 doi: 10.1080/17486700701425870
    [6] Y. Xiao, T. Zhao, S. Tang, Dynamics of an infectious diseases with media/psychology induced non-smooth incidence, Math. Biosci. Eng., 10 (2013), 445–461. https://doi.org/10.3934/mbe.2013.10.445 doi: 10.3934/mbe.2013.10.445
    [7] S. Collinson, J. Heffernan, Modeling the effects of media during an influenza epidemic, BMC Public Health, 14 (2014), 376. https://doi.org/10.1186/1471-2458-14-376 doi: 10.1186/1471-2458-14-376
    [8] Q. Yan, S. Tang, J. Wu, S. Gabriele, Media coverage and hospital notifications: correlation analysis and optimal media impact duration to manage a pandemic, J. Theor. Biol., 390 (2016), 1–13. https://doi.org/10.1016/j.jtbi.2015.11.002 doi: 10.1016/j.jtbi.2015.11.002
    [9] C. Yang, J. Wang, A cholera transmission model incorporating the impact of medical resources, Math. Biosci. Eng., 16 (2019), 5226–5246. https://doi.org/10.3934/mbe.2019261 doi: 10.3934/mbe.2019261
    [10] P. Song, Y. Xiao, Global hopf bifurcation of a delayed equation describing the lag effect of media impact on the spread of infectious disease, J. Math. Biol., 76 (2018), 1249–1267. https://doi.org/10.1007/s00285-017-1173-y doi: 10.1007/s00285-017-1173-y
    [11] T. Zhao, M. Zhao, Global hopf bifurcation analysis of an susceptible-infective-removed epidemic model incorporating media coverage with time delay, J. Biol. Dyn., 11 (2017), 8–24. https://doi.org/10.1080/17513758.2016.1229050 doi: 10.1080/17513758.2016.1229050
    [12] Z. Mukandavire, S. Liao, J. Wang, H. Gaff, D. L. Smith, J. G. Morris, Estimating the reproductive numbers for the 2008–2009 cholera outbreaks in Zimbabwe, Proc. Natl. Acad. Sci. USA, 108 (2011), 8767–8772. https://doi.org/10.1073/pnas.1019712108 doi: 10.1073/pnas.1019712108
    [13] J. Wang, X. Wang, D. Gao, Influence of human behavior on cholera dynamics, Math. Biosci., 267 (2015), 41–52. https://doi.org/10.1016/j.mbs.2015.06.009 doi: 10.1016/j.mbs.2015.06.009
    [14] L. Zhang, Z. Wang, Y. Zhang, Dynamics of a reaction-diffusion waterborne pathogen model with direct and indirect transmission, Comput. Math. Appl., 72 (2016), 202–215. https://doi.org/10.1016/j.camwa.2016.04.046 doi: 10.1016/j.camwa.2016.04.046
    [15] G. Sun, J. Xie, S. Huang, Z. Jin, M. Li, L. Liu, Transmission dynamics of cholera: mathematical modeling and control strategies, Commun. Nonlinear Sci. Numer. Simul., 45 (2017), 235–244. https://doi.org/10.1016/j.cnsns.2016.10.007 doi: 10.1016/j.cnsns.2016.10.007
    [16] J. Wang, J. Wang, Analysis of a reaction-diffusion cholera model with distinct dispersal rates in the human population, J. Dyn. Differ. Equations, 33 (2021), 549–575. http://doi.org/10.1007/s10884-019-09820-8 doi: 10.1007/s10884-019-09820-8
    [17] H. Shu, Z. Ma, X. Wang, Threshold dynamics of a nonlocal and delayed cholera model in a spatially heterogeneous environment, J. Math. Biol., 83 (2021), 41. https://doi.org/10.1007/s00285-021-01672-5 doi: 10.1007/s00285-021-01672-5
    [18] J. Wang, Mathematical models for cholera dynamics-a review, Microorganisms, 10 (2022), 2358. https://doi.org/10.3390/microorganisms10122358 doi: 10.3390/microorganisms10122358
    [19] J. Hale, S. Lunel, Introduction to Functional Differential Equations, Springer, New York, 1993. Available from: https://link.springer.com/book/10.1007/978-1-4612-4342-7.
    [20] H. Smith, Monotone Dynamical Systems: An Introduction to the Theory of Competitive and Cooperative Systems, American Mathematical Society, Providence, RI, 1995. https://doi.org/10.1090/surv/041
    [21] P. Driessche, J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002), 29–48. https://doi.org/10.1016/S0025-5564(02)00108-6 doi: 10.1016/S0025-5564(02)00108-6
    [22] J. Hale, Dynamical systems and stability, J. Math. Anal. Appl., 26 (1969), 39–59. https://doi.org/10.1016/0022-247X(69)90175-9 doi: 10.1016/0022-247X(69)90175-9
    [23] J. Hale, Asymptotic Behavior of Dissipative Systems, American Mathematical Society, Providence, 25 (1988). https://doi.org/10.1090/surv/025
    [24] X. Zhao, Dynamical Systems in Population Biology, 2nd edition, Springer, New York, 2017. https://doi.org/10.1007/978-3-319-56433-3
    [25] X. Zhao, Uniform persistence and periodic coexistence states in infinite-dimensional periodic semiflows with applications, Can. Appl. Math. Q., 3 (1995), 473–495. Available from: https://www.math.mun.ca/~zhao/Selectpapers/Zhao1995CAMQpub.pdf.
    [26] R. Corless, G. Gonnet, D. Hare, D. Jeffrey, D. Knuth, On the LambertW function, Adv. Comput. Math., 5 (1996), 329–359. https://doi.org/10.1007/BF02124750 doi: 10.1007/BF02124750
    [27] M. Y. Li, J. Muldoweny, A geometric approach to the global-stability problems, SIAM J. Math. Anal., 27 (1996), 1070–1083. https://doi.org/10.1137/S0036141094266449 doi: 10.1137/S0036141094266449
    [28] J. Muldowney, Compound matrices and ordinary differential equations, Rocky. Mt. J. Math., 20 (1990), 857–872. Available from: https://www.jstor.org/stable/44237627.
    [29] W. Coppel, Stability and Asymptotic Behavior of Differential Equations, Health, Boston, 1995.
    [30] P. Song, Y. Xiao, Analysis of an epidemic system with two response delays in media impact function, Bull. Math. Biol., 81 (2019), 1582–1612. https://doi.org/10.1007/s11538-019-00586-0 doi: 10.1007/s11538-019-00586-0
    [31] J. Wu, Symmetric functional differential equations and neural networks with memory, Trans. Am. Math. Soc., 350 (1998), 4799–4838. https://doi.org/10.1090/S0002-9947-98-02083-2 doi: 10.1090/S0002-9947-98-02083-2
    [32] J. Wei, M. Y. Li, Hopf bifurcation analysis in a delayed Nicholson blowflies equation, Nonlinear Anal., 60 (2005), 1351–1367. https://doi.org/10.1016/j.na.2003.04.002 doi: 10.1016/j.na.2003.04.002
    [33] H. Shu, L. Wang, J. Wu, Global dynamics of Nicholson's blowflies equation revisited: onset and termination of nonlinear oscillations, J. Differ. Equations, 255 (2013), 2565–2586. https://doi.org/10.1016/j.jde.2013.06.020 doi: 10.1016/j.jde.2013.06.020
    [34] H. Shu, G. Fan, H. Zhu, Global hopf bifurcation and dynamics of a stage–structured model with delays for tick population, J. Differ. Equations, 284 (2021), 1–22. https://doi.org/10.1016/j.jde.2021.02.037 doi: 10.1016/j.jde.2021.02.037
    [35] X. Zhang, F. Scarabel, X. Wang, J. Wu, Global continuation of periodic oscillations to a diapause rhythm, J. Dyn. Differ. Equations, 34 (2022), 2819–2839. https://doi.org/10.1007/s10884-020-09856-1 doi: 10.1007/s10884-020-09856-1
    [36] J. Wei, H. Wang, W. Jiang, Bifurcation Theory of Delay Differential Equations, Science Press, 2012.
    [37] K. Engelborghs, T. Luzyanina, G. Samaey, DDE-BIFTOOL v. 2.00: A Matlab package for bifurcation analysis of delay differential equations, Technical Report TW-330, KU Leuven, Belgium, 2001. Available from: https://www.researchgate.net/publication/245840825.
    [38] K. Engelborghs, T. Luzyanina, D. Roose, Numerical bifurcation analysis of delay differential equations using DDE-BIFTOOL, ACM Trans. Math. Software, 28 (2002), 1–21. https://doi.org/10.1145/513001.513002 doi: 10.1145/513001.513002
    [39] H. Shu, X. Hu, L. Wang, J. Watmough, Delay induced stability switch, multitype bistability and chaos in an intraguild predation model, J. Math. Biol., 71 (2015), 1269–1298. https://doi.org/10.1007/s00285-015-0857-4 doi: 10.1007/s00285-015-0857-4
  • This article has been cited by:

    1. Shenxing Li, Wenhe Li, Dynamical Behaviors of a Stochastic Susceptible-Infected-Treated-Recovered-Susceptible Cholera Model with Ornstein-Uhlenbeck Process, 2024, 12, 2227-7390, 2163, 10.3390/math12142163
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1720) PDF downloads(155) Cited by(1)

Figures and Tables

Figures(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog