Research article Special Issues

Within-host delay differential model for SARS-CoV-2 kinetics with saturated antiviral responses

  • The present study discussed a model to describe the SARS-CoV-2 viral kinetics in the presence of saturated antiviral responses. A discrete-time delay was introduced due to the time required for uninfected epithelial cells to activate a suitable antiviral response by generating immune cytokines and chemokines. We examined the system's stability at each equilibrium point. A threshold value was obtained for which the system switched from stability to instability via a Hopf bifurcation. The length of the time delay has been computed, for which the system has preserved its stability. Numerical results show that the system was stable for the faster antiviral responses of epithelial cells to the virus concentration, i.e., quick antiviral responses stabilized patients' bodies by neutralizing the virus. However, if the antiviral response of epithelial cells to the virus increased, the system became unstable, and the virus occupied the whole body, which caused patients' deaths.

    Citation: Kaushik Dehingia, Anusmita Das, Evren Hincal, Kamyar Hosseini, Sayed M. El Din. Within-host delay differential model for SARS-CoV-2 kinetics with saturated antiviral responses[J]. Mathematical Biosciences and Engineering, 2023, 20(11): 20025-20049. doi: 10.3934/mbe.2023887

    Related Papers:

    [1] A. D. Al Agha, A. M. Elaiw . Global dynamics of SARS-CoV-2/malaria model with antibody immune response. Mathematical Biosciences and Engineering, 2022, 19(8): 8380-8410. doi: 10.3934/mbe.2022390
    [2] A. M. Elaiw, Raghad S. Alsulami, A. D. Hobiny . Global dynamics of IAV/SARS-CoV-2 coinfection model with eclipse phase and antibody immunity. Mathematical Biosciences and Engineering, 2023, 20(2): 3873-3917. doi: 10.3934/mbe.2023182
    [3] Chentong Li, Jinhu Xu, Jiawei Liu, Yicang Zhou . The within-host viral kinetics of SARS-CoV-2. Mathematical Biosciences and Engineering, 2020, 17(4): 2853-2861. doi: 10.3934/mbe.2020159
    [4] 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
    [5] Wenhan Guo, Yixin Xie, Alan E Lopez-Hernandez, Shengjie Sun, Lin Li . Electrostatic features for nucleocapsid proteins of SARS-CoV and SARS-CoV-2. Mathematical Biosciences and Engineering, 2021, 18(3): 2372-2383. doi: 10.3934/mbe.2021120
    [6] Yangyang Yu, Yuan Liu, Shi Zhao, Daihai He . A simple model to estimate the transmissibility of the Beta, Delta, and Omicron variants of SARS-COV-2 in South Africa. Mathematical Biosciences and Engineering, 2022, 19(10): 10361-10373. doi: 10.3934/mbe.2022485
    [7] Sumati Kumari Panda, Abdon Atangana, Juan J. Nieto . New insights on novel coronavirus 2019-nCoV/SARS-CoV-2 modelling in the aspect of fractional derivatives and fixed points. Mathematical Biosciences and Engineering, 2021, 18(6): 8683-8726. doi: 10.3934/mbe.2021430
    [8] Sarafa A. Iyaniwura, Rabiu Musa, Jude D. Kong . A generalized distributed delay model of COVID-19: An endemic model with immunity waning. Mathematical Biosciences and Engineering, 2023, 20(3): 5379-5412. doi: 10.3934/mbe.2023249
    [9] Tahir Khan, Roman Ullah, Gul Zaman, Jehad Alzabut . A mathematical model for the dynamics of SARS-CoV-2 virus using the Caputo-Fabrizio operator. Mathematical Biosciences and Engineering, 2021, 18(5): 6095-6116. doi: 10.3934/mbe.2021305
    [10] Matthew Hayden, Bryce Morrow, Wesley Yang, Jin Wang . Quantifying the role of airborne transmission in the spread of COVID-19. Mathematical Biosciences and Engineering, 2023, 20(1): 587-612. doi: 10.3934/mbe.2023027
  • The present study discussed a model to describe the SARS-CoV-2 viral kinetics in the presence of saturated antiviral responses. A discrete-time delay was introduced due to the time required for uninfected epithelial cells to activate a suitable antiviral response by generating immune cytokines and chemokines. We examined the system's stability at each equilibrium point. A threshold value was obtained for which the system switched from stability to instability via a Hopf bifurcation. The length of the time delay has been computed, for which the system has preserved its stability. Numerical results show that the system was stable for the faster antiviral responses of epithelial cells to the virus concentration, i.e., quick antiviral responses stabilized patients' bodies by neutralizing the virus. However, if the antiviral response of epithelial cells to the virus increased, the system became unstable, and the virus occupied the whole body, which caused patients' deaths.



    Since January 2020, a deadly pandemic known as COVID-19 has spread from person to person, wreaking havoc across the globe. The virus responsible for the illness is SARS-CoV-2, also known as the 2019-ncov [1]. Multiple cases of pneumonia, dry cough, fever, fatigue, breathing difficulty and bilateral lung infiltration have been reported due to the virus's spread within the human body [2,3]. According to the World Health Organization (WHO), as of March 18, 2023 there have been 760 million cases reported worldwide, with 6.8 million fatalities. Vaccines like Pfizer-BioNTech, Moderna (mRNA-1273), and COVAXIN are available for hospitality and to control the pandemic.

    Mathematical models give more insight and knowledge into the dynamics of infectious diseases, which may help control the pandemic. Several researchers [4,5,6,7,8,9] have investigated the human-to-human transmission dynamics of the COVID-19 pandemic using mathematical modeling. Mathematical models incorporating time delays are also used for understanding infectious disease dynamics because they are more realistic and accurate due to the memory effect. Yang et al. [10] proposed a COVID-19 model with a time delay that characterized the viral infection cycle and treatment time. They fitted their model with the data from Beijing and Wuhan and suggested that early detection and isolation are the most important ways to prevent the spread of the epidemic. In [11], the effect of the time delay of the immune response on COVID-19 transmission dynamics has been investigated. A mathematical model has been designed in [12] to explore the importance of time lag in precautionary measures to control the COVID-19 pandemic. Babasola et al. [13] explored the role of time delay with a convex incidence rate in the COVID-19 epidemic. They observed that time delay destabilizes the system and generates periodic solutions. The effects of time delay in using vaccinations for COVID-19 spread dynamics has been investigated in [14].

    Moreover, controlling the COVID-19 pandemic also requires a thorough understanding of the within-host dynamics of the SARS-CoV-2 virus. For this purpose, numerous modeling studies have been performed to predict the behavior of the SARS-CoV-2 virus at the cellular level. Both viral and host factors play a significant role in SARS-CoV-2 infections. Host factors during infection trigger an immune response that combats the virus. The host's innate immune system can identify viral infections by utilizing pattern recognition receptors to identify pathogen-associated molecular patterns. As the innate immune system detects the viruses, it typically triggers the pulmonary and systemic inflammatory reactions linked to the SARS-CoV-2 virus. According to a modeling study [15], the innate immune response may be able to clear the virus more effectively if the adaptive immune response is temporarily suppressed. The kinetics of the human SARS-CoV-2 infection have been mathematically modeled in [16]. Prakash et al. [17] proposed a multi-scale model that incorporates both the intra-host and inter-host dynamics of COVID-19. Their findings suggest that treatment with antiviral drugs, immunotherapies, and improved sanitation and social isolation was proposed as the most effective method for lowering the virus's basic reproduction number and environmental spread in a population. Chhetri et al. [18] studied the role of various drugs at various stages of COVID-19 pathogenesis using a mathematical model of the complex interaction of virus replication and the host immune response. The effector T cell response to SARS-CoV-2 and the dynamics of the virus were studied in [19]. The dynamics of SARS-CoV-2 in infected patients were simulated in [20]. Chowdhury et al. [21] analyzed a mathematical model to investigate the impact of natural killer cells and T cells on SARS-CoV-2 kinetics. Li et al. [22] proposed a within-host SARS-CoV-2 model to observe drug effects on virus growth and the immunity effect of patients. In [23], dynamical analysis of the model [22] was studied. Ghosh et al. [24] observed the within-host dynamics of SARS-CoV-2 in the presence of innate and adaptive immune responses and then observed the effect of vaccination and antiviral treatments. Stochastic dynamics of the within-host COVID-19 epidemic with discrete time delay and noise have been studied in [25] and showed the impact of delay tactics and noise on the extinction of the disease. Staroverov et al. [26] proposed a novel mathematical model for SARS-CoV-2 dynamics that explicitly modeled intracellular events like the exhaustion of cellular resources needed for virus production and innate immune responses. Pillis et al. [27] showed a mathematical model of how SARS-CoV-2 spreads within a host and how neutralizing antibodies react to vaccination.

    It is found in the above literature that during the infection by the virus, the epithelial cells resist the viral infection by activating the immune system. Thus, considering this fact, we modified the model discussed in [22,23]. So, in this study, we consider a model incorporating the antiviral responses of uninfected epithelial cells in terms of Michaelis-Menten, which is suitably described as the "saturated responses". In addition, it takes some time for uninfected epithelial cells to develop a suitable antiviral response after contracting a virus. Therefore, we introduce a time delay into the saturated response term. The rest of the manuscript is assembled as follows: In Section 2, we have formulated our model. The basic characteristics of the solution of the model have been discussed in Section 3. The stability of the model and estimation of the time lag for preserving a stable limit cycle have been investigated in Section 4. The numerical simulation has been performed in Section 5. A concluding remark has been made in Section 6.

    This section will formulate a model that describes the within-host SARS-CoV-2 viral kinetics. The considered model was originally proposed by Li et al. [22], and the model is described by the following equations:

    dE(t)dt=d1(E(0)E(t))βE(t)v(t),dI(t)dt=βE(t)v(t)d2I(t),dv(t)dt=μI(t)d3v(t). (2.1)

    The numbers of pulmonary epithelial cells that are uninfected, infected and the virus concentration are represented here as E(t), I(t), and v(t), respectively. The rate at which virus-free epithelial cells are infected by the virus is indicated by the symbol β. The virus-free pulmonary epithelial cells are produced at a rate of d1E(0) and die at a rate of d1. The term d2 represents the death rate of virus-infected epithelial cells. Also, μ and d3 represent the production rate and death rate of viruses, respectively. The epithelial cells are infected once the virus comes into contact with a human being. In addition, epithelial cells can fight off the virus during infection by triggering the immune system to control antiviral responses. Since uninfected epithelial cells resist viral infection by secreting immune cytokines and chemokines, we modified the model (2.1) to describe this behavior by the nonlinear term βE(t)v(t)1+v(t) [28,29,30]. In addition, uninfected epithelial cells need time to develop a proper antiviral response after infection by the virus. Therefore, there is a time lag to regulate the antiviral response and this time delay can be introduced as a discrete-time delay, τ into the term, βE(t)v(t)1+v(t). Thus, the considered model becomes,

    dE(t)dt=d1(E(0)E(t))βE(tτ)v(tτ)1+v(tτ),dI(t)dt=βE(tτ)v(tτ)1+v(tτ)d2I(t),dv(t)dt=μI(t)d3v(t), (2.2)

    subject to the initial conditions:

    E(θ)=ϕ1(θ),I(θ)=ϕ2(θ),v(θ)=ϕ3(θ),ϕ1(θ)0,ϕ2(θ)0,ϕ3(θ)0,θ[τ,0],also ϕ1(0)>0,ϕ2(0)>0,ϕ3(0)>0,

    where

    ζ+={(ϕ1(θ),ϕ2(θ),ϕ3(θ))S([θ,0]R3+)}, (2.3)

    where the continuous functions ϕi are defined on the interval θ[τ,0].

    In this section, we will discuss the nonnegativity and boundedness of solutions for the considered model.

    Proposition 3.1. Corresponding to initial conditions (2.3), all solutions of the system (2.2) are in R3+ and bounded for all t>0.

    Proof. The epithelial cells are subdivided into uninfected and infected epithelial cells. Thus the first two equations of the system (2.2) give us

    ddt{E(t)+I(t)}=d1(E(0)E(t))d2I(t),d1(E(0)E(t))d2I(t),=d1E(0)δ(E(t)+I(t)),

    where δ=min{d1,d2}. Thus, an upper bound always exists for uninfected and infected epithelial cells. It is easy to see from the model's third equation that the free virus population v is also bounded above.

    Now, the system (2.2) can be re-written in vector notation as

    ˙X=Γ(X), (3.1)

    with X=[E(t),I(t),v(t)]TR3+ and

    Γ(X)=(Γ1(X)Γ2(X)Γ3(X))=(d1(E(0)E(t))βE(tτ)v(tτ)1+v(tτ)βE(tτ)v(tτ)1+v(tτ)d2I(t)μI(t)d3v(t)),

    where, ΓC(R3+) and Γ:R3+R3+. The righthand side of system (2.2) is locally Lipchitz, thus the derivatives are bounded and satisfy

    Γi(X)|Xi(t)=0,XR3+=Γi(0);for alli=1,2,3.

    Also, using the second lemma of [31], the solution of system (2.2) remains positive throughout the domain R3+, t>0 corresponding to initial conditions (2.3). Thus, the solutions of model (2.2) are nonnegative and bounded for time t>0. This completes the proof.

    The basic reproduction number in virus dynamics determines whether the disease will persist or die. It also controls disease spread in the population. We will use the method described in Driessche and Watmough [32] to calculate the basic reproduction number of model (2.2).

    The two infected compartments in the model (2.2) are I and v. If Fi and Vi denote the appearance rate of new infections in the ith compartment and the transfer rate of individuals from the ith compartment, respectively, then we have two matrices F and V as follows:

    F=(βE(t)v(t)1+v(t)0),V=(d2I(t)d3v(t)μI(t)). (4.1)

    The Jacobian matrices of F and V at the virus-free equilibrium point, denoted by F and V respectively, are evaluated as

    F=(0βE(0)00),V=(d20μd3). (4.2)

    Now, we compute the next generation matrix FV1 as follows

    FV1=1d2d3(0βE(0)00)(d30μd2)=1d2d3(μβE(0)βd2E(0)00).

    Following Driessche and Watmough [32], the basic reproduction number χ0 of the model (2.2) is the spectral radius of the matrix FV1 which is given by χ0=μβE(0)d2d3.

    In this section, we will investigate the existence of the equilibrium points for the model (2.2) and study their local stability. As the time delay τ neither affects the number nor the type of equilibrium points; the existing equilibrium points of the model (2.2) are

    P(E(0),0,0), uninfected or virus-free equilibrium point, which always exists.

    P(E,I,v), coexisting equilibrium point; where,

    E=d2I(1+v)βv=E(0)χ0(1+(χ01)d1β+d1),I=d3vμ=d3d1μ(β+d1)(χ01),andv=d1(β+d1)(χ01).

    Clearly, the coexisting equilibrium P(E,I,v) exists if χ0>1.

    For local stability analysis, we consider two cases:

    Case 1: τ=0

    (i) The Jacobian matrix of the system (2.2) at uninfected or virus-free equilibrium point P(E(0),0,0) is

    J0P=(d10βE(0)0d2βE(0)0μd3).

    The characteristic equation of the matrix J0P is

    λ3+A11λ2+A12λ+A13=0, (4.3)

    where

    A11=d1+d2+d3,A12=d1(d2+d3)+d2d3(1χ0),A13=d1d2d3(1χ0).

    The Routh-Hurwitz stability criterion states the equilibrium point P(E(0),0,0) is locally asymptotically stable if A11>0, A12>0, A13>0 and A11A12A13>0. These inequalities imply that P(E(0),0,0) is locally asymptotically stable if χ0<1μ<d2d3βE(0); otherwise, it is unstable. Thus, if the production rate of the virus is less than the ratio between the product of the decay rate of infected epithelial cells and virus concentration and the product of the infection rate of the virus and density of uninfected epithelial cells, then the virus-free equilibrium point P(E(0),0,0) is locally asymptotically stable.

    (ii) Corresponding to the co-existing equilibrium point P(E,I,v), the Jacobian matrix of system (2.2) is

    J0P=(d1βv1+v0βE(1+v)2βv1+vd2βE(1+v)20μd3).

    The characteristic equation of the matrix J0P is

    λ3+B11λ2+B12λ+B13=0, (4.4)

    where

    B11=d1+d2+d3+βv1+v,B12=d2d3μβE(1+v)2+(d2+d3)(d1+βv1+v)=d2d3(v1+v)+(d2+d3)(d1+βv1+v)[substitutingE=d2I(1+v)βvandIv=d3μ],B13=(d1+βv1+v)(d2d3μβE(1+v)2)+βE(1+v)2μβv1+v=(d1+β)d2d3v1+v[substitutingE=d2I(1+v)βvandIv=d3μ].

    According to the Routh-Hurwitz stability criteria, whenever χ0>1, the equilibrium point P(E,I,v) is locally asymptotically stable if B11>0, B12>0, B13>0 and B11B12B13>0; otherwise it is unstable. However, clearly B11>0, B12>0, B13>0 and B11B12B13>0 if χ0>1. Thus, for the stability of the coexisting equilibrium point P(E,I,v), the system (2.2) must have the production rate of the virus greater than the ratio between the product of the decay rate of infected epithelial cells and virus concentration and the product of the infection rate of the virus and density of uninfected epithelial cells.

    Case 2: τ0

    (i) The Jacobian matrix of system (2.2) at uninfected or virus-free equilibrium point P(E(0),0,0) is

    JP=(d10βE(0)eλτ0d2βE(0)eλτ0μd3).

    The one eigenvalue of the matrix JP is λP,1=d1<0, and the other two eigenvalues will be calculated from the following quadratic equation:

    λ2+(d2+d3)λ+(d2d3μβE(0)eλτ)=0. (4.5)

    Let τ>0. Since Eq (4.5) is transcendental with an infinite number of solutions, we assume that λ=ιω, and without loss of generality we may assume that ω=0 is a root of (4.5). Therefore, putting λ=ιω in (4.5), we get

    ω2+(d2+d3)ιω+d2d3μβE(0)(cosωτιsinωτ)=0. (4.6)

    Separating the real and imaginary parts of Eq (4.6), we have the following system:

    {ω2+d2d3=μβE(0)cos(ωτ)(d2+d3)ω=μβE(0)sin(ωτ). (4.7)

    Squaring both the equations of (4.7) and then adding them, we obtain

    ω4+(d22+d23)ω2+d22d23[μβE(0)]2=0, (4.8)

    Since ω0, Eq (4.8) is equivalent to

    ω2=12(d22+d23)+12(d22d23)2+4[μβE(0)]2. (4.9)

    If χ0<1, then

    d22d23[μβE(0)]2>0,

    and

    (d22+d23)24[d22d23[μβE(0)]2]<(d22+d23)2,

    so that

    (d22d23)2+4[μβE(0)]2<d22+d23.

    Therefore, we have ω2<0, which is a contradiction. Hence, whenever χ0<1, according to Rouché's theorem [33,34], Eq (4.5) cannot have pure imaginary roots and the virus-free equilibrium P(E(0),0,0) is locally asymptotically stable, for any strictly positive time-delay τ.

    Suppose χ0>1. We already know that the one eigenvalue of the matrix JP is λP,1=d1<0, and the other two eigenvalues are the zeroes of the following expression:

    Γ(λ):=λ2+(d2+d3)λ+(d2d3μβE(0)eλτ). (4.10)

    As λP,1=d1<0, for stability, we need to check whether the other two eigenvalues have negative real parts or not. However, Γ(0)=d2d3μβE(0)<0 for χ0>1 and also limλ+Γ(λ)=+. Therefore, by continuity of Γ(λ), there is at least one positive zero for the expression (4.10). Hence, we conclude that the virus-free equilibrium P(E(0),0,0) is unstable when χ0>1 i.e., also for the case of τ0, the virus-free equilibrium P(E(0),0,0) is stable if χ0<1μ<d2d3βE(0).

    (ii) At coexisting equilibrium point P(E,I,v), the Jacobian matrix of the system (2.2) is

    JP=(d1βveλτ1+v0βEeλτ(1+v)2βveλτ1+vd2βEeλτ(1+v)20μd3).

    The characteristics equation of the matrix JP is

    Γ(λ,τ)=λ3+(d1+d2+d3+βveλτ1+v)λ2+[d2d3μβEeλτ(1+v)2+(d2+d3)(d1+βveλτ1+v)]λ+(d1+βveλτ1+v)(d2d3μβEeλτ(1+v)2)+βEeλτ(1+v)2βμveλτ1+v=0λ3+(d1+d2+d3)λ2+(d2d3+d1d2+d1d3)λ+d1d2d3+eλτ[(βv1+v)λ2(βμE(1+v)2(d2+d3)βv1+v)λ+(d2d3βv1+vd1βμE(1+v)2)]=0λ3+P11λ2+P12λ+P13+eλτ(Q11λ2+Q12λ+Q13)=0, (4.11)

    where

    P11=d1+d2+d3,P12=d2d3+d1d2+d1d3,P13=d1d2d3,Q11=βv1+v,Q12=βμE(1+v)2+(d2+d3)βv1+v,Q13=d2d3βv1+vd1βμE(1+v)2.

    Since Eq (4.11) is a transcendental equation with an infinite number of solutions, we cannot apply the classical Routh-Hurwitz criterion to Eq (4.11). To determine the stability of P, we consider λ=±ιΩ (where, Ω>0), then Eq (4.11) can be written as

    ιΩ3P11Ω2+P12(ιΩ)+P13+(cosΩτιsinΩτ)(Q11Ω2+Q12ιΩ+Q13)=0.

    Separating real and imaginary parts, we get

    {P11Ω2P13=cosΩτ(Q11Ω2+Q13)+ΩQ12sinΩτΩ3ΩP12=ΩQ12cosΩτsinΩτ(Q11Ω2+Q13). (4.12)

    Squaring both the equations of (4.12) and then adding them, we obtain,

    Ω6+(P2112P12Q211)Ω4+(P2122P11P13+2Q11Q13Q212)Ω2+P213Q213=0.

    For simplification, we can rewrite it as

    Ω6+p1Ω4+p2Ω2+p3=0. (4.13)

    Here,

    p1=P2112P12Q211=(d1+d2+d3)22(d2d3+d1d2+d1d3)(βv1+v)2,p2=P2122P11P13+2Q11Q13Q212=(d2d3+d1d2+d1d3)22d1d2d3(d1+d2+d3)+2(βv1+v)(d2d3βv1+vd1βμE(1+v)2)((d2+d3)βv1+vβμE(1+v)2)2,andp3=P213Q213=(d1d2d3)2(d2d3βv1+vd1βμE(1+v)2)2.

    It can be verified that P2112P12Q211=28.6852>0, and P213Q213=0.010076<0 for the parameter values prescribed in Table (1). Hence, there exists at least one nonnegative real root Ω0 for Eq (4.13). Therefore, we found a purely imaginary root ±Ω0ι for Eq (4.11). So, the stability of system (2.2) may switch at P as τ changes.

    Table 1.  Explanation of the parameters, their units and values.
    Name Definition Range Source
    β virus infection rate (0.44,0.66) [22]
    E(0) number of uninfected epithelial cells (20,30) [22]
    μ production rate of virus (day)1 (0.192,0.288) [22]
    d1 death rate of uninfected pulmonary epithelial cells (day)1 101 assumed
    d2 death rate of infected pulmonary epithelial cells (day)1 (0.088,0.132) [22]
    d3 decay rate of virus concentration (day)1 (4.288,6.432) [22]

     | Show Table
    DownLoad: CSV

    We eliminate sin(Ωτ) from both the equations of (4.12) and obtain

    cosΩτ=(P13P11Ω2)(Q11Ω2Q13)+Ω2Q12(Ω2P12)Ω2Q212+(Q11Ω2Q13)2.

    Then, τn corresponding to Ω0 is given by

    τn=2nπΩ0+1Ω0arccos[(P13P11Ω20)(Q11Ω20Q13)+Ω20Q12(Ω20P12)Ω20Q212+(Q11Ω20Q13)2],wherenis an integer.

    For τ=0, P(E,I,v) is stable. Hence, P will remain stable for τ<τ0, where τ0=τ0 and n=0 [35].

    Now, we will verify the transversality condition

    [dRe(λ(τ))dτ]1τ=τ00,

    for the occurrence Hopf bifurcation.

    As Eq (4.11) has purely imaginary roots λ0=ιΩ0 and ˉλ0=ιΩ0 at τ0

    Γ(λ0,τ0)λ=3λ20+2P11λ0+P12+eτ0λ0[2Q11λ0+Q12τ0(Q11λ20+Q12λ0+Q13)]=3Ω20+2ιP11Ω0+P12+eιτ0Ω0[2ιQ11Ω0+Q12τ0(Q11Ω20+ιQ12Ω0+Q13)].

    So, the implicit function theorem states that in a neighborhood of τ0, there must exist a complex function λ=λ(τ) such that λ0(τ0)=λ0 and Γ(λ(τ),τ)=0, and dλdτ=Γ(λ(τ),τ)/τΓ(λ(τ),τ)/Φ for τ in a neighborhood of τ0. Thus,

    dλdτ=λeτλ(Q11λ2+Q12λ+Q13)3λ2+2P11λ+P12+eτλ[2Q11λ+Q12τ(Q11λ2+Q12λ+Q13)]. (4.14)

    By using Eqs (4.11) and (4.12) in (4.14), we obtain the following expression,

    [dλdτ]1=(3λ2+2P11λ+P12)λ(λ3+P11λ2+P12λ+P13)+2Q11λ+Q12λ(Q11λ2+Q12λ+Q13)τλ. (4.15)

    Let λ(τ)=Re(λ(τ))+ιIm(λ(τ)).

    Then, from Eq (4.15), we have

    [dRe(λ(τ))dτ]1τ=τ0=[3Ω40+2Ω20(P2112P12)+P2122P11P13(P13P11Ω20)2+(Ω30P12Ω0)22Ω20Q211+Q2122Q11Q13(Ω20Q11Q13)2+Ω20Q212].

    By using the prescribed parameter values in Table (1), we get [dRe(λ(τ))dτ]1τ=τ0=27.255>0, which verifies the transversality condition for the Hopf bifurcation.

    Therefore, at τ=τ0 system (2.2) undergoes a Hopf bifurcation around the point P. It implies that an isolated periodic orbit emerges in the neighborhood P. Thus, P is locally asymptotically stable for τ=0. Moreover, if there exists a threshold value of τ (say τ0) then P is locally asymptotically stable for 0<τ<τ0 and unstable for τ>τ0. Also, a Hopf bifurcation occurs around the point P when τ=τ0.

    Now, we will investigate the dynamics of appearing periodic solutions and compute the time lag to preserve the limit cycles. In order to do this, we consider model (2.2) that satisfies the initial conditions (2.3) on the interval [τ,0) and also the space of all continuous real-valued functions defined on [τ,+). Linearizing model (2.2) around P(E,I,v), we get

    {dEdt=d1EβvE(tτ)1+vβEv(tτ)(1+v(tτ))2,dIdt=βvE(tτ)1+vβEv(tτ)(1+v(tτ))2d2I,dvdt=μId3v. (4.16)

    By using Laplace transformation of (4.16), leading to

    {ηLE(η)ˉE(0)=d1LE(η)βv1+veητLE(η)βEeητ(1+v)2Lv(η)βveητ1+vKE(η)βE(1+v)2eητKv(η),ηLI(η)ˉI(0)=βv1++veητLE(η)+βv1+veητKE(η)βE(1+v)2eητLv(η)βE(1+v)2eητKv(η)d2LI(η),ηLv(η)ˉv(0)=μLI(η)d3Lv(η), (4.17)

    with

    KE(η)=0τeηtE(t)dt,Kv(η)=0τeηtv(t)dt,

    where LE(η), LI(η), and Lv(η) are the respective Laplace transformations of E(t), I(t), and v(t). According to the well-known theory described in [35,36] and using classical Nyquist criteria stated in [37], the coexistence equilibria P is asymptotically stable, for

    ReΓ(ιξ0)=0, (4.18)
    ImΓ(ιξ0)>0, (4.19)

    with

    Γ(η)=η3+P11η2+P12η+P13+eητ(Q11η2+Q12η+Q13),

    and ξ0>0 is the smallest root of the expressions (4.18) and (4.19).

    We can rewrite the expressions (4.18) and (4.19) as

    P11ξ20+P13=Q12ξ0sin(ξ0τ)(Q13Q11ξ20)cos(ξ0τ), (4.20)
    ξ30+P12ξ0>(Q13Q11ξ20)sin(ξ0τ)Q12ξ0cos(ξ0τ), (4.21)

    which gives sufficient conditions for the stability of coexisting equilibrium P.

    To calculate time lag τ, we will determine an upper bound ξ+ on ξ0 that is independent from τ. To do this, we assume that at ξ=ξ0, values of ξ satisfy Eq (4.21) such that 0ξξ+.

    Expression (4.20) gives

    P11ξ20=P13+Q12ξ0sin(ξ0τ)+Q13cos(ξ0τ)Q11ξ20cos(ξ0τ). (4.22)

    Now, we maximize the right side of (4.22) as

    P13+Q12ξ0sin(ξ0τ)+Q13cos(ξ0τ)Q11ξ20cos(ξ0τ),

    subject to the conditions,

    |cos(ξ0τ)|1,|sin(ξ0τ)|1.

    Therefore, we obtain that

    |P11|ξ20|P13|+|Q12|ξ0+|Q13|+|Q11|ξ20.

    Hence, it can be expressed as

    ξ+12(|P11||Q11|)[|Q12|+Q212+4(|P11||Q11|)(|P13|+|Q13|)]. (4.23)

    From (4.23), it can be verified that ξ0ξ+.

    Also, (4.21), gives

    ξ20<P12+Q12cos(ξ0τ)+Q11ξ0sin(ξ0τ)Q13sin(ξ0τ)ξ0. (4.24)

    Thus, if τ=0, then (4.21) becomes ξ20<P12+Q12, and from (4.22), P11ξ20=P13+Q13Q11ξ20; that is, ξ20=P13+Q13P11+Q11. Therefore, we confirm that at τ=0, the equilibrium point P where all the populations exist is locally asymptotically stable if (P13+Q13)<(P11+Q11)(P12+Q12) holds. However, for small τ>0, (4.24) must hold.

    Putting (4.22) into (4.24) and rearranging the expressions, we get

    (Q13Q11ξ20P11Q12)[cos(ξ0τ)1]+[(Q12P11Q11)ξ0+P11Q13ξ0]sin(ξ0τ)<P11P12P13+P11Q12Q13+Q11ξ20(Q13Q11ξ20P11Q12)[cos(ξ0τ)1]+[(Q12P11Q11)ξ0+P11Q13ξ0]sin(ξ0τ)<(P11+Q11)(P12+Q12)(P13+Q13). (4.25)

    Using the bounds, we obtain

    (Q13Q11ξ20P11Q12)[cos(ξ0τ)1]=2(Q11ξ20+P11Q12Q13)sin2(ξ0τ2)12ξ2+|(Q11ξ2++P11Q12Q13)|τ2,

    and

    [(Q12P11Q11)ξ0+P11Q13ξ0]sin(ξ0τ)[|(Q12P11Q11|ξ2++|P11||Q13|]τM.

    From (4.25), we obtain that Ψ1τ2+Ψ2τΨ3, with

    Ψ1=12|(Q11ξ2++P11Q12Q13)|ξ2+,Ψ2=|(Q12P11Q11|ξ2++|P11||Q13|,Ψ3=(P11+Q11)(P12+Q12)(P13+Q13).

    Thus, we get the expression

    τ+=12Ψ1[Ψ2+Ψ22+4Ψ1Ψ3],

    then the Nyquist criteria [37] holds for 0ττ+. This τ+ is the estimated maximum length of the time lag to preserve the stability of the limit cycle.

    This section demonstrates a few numerical examples to verify our analytical results. Also, we will visualize the role of time delay τ on the system's stability (2.2). In order to do a numerical study of system (2.2), we will use the values of the parameters given in Table (1) that are obtained in [22] using Chest radiograph score data. We would like to refer to [22] for an explanation of the range, values of the parameters, and their units.

    First, we will verify that whatever the value of time delay τ, the virus-free equilibrium point is locally asymptotically stable if χ0<1, i.e., μ<d2d3βE(0). In order to verify this, we take the fixed parameter set as: d1=101, β=0.65, d2=0.11, μ=0.03<d2d3βE(0), d3=5.36 and the initial condition as: E(0)=22.41, I(0)=2.59, v(0)=0.061. From Figure (1), it is clearly observed that in the presence or absence of time delay, the virus-free equilibrium point is locally asymptotically stable if μ<d2d3βE(0). Thus, whatever the value of time delay in the process of antiviral responses of uninfected epithelial cells against the virus, if the virus production rate μ is lower than a threshold value d2d3βE(0), then the system itself is capable of clearing the virus.

    Figure 1.  Depicting time series plot of the system (2.2) with (a) τ=0 and (b) τ=20 for μ=0.03.

    Now, we will check the stability of system (2.2) at the coexisting equilibrium point for τ=0 and τ0. For this, we take the fixed parameter set as: d1=101, β=0.65, d2=0.11, μ=0.24>d2d3βE(0), d3=5.36 and the initial condition as: E(0)=22.41, I(0)=2.59, v(0)=0.061. For the above parameter set and the initial condition, system (2.2) has two equilibrium points: P(22.41,0,0) and P(6.26356,14.6786,0.65725). In absence of time delay, i.e., τ=0, the eigenvalues associated with virus-free equilibrium P(22.41,0,0) are 5.958, 0.4878, and 0.1, which indicates that P is an unstable saddle-node with no oscillating behavior. Hence, around the equilibrium point P, the system (2.2) shows unstable behavior. On the other hand, corresponding to the equilibrium P(6.265,14.68,0.6578), the eigenvalues are 5.431, 0.284, 0.114, which indicates that P is a stable node with no oscillating behavior. Thus, system (2.2) is stable around the equilibrium P. It is clearly observed from Figure (2) that all the populations exist over finite time with stable natures around the equilibrium P(6.26356,14.6786,0.65725).

    Figure 2.  Depicting (a) time series plot and (b) 3D plot of the system (2.2) with τ=0 and μ=0.24.

    After checking the stability of system (2.2) in the absence of delay, we are now interested in checking the stability by varying the time delay parameter τ. For τ=1, the system (2.2) shows similar stability behavior like the case τ=0 around the coexisting equilibrium P(6.26356,14.6786,0.65725) (see Figure (3)). However, when τ=12, the system exhibits oscillatory behavior at first. It stabilizes as time increases (see Figure (4)), which indicates that as time delay increases, i.e., the time required for uninfected epithelial cells to respond to free virus increases, the cell populations oscillate first and are then able to stabilize themselves.

    Figure 3.  Depicting (a) Time series plot and (b) 3D plot of the system (2.2) with τ=1 and μ=0.24.
    Figure 4.  Depicting (a) time series plot and (b) 3D plot of the system (2.2) with τ=12 and μ=0.24.

    We further increase the value of τ to 12.7646, at which point we have found that the system experiences a Hopf bifurcation, i.e., for this critical value of τ=12.7646 system stability is switched to instability via a limit cycle solution. Thus, we have observed a stable limit cycle around the coexisting equilibrium P (see Figure (5)). Moreover, we numerically computed the values of p1=28.6852, p2=0.4165, and p3=0.010076 for which Eq (4.13) has a real root, Ω0=0.16535 and thus τ0=12.7646. Hence, Eq (4.13) has a purely imaginary eigenvalue. So, an isolated periodic solution, i.e., a limit cycle has appeared at τ0=12.7646. The effect of τ on system stability is shown in Figure (8). Also, from Figure (8) it can be confirmed that system (2.2) bifurcates from one solution to two solutions (called maximum and minimum solutions) at τ0=12.7646. Further, the system collapses for an approximate value of τ>45.5. A limit cycle solution of system (2.2), i.e., periodic oscillation of the population, has been observed for τ=20>12.7646 (see Figure (6)). For τ=30>12.7646, system (2.2), shows unstable behavior with an aperiodic nature (see Figure (7)).

    Figure 5.  Depicting (a) time series plot and (b) 3D plot of the system (2.2) with τ=12.7646 and μ=0.24.
    Figure 6.  Depicting (a) time series plot and (b) 3D plot of the system (2.2) with τ=20 and μ=0.24.
    Figure 7.  Depicting (a) time series plot and (b) 3D plot of the system (2.2) with τ=30 and μ=0.24.
    Figure 8.  Maximum and minimum solutions of each population with respect to τ and μ=0.24.

    Biologically, if the response time of uninfected epithelial cells toward the free virus increases to 12.7646, then the cell population can control the free virus through competition. However, with further increases in the antiviral response time of uninfected epithelial cells to the virus, the cell populations are unable to stabilize the virus infection, and the body loses its stability, which also collapses as response time increases.

    In this study, we have developed a modified delay differential model to describe the within-host viral kinetics of SARS-CoV-2. The response of uninfected epithelial cells to the virus or the interaction of the virus with uninfected epithelial cells was incorporated as a nonlinear process into the model. Furthermore, a discrete-time delay has been introduced to describe the time required for uninfected epithelial cells to activate suitable antiviral responses by generating immune cytokines and chemokines. We have found that all the solutions to the formulated model are nonnegative and bounded. The basic reproduction number, χ0 was computed and found as μβE(0)d2d3. In the absence of time delay, the virus-free equilibrium point P was asymptotically stable for χ0<1, and the co-existing equilibrium P was stable for χ0>1, provided that few certain conditions hold. However, in the presence of a time delay, the virus-free equilibrium point P is locally asymptotically stable for χ0<1, and the coexisting equilibrium P loses its stability via a Hopf bifurcation. In addition, we have calculated the time lag over which the system maintains its stability around the co-existing equilibrium, P. The main finding of this study was that the considered system was stable for faster antiviral responses of epithelial cells to the virus, i.e., quick antiviral responses of epithelial cells stabilized patients' bodies by neutralizing the virus. However, if the antiviral response time of uninfected epithelial cells lengthens, the virus infects all the uninfected epithelial cells and may spread the disease over the body, which may require other treatment strategies to neutralize the virus's efficacy.

    The main limitation of this study was that the model was straightforward and considered only two cell populations along with the virus concentration. However, while the SARS-CoV-2 virus interacted with the human body, other immune cells, such as natural killer cells, CD4+ T cells and CD8+ T cells, were also responsible for their antiviral responses. Thus, the model can be extended to explore the immune system's role by considering the effect of these immune cells [15,20,38,39]. To measure the vaccine or any other treatment effect that can be used in the model for eliminating the virus from the body [17,24]; this will be carried out in our future study.

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

    The authors declare there is no conflict of interest.



    [1] K. Liang, Mathematical model of infection kinetics and its analysis for COVID-19, SARS and MERS, Infect. Genet. Evol., 82 (2020), 104306. https://doi.org/10.1016/j.meegid.2020.104306 doi: 10.1016/j.meegid.2020.104306
    [2] C. Yang, J. Wang, A mathematical model for the novel coronavirus epidemic in Wuhan, China, Math. Biosci. Eng., 17 (2020), 2708–2724. https://doi.org/10.3934/mbe.2020148 doi: 10.3934/mbe.2020148
    [3] S. K. Biswas, J. K. Ghosh, S. Sarkar, U. Ghosh, COVID-19 pandemic in India: a mathematical model study, Nonlinear Dyn., 102 (2020), 537–553. https://doi.org/10.1007/s11071-020-05958-z doi: 10.1007/s11071-020-05958-z
    [4] A. A. Arjani, Md. T. Nasseef, S. M. Kamal, B. V. S. Rao, M. Mahmud, Md. S. Uddin, Application of mathematical modeling in prediction of COVID-19 transmission dynamics, Arab. J. Sci. Eng., 47 (2022), 10163–10186. https://doi.org/10.1007/s13369-021-06419-4 doi: 10.1007/s13369-021-06419-4
    [5] A. J. Kucharski, T. W. Russell, C. Diamond, Early dynamics of transmission and control of COVID-19: a mathematical modelling study, Lancet Infect. Dis., 20 (2020), 553–558. https://doi.org/10.1016/S1473-3099(20)30144-4 doi: 10.1016/S1473-3099(20)30144-4
    [6] M. Zamir, F. Nadeem, M. A. Alqudah, T. Abdeljawad, Future implications of COVID-19 through Mathematical modeling, Results Phys., 33 (2022), 105097. https://doi.org/10.1016/j.rinp.2021.105097 doi: 10.1016/j.rinp.2021.105097
    [7] F. A. Rihan, H. J. Alsakaji, C. Rajivganthi, Stochastic SIRC epidemic model with time-delay for COVID-19, Adv. Differ. Equ., 502 (2020), 502(2020). https://doi.org/10.1186/s13662-020-02964-8 doi: 10.1186/s13662-020-02964-8
    [8] S. R. Bandekar, M. Ghosh, C. Rajivganthi, Impact of vaccination on the dynamics of COVID-19: A mathematical study using fractional derivatives, Int. J. Biomath., 17 (2024), 2350018. https://doi.org/10.1142/S1793524523500183 doi: 10.1142/S1793524523500183
    [9] H. J. Alsakaji, F. A. Rihan, A. Hashish, Dynamics of a stochastic epidemic model with vaccination and multiple time-delays for COVID-19 in the UAE, Complexity, 2022 (2022), 4247800. https://doi.org/10.1155/2022/4247800 doi: 10.1155/2022/4247800
    [10] C. Yang, Y. Yang, Z. Li, L. Zhang, Modeling and analysis of COVID-19 based on a time delay dynamic model, Math. Biosci. Eng., 18 (2020), 154–165. https://doi.org/10.3934/mbe.2021008 doi: 10.3934/mbe.2021008
    [11] M. Radha, S. Balamuralitharan, A study on COVID-19 transmission dynamics: stability analysis of SEIR model with Hopf bifurcation for effect of time delay, Adv. Differ. Equ., 2020 (2020), 523. https://doi.org/10.1186/s13662-020-02958-6 doi: 10.1186/s13662-020-02958-6
    [12] A. Raza, A. Ahmadian, M. Rafiq, M. C. Ang, S. Salahshour, M. Pakdaman, The impact of delay strategies on the dynamics of coronavirus pandemic model with nonlinear incidence rate, Fractals, 30 (2022), 2240121. https://doi.org/10.1142/S0218348X22401211 doi: 10.1142/S0218348X22401211
    [13] O. Babasola, O. Kayode, O. J. Peter, F. C. Onwuegbuche, F. A. Oguntolu, Time-delayed modelling of the COVID-19 dynamics with a convex incidence rate, Inform. Med. Unlocked., 35 (2022), 101124. https://doi.org/10.1016/j.imu.2022.101124 doi: 10.1016/j.imu.2022.101124
    [14] S. M. Al‑Tuwairqi, S. K. Al‑Harbi, A time‑delayed model for the spread of COVID‑19 with vaccination, Sci. Rep., 12 (2022), 19435. https://doi.org/10.1038/s41598-022-23822-5 doi: 10.1038/s41598-022-23822-5
    [15] S. Q. Du, W. Yuan, Mathematical modeling of interaction between innate and adaptive immune responses in COVID-19 and implications for viral pathogenesis, J. Med. Virol., 92 (2020), 1615–1628. https://doi.org/10.1002/jmv.25866 doi: 10.1002/jmv.25866
    [16] E. A. Hernandez-Vargas, J. X. Velasco-Hernandez, In-host Mathematical Modelling of COVID-19 in Humans, Annu. Rev. Control, 50 (2020), 448–456. https://doi.org/10.1016/j.arcontrol.2020.09.006 doi: 10.1016/j.arcontrol.2020.09.006
    [17] D. B. Prakash, D. K. K. Vamsi, D. B. Rajesh, C. B. Sanjeevi, Control Intervention Strategies for within-host, between-host and their efficacy in the treatment, spread of COVID-19 : a multi scale modeling approach, Comput. Math. Biophys., 8 (2020), 198–210. https://doi.org/10.1515/cmb-2020-0111 doi: 10.1515/cmb-2020-0111
    [18] B. Chhetri, V. M. Bhagat, D. K. K. Vamsi, V. S. Ananth, B. Prakash, R. Mandale, et al., Within-host mathematical modeling on crucial inflammatory mediators and drug interventions in COVID-19 identifies combination therapy to be most effective and optimal, Alex. Eng. J., 60 (2021), 2491–2512. https://doi.org/10.1016/j.aej.2020.12.011 doi: 10.1016/j.aej.2020.12.011
    [19] A. E. S. Almocera, G. Quiroz, E. A. Hernandez-Vargas, Stability analysis in COVID-19 within-host model with immune response. Commun. Nonlinear Sci. Numer. Simul., 95 (2021), 105584. https://doi.org/10.1016/j.cnsns.2020.105584 doi: 10.1016/j.cnsns.2020.105584
    [20] R. Ghostine, M. Gharamti, S. Hassrouny, I. Hoteit, Mathematical modeling of immune responses against SARS-CoV-2 using an ensemble Kalman Filter, Mathematics, 9 (2021), 2427. https://doi.org/10.3390/math9192427 doi: 10.3390/math9192427
    [21] S. M. E. K. Chowdhury, J. T. Chowdhury, S. F. Ahmed, P. Agarwal, I. A. Badruddin, S. Kamangar, Mathematical modelling of COVID-19 disease dynamics: interaction between immune system and SARS-CoV-2 within host, AIMS Math., 7 (2021), 2618–2633. https://doi.org/10.3934/math.2022147 doi: 10.3934/math.2022147
    [22] C. Li, J. Xu, J. Liu, Y. Zhou, The within-host viral kinetics of SARS-CoV-2, Math. Biosci. Eng., 17 (2020), 2853–2861. https://doi.org/10.3934/mbe.2020159 doi: 10.3934/mbe.2020159
    [23] B. J. Nath, K. Dehingia, V. N. Mishra, Y. M. Chu, H. K. Sarmah, Mathematical analysis of a within-host model of SARS-CoV-2, Adv. Differ. Equ., 2021 (2021), 113. https://doi.org/10.1186/s13662-021-03276-1 doi: 10.1186/s13662-021-03276-1
    [24] I. Ghosh, Within host dynamics of SARS‑CoV‑2 in humans: modeling immune responses and antiviral treatments, SN comput. sci., 2 (2021), 482. https://doi.org/10.1007/s42979-021-00919-8 doi: 10.1007/s42979-021-00919-8
    [25] I. M. Elbaz, M. A. Sohaly, H. El‑Metwally, Modeling the stochastic within‑host dynamics SARS‑CoV‑2 infection with discrete delay, Theor. Biosci., 141 (2020), 365–374. https://doi.org/10.1007/s12064-022-00379-5 doi: 10.1007/s12064-022-00379-5
    [26] V. Staroverov, S. Nersisyan, A. Galatenko, D. Alekseev, S. Lukashevich, F. Polyakov, et al., Development of a novel mathematical model that explains SARS-CoV-2 infection dynamics in Caco-2 cells, PeerJ, 11 (2023), e14828. https://doi.org/10.7717/peerj.14828 doi: 10.7717/peerj.14828
    [27] L. G. de Pillis, R. Caffrey, G. Chen, et al., A mathematical model of the within-host kinetics of SARS-CoV-2 neutralizing antibodies following COVID-19 vaccination, J. Theor. Biol., 556 (2023), 111280. doi:10.1016/j.jtbi.2022.111280 doi: 10.1016/j.jtbi.2022.111280
    [28] T. A. Miura, Respiratory epithelial cells as master communicators during viral infections, Curr. Clin. Microbiol., 6 (2019), 10–17. https://doi.org/10.1007/s40588-019-0111-8 doi: 10.1007/s40588-019-0111-8
    [29] A. N. Chatterjee, F. A. Basir, M. A. Almuqrin, J. Mondal, I. Khan, SARS-CoV-2 infection with lytic and non-lytic immune responses: a fractional order optimal control theoretical study, Results in Physics, 26 (2021), 104260. https://doi.org/10.1016/j.rinp.2021.104260 doi: 10.1016/j.rinp.2021.104260
    [30] H. J. Alsakaji, S. Kundu, F. A. Rihan, Delay differential model of one-predator two-prey system with Monod-Haldane and holling type Ⅱ functional responses, Appl. Math. Comput., 397 (2021). https://doi.org/10.1016/j.amc.2020.125919 doi: 10.1016/j.amc.2020.125919
    [31] X. Yang, L. Chen, J. Chen, Permanence and positive periodic solution for the single-species non-autonomous delay diffusive models, Comput. Math. Applic., 32 (1996), 109–116. https://doi.org/10.1016/0898-1221(96)00129-0 doi: 10.1016/0898-1221(96)00129-0
    [32] P. V. 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
    [33] J. P. LaSalle, The stability of dynamical systems, SIAM, Philadelphia, Pa, USA, 1976. https://doi.org/10.21236/ADA031020
    [34] Y. Kuang, Delay differential equations with applications in population dynamics, Academic Press, Inc.: Boston, MA, USA, 1993.
    [35] H. I. Freedman, V. S. H. Rao, Stability criteria for a system involving two-time delays, SIAM J. Appl. Math., 46 (1986), 552–560. https://doi.org/10.1137/0146037 doi: 10.1137/0146037
    [36] L. H. Erbe, H. I. Freedman, V. S. H. Rao, Three species food chain models with mutual interference and time delays, Math. Biosci., 80 (1986), 57–80. https://doi.org/10.1016/0025-5564(86)90067-2 doi: 10.1016/0025-5564(86)90067-2
    [37] H. I. Freedman, V. S. H. Rao, The trade-off between mutual interference and time lags in predator-prey systems, Bull. Math. Biol., 45 (2019), 1983,991–1004. https://doi.org/10.1016/S0092-8240(83)80073-1 doi: 10.1016/S0092-8240(83)80073-1
    [38] Y. Fadaei, F. A. Rihan, C. Rajivganthi, Immunokinetic model for COVID-19 patients, Complexity, 2022 (2022), 8321848. https://doi.org/10.1155/2022/8321848 doi: 10.1155/2022/8321848
    [39] G. Li, Y. Fan, Y. Lai, T. Han, Z. Li, P. Zhou, et al., Coronavirus infections and immune responses, J. Med. Virol., 92 (2020), 424–432. https://doi.org/10.1002/jmv.25685 doi: 10.1002/jmv.25685
  • This article has been cited by:

    1. Kottakkaran Sooppy Nisar, Muhammad Farman, Mahmoud Abdel-Aty, Chokalingam Ravichandran, A review of fractional order epidemic models for life sciences problems: Past, present and future, 2024, 95, 11100168, 283, 10.1016/j.aej.2024.03.059
    2. Kaushik Dehingia, Salah Boulaaras, Evren Hinçal, Kamyar Hosseini, Thabet Abdeljawad, M.S. Osman, On the dynamics of a financial system with the effect financial information, 2024, 106, 11100168, 438, 10.1016/j.aej.2024.08.049
    3. Rituparna Bhattacharyya, Brajesh Kumar Jha, Computational Fuzzy Modelling Approach to Analyze Neuronal Calcium Dynamics With Intracellular Fluxes, 2024, 1085-9195, 10.1007/s12013-024-01541-0
    4. Parvaiz Ahmad Naik, Anum Zehra, Muhammad Farman, Aamir Shehzad, Sundas Shahzeen, Zhengxin Huang, Forecasting and dynamical modeling of reversible enzymatic reactions with a hybrid proportional fractional derivative, 2024, 11, 2296-424X, 10.3389/fphy.2023.1307307
    5. Kottakkaran Sooppy Nisar, Muhammad Farman, Anum Zehra, Evren Hincal, Numerical and analytical study of fractional order tumor model through modeling with treatment of chemotherapy, 2024, 0228-6203, 1, 10.1080/02286203.2024.2327659
  • 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(1810) PDF downloads(74) Cited by(5)

Figures and Tables

Figures(8)  /  Tables(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog