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

Dynamics of a malaria infection model with time delay

  • In this paper, a new mathematical model (a system of delay differential equation) is proposed to describe dynamical behaviors of malaria in an infected host with red blood cells (RBCs), infected red blood cells (iRBCs) and immune factors. The basic reproduction number 0 of the malaria infection is derived. If 01, the uninfected equilibrium E0 is globally asymptotically stable. If 0>1, there exists two kinds of infection equilibria. The conditions of these equilibria with respect to the existence, stability and uniform persistence are given. Furthermore, fluctuations occur when the model undergoes Hopf bifurcation, and periodic solution appears near the positive equilibrium. The direction and stability of Hopf bifurcation are also obtained by applying the center manifold method and the normal form theory. Numerical simulations are provided to demonstrate the theoretical results.

    Citation: Qian Ding, Jian Liu, Zhiming Guo. Dynamics of a malaria infection model with time delay[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 4885-4907. doi: 10.3934/mbe.2019246

    Related Papers:

    [1] Yilong Li, Shigui Ruan, Dongmei Xiao . The Within-Host dynamics of malaria infection with immune response. Mathematical Biosciences and Engineering, 2011, 8(4): 999-1018. doi: 10.3934/mbe.2011.8.999
    [2] 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
    [3] Cruz Vargas-De-León . Global analysis of a delayed vector-bias model for malaria transmission with incubation period in mosquitoes. Mathematical Biosciences and Engineering, 2012, 9(1): 165-174. doi: 10.3934/mbe.2012.9.165
    [4] Yu Yang, Gang Huang, Yueping Dong . Stability and Hopf bifurcation of an HIV infection model with two time delays. Mathematical Biosciences and Engineering, 2023, 20(2): 1938-1959. doi: 10.3934/mbe.2023089
    [5] 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
    [6] Tianqi Song, Chuncheng Wang, Boping Tian . Mathematical models for within-host competition of malaria parasites. Mathematical Biosciences and Engineering, 2019, 16(6): 6623-6653. doi: 10.3934/mbe.2019330
    [7] Shunyi Li . Hopf bifurcation, stability switches and chaos in a prey-predator system with three stage structure and two time delays. Mathematical Biosciences and Engineering, 2019, 16(6): 6934-6961. doi: 10.3934/mbe.2019348
    [8] 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
    [9] Liming Cai, Shangbing Ai, Guihong Fan . Dynamics of delayed mosquitoes populations models with two different strategies of releasing sterile mosquitoes. Mathematical Biosciences and Engineering, 2018, 15(5): 1181-1202. doi: 10.3934/mbe.2018054
    [10] 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
  • In this paper, a new mathematical model (a system of delay differential equation) is proposed to describe dynamical behaviors of malaria in an infected host with red blood cells (RBCs), infected red blood cells (iRBCs) and immune factors. The basic reproduction number 0 of the malaria infection is derived. If 01, the uninfected equilibrium E0 is globally asymptotically stable. If 0>1, there exists two kinds of infection equilibria. The conditions of these equilibria with respect to the existence, stability and uniform persistence are given. Furthermore, fluctuations occur when the model undergoes Hopf bifurcation, and periodic solution appears near the positive equilibrium. The direction and stability of Hopf bifurcation are also obtained by applying the center manifold method and the normal form theory. Numerical simulations are provided to demonstrate the theoretical results.


    Malaria is one of the most dangerous global health problems. It causes millions of infections every year, and more than 1.1 million people died from the disease, especially infants, young children and pregnant women (WHO [1]). The mainly affected group is children under the age of five, accounting for 82% of all malaria deaths. It is endemic in the tropical and subtropical regions of the world, and people in Africa are at the risk of such disease accounting for 27% of the world's malaria infections (WHO [1]). It is reported that about every 30 seconds, a child's life will be threatened by such disease. Although some children are survived, they still suffer miserable physical problems, such as hearing impairment and brain damage. Meanwhile, pregnant women are also susceptible to malaria infection, which can cause maternal mortality, low birth weight in infants as well as maternal anaemia. According to the statistics, it caused an estimated 243 million malaria cases led to an estimated 863,000 deaths in 2008 (WHO [1]).

    Malaria infection is mainly induced by the following four malaria parasites: P. falciparum, P. malariae, P. ovale and P. vivax. Among them, P. falciparum is responsible for almost all of the deaths attributed to malaria [2]. And anopheles is a vector of malaria. Once the infected anopheles bites the host, the malaria parasites first penetrate liver cells of the host and then move into the blood, where they multiply and undergo replication cycles in the red blood cells. After 10-14 days or more, the malaria parasites develop and mature in the biting mosquitos until they can infect others. When malaria parasites evolve in the host, they can stimulate the activity of immune cells which produce an immune response.

    In the past few years, lots of efforts have been carried out to control and prevent such disease [3]. However, people living in high epidemic areas are still hardly to obtain effective malaria prevention, diagnosis and treatment. As the emergence of malaria drug resistant strains and insecticide resistant mosquitoes, how to prevent and control malaria infection is still a problem. The main detriment of plasmodium in the host occurs in the blood-stage, where the parasites will develop and reproduce. In order to clarify the mechanism, many mathematical models have been employed to describe the dynamics of malaria infection.

    The first model was proposed by Hetzel and Anderson [4] which studied the performance of a mathematical model for the blood phase of malaria infection with a single strain. Analyzing the cells population dynamics in the absence of a host immune response, the authors demonstrated a relationship between host and parasite parameters, which defined a criterion for the successful invasion and persistence of the parasite. Hetzel and Anderson [4] indicated some important parameters, such as the production rates of merozoite and erythrocyte, the mortality rates of merozoite and erythrocyte as well as the invasion rate of merozoite.

    In 2011, Li et al. [5] studied the dynamics for malaria parasites infection in the host blood-stage. Considered that the clearance of host immune cells was restricted by concentration, the following model was established:

    {dHdt=λd1HαHM,dIdt=αHMδIp1IE1+βI,dMdt=rIμMp2ME1+γM,dEdt=d2E+k1IE1+βI+k2ME1+γM,

    where H,I,M and E are population of red blood cells, infected red blood cells, malaria parasites and immunity effectors, respectively. d1,δ,μ and d2 represent the decay rate of H(t), I(t), M(t) and E(t) respectively. λ, α and r are production rate of H(t), infection of H(t) by M(t), product rate of M(t) respectively. p1,p2,k1,k2,β and γ represent the removal rate of I(t) and M(t) by E(t), proliferation rate of E(t) by I(t) and M(T), 1β means half saturation constant for I(t) and 1γ is half saturation constant for M(t) respectively. In [5], the authors showed that there existed a threshold value 0 and the malaria-free equilibrium was globally asymptotically stable if 0<1, if 0>1, there existed two kinds of infection equilibria: malaria infection equilibrium (without specific immune response) and positive equilibrium (with specific immune response). Conditions on the existence and stability of both infection equilibria are given. Moreover, it indicated that the model could undergo Hopf bifurcation at the positive equilibrium and exhibit periodic oscillations.

    A malaria model between human and mosquitoes was established by Xiao and Zou [6] which presented an important conclusion that two malaria parasites may co-exist in the same region but not in the same host.

    In 2016, Chang [7] made a great progress in the study of the innate immune mechanism of plasmodium falciparum escaping the immune clearance in the host. Innate immune response is the first line for host to defense pathogen invasion. After being stimulated by the red blood cells, the host neutrophils release chromatin and lysosomal granules in the cytoplasm to form a reticular structure (NETs) in a way of active death (Netosis). NETs can capture and kill infected red blood cells. At the same time, plasmodium degrades NETs by secreting DNA enzymes (TatD-like DNase) to escape host immune elimination. Chang [7] indicated that TatD-like DNase was an important factor in the survival of malaria parasites as well as a potential candidate for malaria vaccine.

    In terms of a great deal of literatures in malaria infection [8,9,10,11,12], we focus on some more important factors and hope to establish new models that can be used to study malaria infection in a host from different perspectives. Liu [13] considered three dynamical variables of populations: RBCs T(t), iRBCs I(t) and the immunity effectors (NETs) M(t). Here the model reads as

    {dT(t)dt=λd1T(t)μT(t)I(t),dI(t)dt=μT(t)I(t)d2I(t)αI(t)M(t),dM(t)dt=βI(t)M(t)d3M(t). (1.1)

    The variables and parameters are given in the table below.

    The main results of model (1.1) have been shown in [13]. And here, we just simply introduce the results as below.

    (ⅰ) If 0<1, then the infection-free equilibrium is globally asymptotically stable.

    (ⅱ) If 1<0<1+d3μd1β, then the infection equilibrium (without specific immune response) is globally asymptotically stable.

    (ⅲ) If 0>1+d3μd1β, then the infection equilibrium (with specific immune response) is globally asymptotically stable.

    However, in real life, malaria infection does not tend to a fixed state over time, it turns out that some diseases will recur over a period of time or even change periodically. That means, it is not a good way to study the dynamics of malaria in host by using an ordinary differential equation model. Hence, in order to describe the time lag of the generation of immune factors, we consider adding a delay term to model (1.1). The delay differential equation model is as follows.

    {dT(t)dt=λd1T(t)μT(t)I(t),dI(t)dt=μT(t)I(t)d2I(t)αI(t)M(t),dM(t)dt=βI(tτ)M(tτ)d3M(t).

    Compared with model (1.1), this model adds a time delay τ in the production item of immune factors βI(t)M(t). In the following analysis, we calculated the characteristic root of the characteristic transcendental equation and found that the stability of positive equilibrium is changed from stable to unstable, and thus the Hopf bifurcation occurs.

    This paper is organized as follows. In section 2, we propose a delay differential equation model for the within-host dynamics of malaria infection based on the basic reproduction number, existence of equilibria, stability of infection-free equilibrium E0 and infection equilibria E1,E as well as persistence of positive equilibrium E. Section 3 is devoted to exhibit periodic oscillations due to the Hopf bifurcation at the positive equilibrium E. Specifically, by choosing immune response delay as bifurcation parameter, we can demonstrate that a limit cycle occurs via Hopf bifurcation when the time delay τ passes through the critical value. The direction and stability of Hopf bifurcation are derived by applying the center manifold method and the normal form theory. In section 4, some numerical simulations are presented to interpret our main results biologically. A short discussion on research results is also given in this section.

    Consider the delay differential equation model

    {dT(t)dt=λd1T(t)μT(t)I(t),dI(t)dt=μT(t)I(t)d2I(t)αI(t)M(t),dM(t)dt=βI(tτ)M(tτ)d3M(t). (2.1)

    The initial conditions of (2.1) are given as

    T(θ)=φ1(θ)0,I(θ)=φ2(θ)0,M(θ)=φ3(θ)0,θ[τ,0],φ1(0)>0,φ2(0)>0,φ3(0)>0, (2.2)

    where (φ1,φ2,φ3)C([τ,0],R3+), the Banach space of continuous functions map the interval [τ,0] into R3+. By fundamental theory of functional differential equation (Kuang [14]), there exists a unique solution (T(t),I(t),M(t)) of (2.1) with initial conditions (2.2).

    Upon the positivity and boundedness of solutions for model (2.1), we claim the following result.

    Theorem 2.1. Let (T(t),I(t),M(t)) be the solution of model (2.1) with initial conditions (2.2), then T(t),I(t),M(t) are positive and ultimately bounded for all t0.

    Proof. The positivity of the model (2.1) follows directly from Theorem 3.4 [15]. As for the boundedness of solutions to the model (2.1) with initial conditions (2.2), we denote

    L(t)=T(t)+I(t)+αβM(t+τ).

    By the positivity of solutions, we have

    dL(t)dt=λd1T(t)μT(t)I(t)+μT(t)I(t)d2I(t)αI(t)M(t)+αI(t)M(t)αd3βM(t+τ)=λd1T(t)d2I(t)αd3βM(t+τ)λmL(t), (2.3)

    where m=min{d1,d2,d3}. Hence, it comes

    lim suptL(t)λm.

    This implies that T(t),I(t),M(t) are ultimately bounded.

    Next, we define threshold values 0 and 1. For model (2.1), the basic reproduction number of iRBCs, which describes the average number of newly infected cells generated by one infected cell during its lifespan, is defined by

    0=λμd1d2,

    and the immune response threshold value is defined as below

    1=1+d3μd1β.

    We will see that 0 and 1 play key roles in determining the existence and stability of equilibria of the model (2.1). Actually, let

    {λd1TμTI=0,μTId2IαIM=0,βIMd3M=0. (2.4)

    The following results can be verified by direct calculations.

    (ⅰ) If 01, then (2.1) always has a unique infection-free equilibrium E0=(T0,0,0), where T0=λd1.

    (ⅱ) If 1<01, then (2.1) has two equilibria, E0 and infection equilibrium (without specific immune response) E1=(T1,I1,0), where T1=d2μ and I1=λμd1d2d2μ=d1(01)μ.

    (ⅲ) If 0>1, then (2.1) has three equilibria, E0,E1 and infection equilibrium (with specific immune response) E=(T,I,M) respectively, where T=λβd1β+d3μ,I=d3β and M=λμβd1d2βd2d3μd1αβ+d3αμ=d1d2β(01)d1αβ+d3αμ.

    Now, we investigate the stability of the equilibria of model (2.1). Consider the linearization of the model (2.1) at E=(T(t),I(t),M(t)),

    x(t)=Ax(t)+Bx(tτ), (2.5)

    where x(t)=(T(t),I(t),M(t))T,

    A=(d1μI(t)μT(t)0μI(t)μT(t)d2αM(t)αI(t)00d3). (2.6)
    B=(0000000βM(tτ)βI(tτ)). (2.7)

    The characteristic equation of (2.5) is formulated as (Hale [16])

    det[ΛIAeΛτB]=0. (2.8)

    Theorem 2.2. If 0<1, then the infection-free equilibrium E0 of model (2.1) is locally asymptotically stable. Moreover, E0 is unstable if 0>1.

    Proof. By substituting E0 in (2.5), (2.6), (2.7) and (2.8), we obtain the following characteristic equation

    (Λ+d1)(Λ+d3)(Λλμd1+d2)=0.

    Clearly, Λ1=d1, Λ2=d3 and

    Λ3=λμd1d2=d2(01),

    which dominates the stability of E0. If 0<1, then Λ3<0. Therefore, when 0<1, all eigenvalues of model (2.1) have negative real parts and hence, equilibrium E0 is locally asymptotically stable. In addition, when 0>1, we have Λ3>0, which means E0 is unstable.

    Theorem 2.3. If 01, then the infection-free equilibrium E0 of model (2.1) is globally asymptotically stable.

    Proof. Define Lyapunov functional V0 as follows

    V0=TT0T0lnTT0+I+αβM+αttτI(θ)M(θ)dθ.

    Differentiating V0 with time t along the model (2.1), then

    dV0dt=(1T0T)(d1T0d1TμTI)+(μTId2IαIM)+αβ(βIMd3M)=d1T(TT0)2μTI+μT0I+μTId2Id3αβM=d1T(TT0)2+(λμd1d2)Id3αβM=d1T(TT0)2+d2(01)Id3αβM.

    Obviously, if 01, then dV0dt0 for any T(t),I(t) and M(t). In addition, dV0dt=0 if and only if T(t)=T0,I(t)=0 and M(t)=0. Let Γ be the largest invariant set of {(T(t),I(t),M(t)R3+dV0dt=0}, we easily get Γ={E0}. According to LaSalles invariance principle (Kuang [14]), it follows that equilibrium E0 is globally asymptotically stable when 01. This completes the proof.

    Theorem 2.4. If 1<0<1, then the infection equilibrium (without specific immune response) E1 of model (2.1) is locally asymptotically stable.

    Proof. By substituting E1 into (2.5), (2.6), (2.7) and (2.8), we have the following characteristic equation

    [Λd3(eΛτd1βd3μ(01)1)][Λ2+d10Λ+d1d2(01)]=0. (2.9)

    It is easy to check that Λ2+d10Λ+d1d2(01)=0 has two eigenvalues Λ1 and Λ2 with negative real parts if 1<0<1. We assume that Λ=a+ib satisfies

    Λd3[eΛτd1βd3μ(01)1]=0. (2.10)

    Substituting Λ=a+ib into (2.10) and separating the real and imaginary parts, we obtain

    {a=d3[βd1(01)d3μeaτcos(bτ)1],b=d3[βd1(01)d3μeaτsin(bτ)]. (2.11)

    Assume that a0, when 1<0<1, then a<d3[eaτcos(bτ)1] from the right hand side of the first equation of (2.11). This contradicts the assumption. So we get a<0, i.e. ReΛ3<0. Thus, all eigenvalues of equation (2.9) have negative real parts and equilibrium E1 is locally asymptotically stable.

    Theorem 2.5. If 1<01, then the infection equilibrium (without specific immune response) E1 of model (2.1) is globally asymptotically stable.

    Proof. Define Lyapunov functional V1 as follows

    V1=TT1T1lnTT1+II1I1lnII1+αβM+αttτI(θ)M(θ)dθ.

    Differentiating V1 with time t along model (2.1), then

    dV1dt=(1T1T)dTdt+(1I1I)dIdt+αβ(βIMd3M)=d1T(TT1)2+μT1I1(1TIT1I1T1T+I1I)+(1I1I)(μTIμT1I+αIM1αIM)+αIMαI1M=d1T(TT1)2+μT1I1(1TIT1I1T1T+I1I)+μT1I1(1I1I)(TIT1I1II1)=d1T(TT1)2+μT1I1(1TIT1I1T1T+II1)+μT1I1(TIT1I1II1TT1+1)=d1T(TT1)2+μT1I1(2TT1T1T).

    Obviously, if 1<01, then dV1dt0 for any T(t),I(t) and M(t). In addition, dV1dt=0 if and only if T(t)=T1,I(t)=I1 and M(t)=0. By LaSalles invariance principle (Kuang [14]), equilibrium E1 is globally asymptotically stable when 1<01. The proof is completed.

    Next, we study the stability of the infected equilibrium (with specific immune response), which is denoted by positive equilibrium E. Recall that the positive equilibrium E exists if and only if 0>1. At equilibrium E, the characteristic equation for the corresponding linearized model of (2.1) is

    Λ3+AΛ2+BΛ+C+(A1Λ2+B1Λ+C1)eΛτ=0, (2.12)

    where

    A=d1+d3+μI,A1=d3,B=d1d3+d3μI+μ2TI,B1=d1d3d3μI+d3αM,C=d3μ2TI,C1=d1d3αMd2d3μI.

    When τ=0, equation (2.12) becomes

    Λ3+(A+A1)Λ2+(B+B1)Λ+(C+C1)=0, (2.13)

    where

    A+A1=d1+μI>0,B+B1=μ2TI+d3αM>0,C+C1=d3αM(μI+d1)>0. (2.14)

    Clearly, (A+A1)(B+B1)>C+C1. By Hurwitz criteria, all eigenvalues of equation (2.13) have negative real parts, which leads to the local stability of equilibrium E when τ=0. In fact, in this case, E is also globally stable.

    Theorem 2.6. If 0>1 and τ=0, then the infection equilibrium (with specific immune response) E of model (2.1) is globally asymptotically stable.

    Proof. Define Lyapunov functional V that

    V=TTTlnTT+IIIlnII+αβ(MMMlnMM).

    Differentiating V with respect to t along model (2.1), we get

    dVdt=(1TT)dTdt+(1II)dIdt+αβ(1MM)dMdt=(1TT)(d1T+μTId1TμTI)+(1II)[μTI(μTαM)IαIM]+αβ(1MM)(βIMβIM)=d1T(TT)2+(μTId1TμTI)+(1II)(μTIμTI)+(1II)(αMIαIM)+αMI(1MM)(IMIMMM)=d1T(TT)2+μTI(1TT)(1TITI)+μTI(1II)(TITIII)+αMI(1II)(IIIMIM)+αMI(IMIMMMII+1)=d1T(TT)2+μTI(1TITITT+II)+μTI(TITIIITT+1)+αMI(IIIMIM1+MM)+αMI(IMIMMMII+1)=d1T(TT)2+μTI(2TTTT).

    If 0>1, then dVdt0 for any T(t),I(t) and M(t). In addition, dVdt=0 if and only if T(t)=T,I(t)=I and M(t)=M. By using LaSalles invariance principle (Kuang [14]), equilibrium E is globally asymptotically stable when 0>1. This completes the proof.

    Now we consider the stability of the equilibrium E if 0>1 and τ>0. By theorem 4.4 of Hal Smith [15], for small enough delay, the characteristic roots of (2.12) are either very near the eigenvalues of (2.13) or having more negative real parts than any of the eigenvalues of (2.13). Hence, when the delay is small the equilibrium E is locally asymptotically stable.

    When 0>1, for any τ>0, zero is not a root of (2.12). Note that any complex roots to the equations (2.12) appear in pairs, and all roots of (2.12) have negative real parts if τ=0. Therefore, any root of (2.12) has negative real part for sufficiently small τ. Assume that there exists τ=τ, such that (2.12) has a pair of pure imaginary roots, denoted by Λ=±ωi,(ω>0). Substituting Λ=ωi into (2.12), we have,

    i(ω3Bω)Aω2+C+(A1ω2+iB1ω+C1)(cosωτisinωτ)=0.

    Separate the real and imaginary parts,

    {(C1A1ω2)cosωτ+B1ωsinωτ=Aω2C,B1ωcosωτ(C1A1ω2)sinωτ=ω3Bω.

    It follows that

    ω6+(A22BA21)ω4+(B2+2A1C12ACB21)ω2+C2C21=0.

    Let z=ω2, it follows that

    z3+p1z2+q1z+r1=0, (2.15)

    where

    p1=A22BA21,q1=B2+2A1C12ACB21,r1=C2C21.

    By calculating p1,q1 and r1, we obtain

    p1=(d1+d3+d3μβ)22d1d3d232d23μβ2d3λμ2β(d1+d3μβ)=d21+(d3μβ)2+2d1d3μβ2d3λμ2d1β+d3μ=d21+d21(11)2+2d21(11)2d1d2(11)01=d21212d1d2(001),
    q1=[d1d3+d23μβ+d23λμ2β(d1+d3μβ)]2[d1d3+d3(d2λμd11)+d23μβ]2+2d3[d2d23μβ+d1d3(d2λμd11)]2d23λμ2(d1+d3+d3μβ)d1β1=[d1d31+d1d201(11)]2[d1d31+d2d3(101)]2+2d3(d1d2d31d1d2d301)2d1d2d3(11)01(d11+d3)=2d21d2d30(11)+[d1d201(11)]2d1d2d231(101)[d2d3(101)]22d1d2d3[d3(01)+d10(11)]=[d1d201(11)+d2d3(101)][d1d201(11)d2d3(101)],
    r1=[d23λμ2β(d1+d3μβ)]2[d1d3(d2λμd1+d3μβ)+d2d23μβ]2=[d1d2d3(001)]2[d1d2d3(101)+d1d2d3(11)]2=d21d22d23[20(121)21+20].

    Equation (2.15) has no positive roots if it satisfies p1>0,r1>0,p1q1r1>0 (Hurwitz criterion). Hence, τ doesn't exist. It follows that all solutions of equation (2.15) have negative real part and E is locally asymptotically stable for any τ>0. Therefore, we have the following theorem.

    Theorem 2.7. If 0>1 and p1>0,r1>0,p1q1r1>0, then the infection equilibrium E is locally asymptotically stable for any τ>0.

    Theorem 2.8. If 0>1 and r1<0, there exists τ>0, such that equation (2.15) has a pair of conjugate purely imaginary roots ±ωi when τ<τ. Moreover, the infection equilibrium E is locally asymptotically stable if τ<τ.

    Proof. We know that if r1<0, then equation (2.15) has at least one positive root and not more than three positive roots. Assume that equation (2.15) has three positive roots, denoted by zk,k=1,2,3, it follows that

    τnk=1ωk[arccos(B1AA1)ω4k+(C1A+A1CBB1)ω2kCC1A21ω4k+(B212A1C1)ω2k+C21+2nπ],

    where k=1,2,3,nZ+. Suppose that

    τ=mink=1,2,3{τ0k}.

    When τ=τ, equation (2.15) has a pair of conjugate purely imaginary roots. Note that all roots of equation (2.15) have negative real part if τ=0, and all roots of equation (2.15) have negative real part if τ<τ, which follows from the continuous dependence of the solution on τ.

    In what follows, we study the persistence of equilibrium E, following the analysis of Wang et al. [17].

    Lemma 2.1. Let X be a locally compact metric space and it is the union of two disjoint subsets X1 and X2, with X2 be compact in X, X1 be open and forward invariant under the continuous semiflow Φ on X. Assume that

    Ω2=yY2ω(y),Y2={xX2;Φt(x)X2,t>0},

    has an acyclic isolated covering M=mk=1Mk. If each part Mk of M is a weak repeller for X1, then X2 is a uniform strong repeller for X1.

    Theorem 2.9. If 0>1, then M is uniformly persistent, that is, there exists a positive constant σ, such that the positive solutions of model (2.1) satisfy lim inftM(t)>σ.

    Proof. By Theorem 2.1, model (2.1) is point dissipative. Let

    X=R3+={(T,I,M)|T0,I0,M0,T+I+αβMλm},
    X1={(T,I,M)L|M>0,T+I+αβMλm},X2=XX1

    To prove Theorem 2.9, we need to show that X2 is uniformly strong repeller for X1. It suffice to verify the conditions of Lemma 2.1. Obviously, X is locally compact, X2 is compact and X1 is forward invariant. In addition, there are two equilibria E0=(λd1,0,0) and E1=(d2μ,λμd1d2d2μ,0) in X2. Consider

    {dT(t)dt=λd1T(t)μT(t)I(t),dI(t)dt=μT(t)I(t)d2I(t). (2.16)

    By Theorem 2.5, E11=(T1,I1)=(d2μ,λμd1d2d2μ) is globally asymptotically stable for solutions (T(t),I(t)) to (2.16) with positive initial value if 1<01. In addition, let Y={(T,0)|T0}, then Y is positively invariant for (2.16) and the solutions to (2.16) in Y approach E00=(T0,0)=(λd1,0) as t, it becomes zero or unbounded as t. Therefore, it follows that

    Ω2={E11,E00},M=M1M2=E11E00.

    Thus, we know that M is acyclic in X2. From theorem 2.4, 2.5 and the Jacobian matrix at E00, it is easy to check that E00 and E11 are hyperbolic if 0>1, which implies that E00 and E11 are isolated in X2. From what has been discussed above, Ω2 has an acyclic isolated covering M=E00E11.

    Next, we need to show that each part Mk(k=1,2) of M is a weak repeller for X1. It suffice to show that Ws(E11)X1= and Ws(E00)X1=, where Ws(E11) denotes the stable manifold of E11, and Ws(E00) denotes the stable manifold of E00. Suppose Ws(E11)X1, then there is a solution of model (2.1) such that limt(T(t),I(t),M(t))E11. Therefore, from the third equation of model (2.1) that for any ε>0, there exists t0>0, then we obtain

    M(t)M[β(λμd1d2)d2μεd3],tt0. (2.17)

    Note that

    g(ε)=β(λμd1d2)d2μεd3,

    then for tt0 and M(t)g(ε)M. Since 0>1, we can choose ε sufficiently small such that g(ε)>0. Thus, (2.17) shows that M(t) as t. This contradicts the ultimate boundedness of Theorem 2.1. Hence, Ws(E11)X1=. Similarly, we can prove Ws(E00)X1=.

    In the following, we shall determine when the equilibrium E becomes unstable and Hopf bifurcation occurs. We will seek conditions which guarantee characteristic equation (2.15) to have a root with negative real part as well as a pair of conjugate purely imaginary roots.

    Theorem 3.1. Suppose 0>1 and r1<0, then there exists τ>0. If τ>τ, there is a Hopf bifurcation of model (2.1) from equilibrium E as τ passes through the critical value τ.

    Proof. Consider the derivative of eigenvalue on real axis across section of positive equilibrium point E on τ=τ ([18]). Assume that Λ(τ)=ξ(τ)+iω(τ) is the eigenvalue of equation (2.15) at τ near τ and calculate the derivative of equation (2.15) on τ, we obtain ([19])

    (3Λ2+2AΛ+B)dΛdτ+(2A1Λ+B1)eΛτdΛdτ(A1Λ2+B1Λ+C1)eΛτ(Λ+τdΛdτ)=0.

    This gives

    dΛdτ=Λ(A1Λ2+B1Λ+C1)eΛτ3Λ2+2AΛ+B+[2A1Λ+B1τ(A1Λ2+B1Λ+C1)]eΛτ.

    For ξ(τ)=0 and ω(τ)=ω, then

    [dReΛ(τ)dτ]1=Re[(3Λ2+2AΛ+B)eΛτ+2A1Λ+B1Λ(A1Λ2+B1Λ+C1)]τ=τ=1Z{B1ω2[(B3ω2)cosωτ2Aωsinωτ]+(C1ωA1ω3)[2Aωcosωτ+(B3ω2)sinωτ]B21ω2+2A1ω(C1ωA1ω3)}=1Z[3ω6+2(A22BA21)ω4+(B2+2A1C12ACB21)ω2]=zZ(3z2+2p1z+q1),

    where Z=[b21ω2+(C1A1ω2)2]ω2>0. Since z3+p1z2+q1z+r1=0 and r1<0, then 3z2+2p1z+q1>0 if either p10 or q10. So we have

    dReΛ(τ)dτ>0,

    if either p10 or q10.

    Remark 3.1. By theorem 3.1, we can extend that if there exists p1,q1 and r1, which makes z3+p1z2+q1z+r1=0 and zZ(3z2+2p1z+q1)>0 hold. Then periodic solutions are bifurcated near the positive equilibrium E.

    We have already shown the existence of Hopf bifurcation. Next, we will give a formula by using the center manifold theorem and the normal form theory to determine the direction of Hopf bifurcation and stability of periodic solutions.

    Let v(t)=T(τt)T,x(t)=I(τt)I and y(t)=M(τt)M. Then (2.1) can be rewritten as

    (˙v(t)˙x(t)˙y(t))=τA1(v(t)x(t)y(t))+τB(v(t1)x(t1)y(t1))+F(vt,xt,yt,τ), (3.1)

    where

    A=(d1μIμT0μI0αI00d3),

    B=(0000000βMβI),

    F(vt,xt,yt,τ)=(τμv(t)x(t)τμv(t)x(t)ταx(t)y(t)τβx(t1)y(t1)τd3M).

    Let ˆτ be the critical value of τ where model (2.1) undergoes a Hopf bifurcation at E. Assume τ=ˆτ+h, then h=0 is the Hopf bifurcation value of (2.1).

    Choose the phase space C=C([1,0],R3). Define L(h):CR3 by

    L(h)ϕ=(ˆτ+h)A1ϕ(0)+(ˆτ+h)Bϕ(1),ϕC.

    From the Riesz representation theorem, there exists a matrix whose components are bounded variation functions η(θ,h):[1,0]R3,θ[1,0], such that

    L(h)ϕ=01dη(θ,h)ϕ(θ),ϕC.

    We select η(θ,h)=(ˆτ+h)Aδ(θ)(ˆτ+h)Bδ(θ+1), where

    δ(θ)={1,θ=0,0,θ0.

    For ϕC1([1,0],R3), we give

    A(h)ϕ={˙ϕ(θ),θ[1,0),01dη(t,h)ϕ(t),θ=0,

    such that

    R(h)ϕ={0,θ[1,0),F(ϕ,ˆτ+h),θ=0.

    Then (3.1) can be rewritten as

    ˙μt=A(h)μt+R(h)μt. (3.2)

    For φC1([0,1],(C3)), define

    Aφ(s)={˙φ(s),s(0,1],01φ(t)dη(t,0)=ˆτφ(0)A1+ˆτφ(1)B,s=0.

    And for ϕC([1,0],C3) with φC([0,1],(C3)), define

    φ,ϕ=¯φ(0)ϕ(0)01θ0¯φ(ξθ)dη(θ,0)ϕ(ξ)dξ,

    then A and A(0) are adjoint operators, such that φ,Aϕ=Aφ,ϕ. Let q(θ) and q(s) be the eigenvectors of A and A corresponding to iwˆτ and iwˆτ respectively. Via direct calculation, we obtain following result.

    Lemma 3.1. q(θ)=(1,α1,β1)Teiˆwˆτθ is the eigenvector of operator A on iˆwˆτ, q(s)=¯D(1,α2,β2)Teiˆwˆτs is the eigenvector of operator A on iˆwˆτ, and q,q=1,q,ˉq=0, where α1=iˆw+d1+μIμT, β1=iˆwα1μIαI, α2=iˆw+d1+μIμI, β2=iˆwα2+μTeiˆwˆτβM.

    Proof. Let q(θ)=(1,α1,β1)Teiˆwˆτθ be an eigenvector of operator A on iˆwˆτ, it follows from Aq(0)=iˆwˆτq(0), then

    ˆτA1(1α1β1)+ˆτB(1α1β1)=iˆwˆτ(1α1β1).

    The calculation shows that α1=iˆw+d1+μIμT, β1=iˆwα1μIαI. Similarly, we can obtain α2=iˆw+d1+μIμI, β2=iˆwα2+μTeiˆwˆτβM.

    q,q=¯q(0)q(0)010ξ=0¯q(ξθ)dη(θ,0)q(ξ)dξ=D(1+α1¯α2+β1¯β2)010ξ=0D(1,¯α2,¯β2)eiˆwˆτ(ξθ)dη(θ,0)(1,α1,β1)Teiˆwˆτξdξ=D(1+α1¯α2+β1¯β2)D(1,¯α2,¯β2)01dη(θ,0)θeiˆwˆτθ(1,α1,β1)T=D(1+α1¯α2+β1¯β2)+Dˆτ(α1¯β2βM+β1¯β2βI)eiˆwˆτ=D(1+α1¯α2+β1¯β2+ˆτ¯β2β(α1M+β1I)eiˆwˆτ).

    To ensure q,q=1, we only need

    D=11+α1¯α2+β1¯β2+ˆτ¯β2β(α1M+β1I)eiˆwˆτ.

    The details have been given in Hassard et al. [20] corroborated q,ˉq=0. Here we are not going to repeat it. The proof is completed.

    Following the algorithm in Hassard et al. [20], we first compute the coordinates to describe the center manifold C0 at h=0. Let μt=(vt,xt,yt)T be the solution of (3.1) when τ=ˆτ, i.e. when h=0. Define

    z(t)=q,μt,W(t,θ)=μt(θ)2Re{z(t)q(θ)},

    On the center manifold C0, it shows

    W(t,θ)=W(z(t),ˉz(t),θ),

    where

    W(z,ˉz,θ)=W20(θ)z22+W11(θ)zˉz+W02(θ)¯z22+W30(θ)z36+.

    z and ˉz represent the local coordinates for the center manifold C0 in the direction of q and ˉq. Note that W is real if ut is real, and here, we only consider real solutions. For the solution utC0 of (3.1), since h=0,

    ˙z(t)=iˆwˆτz(t)+q(s),F(W(t,)+2Re{z(t)q()})=iˆwˆτz(t)+ˉq(0)F(W(z,ˉz,0)+2Re{z(t)q(0)})iˆwˆτz+ˉq(0)F0(z,ˉz).

    We rewrite the above equation as

    ˙z(t)=iˆwˆτz(t)+g(z(t),ˉz(t)), (3.3)

    where

    g(z,ˉz)=ˉq(0)F(W(z,ˉz,0)+2Re{z(t)q(0)})=g20z22+g11zˉz+g02¯z22+g21z2ˉz2+. (3.4)

    From (3.2) and (3.3), we have

    ˙W=˙μt˙zq˙ˉzˉq={AW2Re{ˉq(0)F0q(θ)},if1θ<0AW2Re{ˉq(0)F0q(θ)}+F0,ifθ=0AW+H(z,ˉz,θ),

    where

    H(z,ˉz,θ)=H20(θ)z22+H11(θ)zˉz+H02(θ)ˉz22+. (3.5)

    Expanding the above series and comparing the coefficients, it yields

    (A2iˆwˆτ)W20(θ)=H20(θ),AW11(θ)=H11(θ),(A+2iˆwˆτ)W02(θ)=H02(θ), (3.6)

    Note that

    q(0)=D(1,α2,β2),v(t)=z+ˉz+W(1)20(0)z22+W(1)11(0)zˉz+W(1)02(0)¯z22+,x(t)=zα1+ˉz¯α1+W(2)20(0)z22+W(2)11(0)zˉz+W(2)02(0)¯z22+,y(t)=zβ1+ˉz¯β1+W(3)20(0)z22+W(3)11(0)zˉz+W(3)02(0)¯z22+,x(t1)=zα1eiˆwˆτ+ˉzeiˆwˆτα1+W(2)20(1)z22+W(2)11(1)zˉz+W(2)02(1)¯z22+,y(t1)=zβ1eiˆwˆτ+ˉz¯β1eiˆwˆτ+W(3)20(1)z22+W(3)11(1)zˉz+W(3)02(1)¯z22+.

    and

    F0=(ˆτμv(t)x(t)ˆτμv(t)x(t)ˆταx(t)y(t)ˆτβx(t1)y(t1)ˆτd3M).

    Direct substitution and comparing the coefficients with (3.4) shows that

    g202=[μα1+¯α2(μα1αα1β1)+¯β2α1β1e2iˆwˆτ]Dˆτ,g11=μ(α1+¯α1)+¯α2[μ(α1+¯α1)α(α1¯β1+¯α1β1)]+¯β2[α1¯β1+¯α1β1]Dˆτ,g022=[μ¯α1+¯α2(μ¯α1α¯α1¯β1)+¯β2(¯α1¯β1e2iˆwˆτ)]Dˆτ,g212=[μ(W(2)11(0)+12W(2)20+12W(1)20(0)¯α1+W(1)11(0)α1)+¯α2μ(W(2)11(0)+12W(2)20+12W(1)20(0)¯α1+W(1)11(0)α1)¯α2α(α1W(3)11(0)+12¯α1W(3)20(0)+12W(2)20¯β1+β1W(2)11(0))+¯β2β(α1eiˆwˆτW(3)11(1)+¯α1eiˆwˆτ12W(3)20(1)+12W(2)20(1)¯β1eiˆwˆτ+β1eiˆwˆτW(2)11(1))]Dˆτ.

    We still need to compute W20(θ) and W11(θ). For θ[1,0], it comes

    H(z,ˉz,θ)=2Re{ˉz(0)F0q(θ)}=gq(θ)ˉgˉq(θ)=(g20z22+g11zˉz+g02¯z22+)q(θ)(ˉg20ˉz22+g11zˉz+¯g02z22+)ˉq(θ).

    Comparing the coefficients with (3.5), we found that

    H20(θ)=g20q(θ)¯g02ˉq(θ),

    and

    H11(θ)=g11q(θ)¯g11ˉq(θ).

    It follows from (3.6) that

    ˙W20(θ)=2iˆwˆτW20(θ)+g20q(0)eiˆwˆτθ+¯g02ˉq(0)eiˆwˆτθ.

    Then solving the above equation, we get

    W20(θ)=e2iˆwˆτdθ((g20q(θ)+¯g02ˉq(θ))e2iˆwˆτdθdθ+E1)=ig20q(0)eiˆwˆτθˆwˆτ+i¯g02ˉq(0)eiˆwˆτθ3ˆwˆτ+E1e2iˆwˆτθ. (3.7)

    Similarly,

    W11(θ)=ig11q(0)ˆwˆτeiˆwˆτθ+i¯g11ˉq(0)ˆwˆτeiˆwˆτθ+E2, (3.8)

    where E1 and E2 are both two-dimensional vectors and can be determined by setting θ=0 in H. In fact, since

    H(z,ˉz,0)=2Re{ˉq(0)F0q(0)}+F0,

    we have

    H20(0)=g20q(0)¯g02ˉq(0)+2ˆτ(μα1μα1αα1β1α1β1e2iˆwˆτ),

    and

    H11(0)=g11q(0)¯g11ˉq(0)+ˆτ(μ(α1+¯α1)μ(α1+¯α1)α(α1^β1+¯α1β1)α1¯β1+¯α1β1).

    It follows from (3.6) and the definition of A that

    ˆτA1W20(0)+ˆτBW20(1)=2iˆwˆτW20(0)H20(0),

    and

    ˆτA1W11(0)+ˆτBW11(1)=H11(0).

    Substituting (3.7), (3.8) into the above two equations respectively yields

    E1=1ˆτ(A1+B2iˆwˆτ2iˆwI)1[H20(0)+4Re(g20q(0))iˆwg20A1ˉq(0)+i3ˆwˉg02A1ˉq(0)+iˆwg20eiˆwˆτBq(0)+i3ˆwˉg02eiˆwˆτBˉq(0)],
    E2=1ˆτ(A1+B)1(H11(0)+2ˆwA1Im(g11(0)q(0))+2ˆwBIm(g11q(0)eiˆwˆτ)),

    where I is the 3×3 identity matrix. Consequently, g21 can be expressed in terms of the parameters and delay ˆτ. Then the following values can be computed,

    C1(0)=i2ˆwˆτ(g20g112|g11|2|g02|23)+g212,μ2=Re(C1(0))Reλ(ˆτ),β2=2Re(C1(0)).

    Utilizing the results of Hassard et al. [20], we make a conclusion as follows.

    Theorem 3.2.

    (ⅰ) If μ2>0(<0), then the Hopf bifurcation is supercritical (subcritical).

    (ⅱ) If β2<0(>0), then the bifurcating periodic solutions are stable (unstable).

    In order to interpret the conclusions from a quantitative perspective, the dynamics of malaria infection in the host by numerical simulations will be analyzed in the following. In this section, we use matlab to find the numerical solutions of the model (2.1) and analyze the effect of basic reproduction number 0 and the immune response reproduction number 1 on malaria infection.

    With parameters values given in Table 1 and Table 2, we get 0=0.0025<1. Hence, Theorem 2.3 indicates that the infection-free equilibrium E0 is globally asymptotically stable. Meanwhile, this result shows that iRBCs can be eliminated by immune response in host infected with malaria, such that malaria infection can not be established within a host if 01 (see Figure 1).

    Table 1.  Variables in model (1.1).
    Symbols Variables Initial Values Reference
    T population of red blood cells (RBCs) 5×106cells/ul [4,5,10]
    I population of infection red blood cells (iRBCs) 104cells/ul [4,5,10]
    M population of immune factors (NETs) 104cells/ul [4,5,10]

     | Show Table
    DownLoad: CSV
    Table 2.  Parameters in model (1.1).
    Symbols Parameters Initial Values Reference
    λ production rate of T(t) 4.15×104cells/ul/day [4,5]
    d1 decay rate of T(t) 8.3×103/day [4,5]
    d2 decay rate of I(t) 1.0/day [4,5]
    d3 decay rate of M(t) 0.05/day [4]
    μ infection of T(t) by plasmodium 2×109ul/cell/day [4,5]
    α removal rate of I(t) by M(t) 108cells/ul/day [4,5]
    β production rate of M(t) by I(t) 2.5×105cells/ul/day [4,5]
    g 1g half saturation constant for I(t) 5×104cells/ul/day [5]

     | Show Table
    DownLoad: CSV
    Figure 1.  When 0=0.0025<1, the infection-free equilibrium E0=(5×106,0,0) is globally asymptotically stable for any τ>0.

    Taking u=3×107,β=1.5×108 and the other parameters values as in Table 1 and Table 2, we get 0=1.5 and 1121. The infection equilibrium (without specific immune response) E1 is globally asymptotically stable. It means that the generation of immune factors in the host are degraded immediately, which ultimately leads to the failure of the immune factors in the host, that is, immune factors do not remove the iRBCs in the host completely, such that malaria infection can be established in the host if 0>1 (see Figure 2).

    Figure 2.  Take u=3×107 and β=1.5×108, then 0=1.5 and 1121. The infection equilibrium (without specific immune response) E1=(3.33×106,13833,0) is globally asymptotically stable for any τ>0.

    Taking τ=0,u=3×107,β=7×106 and the other parameters values as in Table 1 and Table 2, we get 0=1.5 and 11.26. The infection equilibrium (with specific immune response) E is globally asymptotically stable. It shows that immune factors play roles in the host, but the number of immune factors and iRBCs will eventually tend to a dynamic balance. Immune factors can not completely eliminate the iRBCs in the host, such that malaria infection can be established in the host if τ=0 and 01 (see Figure 3).

    Figure 3.  Take τ=0,u=3×107 and β=7×106, then 0=1.5 and 11.26. The infection equilibrium (with specific immune response) E=(3.974×106,7143,3.44×106) is globally asymptotically stable.

    When all parameters values are as in Figure 3 with τ=0.1, we can find the solution curve is the same as Figure 3. It means that there is no great change in the stability of the equilibrium E when the time delay τ is small. Immune factors can not completely eliminate the iRBCs in the host, so malaria infection can be established in the host if τ=0.1 and 0>1 (see Figure 4).

    Figure 4.  Take τ=0.1,u=3×107 and β=7×106, then 0=1.5 and 11.26. The infection equilibrium (with specific immune response) E=(3.974×106,7143,3.44×106) is locally asymptotically stable.

    When τ increases and passes through τ0.238. For example, taking τ=2, we get 0=1.5,11.26 and E becomes unstable. Theorem 3.1 implies that model (2.1) undergoes Hopf bifurcation and a periodic solution appears. It means that in the previous period of relative stability, the number of iRBCs is very small. Meanwhile, the number of immune factors is small as well, the iRBCs in the host has not been cleared completely. We consider that the incubation period for malaria leads to the number of iRBCs rising again after a period of time. Maybe that is why malaria might rekindle in the host (see Figure 5).

    Figure 5.  Take τ=2,u=3×107 and β=7×106, then 0=1.5 and 11.26. The infection equilibrium (with specific immune response) E=(3.974×106,7143,3.44×106) is unstable and there is a periodic solution bifurcated from E.

    In this paper, we have discussed a malaria infection model with time delay.

    For the model (2.1), immune response delay τ plays a important role in the stability of the equilibrium E. We derive the basic reproduction number 0 for the malaria infection and establish that the dynamics are completely determined by the basic reproduction number 0. The results show that when 01, the infection-free equilibrium E0 is globally asymptotically stable for any delay τ. It signifies that the malaria is cleared and immune response is not active. When 1<01, the infection equilibrium E1 without immune response exists and it is globally asymptotically stable for any delay τ, which means that the immune response would not be activated and malaria infection can be established in the host. When 0>1, some results are given as below:

    (ⅰ) If τ=0, the infection equilibrium with immune response E is globally asymptotically stable.

    (ⅱ) If τ<τ, the infection equilibrium with immune response E is locally asymptotically stable.

    (ⅲ) If model (2.1) satisfies theorem 3.1 or remark 3.1, the dynamical behaviors of equilibrium E will occur and locally asymptotically stable becomes unstable and Hopf bifurcation appears. By choosing immune response delay as bifurcation parameter, we have demonstrated that a limit cycle occurs via Hopf bifurcation, when the delay passes through the critical value τ. This explains the fact that the immune response delay plays negative role in controlling disease progression. The direction and stability of Hopf bifurcation is derived by applying the center manifold method and the normal form theory.

    Moreover, we study the uniform persistence of infection equilibrium E.

    Numerical simulations are also provided to demonstrate these theoretical results. Finally, we hope that our work will be helpful to the study of malaria infection.

    The authors wish to thank two anonymous reviewers for their very helpful comments. QD, JL and ZG were supported by National Science Foundation of China (No. 11771104), Program for Chang Jiang Scholars and Innovative Research Team in University (IRT-16R16).

    All authors declare no conflicts of interest in this paper.



    [1] WHO, "Malaria", 2008. Available from: http://www.who.int/malaria/en.
    [2] F. McKenzie and H. Bossert, An integrated model of Plasmodium falciparum dynamics, J. Theoret. Biol., 232 (2005), 411–426.
    [3] P. Streatfield, W. Khan and A. Bhuiya, Malaria mortality in Africa and Asia: evidence from INDEPTH health and demographic surveillance model sites, Global Health Action, 7 (2014), 25369–25369.
    [4] C. Hetzel and R. Anderson, The within-host cellular dynamics of bloodstage malaria: theoretical and experimental studies, Parasitology, 113 (1996), 25–38.
    [5] Y. Li, S. Ruan and D. Xiao, The within-host dynamics of malaria infection with immune response, Math. Biosci. Eng., 8 (2011), 999–1018.
    [6] Y. Xiao and X. Zou, Can Multiple Malaria Species Co-persist?, SIAM J. Appl. Math., 73 (2013), 351–373.
    [7] Z. Chang, J. Ning and Y. Zhang, The TatD-like DNase of Plasmodium is a virulence factor and a potential malaria vaccine candidate, Nat. Commun., 7 (2016), 11537–11546.
    [8] L. Molineaux and K. Dietz, Review of intra-host models of malaria, Parassitologia, 41 (1999), 221–231.
    [9] R. Anderson, R. May and S. Gupta, Non-linear phenomena in host-parasite interactions, Parasitology, 99 (1989), S59–S79.
    [10] C. Chiyaka, W. Garira and S. Dube, Modelling immune response and drug therapy in human malaria infection, Comput. Math. Methods Med., 9 (2008), 143–163.
    [11] G. Johnston, P. Gething and S. Hay, Modeling within-host effects of drugs on Plasmodium falciparum transmission and prospects for malaria elimination, PLoS Comput. Biol., 10 (2014), e1003434.
    [12] K. Landman, K. Tan and P. Arguin, Adherence to malaria prophylaxis among Peace Corps Volunteers in the Africa region, Travel. Med. Infect. Dis., 13 (2015), 61–68.
    [13] J. Liu and Z. Guo, The dynamic behavior of mathematical model of pathogenesis of plasmodium falciparum, J. Guangzhou Univ., 17 (2018), 13–22.
    [14] Y. Kuang, Delay differential equations with applications in population dynamics, Academic Press, Boston, 1993.
    [15] H. L. Smith, An introduction to delay differential equations with applications to the life sciences, Springer, New York, 2011.
    [16] J. Hale, Oscillations in neutral functional differential equations, C.i.m.e.summer Schools, 1972, 97–111.
    [17] Z. Wang, Z. Guo and H. Peng, A mathematical model verifying potent oncolytic efficacy of M1 virus, Math. Biosci., 276 (2016), 19–27.
    [18] S. Ruan and W. Wang, Dynamical behavior of an epidemic model with a nonlinear incidence rate, J. Differential Equations, 188 (2003), 135–163.
    [19] D. Xiao and Y. Yang, Influence of latent period and nonlinear incidence rate on the dynamics of SIRS epidemiological models, Discrete Contin. Dyn. Syst. Ser. B, 13 (2012), 195-211.
    [20] B. D. Hassard, N. D. Kazarinoff and Y. Wan, Theory and applications of Hopf bifurcation, Cambridge University Press, Cambridge, 1981.
  • This article has been cited by:

    1. A. M. Elaiw, A. D. Al Agha, Global analysis of a reaction–diffusion blood-stage malaria model with immune response, 2020, 13, 1793-5245, 2050029, 10.1142/S1793524520500291
    2. Yanyuan Xing, Zhiming Guo, Jian Liu, Backward bifurcation in a malaria transmission model, 2020, 14, 1751-3758, 368, 10.1080/17513758.2020.1771443
    3. Song-bai Guo, Min He, Jing-an Cui, Global Stability of a Time-delayed Malaria Model with Standard Incidence Rate, 2023, 0168-9673, 10.1007/s10255-023-1042-y
    4. Jian Liu, Zhiming Guo, Hongpeng Guo, The blood-stage dynamics of malaria infection with immune response, 2022, 16, 1751-3758, 294, 10.1080/17513758.2021.2017033
    5. Songbai Guo, Xin Yang, Zuohuan Zheng, Global dynamics of a time-delayed malaria model with asymptomatic infections and standard incidence rate, 2023, 31, 2688-1594, 3534, 10.3934/era.2023179
  • Reader Comments
  • © 2019 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(5074) PDF downloads(669) Cited by(5)

Figures and Tables

Figures(5)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog