Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article

Mathematical modeling of intervention and low medical resource availability with delays: Applications to COVID-19 outbreaks in Spain and Italy

  • Received: 29 April 2021 Accepted: 08 June 2021 Published: 30 June 2021
  • Infectious diseases have been one of the major causes of human mortality, and mathematical models have been playing significant roles in understanding the spread mechanism and controlling contagious diseases. In this paper, we propose a delayed SEIR epidemic model with intervention strategies and recovery under the low availability of resources. Non-delayed and delayed models both possess two equilibria: the disease-free equilibrium and the endemic equilibrium. When the basic reproduction number R0=1, the non-delayed system undergoes a transcritical bifurcation. For the delayed system, we incorporate two important time delays: τ1 represents the latent period of the intervention strategies, and τ2 represents the period for curing the infected individuals. Time delays change the system dynamics via Hopf-bifurcation and oscillations. The direction and stability of delay induced Hopf-bifurcation are established using normal form theory and center manifold theorem. Furthermore, we rigorously prove that local Hopf bifurcation implies global Hopf bifurcation. Stability switching curves and crossing directions are analyzed on the two delay parameter plane, which allows both delays varying simultaneously. Numerical results demonstrate that by increasing the intervention strength, the infection level decays; by increasing the limitation of treatment, the infection level increases. Our quantitative observations can be useful for exploring the relative importance of intervention and medical resources. As a timing application, we parameterize the model for COVID-19 in Spain and Italy. With strict intervention policies, the infection numbers would have been greatly reduced in the early phase of COVID-19 in Spain and Italy. We also show that reducing the time delays in intervention and recovery would have decreased the total number of cases in the early phase of COVID-19 in Spain and Italy. Our work highlights the necessity to consider the time delays in intervention and recovery in an epidemic model.

    Citation: 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[J]. Mathematical Biosciences and Engineering, 2021, 18(5): 5865-5920. doi: 10.3934/mbe.2021295

    Related Papers:

    [1] Juan Wang, Chunyang Qin, Yuming Chen, Xia Wang . Hopf bifurcation in a CTL-inclusive HIV-1 infection model with two time delays. Mathematical Biosciences and Engineering, 2019, 16(4): 2587-2612. doi: 10.3934/mbe.2019130
    [2] 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
    [3] Yuan Ma, Yunxian Dai . Stability and Hopf bifurcation analysis of a fractional-order ring-hub structure neural network with delays under parameters delay feedback control. Mathematical Biosciences and Engineering, 2023, 20(11): 20093-20115. doi: 10.3934/mbe.2023890
    [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] Jiawei Deng, Ping Jiang, Hongying Shu . Viral infection dynamics with mitosis, intracellular delays and immune response. Mathematical Biosciences and Engineering, 2023, 20(2): 2937-2963. doi: 10.3934/mbe.2023139
    [6] Anuj Kumar, Yasuhiro Takeuchi, Prashant K Srivastava . Stability switches, periodic oscillations and global stability in an infectious disease model with multiple time delays. Mathematical Biosciences and Engineering, 2023, 20(6): 11000-11032. doi: 10.3934/mbe.2023487
    [7] Toshikazu Kuniya, Hisashi Inaba . Hopf bifurcation in a chronological age-structured SIR epidemic model with age-dependent infectivity. Mathematical Biosciences and Engineering, 2023, 20(7): 13036-13060. doi: 10.3934/mbe.2023581
    [8] Hongfan Lu, Yuting Ding, Silin Gong, Shishi Wang . Mathematical modeling and dynamic analysis of SIQR model with delay for pandemic COVID-19. Mathematical Biosciences and Engineering, 2021, 18(4): 3197-3214. doi: 10.3934/mbe.2021159
    [9] Zhichao Jiang, Xiaohua Bi, Tongqian Zhang, B.G. Sampath Aruna Pradeep . Global Hopf bifurcation of a delayed phytoplankton-zooplankton system considering toxin producing effect and delay dependent coefficient. Mathematical Biosciences and Engineering, 2019, 16(5): 3807-3829. doi: 10.3934/mbe.2019188
    [10] 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
  • Infectious diseases have been one of the major causes of human mortality, and mathematical models have been playing significant roles in understanding the spread mechanism and controlling contagious diseases. In this paper, we propose a delayed SEIR epidemic model with intervention strategies and recovery under the low availability of resources. Non-delayed and delayed models both possess two equilibria: the disease-free equilibrium and the endemic equilibrium. When the basic reproduction number R0=1, the non-delayed system undergoes a transcritical bifurcation. For the delayed system, we incorporate two important time delays: τ1 represents the latent period of the intervention strategies, and τ2 represents the period for curing the infected individuals. Time delays change the system dynamics via Hopf-bifurcation and oscillations. The direction and stability of delay induced Hopf-bifurcation are established using normal form theory and center manifold theorem. Furthermore, we rigorously prove that local Hopf bifurcation implies global Hopf bifurcation. Stability switching curves and crossing directions are analyzed on the two delay parameter plane, which allows both delays varying simultaneously. Numerical results demonstrate that by increasing the intervention strength, the infection level decays; by increasing the limitation of treatment, the infection level increases. Our quantitative observations can be useful for exploring the relative importance of intervention and medical resources. As a timing application, we parameterize the model for COVID-19 in Spain and Italy. With strict intervention policies, the infection numbers would have been greatly reduced in the early phase of COVID-19 in Spain and Italy. We also show that reducing the time delays in intervention and recovery would have decreased the total number of cases in the early phase of COVID-19 in Spain and Italy. Our work highlights the necessity to consider the time delays in intervention and recovery in an epidemic model.



    In the past decade, infectious diseases have been frequently threatening human lives on a large scale. After the first epidemic model of smallpox was proposed by a pioneering mathematician Daniel Bernoulli in the eighteenth century [1,2], a huge number of epidemic models have been proposed and studied to uncover the underlying spread mechanisms and to control the spread of infectious diseases. Kermack and Mckendrick [3] introduced a seminal SIR compartmental model in 1927 to study the plague disease in Mumbai and succeeded in revealing its epidemiological transition. Since then, mathematical modeling has been growing as a vital tool to suggest public health responses against the spread and transmission of infectious diseases. The SIR and SEIR type models have been performed a crucial role in modern mathematical epidemiology [4,5].

    Numerous factors significantly affect the dynamical behavior of a particular epidemic model, such as the demographics, recovery or treatment rates, and incidence rates [6,7]. In particular, different types of incidence functions can ensure appropriate dynamical characteristics of the associated infectious diseases [8,9]. In classical SIR/SEIR models, the general incidence rate βSIN, bilinear incidence rate βSI and linear recovery rate γI are commonly used. The constant β represents the average number of adequate contacts between susceptible and an infected individual per unit time, and γ represents the per capita recovery rate of infected individuals. Since the infection risk increases with the increase in the number of infective individuals, therefore, the mobility of individuals probably influences these numbers. When a particular infectious disease appears and spreads in a region/community, intervention strategies play a crucial role in controlling the disease [10,11].

    The impacts of intervention strategies, such as mask-wearing, border screening, isolation, quarantine, or communications through the mass media (to convey the information to the public about the epidemic/outbreak and probably risk-reducing behavior), play their significant roles in managing effectual interventions to control the spread of disease and expectantly remove epidemic diseases [12,13,14,15]. For instance, during the outbreak of the H1N1 influenza pandemic in 2009 [16,17] and the outbreak of SARS in 2003 [18,19], the intervention strategies (such as postponing conferences, closing restaurants/schools, isolating infectives, etc.) were chosen by the Chinese government. Recently, a new pandemic COVID-19 appeared and affected more than 200 countries and territories worldwide. To control the spread of COVID-19, governments of different countries strictly implemented the lockdown [20]. Besides lockdown, various governments are also adopting numerous steps and implementing different intervention strategies, such as social distancing, tracing close contacts, washing hands for 30 seconds, etc [21]. In many countries, these strategies impressively worked so that the contacts between susceptible and infected individuals per unit time are being decreased sufficiently, and therefore, the reduced incidence rate [22,23,24,25]. This showed the importance of considering the infection forces that comprises the adaptation of individuals to infection risks under intervention strategies.

    In classical epidemic models, the recovery rate of infected individuals is commonly considered to be proportional to the number of infectives [26,27]. However, in general, this is unacceptable because the implicit assumption is that treatment resources are sufficient [28]. Every society should have a limited capacity of resources for treatment. If medical resources are available in large quantities, the community needs to pay unnecessary costs during non-epidemic periods, and if available resources are low, the community has the risk of a disease outbreak without needed treatment [26]. The recovery rate depends on the resources of the health system available to the community, particularly, the capacity of the hospital settings, efficiency, and effectiveness of treatment [29,30]. The prime factor is the number of health employees, including nurses, physicians, pharmacists, and other health care workers (HCW). The amenities of the hospital, such as equipment and medical apparatus, medicines, and the number of hospital beds are other noteworthy factors that are important for effective and safe prevention, treatment, and diagnosis of patients [28].

    Shan et al. [28] studied an SIR model with the impact of the number of beds in hospitals. The authors considered the standard incidence rate and nonlinear recovery rate. They studied the complex dynamics of the model including numerous bifurcations such as backward bifurcation, saddle-node bifurcation, Hopf bifurcation, Bogdanov-Takens bifurcation. The study recommends that maintaining an adequate number of beds in the hospital is crucial to control the disease. Li et al. [29] investigated the dynamics of an SIR model using nonlinear incidence rate and nonlinear recovery rate. The model exhibited complicated dynamics and suggested that a sufficient number of beds is critical to control the disease. Mu et al. [30] proposed an SI-SIR model for avian influenza with a nonlinear recovery rate under available hospital resources. Zhao et al. [31] investigated the complex dynamics of Zika virus transmission via mathematical modeling with limited medical resources. In [28,29,30,31], the authors considered a similar recovery rate that decreases with a decrease in the number of beds/resources in the hospitals. In the case of low availability of resources for treatment, the recovery rate primarily grows with an increase in the number of infected individuals and approaches the maximum and then starts decreasing. When deliveries of treatment (immunization, medicine, etc.) are depleted due to a large number of infected individuals, the available resources for treatment become very low.

    Time delays are inevitable in almost all realistic systems. Introducing time delays in epidemic models can be important, and time delays include the latency period [32,33], the delay in recovery [34], the delay in media broadcast [35], the delay in awareness programs [36] etc. Different types of delay differential equations (DDEs) have fruitfully been utilized to study infectious diseases [37,38,39]. Greenhalgh et al. [40] proposed a mathematical model with multiple time delays, one delay due to the memory disappearance of aware individuals, and the other delay between the rising awareness and the disease appearance. Zhou et al. [41] analyzed an SIR model with media coverage incorporating time delay and studied delay induced local and global Hopf-bifurcation. The authors found that the delay in media coverage does not influence the stability of the disease-free equilibrium. In [42], the authors proposed a mathematical model with media coverage incorporating time delay. Recently, Liu et al. [43] studied the delay induced local Hopf bifurcation of an SEIR model with multiple time delays caused by the latent period and the recovery period.

    A common assumption about intervention strategies in mathematical models [10,11] is that the impact of intervention strategies on the transmission dynamics is effective, i.e., the number of infected individuals instantly decrease. The recovery rate is assumed to be proportional to the number of infectives, but many researchers argued that this assumption is questionable in reality as we have discussed earlier. As an improvement, we propose an SEIR model with infection force under intervention strategies and a nonlinear recovery function (under low availability of treatment resources), where treatment declines (due to resources limitation) after attaining its maximum value. Further, we incorporate two time delays, the first delay represents the latent period of interventions and the second delay represents the period for curing the infected individuals. The two main objectives of the study are (i) to investigate the impact of low availability of resources for treatment in the presence of intervention strategies, and (ii) to assess the role of time delays in the transmission dynamics. Furthermore, we validate our model and estimate the parameter values based on the available COVID-19 data in Spain and Italy.

    The remaining paper is organized as follows. In Section 2, we formulate an SEIR model (both non-delayed and delayed) with infection force under intervention strategies and recovery rate (under low treatment resources). In Section 3, we analyze the non-delayed system (2.2). In Section 4, we analyze the delayed system (2.3). Persistence has been shown in Section 5. In Section 6, numerical simulations have been presented for both non-delayed system (2.2) and delayed system (2.3). In Section 7, a case study has been discussed for COVID-19 outbreaks in Spain and Italy. The paper ends with conclusion and future developments in Section 8.

    We start with the classical SEIR model in which the total population is divided into four compartments: susceptible individuals (S), exposed individuals (E), infected individuals (I), and recovered individuals (R). Susceptible individuals are healthy but can be infected via contacting with infectives. Individuals who are exposed to disease pathogen and may be infected via contacts are called exposed individuals. Exposed individuals may or may not develop the disease, and these individuals are typically not infectious. Infected individuals are those who have been infected and are capable of transmitting the disease to susceptible individuals. Recovered individuals are those who were infected but have been healthy now, and they have immune protection for a long time. Under the above assumptions, the SEIR model with incidence rate under intervention strategies and recovery rate under low availability of medical resources is provided by

    dSdt=AμSF(I)S,dEdt=(μ+q)E+F(I)S,dIdt=qE(μ+δ)IG(I),dRdt=G(I)μR, (2.1)

    associated with the state space R4+={(S,E,I,R):S>0,E>0,I>0,R>0}. All parameters are assumed to be positive. Here A is the rate at which new individuals (including immigrants and newborns) enter into the susceptible population; μ is the natural mortality rate; q is the rate at which exposed individuals become infectious; δ is the disease induced death rate. The infection force F(I) and recovery function G(I) in system (2.1) are the functions of infective individuals. These two functions play important roles in governing the disease transmission dynamics. There are various nonlinear transmission rates and nonlinear recovery rates proposed by researchers, for instance, [10,11,13,14,28,44,45] and references therein.

    System (2.1) comprises the adaptation of human behavior under intervention strategies. In fact, when a new infectious disease emerges, both the infection probability and contact rate increase since people have less knowledge about the disease. Further, when the number of infective individuals get larger and the disease becomes more serious, psychological factors lead people to modify their behavior and implement suitable measures/intervention to reduce the opportunities of contact and the probability of infection. For instance, F(I) may decrease when the number of infectives increases since the government may tend to contain the infectious disease by intervention strategies. It has also been interpreted as the psychological effect [14]. Mathematically, this idea could be modeled as follows: the infection force F(I) increases when I is small and decreases when I is large (shown in Figure 1). For notational convenience, we assume that the infection force F(I) could be factorized into βIf(I), where 1f(I) signifies the impact of intervention strategies on the decrease of effective contact coefficient β [27]. In the absence of intervention strategies, i.e., f(I)=1, it is notable that the infection force takes the form of well-known bilinear transmission βSI. Here, we have the following assumptions for f(I):

    Figure 1.  The graphs show that both infection force and recovery function first increase with respect to the number of infected individuals, and after a critical number of infected individuals, both functions decrease.

    H(1):f(0)>0 and f(I)>0 for I>0,

    H(2): There is a ξ such that (If(I))>0 for 0<I<ξ and (If(I))<0 for I>ξ.

    These assumptions explain the impacts of intervention strategies determined by a critical value ξ: if 0<I<ξ, the infection force increases, while if I>ξ, the infection force decreases.

    In case of low availability of medical resources for a large number of infectives, the recovery function primarily grows with increase of the number of infectives and approaches the maximum and then declines (shown in Figure 1). We have the following assumptions for g(I):

    H(3): g(0)>0 and g(I)>0 for I>0,

    H(4): There is a η such that (Ig(I))>0 for 0<I<η and (Ig(I))<0 for I>η.

    These assumptions explain the impacts of treatment policies determined by a critical value η: if 0<I<η, the recovery function increases, while if I>η, the recovery function decreases.

    Under the above assumptions on the infection force and recovery function, the new SEIR model (2.1) takes the following form:

    dSdt=AμSβIf(I)S,dEdt=βIf(I)S(μ+q)E,dIdt=qE(μ+δ)IγIg(I),dRdt=γIg(I)μR. (2.2)

    Furthermore, the functions f(I) and g(I) could be chosen from a combination of any of the following formats:

    1. f(I)org(I)=1+αIp,p>1,

    2. f(I)org(I)=exp(αI),

    3. f(I)org(I)=1+I+αIp,p>1,

    where α and p are positive constants. For the case of intervention strategies, the parameter α could be understood as the strength of interventions, and for the case of treatment, the parameter α could be understood as limitation on the treatment availability. In particular, if we take g(I)=1, then model (2.2) becomes same as the model discussed in [11].

    In Figure 2, the infection force has been plotted with respect to infected individuals and the strength of interventions. It depicts that when the strength of interventions is zero, then incidence function always increases. However, in presence of interventions, the incidence function first increases and then decreases. Moreover, the maximum of F(I) becomes smaller with an increase in the strength of interventions. Similarly in Figure 3, the recovery function has been plotted with respect to infected individuals and limitation on treatment resources. We see that when the resource limitation is zero, then the recovery function increases. However, in presence of resource limitations, the recovery function first increases and then decreases. In addition, the maximum of G(I) becomes smaller with an increase in resource limitations.

    Figure 2.  The graph shows the surface plot of the infection force with respect to infected individuals and strength of interventions.
    Figure 3.  The graph shows the surface plot of the recovery function with respect to infected individuals and limitation on treatment resources.
    Figure 4.  Figure shows the existence of a unique endemic equilibrium D when R0>1.

    Delays naturally exist and can play an important role in disease dynamics. We include two time delays in system (2.2), the first delay τ1 represents the latent period of the intervention strategies and the second delay τ2 represents the period for curing the patients. Delays in intervention and curing processes are significant in infectious disease modeling. For example, the delay in intervention strategies, including the time duration for individuals' responses to the reported infection and the reporting delay, have been observed in H1N1-2009 [35]. The model (2.2) with these time delays is provided by

    dSdt=AμSβIf(I(tτ1))S,dEdt=βIf(I(tτ1))S(μ+q)E,dIdt=qE(μ+δ)IγI(tτ2)g(I(tτ2)),dRdt=γI(tτ2)g(I(tτ2))μR. (2.3)

    The initial conditions for the model (2.3) are defined as

    S(θ)=ϕ1(θ),E(θ)=ϕ2(θ),I(θ)=ϕ3(θ),R(θ)=ϕ4(θ), (2.4)

    where θ[τ,0], τ=max{τ1,τ2}, (ϕ1,ϕ2,ϕ3,ϕ4)C([τ,0],R4), ϕi(0)>0,i=1,2,3,4, and C([τ,0],R4) is the Banach space of continuous functions from [τ,0] to R4 equipped with the sup-norm. It could easily be established from the elementary theory of functional differential equations [49] that the system (2.3) possesses a unique solution with initial conditions (2.4).

    It is essential to show that all the population variables are nonnegative for all t0, which indicates that any trajectory that begins with positive initial condition will stay positive for t0.

    Theorem 3.1. The closed region

    ={(S,E,I,R)R4+:0<S+E+I+RAμR4+} (3.1)

    is positively invariant for system (2.2).

    Proof. See Appendix A.

    The system (2.2) has the following two equilibria:

    1. The disease-free equilibrium D0(Aμ,0,0,0), which always exists. Here we define the basic reproduction number R0 for system (2.2) as

    R0=βAqμf(0)(μ+q)(μ+δ+γg(0)), (3.2)

    where 1μ+δ+γg(0) is the life expectancy of infectious persons. Aμ signifies the number of susceptible persons at the starting of the infectious process and βAμf(0) represents the value when all the individuals are susceptible. qμ+q represents the fraction of exposed individuals who survive in class E and become infected. Hence the basic reproduction number R0 is biologically well interpreted.

    2. The endemic equilibrium D(S,E,I,R), where

    S=Af(I)βI+μf(I),E=βAI(μ+q)(βI+μf(I)),R=γμIg(I),

    and I is a unique positive root of the following equation

    AββI+μf(I)γq(μ+q)g(I)(μ+q)(μ+δ)q=0. (3.3)

    Let

    Q(I)=AββI+μf(I)γq(μ+q)g(I)(μ+q)(μ+δ)q,

    then we can see that Q(0)=Aβμf(0)γq(μ+q)g(0)(μ+q)(μ+δ)q>0, if R0>1. Let Q1(I)=AββI+μf(I) and Q2(I)=(μ+q)(μ+δ)q+γq(μ+q)g(I), then we have limIQ1(I)=0, limIQ2(I)=(μ+q)(μ+δ)q, Q1(I)<0 and Q2(I)<0. We also observe that Q1(0)>Q2(0) when R0>1. Hence Q1(I) and Q2(I) intersect each other only once when R0>1, which proves the existence and uniqueness of D.

    We observe that Q1(0)<Q2(0) when R0<1 and both Q1(I) and Q2(I) are monotonically decreasing. Thus, Q1(I) and Q2(I) do not intersect each other when R0<1. Therefore, there is no endemic equilibrium when R0<1.

    On the other hand, by choosing f(I)=1+α1I2 and g(I)=1+α2I2, we see the existence of D by plotting the two nullclines in SI-plane.

    Theorem 3.2. 1. The disease-free equilibrium D0 is locally asymptotically stable when R0<1 and unstable when R0>1.

    2. System (2.2) undergoes a transcritical bifurcation when R0=1 and the endemic equilibrium D is locally asymptotically stable when R0>1.

    Proof. For the proof of Theorem 3.2, interested readers may refer to Appendix A.

    The threshold quantity R0 denotes the average number of new disease infection generated by a single infection introduced into a community of susceptible individuals. The result of Theorem 3.2 shows that the disease can be eliminated in the community when the basic reproduction number R0<1 and the initial population sizes are within the attraction basin of D0.

    Theorem 3.3. The disease-free equilibrium D0 is globally asymptotically stable when R0<1.

    Proof. For the proof of Theorem 3.3, one may refer to Appendix A.

    Theorem 3.4. The endemic equilibrium D is globally asymptotically stable when R0>1.

    Proof. Refer to Appendix A.

    Remark 3.1. 1. One can observe that in the absence of intervention strategies, i.e., f(I)=1, the basic reproduction number is given by R0=βAqμ(μ+q)(μ+δ+γg(0)). Thus if f(0)=1, which means that if we perform intervention strategies at a suitable level of infection, R0 remains same. However, by our assumption, f is an increasing function, which means R0 is greater than that in the presence of interventions. From Theorem 3.4, D is globally asymptotically stable, thus intervention strategies decreases the endemic level of disease.

    2. Further, if there is no limitation on resources of treatment for large infective individuals, i.e., g(I)=1, the basic reproduction number is given by R0=βAqμf(0)(μ+q)(μ+δ+γ). Thus if g(0)=1, which means that if there are low availability of resources for treatment at some level of infection, R0 remains same. However, by our assumption, g is an increasing function, which means R0 would be less than that in case of low availability of resources for treatment. Since from Theorem 3.4, D is globally asymptotically stable, therefore, low resources of treatment increase the endemic level of disease. Clearly, it could be concluded that the disease persists along with the low availability of resources for treatment when the basic reproduction number is greater than one.

    Define L(t)=mint0{S(t),E(t),I(t),R(t)}. Clearly, L(0)=min{S(0),E(0),I(0),R(0)}>0. We need to show that L(t)>0 for all t0. Suppose that there exists a t0>0 such that L(t0)=0 and L(t)>0 for all t[0,t0). Here, we need to discuss the following four cases: (i) L(t0)=S(t0); (ii) L(t0)=E(t0); (iii) L(t0)=I(t0); (iv) L(t0)=R(t0). We only give the proof of case (iv). The remaining cases could be discussed similarly.

    Let L(t0)=R(t0). Since L(t)>0 for all t[0,t0), I(tτ2)>0,

    dRdtμR(t)t[0,t0),

    integrating from 0 to t0, we obtain

    0=R(t0)R(0)eμt0>0,

    which leads a contradiction. Hence S(t),E(t),I(t),R(t) are positive for all t0. For the boundedness of system (2.3), one can follow Section 3. The closed region defined in (3.1) is also invariant for system (2.3).

    The first three equations of system (2.3) do not include the variable R. Therefore, for simplicity, we ignore the forth equation and analyze the following reduced system containing only first three equations:

    dSdt=AμSβIf(I(tτ1))S,dEdt=βIf(I(tτ1))S(μ+q)E,dIdt=qE(μ+δ)IγI(tτ2)g(I(tτ2)). (4.1)

    Since delay does not affect the number of equilibria, system (4.1) always has a disease-free equilibrium D0 and a unique endemic equilibrium D, which have already been defined in Section 3. The basic reproduction number R0 has also been defined in Section 3.

    Theorem 4.1. If V211V2132V12>0 and V212V214>0, D0 is locally asymptotically stable when τ1>0,τ2>0 and R0<1.

    2. D0 is unstable when R0>1.

    Proof. For the proof of Theorem 4.1, one may refer to Appendix A.

    Remark 4.1. From the proof of Theorem 4.1 in Appendix A, we observe that if τ1>0,τ2=0, D0 is locally asymptotically stable when R0<1. Further if τ1=0 and τ2>0, then D0 is not locally asymptotically stable. Thus the result of Theorem 4.1 shows that time delay in the interventions (τ1) does not affect the stability of D0, however, the result for the time delay in the recovery (τ2) suggests that only a restriction on the basic reproduction number R0 will not be sufficient for the local stability of disease free equilibrium.

    In this subsection, we discuss the stability of D and establish the local Hopf bifurcation of system (4.1). The characteristic equation of system (4.1) at D is given by

    det(J11λ0J12+J18eλτ1J13J14λJ15+J19eλτ10J16J17+J20eλτ2λ)=0, (4.2)

    where

    J11=μβIf(I),J12=βSf(I),J13=βIf(I),J14=(μ+q),J18=βISf(I)2f(I),J19=βISf(I)2f(I),J15=βSf(I),J16=q,J17=(μ+δ),J20=γ[Ig(I)g(I)]g(I)2.

    Eq. (4.2) is equivalent to

    λ3+P11λ2+P12λ+P13+(P14λ+P15)eλτ1+(P16λ2+P17λ+P18)eλτ2=0, (4.3)

    where

    P11=(J11+J14+J17),P12=J11(J14+J17)+J14J17J15J16,P13=J11(J14J17+J15J16)J12J13J16,P14=J16J19,P15=J16(J11J19J13J18),P16=J20,P17=J20(J11+J14),P18=J11J14J20.

    For the stability and Hopf bifurcation, we have the following results:

    Remark 4.2. If τ1=τ2=0. In this case, system (4.1) becomes the non-delayed system (2.2). We have already discussed this case in Section 3.

    Lemma 4.2. [54] For the transcendental equation

    x(λ,eλτ1,,eλτm)=λn+x(0)1λn1++x(0)n1λ+x(0)n+[x(1)1λn1++x(1)n1λ+x(1)n]eλτ1++[x(m)1λn1++x(m)n1λ+x(m)n]eλτm=0,

    as τ1,τ2,τm vary, the sum of the orders of the zeros of x(λ,eλτ1,,eλτm) in the open right half plane can change only if a zero occurs on or crosses the imaginary axis.

    Lemma 4.3. If M11M13+M12M14>0, then (dλdτ1)1λ=iω1,τ1=τ1j>0(j=0,1,2,) holds.

    Similarly one can prove (dλdτ1)1λ=iω±1,τ1=τ1k>0(k=0,1,2,) and (dλdτ1)1λ=iω1,2,31,τ1=τ1l>0(l=0,1,2,).

    Theorem 4.4. For reduced system (4.1), the following conclusions hold:

    1. If c11>0,c12>0,c13>0, then D is locally asymptotically stable for τ1(0,+).

    2. If c13<0,c11>0,c12>0 or c13<0,c11>0,c12<0 or c13<0,c11<0,c12<0, there exists a ξ1j such that 0<ξ1jτ1j. Then D is locally asymptotically stable for ξ1j<τ1<τ1j and unstable for ξ1j+1>τ1>τ1j. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ1=τ1j.

    3. If c11>0, c12<0,c13>0 or c11<0, c12>0,c13>0 or c11<0, c12<0,c13>0, there exists a σ1k such that 0<σ1kτ1k. Then D is locally asymptotically stable for σ1k<τ1<τ1k and unstable for σ1k+1>τ1>τ1k. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ1=τ1k.

    4. If c11<0, c12>0,c13<0, there exists a σ1l such that 0<σ1lτ1l. Then D is locally asymptotically stable for σ1l<τ1<τ1l and unstable for σ1l+1>τ1>τ1l. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ1=τ1l.

    Lemma 4.5. If M21M23+M22M24>0, then Re(dλdτ2)1λ=iω2,τ2=τ2j>0(j=0,1,2,) holds.

    Similarly one can also prove (dλdτ2)1λ=iω±2,τ2=τ2k>0(k=0,1,2,) and (dλdτ2)1λ=iω1,2,32,τ2=τ2l>0(l=0,1,2,).

    Theorem 4.6. For reduced system (4.1), the following conclusions hold:

    1. If d11>0,d12>0,d13>0, then D is locally asymptotically stable for τ2(0,+).

    2. If d13<0,d11>0,d12>0 or d13<0,d11>0,d12<0 or d13<0,d11<0,d12<0, there exists a ξ2j such that 0<ξ2jτ2j. Then D is locally asymptotically stable for ξ2j<τ2<τ2j and unstable for ξ2j+1>τ2>τ2j. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ2=τ2j.

    3. If d11>0, d12<0,d13>0 or d11<0, d12>0,d13>0 or d11<0, d12<0,d13>0, there exists a σ2k such that 0<σ2kτ2k. Then D is locally asymptotically stable for σ2k<τ2<τ2k and unstable for σ2k+1>τ2>τ2k. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ2=τ2k.

    4. If d11<0, d12>0,d13<0 there exists a σ2l such that 0<σ2lτ2l. Then D is locally asymptotically stable for σ2l<τ2<τ2l and unstable for σ2l+1>τ2>τ2l. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ2=τ2l.

    Lemma 4.7. If M31M33+M32M34>0, then Re(dλdτ)1λ=iω3,τ=τj>0(j=0,1,2,) holds.

    Similarly we can also prove (dλdτ)1λ=iω±3,τ=τk>0(k=0,1,2,) and (dλdτ)1λ=iω1,2,33,τ=τl>0(l=0,1,2,).

    Theorem 4.8. For reduced system (4.1), the following conclusions hold:

    1. If e11>0,e12>0,e13>0, then D is locally asymptotically stable for τ(0,+).

    2. If e13<0,e11>0,e12>0 or e13<0,e11>0,e12<0 or e13<0,e11<0,e12<0, there exists a ξ3j such that 0<ξ3jτj. Then D is locally asymptotically stable for ξ3j<τ<τj and unstable for ξ3j+1>τ>τj. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ=τj.

    3. If e11>0, e12<0,e13>0 or e11<0, e12>0,e13>0 or e11<0, e12<0,e13>0, there exists a σ3k such that 0<σ3kτk. Then D is locally asymptotically stable for σ3k<τ<τk and unstable for σ3k+1>τ>τk. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ=τk.

    4. If e11<0, e12>0,e13<0, there exists a σ3l such that 0<σ3lτl. Then D is locally asymptotically stable for σ3l<τ<τl and unstable for σ3l+1>τ>τl. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ=τl.

    Lemma 4.9. If M41M43+M42M44>0, then Re(dλdτ1)1λ=iω4,τ1=τ1j>0(j=0,1,2,) holds.

    Theorem 4.10. For reduced system (4.1), If H1 holds, there exists a ξ4j such that 0<ξ4jτ1j. Then D is locally asymptotically stable for ξ4j<τ1<τ1j and unstable for ξ4j+1>τ1>τ1j. Additionally, system (4.1) undergoes a supercritical Hopf bifurcation at D when τ1=τ1j.

    Remark For the proofs of Lemmas 4.3, 4.5, 4.7, 4.9, and Theorems 4.4, 4.6, 4.8, 4.10, interested readers may refer to Appendix A.

    Remark 4.3. Without loss of generality, in Subsection 4.3 we only discussed the case Re(dλdτj)1λ=iω,τ=τj>0(j=1,2). Similarly, for the case of Re(dλdτj)1λ=iω,τ=τj<0(j=1,2), subcritical Hopf bifurcation will appear instead of supercritical Hopf bifurcation. We do not provide details here.

    In Subsection 4.3, we have established that the reduced system (4.1) possesses a family of periodic solutions bifurcating from D at the different critical values of delay parameters τ1 and τ2. In this subsection, by using the center manifold theorem and normal form theory [57], the properties of the Hopf bifurcation at the critical value τ1 are determined. Throughout the section, we consider that τ2<τ1, where τ2(0,τ2j). For detailed discussion related to properties of the Hopf bifurcation (direction and stability of Hopf bifurcation), one may refer to Appendix B.

    Remark 4.4. The disappearance or appearance of a periodic orbit through a local change in the stability properties of a steady point is called Hopf bifurcation. There are two types of Hopf bifurcation, the bifurcation is known as supercritical if the bifurcated periodic solutions are stable and subcritical if they are unstable. The parameter μ2 determines the directions of the Hopf bifurcation: if μ2>0(μ2<0), then the Hopf bifurcation is supercritical (subcritical).

    Theorem 4.11. If Re{c1(0)}<0(Re{c1(0)}>0), then the reduced system (4.1) undergoes a supercritical (subcritical) Hopf bifurcation at D when τ1 crosses its critical value τ1. Furthermore, the bifurcated periodic solutions occurring through Hopf bifurcations are orbitally asymptotically stable on the center manifold if Re{c1(0)}<0 and unstable if Re{c1(0)}>0.

    In this subsection, we investigate the global continuation of periodic solutions bifurcating from D for system (2.3) for fixed τ2 in the interval (0,τ2j). Throughout this subsection, we follow the notations in Wu [59]. For simplification, we denote τ=τ1. Let z(t)=(z1(t),z2(t),z3(t))=(S(t),E(t),I(t)), then system (2.3) can be rewritten as

    ˙z=F(zt,τ,p), (4.4)

    where zt(θ)=(z1(t+θ),z2(t+θ),z3(t+θ))T=(z1t(θ),z2t(θ),z3t(θ))TC([τ,0],R3). It is clear that if R0>1, then system (2.3) has a disease-free equilibrium (D0) and an endemic equilibrium (D). Following Wu [59], we define

    X=C([τ,0],R3),=Cl{(z,τ,p)X×R×R+;zisanonconstantperiodicsolutionof(4.4)},M={(ˉz,τ,p);F(ˉz,τ,p)=0}.

    Lemma 4.12. Assume that (ˉz,τ,p) is an isolated center satisfying (A1-A4) in [59]. Denote by l(ˉz,τ,p) the connected component of (ˉz,τ,p) in . Then either

    (i) l(ˉz,τ,p) is unbounded, or

    (ii) l(ˉz,τ,p) is bounded, l(ˉz,τ,p)M is finite, and (ˉz,τ,p)l(ˉz,τ,p)Mγm(ˉz,τ,p)=0 for all m=1,2,3,, where γm(ˉz,τ,p) is the mth crossing number of (ˉz,τ,p), if mJ(ˉz,τ,p), or it is zero otherwise.

    It is well recognized that if the condition (ii) of above lemma is not true, then l(ˉz,τ,p) is unbounded. Thus, if the projections of l(ˉz,τ,p) onto p-space and onto z-space are bounded, then the projection onto τ-space is unbounded. Further, if we can show that the projection of l(ˉz,τ,p) onto τ-space is away from zero, then the projection of l(ˉz,τ,p) onto τ-space must include interval [τ,+). Following this concept, we can prove our consequences on the global continuation of local Hopf bifurcation.

    Lemma 4.13. If R0>1, then all non-constant periodic solutions of system (2.3) with initial conditions (2.4) are uniformly bounded.

    Lemma 4.14. If R0>1, then there does not exist any non-constant periodic solution of system (2.3) with period τ.

    Proof. Suppose there exists a non-constant periodic solution of system (2.3) with period τ. Then system (2.2) also has a non-constant periodic solution. Both systems (2.2) and (2.3) have same equilibria, i.e., D0 and D. Note that E-axis and I-axis are the invariable manifolds of system (2.2) and the orbits of system (2.2) do not cross each other. Thus, there is no solution that intersects the coordinate axis.

    On the other hand, if system (2.2) has a periodic solution, then there must be an equilibrium in its interior [41] and D0 is located on the coordinate axis. Thus, it can be concluded that the periodic orbit of system (2.2) must lie in the first octant. From Theorem 3.4, the positive equilibrium is globally asymptotically stable in R4+, thus, there is no periodic orbit in the first quadrant. This completes the proof.

    Theorem 4.15. Let ω4 and τ1j(j=0,1,2,) be defined in case 5 in subsection 4.3. Then for each τ>τ1j(j1), system (2.3) has at least j+1 periodic solutions.

    Proof. It is sufficient to prove that the projection of l(D,τ1j,2π/ω4) onto τ-space is [τ,+) for each j>0, where ˉττ1j. The characteristic matrix of (4.4) at positive equilibrium D takes the following form

    Δ(D,τ,p)(λ)=(J11λ0J12+J18eλτJ13J14λJ15+J19eλτ0J16J17+J20eλτ2λ), (4.5)

    where J1i,i=1,2,3,4,5,6,7,8,9 and J20 are defined in subsection 4.3. From the discussion of local Hopf bifurcation, it is easy to verify that (D,τ1j,2π/ω4),j=1,2,... are isolated centers. There exists ϵ>0,δ>0 and a smooth curve λ:(τ1jδ,τ1j+δ)C, such that det(Δ(λ(τ)))=0,|λ(τ)ω4|<ϵ for all τ[τ1jδ,τ1j+δ] and

    λ(τ1j)=iω4,dRe(λ(τ))dτ|τ=τ1j>0.

    Let

    Ωϵ,2π/ω4={(η,p):0<η<ϵ,|p2πω4|<ϵ}.

    It is easy to verify that on [τ1jδ,τ1j+δ]×Ωϵ,2π/ω4,

    det(Δ(D,τ,p)(η+2πpi))=0

    if and only if η=0,τ=τ1j,p=2π/ω4. Thus, the hypotheses (A1-A4) in [59] are satisfied. Moreover, if we define

    H±(D,τ1j,2πω4)(η,p)=det(Δ(D,τ1j±δ,p)(η+2πpi)),

    then we have the crossing number of isolated center (D,τ1j,2πω4) as follows:

    γ(D,τ1j,2πω4)=degB(H(D,τ1j,2πω4),Ωϵ,2π/ω4)degB(H+(D,τ1j,2πω4),Ωϵ,2π/ω4)=1.

    By Theorem 3.2 in Wu [59], we conclude that the connected l(D,τ1j,2π/ω4) through (D,τ1j,2π/ω4) in is nonempty. We have

    (ˉz,τ,p)l(D,τ1j,2π/ω4)γ(ˉz,τ,p)<0.

    Hence l(D,τ1j,2π/ω4) is unbounded.

    From Eq. (.34), we see that, for j1,2π/ω4<τ1j. Then, we are in a position to show that the projection of l(D,τ1j,2π/ω4) onto τ-space is [ˉτ,+), where ˉτ<τ1j. Clearly, it follows from the proof of Lemma 4.14 that system (2.3) with τ=0 has no non-constant periodic solution. Hence, the projection of l(D,τ1j,2π/ω4) onto τ-space is away from zero.

    For a contradiction, we suppose that the projection of l(D,τ1j,2π/ω4) onto τ-space is bounded. This means that the projection of l(D,τ1j,2π/ω4) onto τ-space is included in an interval (0,τ). Note that 2π/ω4<τ1j and applying Lemma 4.14, we have p<τ for (z,τ,p) belonging to l(D,τ1j,2π/ω4). This implies that the projection of the connected component l(D,τ1j,2π/ω4) onto p-space is bounded. In addition, from Lemma 4.13, we obtain that the projection of l(D,τ1j,2π/ω4) onto z-space is bounded if the projection of l(D,τ1j,2π/ω4) onto τ-space is bounded. Thus, the connected component l(D,τ1j,2π/ω4) crossing through (D,τ1j,2π/ω4) is bounded, which is a contradiction. This implies that the projection of l(D,τ1j,2π/ω4) onto τ-space is [ˉτ,+) for each j1, where ˉτ<τ1j. This completes the proof.

    In this subsection, we determine the length of delay to preserve the stability of system (4.1) by the Nyquist criterion [55]. We consider the system (4.1) and space of all real-valued continuous functions on [τ,+) satisfying the initial conditions on [τ,0]. We use the transformations S=S+u1,E=E+u2,I=I+u3 and obtain the following linearized system:

    du1dt=a11u1+a12u3+a13u3(tτ1),du2dt=a21u1+a22u2+a23u3+a24u3(tτ1),du3dt=a31u2+a32u3+a33u3(tτ2), (4.6)

    where

    a11=(μ+βIf(I)),a12=βSf(I),a13=βSIf(I)f(I)2,a21=βIf(I),a22=(μ+q),a23=βSf(I),a24=βSIf(I)f(I)2,a31=q,a32=(μ+δ+γg(I)),a33=γg(I)(g(I))2.

    Now taking Laplace transform of system (4.6), we obtain

    (sa11)ˉu1(s)=u1(0)+a12ˉu3(s)+a13esτ1ˉu3(s)+a13esτ1K1(s),(sa22)ˉu2(s)=u2(0)+a21ˉu1(s)+a23ˉu3(s)+a24esτ1ˉu3(s)+a24esτ1K1(s),(sa32a33esτ2)ˉu3(s)=u3(0)+a31ˉu2(s)+a33esτ2K2(s), (4.7)

    where K1(s)=0τ1esτ1u3(t)dt, K2(s)=0τ2esτ2u3(t)dt, and ˉui(s) are the Laplace transforms of ui(t),i=1,2,3, respectively. The characteristic equation of Eq. (4.7) is

    Ψ(s)=s3+P11s2+P12s+P13+(P14s+P15)esτ1+(P16s2+P17s+P18)esτ2=0, (4.8)

    where P1i,i=1,2,3,4,5,6,7,8 are given in subsection 4.3. The conditions for local asymptotic stability of D are given by

    Re(Ψ(iη0))=0, (4.9)
    Im(Ψ(iη0))>0, (4.10)

    where η0 is the smallest positive root of Eq. (4.9).

    We have already revealed that D is locally asymptotically stable when τ1=τ2=0. Hence, by continuity, all eigenvalues will have negative real parts for sufficiently small τ1>0,τ2>0 provided one can assure that no eigenvalues with positive real parts bifurcates from infinity as τ1 and τ2 increase from zero. This can be proved by Butler's lemma [56]. Now, we consider the conditions of Case 5, i.e., τ1>0 and τ2 is in stable interval. Eq. (4.9) and (4.10) gives

    P11η20P13(P18P16η20)cos(η0τ2)P17η0sin(η0τ2)=P15cos(η0τ1)+P14η0sin(η0τ1), (4.11)
    P12η0+η30P17η0cos(η0τ2)+(P18P16η20)sin(η0τ2)>P14η0cos(η0τ1)P15sin(η0τ1). (4.12)

    If Eq. (4.11) and (4.12) hold simultaneously, these are sufficient conditions to assure stability. We shall use them to obtain an estimation of the length of delay. Our goal is to discover an upper bound η+ on η0 free of τ1 and τ2, and then estimate τ1 so that Eq. (4.12) holds for all values of η, 0ηη+ and hence in particular at η=η0. From (4.11), we obtain

    P11η20=P13+(P18P16η20)cos(η0τ2)+P17η0sin(η0τ2)+P15cos(η0τ1)+P14η0sin(η0τ1). (4.13)

    Maximize P13+(P18P16η20)cos(η0τ2)+P17η0sin(η0τ2)+P15cos(η0τ1)+P14η0sin(η0τ1) subject to |sin(η0τ1)|1,|sin(η0τ2)|1,|cos(η0τ1)|1,|cos(η0τ2)|1, we obtain

    P11η20|P13|+|(P18P16η20)|+|P17|η0+|P15|+|P14|η0. (4.14)

    Assume that Eq. (4.14) has a positive root such η+, then obviously from inequality (4.14), we have η0η+.

    Now, rearranging the inequality (4.12), using Eq. (4.11), and assuming |sin(η0τ2)|1, |cos(η0τ2)|1, we obtain

    (P14P15P11)η0[cos(η0τ1)1](P15+P14P11)η20sin(η0τ1)<|P13P11P12|η0+|P17+P18P16η20P11|η0+|P18(P17P11P16)η20|(P14P15P11). (4.15)

    Using the bounds, we obtain

    (P14P15P11)η0[1cos(η0τ1)]=(P14P15P11)η02sin2(η0τ12)12|(P14P15P11)η+|η+2τ21,

    and

    (P15+P14P11)η20sin(η0τ1)|(P15+P14P11)|η+3τ1.

    Hence, from inequality (4.15), we obtain

    L11τ21+L12τ1<L13, (4.16)

    where

    L11=12|(P14P15P11)η+|η+2,L12=|(P15+P14P11)|η+3,L13=|P13P11P12|η0+|P17+P18P16η20P11|η0+|P18(P17P11P16)η20|(P14P15P11).

    Hence, if

    τ+1=12L11(L12+L212+4L11L13) (4.17)

    then the stability is preserved for 0τ1<τ+1.

    Similarly, one can follow the above calculation to determine the length of delay τ2 where the stability is preserved, taking τ1 in its stable interval.

    To study stability switching, we need to seek purely imaginary characteristic roots. For this, we assume λ=iω(ω>0) and substituting this in Eq. (4.3), we obtain

    m1(iω)+m2(iω)eiωτ1+m3(iω)eiωτ2, (4.18)

    with

    m1(iω)=(iω)3+P11(iω)2+P12(iω)+P13,m2(iω)=P14(iω)+P15,m3(iω)=P16(iω)2+P17(iω)+P18.

    Since |eiωτ2|=1, we have

    |m1+m2eiωτ1|=|m3|, (4.19)

    which is equivalent to

    (m1+m2eiωτ1)(ˉm1+ˉm2eiωτ1)=m3ˉm3.

    After simplification, we obtain

    |m1|2+|m2|2+2Re(m1ˉm2)cos(ωτ1)2Im(m1ˉm2)sin(ωτ1)=|m3|2.

    Thus,

    |m3|2|m1|2|m2|2=2K1(ω)cos(ωτ1)2K2(ω)sin(ωτ1), (4.20)

    where K1(ω)=Re(m1ˉm2) and K2(ω)=Im(m1ˉm2).

    If there is some ω such that K1(ω)2+K2(ω)2=0, then

    K1(ω)=K2(ω)=0m1ˉm2=0. (4.21)

    The right hand side of Eq. (4.20) is zero with any τ1, and

    |m3|2=|m1|2+|m2|2. (4.22)

    Therefore, if there is an ω such that both (4.21) and (4.22) are satisfied, then all τ1R+ are solution of (4.19).

    If K1(ω)2+K2(ω)2>0, then there exists some continuous function χ1(ω) such that

    K1(ω)=K1(ω)2+K2(ω)2cos(χ(ω)),K2(ω)=K1(ω)2+K2(ω)2sin(χ(ω)).

    Indeed,

    χ1(ω)=arg(m1ˉm2).

    Therefore, Eq. (4.20) becomes

    |m3|2|m1|2|m2|2=2K1(ω)2+K2(ω)2cos(χ1(ω)+ωτ1). (4.23)

    Obviously, a sufficient and necessary condition for the existence of τ1R+ satisfying the above equation is

    ||m3|2|m1|2|m2|2|2K21+K22. (4.24)

    Denote Ω1 to be ωR+ satisfying (4.24). We note that (4.24) includes the case K21+K22=0 which leads to (4.21) and (4.22).

    Let

    cos(Θ1(ω))=|m3|2|m1|2|m2|22K21+K22,Θ1[0,π].

    We have

    τ±1,k1(ω)=±Θ1(ω)χ1(ω)+2k1πω,k1Z. (4.25)

    By substituting the value of τ1(ω) given by (4.25), into (4.18) and we obtain an explicit formula for τ2(ω) unconditionally with each ωΩ1, i.e.,

    τ±2,k2(ω)=1ωarg{m3m1+m2eiωτ±1}+2k2π,k2Z. (4.26)

    Hence, the stability crossing curves are

    Γ={(τ±1,k1(ω),τ±2,k2(ω))R2+:ωΩ1,k1,k2Z}. (4.27)

    Another way to find τ2 is similarly to the analysis of τ1, which gives

    τ±2,k2(ω)=±Θ2(ω)χ2(ω)+2k2πω,k2Z, (4.28)

    where

    cos(Θ2(ω))=|m2|2|m1|2|m3|22K23+K24,Θ2[0,π],K3(ω)=K3(ω)2+K4(ω)2cos(χ2(ω)),K4(ω)=K3(ω)2+K4(ω)2sin(χ2(ω)),K3(ω)=Re(m1ˉm3),K4(ω)=Im(m1ˉm3),

    with the condition on ω:

    ||m2|2|m1|2|m3|2|2K23+K24, (4.29)

    which defines a region Ω2.

    By squaring both sides of the conditions (4.24) and (4.29), one can observe that (4.24) and (4.29) are equivalent. Hence

    Ω=Ω1=Ω2,

    and Ω is called the crossing set.

    Lemma 4.16. Ω consists of a finite number of intervals of finite length.

    Proof. The theoretical proof of this lemma can easily be established similarly as Lemma 3.2 of [64]. Numerically, we show that F(ω)=(|m3|2|m1|2|m2|2)24(K21+K22) has finite number of roots on R+ in Figure 21. If F(0)>0, then F(ω) has roots 0<a1<b1<a2<b2<...<aN<bN<+ and

    Ω=Nn=1Ωn,Ωn=[an,bn].

    If F(0)<0, then F(ω) has roots 0<b1<a2<b2<...<aN<bN<+ and

    Ω=Nn=1Ωn,Ω1=(0,b1],Ωn=[an,bn](n2).

    For any Ωn, we have a restriction on the range of χi(ω),i=1,2. We require χi(ω) to be the smallest continuous branch with the property that there exists an ωiΩn, such that

    χi(ωi)>0.

    Therefore, ki has a lower bound, denoted by Mi,n.

    By (4.26), one can obtain either τ+2,k2 or τ2,k2 (but not both) for a given τ+1,k1, and similarly for τ1,k1. By knotty computation, one can verify that when τ1=τ+1,k1(ω), we have τ2=τ2,k2(ω), and when τ1=τ1,k1(ω), we have τ2=τ+2,k2(ω). Therefore

    Γ=n=1,2,...Nk1M1,n,M1,n+1,...k2M2,n,M2,n+1,...Γ±nk1,k2R2+, (4.30)
    Γ±nk1,k2={(±Θ1(ω)χ1(ω)+2k1πω,Θ2(ω)χ2(ω)+2k2πω):ωΩn}. (4.31)

    Γ defined in (4.30) is the set of all stability switching curves on the (τ1,τ2)-plane and Γ±nk1,k2 is continuous in R2.

    For the crossing directions, we directly use Theorem 4.1 of [64] which is shown in Figure 22.

    In this section, we prove the persistence result for both systems (2.2) and (2.3). Biologically, persistence means the survival of all populations in the future time. Mathematically, persistence means positive solutions do not have omega limit points on the boundary of the non-negative set. A population x(t) is called uniformly persistent if there is an δ>0, free of x(0)>0 such that limtx(t)>δ. A system is called uniformly persistent if each component persists uniformly.

    Theorem 5.1. Both systems (2.2) and (2.3) are uniformly persistent in the interior of whenever R0>1.

    Proof. Refer the Appendix C.

    In this section, we present some numerical evaluations to verify our theoretical results and to exhibit different dynamics of systems (2.2) and (2.3).

    First, we numerically validate the theoretical results obtained for non-delayed system (2.2). We consider f(I)=1+α1I2 and g(I)=1+α2I2 and set the following parameter values (there may exist another set of parameter values):

    A=5,μ=0.5,β=0.2,δ=0.05,q=0.2,γ=0.2,α1=0.02,α2=0.02. (6.1)

    In this way, the system (2.2) can be rewritten as

    dSdt=50.5S0.2I1+0.02I2S,dEdt=0.2I1+0.02I2S(0.5+0.2)E,dIdt=0.2E(0.5+0.05)I0.2I1+0.02I2,dRdt=0.2I1+0.02I20.5R. (6.2)

    In this case, R0=0.7619<1 and the system (6.2) has a disease-free equilibrium D0(10.00,0,0,0), which is globally asymptotically stable as shown in Figure 5(a). We observe that curve S starts up very quickly, attains its maximum value and stabilize. The curves E,I,R go down to zero.

    Figure 5.  The graphs show the stability behavior of disease free equilibrium (D0) and endemic equilibrium (D) of system (2.2). Panel (a) shows that D0 is stable when R0<1 and Panel (b) shows that D is stable when R0>1.

    Further, we increase the incidence rate as β=0.5. By simple calculation, we obtain R0=1.9048>1 and endemic equilibrium D(5.313,3.348,0.896,0.353), which is globally asymptotically stable as shown in Figure 5(b). We observe that curve S starts up very quickly, attains its maximum value and then goes down and stabilize at a certain value. The curves E,I,R go up to a certain value and stabilize. We plot a transcritical bifurcation diagram in Figure 6.

    Figure 6.  The graph demonstrates the transcritical bifurcation with respect to β. The blue line depicts the stable equilibrium points and red line depicts the unstable equilibrium. Before a critical value of β=β, D0 is locally asymptotically stable and after a critical value of β=β, D0 is unstable and D exists and locally asymptotically stable.

    In Figure 7, we show that by increasing the value of recovery rate (γ), the level of infected population decreases and by increasing the value of contact rate (β), the level of infected population increases. From this observation, it could also be concluded that if we increase the recovery rate (by increasing the availability of resources for treatment), the level of infected population could be reduced. Similarly, if we decrease the contact rate (by increasing the intervention strategies), the level of infected population could be reduced. In Figure 8, we plot the infected population with respect to time for different values of α1 and α2. We observe that for greater values of α1, the trajectory of the infected population settles down at a lower level, which means by imposing more intervention, the infection level can be reduced. For higher values of α2, the infected population settles down at a high level, which means the level of infection increases with the limitation of treatment. In Figure 9, we show that the value of R0 with respect to β and γ. Here we observe that β increases the value of R0 and γ decreases the value of R0, which means the disease can be extinct by reducing the value of β (by reducing contact between susceptible and infected individuals) and by increasing the value of γ (by increasing the treatment). From Figure 9, it can also be observed that β has more influence than γ on R0.

    Figure 7.  The graphs show the variation of infected population with respect to recovery rate (γ) and contact rate (β). Panel (a) shows that the level of infected population decreases as γ increases. Panel (b) shows that the level of infected population increases as β increases.
    Figure 8.  The graphs depict the variation of infected population with respect to measure of interventions (α1) and limitation of resources for treatment (α2). Panel (a) shows that the level of infected population decreases as α1 increases. Panel (b) shows that the level of infected population increases as α2 increases.
    Figure 9.  Surface plot shows the impact of β and γ on R0.

    In Figure 10(a)(b), 11(a)(b) and 12(a)(b), we show the combined effect of various parameters on endemic level. Figure 10(a) represents the variation of endemic level with respect to two parameters β and γ. We observe that by decreasing the value of contact rate β and by increasing the value of recovery rate (γ), endemic level could be reduced. Figure 10(b) represents the variation of endemic level with respect to two parameters β and α1. We observe that the parameters β and α1 have equal impact on endemic level. From this observation, it can be concluded that by increasing the interventions and reducing the contact rate between susceptible and infected individuals, endemic level can be reduced. Figure 11(a) shows that by increasing the value of recovery rate γ and increasing the strength of interventions (α1), number of infected individuals could be reduced. Figure 11(b) demonstrates that by decreasing the value of contact rate β and reducing the limitation on resources availability (α2), number of infected individuals could be reduced. Figure 12(a) depicts the variation of endemic level with respect to two parameters γ and α2. We note that by increasing the recovery rate (γ) and by reducing the limitation on resources availability (α2), endemic level can be reduced. Figure 12(b) declares that number of infected individuals could be reduced by increasing the value of α1 (by increasing the strength of interventions) and decreasing the value of α2 (by decreasing the limitation on resources availability). These results are suitable in real life when the disease spread fast and the infection level is high, for example, COVID-19, H1N1 influenza, SARS. Our study suggests that treatment resources should be available with intervention strategies.

    Figure 10.  The graphs show the impact of different parameters on prevalence. Panel (a) shows the variation of infected individuals with respect to β and γ. Panel (b) shows the variation of infected individuals with respect to β and α1.
    Figure 11.  The graphs show the impact of different parameters on prevalence. Panel (a) shows the variation of infected individuals with respect to γ and α1. Panel (b) shows the variation of infected individuals with respect to β and α2.
    Figure 12.  The graphs show the impact of different parameters on prevalence. Panel (a) shows the variation of infected individuals with respect to γ and α2. Panel (b) shows the variation of infected individuals with respect to α1 and α2.

    In this section, we perform some numerical simulations to verify our theoretical results of system (2.3) in Section 3. We set the parameter values as

    A=500,μ=0.7,β=0.9,δ=0.05,q=0.2,γ=1.45,α1=0.02,α2=2×1012, (6.3)

    and then the system (2.3) can be rewritten as

    dSdt=5000.7S0.9I1+0.02I(tτ1)2S,dEdt=0.9I1+0.02I(tτ1)2S(0.7+0.2)E,dIdt=0.2E(0.7+0.05)I1.45I(tτ2)1+2×1012I(tτ2)2,dRdt=1.45I(tτ2)1+2×1012I(tτ2)20.7R. (6.4)

    By simple calculation, we obtain that R0=64.9351 and the endemic equilibrium D(248, 362, 32.93, 68.12).

    For τ2=0,τ1>0, by simple calculations, we obtain the critical value of τ1=3.2. By Theorem 4.4, we can deduce that D is locally asymptotically stable when τ1=[0,3.2) and a supercritical Hopf bifurcation occur at the critical value τ1=3.2. Figure 13 also depicts that D is locally asymptotically stable for τ1=2<τ1=3.2 and bifurcated periodic solutions feasible for τ1=10>τ1=3.2.

    Figure 13.  The graphs show the stability behavior of system (2.3) for τ1=2<τ1=3.2 and oscillatory behavior for τ1=10>τ1=3.2 when τ2=0.

    For τ1=0,τ2>0, we obtain the critical value of τ2=3.74. By Theorem 4.6, we can deduce that D is locally asymptotically stable when τ2=[0,3.74) and a supercritical Hopf bifurcation occur at the critical value τ2=3.74. It has been presented in Figure 14 that D is locally asymptotically stable for τ2=1<τ2=3.74 and bifurcated periodic solutions feasible for τ2=10>τ1=3.74.

    Figure 14.  The graphs show the stability behavior of system (2.3) for τ2=1<τ2=3.74 and oscillatory behavior for τ2=10>τ2=3.74 when τ1=0.

    For τ1=τ2=τ>0, we obtain the critical value of τ=0.63. By Theorem 4.8, we can deduce that D is locally asymptotically stable when τ=[0,0.63) and a supercritical Hopf bifurcation occur at the critical value τ=0.63. It has been shown in Figure 15 that D is locally asymptotically stable for τ=0.5<τ=0.63 and bifurcated periodic solutions feasible for τ=0.65>τ=0.63.

    Figure 15.  The graphs show the stability behavior of system (2.3) for τ=0.5<τ=0.63 and oscillatory behavior for τ=0.65>τ=0.63 when τ1=τ2=τ.

    For τ1>0,τ2(0,3.74], by simple calculations we obtain the critical value of τ1=0.05. By Theorem 4.10, we can deduce that D is locally asymptotically stable when τ1=[0,0.05) and a supercritical Hopf bifurcation occur at the critical value τ1=0.05. Figure 16 illustrates that D is locally asymptotically stable for τ1=0.04<τ1=0.05 and bifurcated periodic solutions feasible for τ1=0.16>τ1=0.05.

    Figure 16.  The graphs show the stability behavior of system (2.3) for τ1=0.04<τ1=0.05 and oscillatory behavior for τ1=0.16>τ1=0.05 when τ2=1.5<τ2.

    Furthermore, we plot bifurcation diagrams using the time delays as bifurcation parameters for different cases in Figure 17 and obtain that D is locally asymptotically stable before the critical values of delays and bifurcated periodic solutions occur after the critical values of delays. We also plot global branches of Hopf bifurcation in Figure 18 and obtain the global existence of periodic solutions.

    Figure 17.  The graphs demonstrate the bifurcation diagrams of system (2.3) with respect to time delays. Panel (a) depicts the infected population with respect to τ1 when τ2=0. Panel (b) depicts the infected population with respect to τ2 when τ1=0. Panel (c) depicts the infected population with respect to τ when τ2=τ1=τ. Panel (d) depicts the infected population with respect to τ1 when τ2=1.5<τ2.
    Figure 18.  The graphs demonstrate the global Hopf branches of system (2.3) with respect to time delays. Panel (a) shows the global Hopf branches when τ1=τ1=3.2 and τ2=0. Panel (b) shows the global Hopf branches when τ2=τ2=3.74 and τ2=0. Panel (c) shows the global Hopf branches when τ=τ=0.63 and τ2=τ1=τ. Panel (d) shows the global Hopf branches when τ1=τ1=0.05 and τ2=1.5<τ2.

    Figure 19 represents the changes in critical value of τ1 with respect to β and α1. The critical value of τ1 decreases with an increase in β while increases with an increase in α1. Figure 20 represents the changes in critical value of τ2 with respect to γ and α2. The critical value of τ2 increases with an increase in γ while decreases with an increase in α2. At the same time, we give the variation values of τ1 and τ2 with respect to parameters β,α1,γ and α2 in Table 1 and 2. These results indicate that these parameters also significantly influence the dynamics of the system with time delays such as critical value of bifurcation parameters.

    Figure 19.  The graphs depict the relation of critical value of delay τ1 with β and α1. Other parameter values remain same as (6.3). Panel (a) depicts the changes in critical value of τ1 with respect to β. Panel (b) depicts the changes in critical value of τ1 with respect to α1.
    Figure 20.  The graphs depict the relation of critical value of delay τ2 with γ and α2. Other parameter values remain same as (6.3). Panel (a) depicts the changes in critical value of τ2 with respect to γ. Panel (b) depicts the changes in critical value of τ2 with respect to α2.
    Table 1.  The changes in critical value of τ1 with respect to β and α1.
    β Critical value of τ1 α1 Critical value of τ1
    0.3 3.32 0.1 3.76
    0.5 3.25 0.2 3.86
    0.7 3.22 0.3 4.28
    0.9 3.20 0.4 4.50

     | Show Table
    DownLoad: CSV
    Table 2.  The changes in critical value of τ2 with respect to γ and α2.
    γ Critical value of τ2 α2 Critical value of τ2
    1.4 0.474 3×107 0.5426
    1.5 0.601 5.3×106 0.5093
    1.6 0.676 1.13×105 0.4625
    1.7 0.720 1.74×105 0.4057

     | Show Table
    DownLoad: CSV

    For stability switching curves, we obtain that F(0)=158.96<0 and F(ω) has three roots b1=0.01575,a2=0.02216, and b2=0.0667 (see Figure 21). Thus

    Ω=2n=1Ωn,Ω1=(0,b1],Ω2=[a2,b2].
    Figure 21.  Graph of F(ω). Other parameter values remain same as (6.3).

    We have

    δa1=0,δa2=1,δb1=1,δb2=1.

    Since (δa1,δa2)(δb1,δb2), therefore by Theorem 3.1 of [64], Γ is a set of continuous curves with their two end points on the axises (class Ⅱ) (see Figure 22).

    Figure 22.  Plot of stability switching curves. Arrows are used to depict the crossing directions, i.e., the region on the end of an arrow has two more characteristic roots with positive real parts. Other parameter values remain same as (6.3).

    We know that for system (2.3), the interior equilibrium is stable when τ1=τ2=0, i.e., no characteristic roots have positive real parts. Hence from Figure 22, the interior equilibrium is stable if and only if (τ1,τ2) is on the small bottom-left region of the (τ1,τ2)-plane. As τ1 and τ2 increase, there is a trend that there are more and more characteristic roots with positive real parts.

    Recently, a novel coronavirus (namely COVID-19) spread worldwide. Several countries are in the list of top countries, including the USA, India, Brazil, Russia [61]. But early on the disaster, Spain and Italy emerged as a major global coronavirus hotspot. Both countries have implemented strict lockdown from March 2020. Restrictions started in Spain on March 14 and in Italy on March 9 [62]. Following this, the daily new cases are decreased in Spain and Italy after March 2020 [61,63]. In Spain, the maximum number of reported cases was observed on April 2, 2020, which implies the number of infections was on an upward trend till April 2, 2020. After that day, the number of daily cases was observed to decrease asymptotically after imposing interventions, which is the most significant interpretation of the mathematical model. A similar case can be seen in Italy, the maximum number of reported cases was observed on March 21, 2020, after which the number of daily cases was observed to decrease asymptotically. Without loss of generality, we have chosen the data of Spain and Italy during March 1, 2020 to May 31, 2020. We have plotted the bar graph of new daily cases in Spain and Italy from March 1, 2020 to May 31, 2020, i.e., the real cases for 92 days in Figure 23. We calibrate the system (2.3) to fit the new daily cases and cumulative cases of COVID-19 in Spain and Italy and estimate some parameters of the system (2.3) by the least square method using real data of Spain and Italy from March 1, 2020 to May 31, 2020, which is shown in Figure 25. The cumulative number of cases is represented by C(t), which is given as

    dCdt=qE, (7.1)
    Figure 23.  New daily cases of COVID-19 in Italy and Spain from from March 1, 2020 to May 31, 2020. Panel (a) shows the bar plot of data of COVID-19 in Italy. Panel (b) shows the bar plot of data of COVID-19 in Spain.
    Figure 24.  Outputs of the system (2.3) and new daily cases in Italy and Spain. Red stars represent the data points and blue line represents the output of the system (2.3). Panel (a) shows the fitting of system (2.3) with respect to new daily cases in Italy. Panel (b) shows the fitting of system (2.3) with respect to new daily cases in Spain.
    Figure 25.  Outputs of the system (2.3) and cumulative cases in Italy and Spain. Red stars represent the data points and blue line represents the output of the system (2.3). Panel (a) shows the fitting of system (2.3) with respect to cumulative cases in Italy. Panel (b) shows the fitting of system (2.3) with respect to cumulative cases in Spain.

    and the daily new cases are given by Cd(t), where

    Cd(t)=C(t)C(t1). (7.2)

    The demographic effects are not considered in this discussion because of the short epidemic time scale in comparison to demographic time scale, i.e., we take A=0,μ=0 here. Other estimated parameters and initial conditions for system (2.3) are given in Table 3 and Table 4, respectively.

    Table 3.  Estimated values of parameters of system (2.3) with respect to COVID-19 in Spain and Italy.
    Parameters Estimated values in Spain Estimated values in Italy Reference
    A 0 0 -
    μ 0 0 -
    q 0.07112 0.0665 Estimated
    β 3.1114×108 4.9092×108 Estimated
    γ 0.273 0.3026 Estimated
    δ 3.2720×104 2.0102×104 Estimated
    α1 1×108 2×108 Assumed
    α2 2×108 4×109 Assumed
    τ1 3.5 1 Estimated
    τ2 2 3 Estimated

     | Show Table
    DownLoad: CSV
    Table 4.  Initial population for system (2.3) with respect to COVID-19 in Spain and Italy.
    Initial population S(0) E(0) I(0) R(0)
    Spain 4.69×107 2.18×103 13 10
    Italy 6.04×107 3.10×102 561 33

     | Show Table
    DownLoad: CSV

    The basic reproduction number R0 can be explicitly calculated as R0=βNγ+δ. Based on the parameter values in Table 3, we calculated the value of R0 for both countries Spain and Italy are 5.3391 and 9.7921, respectively.

    This section aims to compare the outcomes of cumulative cases over time in Spain and Italy for different values of different parameters. Here, using the parameter values given in Table 3 as the baseline values, we have shown that how a model could be used to observe the effect of changing strength of interventions and time delays on epidemic dynamics. In particular, we illustrate the impact of the strength of interventions (α1), time delay in intervention (τ1), and time delay in recovery (τ2) on the early phase of pandemic COVID-19. In mathematical terms, we explore how our model dynamics depends on the parameters α1,τ1, and τ2. We compare the total cases generated by the model for different values of strength of interventions (α1) and time delays τ1, and τ2. We observe that if the strength of interventions and time delays in the model alter, total cases generated via models are different.

    From Figure 26(a), we observe that the strength of intervention (α1) could have reduced the total number of cases at the early stage of epidemic in Italy. We note that Italy had recorded 2.32×105 total number of cases by May 31, 2020. We analyze that if the strength of interventions (α1) had increased, the total cases of COVID-19 would have decreased than that the actual cases. Figure 26(a) shows that if the strength of interventions had increased by 50%(α1=3×108) from its baseline value (α1=2×108), then the cumulative cases would have decreased by 12%(2.04×105) and if the strength of interventions had increased by 100%(α1=4×108) from its baseline value, then the cumulative cases would have decreased by 18%(1.89×105). If the strength of interventions had decreased by 50%(α1=1×108) from its baseline value, then the cumulative cases would have increased by 34%(3.12×105). We observe that the strength of interventions would have been beneficial in the early phase of COVID-19 (measured in terms of total confirmed cases of COVID-19).

    Figure 26.  The plots show the dynamics of cumulative cases over time in Italy for different values of α1 and τ1. (a) Illustration of what would have happened if time delay τ1 had changed. (b) Illustration of what would have happened if the strength of intervention α1 had changed.

    We also illustrate that how time delays in intervention and recovery influence the outcomes of early cumulative cases in Italy. The results in Figure 26(b) show that the sooner the interventions are imposed, the greater reduction in the cumulative number of cases. We analyzed that if time delay τ1 increased by 2 times (τ1=2) by its baseline value (τ1=1), then cumulative number of cases would have been increased by 8%(2.52×105) and if τ1 increased by 3 times (τ1=3) by its baseline value, then cumulative number of cases would have been increased by 25%(2.91×105). Thus the timing of implementing the interventions is crucial and analyzing the impact of time delay in the system is significant. Imposing the interventions sufficiently early, may have the potential to reduce the total number of infected cases. Figure 27 shows that if the time delay (τ2) had increased by 3/2 times (τ2=3) by its baseline value (τ2=2) then number of cumulative cases would have been increased by 4% and if the time delay (τ2) had decreased by 2 times (τ2=4) by its baseline value (τ2=2) then number of cumulative cases would have been decreased by 2%.

    Figure 27.  The plot shows the dynamics of cumulative cases over time in Italy for different values of τ2.

    We find the same qualitative result (Figures 28 and 29) when we parameterize the model using the data of COVID-19 in Spain. These results also demonstrate that model system can produce observed disease dynamics, but total cases are different when strength of interventions and time delays are altered. From Figure 28(a), we observe that the strength of intervention (α1) would have reduced the total number of cases at the early stage of epidemic in Spain. We note that Spain had recorded 2.39×105 total number of cases by May 31, 2020. We analyze that if the strength of interventions (α1) had increased, the total cases of COVID-19 would have decreased. Figure 28(a) shows that if the strength of interventions had increased by 100%(α1=2×108) from its baseline value (α1=1×108), then the cumulative cases would have decreased by 25%(1.796×105) and if the strength of interventions had increased by 200%(α1=3×108) from its baseline value, then the cumulative cases would have decreased by 38%(1.48×105). If the strength of interventions had increased by 300%(α1=4×108) from its baseline value, then the cumulative cases would have decreased by 45%(1.30×105). Figure 28(b) illustrates that how time delay in intervention affects the outcomes of early cumulative cases in Spain. We analyzed that if time delay τ1 increased by 1.28 times (τ1=4.5) by its baseline value (τ1=3.5), then cumulative number of cases would have been increased by 21%(2.91×105) and if τ1 decreased by 1.4 times (τ1=2.5) by its baseline value, then cumulative number of cases would have been decreased by 7%(2.22×105). Figure 29 shows that if the time delay (τ2) had increased by 3/2 times (τ2=3) by its baseline value (τ2=2) then number of cumulative cases would have been increased by 5% and if the time delay (τ2) had decreased by 2 times (τ2=1) by its baseline value (τ2=2) then number of cumulative cases would have been decreased by 2%.

    Figure 28.  The plots show the dynamics of cumulative cases over time in Spain for different values of α1 and τ1.
    Figure 29.  The plot shows the comparison of the effect of τ2 on the cumulative cases over time in Spain.

    In the case of many infectious diseases, either we are unable to perform sufficient experiments, or a deeper understanding of disease dynamics is essentially required. Therefore, modeling of epidemics (mathematical and computational methods) is one of the important tools to understand the disease dynamics. Mathematical modeling along with precise computations gives new insights about different important issues of disease dynamics and their prevention, e.g., effects of different important parameters: delay, treatment, epidemic trends, intervention strategies, control, and spread of infections. Intervention strategies and treatment have great impression to control different contagious diseases. In the last few years, models accompanying the impact of intervention strategies on the spread of infectious diseases showed impressive popularity. But these models overlooked the impact of the availability of treatment resources on the dynamics of infectious disease. These models also do not involve different time delays, such as time delay in interventions and cure rate. However, in many biological systems, time delays are inevitably present, and several epidemic model systems have been explored incorporating the delays [35,46,47,49]. Here we put our efforts to understand the impact of intervention strategies and available resources for the treatment/control of infectious diseases, with special emphasis on the current pandemic COVID-19.

    In this work, we have studied an SEIR model system (2.2) with infection force under intervention strategies and recovery rate under the low availability of resources for treatment. Furthermore, we included the two time delays in system (2.2), first-time delay τ1 represents the latent period of the intervention strategies, and second-time delay τ2 is due to the period that is used to cure the infected individuals. For systems (2.2) and (2.3), we obtained the various theoretical and epidemiological findings.

    (ⅰ) We investigated that the system (2.2) has two equilibria, namely, disease-free equilibrium (D0) and endemic equilibrium (D). The existence of equilibria depends on the basic reproduction number (R0). For the non-delayed model, stability analysis for both disease-free and endemic equilibriums are performed. In particular, it has been shown that for disease spread, the basic reproduction number (R0) acts as a threshold value. When R0<1, only disease-free equilibrium (D0) exists and locally asymptotically stable. When R0>1, there exists two equilibria, disease-free equilibrium (D0) and endemic equilibrium (D).

    (ⅱ) System (2.2) undergoes a transcritical bifurcation at R0=1, which means disease die out whenever R0<1 and disease persists whenever R0>1. We have obtained that D0 is globally stable in its feasible region R0<1 and D is globally stable if R0>1. In numerical analysis, by choosing a particular combination of f(I) and g(I), we have performed the impact of various parameters in the disease dynamics (Figure 7, 8). We have shown the relation of the different parameters with R0 (Figure 9). We have also illustrated the combined effect of various parameters on the endemic level (Figure 10, 11, 12). Our observations revealed that increasing the interventions, disease level settle down at low level, but limitation on treatment sets the infected population at higher level. We additionally observed that R0 can be reduced by increasing interventions and decreasing the limitation on resources availability.

    (ⅲ) For delayed system (2.3), we have obtained that when R0<1, D0 is locally asymptotically stable with certain conditions (Theorem 4.1) when τ1>0 and τ2>0. Further, when R0>1, D0 remain unstable when τ1>0 and τ2>0. The system (2.3) seems sensitive to delay lengths. It is observed by both theoretical results and numerical simulations. When R0>1, the effect of two-time delay has been investigated by taking different cases and the threshold values of both the delays have been obtained in section 4.3. Analytically and numerically, we have shown that system (2.3) is stable when delay parameters are less than their threshold values (Figure 13, 14, 15, 16), which means the disease could be controlled easily. Whenever these parameters cross these threshold values, periodic solutions will occur through Hopf bifurcation. It can also be understood by the bifurcation diagram (Figure 17). In additional, the direction and stability of Hopf bifurcation are studied by normal form theory and center manifold theorem. We have also studied the global continuation of local Hopf bifurcation via both analytically and numerically (section 4.5 and Figure 18). We concluded that the time delays in the system (2.3) affect the structure of the solutions substantially. The time delays play the roles to destabilize the system. Indeed, it is evidenced by our theoretical results. The length of delay to preserve stability has also been estimated by employing the Nyquist criterion. More importantly, the stability crossing curves along with crossing directions [47,48,64] have been discussed in two-delay parameter space (refer to the stability switching curve plots in Figure 22). Our system (2.3) demonstrates rich dynamics, such as Hopf bifurcation, and is suitable for large population densities.

    (ⅳ) The dynamics of system (2.3) is also affected by the parameters β,α1,γ,α2. Numerical simulations indicate that the critical value of τ1 decreases with an increase in β (cf. Figure 19 and Table 1). Therefore, we conclude that the impact of an increase in transmission rate is to early the occurrence of Hopf bifurcation. Similarly, the critical value of τ1 increases with an increase in α1, which means the impact of increasing the value of the parameter α1 (imposing more interventions) is to postpone the occurrence of Hopf bifurcation. It can also be concluded that the epidemic cycle is more difficult to occur as the intervention strategies increases. On the other hand, the critical value of τ2 increases with an increase in γ (cf. Figure 20 and Table 2). Thus, the impact of an increase in recovery rate is to postpone the occurrence of Hopf bifurcation while the impact of the increasing the value of the parameter α2 (increasing limitations on treatment) is to early the occurrence of Hopf bifurcation because the critical value of τ2 decreases with increasing α2.

    (ⅴ) Furthermore, we have applied our model to a case study based on the data of COVID-19 in Spain and Italy. We estimated the parameters of system (2.3) by using the least-square method. It has also been depicted that if we could be able to increase the strength of interventions sufficiently, total number of cases in Spain and Italy would have reduced in early phase of COVID-19 (refer to Subsection 7.1). In both countries (Spain and Italy), the similar effects could also be achieved by reducing time delays in intervention and recovery. Further, it is expected that the similar scenarios could also be observed in many other European countries such as UK, Germany, and France. As such, the outcomes related to different significant values may differ from actual ones. Here we should note that many attributes of the system also depend on the published/collected data. The data have been collected for a short time span and do not have any future predictions. The quantitative results have been illustrated for the early phase of the pandemic COVID-19 in only two countries, Spain and Italy. Our model could also be applicable for other countries and also for long time periods.

    Our proposed model systems generalize the study [10,11] and become more realistic via investigation of recovery rate and delay parameters on the system dynamics. Treatment resources play a significant role in protecting the individuals against a particular disease and to control a contagious disease, for example, COVID-19. Thus, our results are the complement to the study by Zhou et al. [11]. The results found here could be favorable for measuring the effect of the intervention strategies and treatment in the control of infectious diseases. The existence of periodic solutions in epidemic models is significant and interesting both in mathematics and applications since it provides a satisfactory explanation for the breakout and recurrence of a disease, which may have profound implications for the prediction, prevention, and control of disease transmission. If the delay due to the latent period of the intervention strategies is large enough, it will lead to repetitive episodes of endemic and in this situation, the disease would be out of control. Our study suggests that imposing intervention strategies timely may be helpful in controlling the outbreak of disease. Further, our study also suggests that to control and predict the disease spread, we should trim the time delays. When the delay is not too large, ultimately, the epidemic disease will become endemic. Similarly, the time delay due to the period used to cure the infected individuals also leads to periodic solutions. Appropriately, we should truncate the delays in the model as much as possible so that we could control and predict the disease spread. The outcomes of our results in Section 7 emphasized the importance of the intervention strategies and time delays for the case of COVID-19. Policy makers should take care of reducing the time delays in such biological processes. These policies can be used by policymakers and other professionals to make decisions on how certain interventions such as lockdown, quarantine, isolation, and social distancing can effectively be implemented.

    By considering the incidence rate under intervention strategies and recovery rate under low resources availability treatment, the system (2.2) does not have complex dynamics. It will be intriguing to consider the incidence and recovery rate in which the infection and recovery function first increases and attain its maximum, and then eventually tends to a saturation level due to crowding effect and limitation of treatment, respectively. It will be insightful to compare the dynamics of these cases with our results as future work. In future studies, according to the actual situation of COVID-19 in Spain and Italy, the pre-stage exposed and post-stage exposed period [65] can be further considered in the model to analyze the impact of intervention and medical resources on COVID-19 prevention and control.

    The research work of Sarita Bugalia is supported by the Council of Scientific & Industrial Research (CSIR), India [File No. 09/1131(0025)/2018-EMR-I]. The research work of Jai Prakash Tripathi is supported by Science and Engineering Research Board (SERB), India [File No. ECR/2017/002786] and UGC-BSR Research Start-Up-Grant, India [No. F.30-356/2017(BSR)]. The research work of Hao Wang is partially supported by The Natural Sciences and Engineering Research Council of Canada (NSERC) Individual Discovery Grant RGPIN-2020-03911 and Discovery Accelerator Supplement Award RGPAS-2020-00090. The authors are grateful to the handling editor and reviewers for their helpful comments and suggestions that have improved the quality of the manuscript.

    All authors declare no conflicts of interest in this paper.

    Proof of Theorem 3.1: From the first equation of system (2.2), we have

    dSdtμSγIf(I)S,

    and

    S(t)S(0)e(μ+γIf(I)).

    Hence, if S(0)>0, then S(t)>0. In the similar fashion, one can also check the positivity of E(t),I(t), and R(t).

    Now, we shall prove the boundedness of all the solutions of system (2.2). For this, we define a function

    N(t)=S(t)+E(t)+I(t)+R(t).

    Taking the time derivative of N(t) and using (2.2), we obtain

    dNdt=AμSμEμIδIμR,AμN,

    which implies

    A(μ+δ)NdNdtAμN,

    and

    Aμ+δlim inftN(t)lim suptN(t)Aμ,Aμ+δN(t)Aμ.

    This implies that all the solutions of system (2.2) are bounded. Thus we proved the Theorem 3.1.

    Proof of Theorem 3.2: 1. The Jacobian matrix of right hand function in system (2.2) at D0 is given by

    JD0=[μ0βAμf(0)00(μ+q)βAμf(0)00q(μ+δ+γg(0))000γg(0)μ],

    which has two eigenvalues μ,μ (which are negative) and the other eigenvalues are the eigenvalues of the following matrix

    B=[(μ+q)βAμf(0)q(μ+δ+γg(0))].

    The trace and determinant of matrix B are given by tr(B)=(μ+δ+γg(0)+μ+q)<0 and det(B)=(μ+δ+γg(0))(μ+q)qβAμf(0). We observe that det(B)>0 if R0<1 and det(B)<0 if R0>1. Therefore, by Routh-Hurwitz criterion [50], it can be concluded that D0 is locally asymptotically stable when R0<1 and unstable when R0>1.

    2. When R0=1, i.e., β=β=μf(0)(μ+q)(δ+μ+γg(0))Aq, then

    J(D0,β)=[μ0(μ+q)(δ+μ+γg(0))q00(μ+q)(μ+q)(δ+μ+γg(0))q00q(μ+δ+γg(0))000γg(0)μ].

    One can easily observe that J(D0,β) has a simple zero eigenvalue and all other three eigenvalues are μ,μ,(2μ+q+δ+γg(0)), which are negative. Computing the right eigenvector w and left eigenvector v of JD0 with respect to zero eigenvalue, we have

    w1=(μ+q)(γ+g(0)(δ+μ))γq,w2=μ(γ+g(0)(δ+μ))γq,w3=μg(0)γ,w4=1,v1=0,v2=qq+μ,v3=1,v4=0.

    Now from Theorem 4.1 of [51], we need to calculate the bifurcation constants a and b. Denote S=x1,E=x2,I=x3, and R=x4. After some algebraic calculation, we obtain

    a=4k,i,j=1vkωiωj2fk(D0,β)xixj=μ2(μ+q)g(0)2Aqγ2(μ+γ+γg(0))22f(0)(μ+δ)f(0)(μg(0)γ)22μ2g(0)γ(f(0)f(0)g(0)g(0))<0,b=4k,i=1vkωi2fk(D0,β)xiβ=Aqμg(0)μγ(q+μ)f(0)>0.

    Therefore, from Theorem 4.1 of [51], the disease-free equilibrium (D0) changes its stability from stable to unstable at R0=1. Moreover, there exists a positive equilibrium as R0 crosses value 1, which is locally asymptotically stable. Therefore, system (2.2) undergoes a transcritical bifurcation at R0=1.

    Proof of Theorem 3.3: To prove the global stability of D0, we follow the approach given by Castillo-Chavez et al. [52]. We rewrite the system (2.2) as follows

    dXdt=F(X,V),dVdt=G(X,V),G(X,0)=0, (.1)

    where X=(S,R)R2 signifies the number of uninfected individuals and V=(E,I)R2 signifies the number of infected individuals. Disease-free equilibrium (D0) is globally stable if the following two conditions are fulfilled:

    (H1) For dXdt=F(X,V), X is globally asymptotically stable,

    (H2) G(X,V)=MVˆG(X,V),ˆG(X,V)>0 for (X,V),

    where M=DVG(X,0) is an M-matrix. For the system (2.2), we have

    F(X,0)=(ΛμS0). (.2)

    It is obvious that the equilibrium X=(Aμ,0) is globally asymptotically stable of system (.2). Further, for system (2.2), we obtain

    M=((μ+q)βAμf(0)q(μ+δ+γg(0))),ˆG(X,V)=(βIf(I)(AμS)0).

    It is clear that ˆG(X,V)0. Hence, D0 is globally stable.

    Proof of Theorem 3.4: Define the following Lyapunov function:

    L(S,E,I,R)=|SS|+|EE|+|II|+|RR|.

    Clearly, L(D)=0 and L(D)0 when DD. The upper right derivative of L(S,E,I,R) can be calculated as:

    D+L=sgn(SS)[(βSIf(I)βSIf(I))μ(SS)]+sgn(EE)[(μ+q)(EE)+(βSIf(I)βSIf(I))]+sgn(II)[q(EE)(μ+δ)(II)(γIg(I)γIg(I))]+sgn(RR)[(γIg(I)γIg(I))μ(RR)].

    In the above equation, there are 8 different types of situations depending on the size of S and S, E and E, I and I, R and R. It would be sufficient to show for S>S,E>E,I>I,R>R, similarly, one can also do for the other situations. Thus we obtain

    D+Lμ|SS|μ|EE|(μ+δ)|II|μ|RR|<μ|SS|μ|EE|μ|II|μ|RR|<μL.

    Integrating the above inequality from t0 to t both sides, we obtain

    L(t)+μtt0LdtL(t0)<+.

    Since S,E,I,R are bounded and their derivatives are also bounded, therefore, L is uniformly continuous on [t0,+). From Barbalat's Lemma [53], limt+L(t)=0. Hence D+L<μL<0, which implies D is globally stable.

    Proof of Theorem 4.1: The characteristic equation of the Jacobian of system (4.1) at D0 is given by

    det(K0+K1eλτ1+K2eλτ2λI)=0, (.3)

    where

    K0=(μ0βAμf(0)0(μ+q)βAμf(0)0q(μ+δ)),K1=(000000000),K2=(00000000γg(0)),

    and I is a 3×3 identity matrix. So the characteristic equation (.3) becomes in the following form

    (λ+μ){(λ+(μ+δ)+γg(0)eλτ2)(λ+μ+q)βAqμf(0)}=0. (.4)

    One eigenvalue is μ, which is negative and other eigenvalues are solution of the following equation

    p(λ)=(λ+(μ+δ)+γg(0)eλτ2)(λ+μ+q)βAqμf(0)=0 (.5)

    or

    λ2+V11λ+V12+(V13λ+V14)eλτ2=0,

    where

    V11=(2μ+q+δ),V12=(μ+q)(μ+δ)βAqμf(0),V13=γg(0),V14=(μ+q)γg(0).

    Let λ=iω,(ω>0) be a root of Eq. (.5). Putting λ=iω in Eq. (.5) and separating real and imaginary parts, we have

    ω2+V12=V14cos(ωτ2)V13ωsin(ωτ2),V11ω=V13ωcos(ωτ2)+V14sin(ωτ2).

    Squaring and adding above equations, we obtain

    ω4+(V211V2132V12)ω2+V212V214=0. (.6)

    Let ω2=v, then we have

    v2+(V211V2132V12)v+V212V214=0. (.7)

    By Descarte's rule of sign, if V211V2132V12>0 and V212V214>0, then Eq. (.6) has no positive roots. Hence all roots of Eq. (.6) have negative real parts.

    If τ2=0, then Eq. (.5) becomes in the following form

    λ2+(2μ+q+δ+γg(0))λ+(μ+δ+γg(0))(μ+q)βAqμf(0)=0,λ2+(2μ+q+δ+γg(0))λ+(μ+δ+γg(0))(μ+q)(1R0)=0.

    If R0<1 then by Descarte's rule of sign, above equation has all its roots with negative real parts. Hence D0 is locally asymptotically stable.

    From p(λ)=(λ+μ+q)(λ+(μ+δ)+γg(0)eλτ2)βAqμf(0), since limλ+p(λ)=+ and

    p(0)=(μ+δ+γg(0))(μ+q)βAqμf(0),=(μ+δ+γg(0))(μ+q)(1R0),<0,whenR0>1.

    Therefore, p(λ)=0 has a positive real root. Hence, D0 is unstable when R0>1.

    Proof of Lemma 4.3 and Theorem 4.4: If τ1>0,τ2=0, then the characteristic Eq. (4.3) becomes

    λ3+(P11+P16)λ2+(P12+P17)λ+(P13+P18)+(P14λ+P15)eλτ1=0. (.8)

    Let λ=iω1,(ω1>0) be a root of Eq. (.8). Putting λ=iω1 in Eq. (.8) and separating imaginary and real parts, we have

    P15cos(ω1τ1)+P14ω1sin(ω1τ1)=(P11+P16)ω21(P13+P18),P14ω1cos(ω1τ1)P15sin(ω1τ1)=ω31(P12+P17)ω1. (.9)

    Squaring and adding above equations, we obtain

    ω61+c11ω41+c12ω21+c13=0, (.10)

    where

    c11=(P11+P16)22(P12+P17),c12=(P12+P17)22(P11+P16)(P13+P18)P214,c13=(P13+P18)2P215.

    Let ω21=v1, then Eq. (.10) becomes

    v31+c11v21+c12v1+c13=0. (.11)

    For the sign of the roots of Eq. (.10), by Descartes rule of sign, we have the following cases:

    1. If c11>0,c12>0,c13>0, then Eq. (.10) has no positive roots. Hence all the roots of Eq. (.10) have negative real parts for τ1(0,+).

    2. If c13<0,c11>0,c12>0 or c13<0,c11>0,c12<0 or c13<0,c11<0,c12<0, then Eq. (.10) has a unique positive root ω1=v1. Eliminating sin(ω1τ1) from Eq. (.9), we obtain

    τ1j=1ω1cos1[P14ω14+((P11+P16)P15(P12+P17)P14)ω12P15(P13+P18)P215+(P14ω12)2+2jπ],j=0,1,2,. (.12)

    3. If c11>0, c12<0,c13>0 or c11<0, c12>0,c13>0 or c11<0, c12<0,c13>0, then Eq. (.10) has two positive roots ω±1=v±1. Eliminating sin(ω1τ1) from Eq. (.9), we obtain

    τ±1k=1ω±1cos1[P14ω±14+((P11+P16)P15(P12+P17)P14)ω±12P15(P13+P18)P215+(P14ω±12)2+2kπ],k=0,1,2,. (.13)

    4. If c11<0, c12>0,c13<0, then Eq. (.10) has three positive roots ω1,2,31=v1,2,31. Eliminating sin(ω1τ1) from Eq. (.9), we obtain

    τ1,2,31l=1ω1,2,31cos1[P14ω1,2,314+((P11+P16)P15(P12+P17)P14)ω1,2,312P15(P13+P18)P215+(P14ω1,2,312)2+2lπ],l=0,1,2,. (.14)

    Hence, Eq. (.8) has a pair of purely imaginary roots ±iω1 with τ1=τ1j,j=0,1,2,, a pair of purely imaginary roots ±iω±1 with τ1=τ±1k,k=0,1,2,, and a pair of purely imaginary roots ±iω1,2,31 with τ1=τ1,2,31l,l=0,1,2,. Let {τ1j}+j=0={τ1j}+j=0 such that τ10<τ11<τ12<<τ1j<, {τ1k}+k=0={τ±1k}+k=0 such that τ10<τ11<τ12<<τ1k<, where τ10=min{τ+10,τ10}, and {τ1l}+l=0={τ1,2,31l}+l=0 such that τ10<τ11<τ12<<τ1l<, where τ10=min{τ110,τ210,τ310}.

    Let λ(τ1)=α(τ1)+iω1(τ1) be a root of Eq. (.8) near τ1=τ1j and α(τ1j)=0, ω1(τ1j)=ω1, j=0,1,2,. Further, differentiating Eq. (.8) with respect to τ1 and substituting λ=iω1, τ1=τ1j, j=0,1,2,, we obtain

    Re(dλdτ1)1λ=iω1,τ1=τ1j=M11M13+M12M14M213+M214, (.15)

    where

    M11=P15ω1sin(τ1jω1)P14ω21cos(τ1jω1),M12=P14ω21sin(τ1jω1)+P15ω1cos(τ1jω1),M13=P14τ1jω1sin(τ1jω1)+P14cos(τ1jω1)P15τ1jcos(τ1jω1)+P12+P173ω21,M14=P15τ1jsin(τ1jω1)P14(sin(τ1jω1)+τ1jω1cos(τ1jω1))+2(P11+P16)ω1.

    Proof of Lemma 4.5 and Theorem 4.6: If τ1=0,τ2>0, then the characteristic Eq. (4.3) becomes

    λ3+P11λ2+(P12+P14)λ+(P13+P15)+(P16λ2+P17λ+P18)eλτ2=0. (.16)

    Let λ=iω2,(ω2>0) be a root of Eq. (.16). Putting λ=iω2 in Eq. (.16) and separating imaginary and real parts, we have

    (P18P16ω22)cos(ω2τ2)+P17ω2sin(ω2τ2)=P11ω22(P13+P15),P17ω2cos(ω2τ2)(P18P16ω22)sin(ω2τ2)=ω32+(P12+P14)ω2. (.17)

    Squaring and adding above equations, we obtain

    ω62+d11ω42+d12ω22+d13=0, (.18)

    where

    d11=P2112(P12+P14)P216,d12=(P12+P14)2P11(P13+P15)P217+2P16P18,d13=(P13+P15)2P218.

    Let ω22=v2, then Eq. (.18) becomes

    v32+d11v22+d12v2+d13=0. (.19)

    1. If d11>0,d12>0,d13>0, then Eq. (.18) has no positive roots. Hence all roots of Eq. (.18) have negative real parts when τ2(0,+).

    2. If d13<0,d11>0,d12>0 or d13<0,d11>0,d12<0 or d13<0,d11<0,d12<0, then Eq. (.18) has a unique positive root ω2=v2. Eliminating sin(ω2τ2) from Eq. (.17), we obtain

    τ2j=1ω2cos1[(P16(P11+P15)+P11P18+P17(P12+P14))ω22(P17+P11P16)ω24(P11+P15)P18(P18P16ω22)2+(P17ω2)2+2jπ],j=0,1,2,. (.20)

    3. If d11>0, d12<0,d13>0 or d11<0, d12>0,d13>0 or d11<0, d12<0,d13>0, then Eq. (.18) has two positive roots ω±2=v±2. Eliminating sin(ω2τ2) from Eq. (.17), we obtain

    τ±2k=1ω±2cos1[(P16(P11+P15)+P11P18+P17(P12+P14))ω±22(P17+P11P16)ω±24(P11+P15)P18(P18P16ω±22)2+(P17ω±2)2+2kπ],k=0,1,2,. (.21)

    4. If d11<0, d12>0,d13<0, then Eq. (.18) has two positive roots ω1,2,32=v1,2,32. Eliminating sin(ω2τ2) from Eq. (.17), we obtain

    τ1,2,32l=1ω1,2,32cos1[(P16(P11+P15)+P11P18+P17(P12+P14))ω1,2,322(P17+P11P16)ω1,2,324(P11+P15)P18(P18P16ω1,2,322)2+(P17ω1,2,32)2+2lπ],l=0,1,2,. (.22)

    Hence, Eq. (.18) has a pair of purely imaginary roots ±iω2 with τ2=τ2j,j=0,1,2,, a pair of purely imaginary roots ±iω±2 with τ2=τ±2k,k=0,1,2,, and a pair of purely imaginary roots ±iω1,2,32 with τ2=τ1,2,32l,l=0,1,2,. Let {τ2j}+j=0={τ2j}+j=0 such that τ20<τ21<τ22<<τ2j<; {τ2k}+k=0={τ±2k}+k=0 such that τ20<τ21<τ22<<τ2k<, where τ20=min{τ+20,τ20}; and {τ2l}+l=0={τ1,2,32l}+l=0 such that τ20<τ21<τ22<<τ2l<, where τ20=min{τ120,τ220,τ320}.

    Let λ(τ2)=α(τ2)+iω2(τ2) be a root of Eq. (.18) near τ2=τ2j and α(τ2j)=0, ω2(τ2j)=ω2, j=0,1,2,. Further, differentiating Eq. (.18) with respect to τ2 and substituting λ=iω2, τ2=τ2j, j=0,1,2,, we obtain

    Re(dλdτ2)1λ=iω2,τ2=τ2j=M21M23+M22M24M223+M224, (.23)

    where

    M21=P16ω23(sin(τ2jω2))+P18ω2sin(τ2jω2)P17ω22cos(τ2jω2),M22=P17ω22sin(τ2jω2)+P16ω23(cos(τ2jω2))+P18ω2cos(τ2jω2),M23=2P16ω2sin(τ2jω2)P17τ2jω2sin(τ2jω2)+P16τ2jω22cos(τ2jω2)+P17cos(τ2jω2)P18τ2jcos(τ2jω2)+P12+P143ω22,M24=P16τ2jω22(sin(τ2jω2))P17sin(τ2jω2)+P18τ2jsin(τ2jω2)+2P16ω2cos(τ2jω2)P17τ2jω2cos(τ2jω2)+2P11ω2.

    Proof of Lemma 4.7 and Theorem 4.8: If τ1=τ2=τ, then the characteristic Eq. (4.3) becomes

    λ3+P11λ2+P12λ+P13+(P16λ2+(P14+P17)λ+P15+P18)eλτ=0, (.24)

    Let λ=iω3,(ω3>0) be a root of Eq. (.24). Putting λ=iω3 in Eq. (.24) and separating imaginary and real parts, we have

    (P16ω23+P15+P18)cos(ω3τ)+(P14+P17)ω3sin(ω3τ)=P11ω23P13,(P14+P17)ω3cos(ω3τ)(P16ω23+P15+P18)sin(ω3τ)=ω33P12ω3. (.25)

    Squaring and adding above equations, we obtain

    ω63+e11ω43+e12ω23+e13=0, (.26)

    where

    e11=P2112P12P216,e12=P212(P14+P17)22P11P13+2P16(P15+P18),e13=P213(P15+P18)2

    Let ω23=v3, then Eq. (.26) becomes

    v33+e11v23+e12v3+e13=0. (.27)

    1. If e11>0,e12>0,e13>0, then Eq. (.26) has no positive roots. Hence all roots of Eq. (.26) have negative real parts when τ(0,+).

    2. If e13<0,e11>0,e12>0 or e13<0,e11>0,e12<0 or e13<0,e11<0,e12<0, then Eq. (.26) has a unique positive root ω3=v3. Eliminating sin(ω3τ) from Eq. (.25), we obtain

    τj=1ω3cos1[(P14+P17P11P16)ω34+(P13P16+P11(P15+P18)P12(P14+P17))ω32P13(P15+P18)((P14+P17)ω3)2+(P16ω32+P15+P18)2+2jπ],j=0,1,2,. (.28)

    3. If e11>0, e12<0,e13>0 or e11<0, e12>0,e13>0 or e11<0, e12<0,e13>0, then Eq. (.26) has two positive roots ω±3=v±3. Eliminating sin(ω3τ) from Eq. (.25), we obtain

    τ±k=1ω±3cos1[(P14+P17P11P16)ω±34+(P13P16+P11(P15+P18)P12(P14+P17))ω±32P13(P15+P18)((P14+P17)ω±3)2+(P16ω±32+P15+P18)2+2kπ],k=0,1,2,. (.29)

    4. f e11<0, e12>0,e13<0, then Eq. (.26) has two positive roots ω1,2,33=v1,2,33. Eliminating sin(ω3τ) from Eq. (.25), we obtain

    τ1,2,3l=1ω1,2,33cos1[(P14+P17P11P16)ω1,2,334+(P13P16+P11(P15+P18)P12(P14+P17))ω1,2,332P13(P15+P18)((P14+P17)ω1,2,33)2+(P16ω1,2,332+P15+P18)2+2lπ],l=0,1,2,. (.30)

    Hence, Eq. (.24) has a pair of purely imaginary roots ±iω3 with τ=τj,j=0,1,2,, a pair of purely imaginary roots ±iω±3 with τ=τ±k,k=0,1,2,, and a pair of purely imaginary roots ±iω1,2,33 with τ=τ1,2,3l,l=0,1,2,. Let {τj}+j=0={τj}+j=0 such that τ0<τ1<τ2<<τj<, {τk}+k=0={τ±k}+k=0 such that τ0<τ1<τ2<<τk<, where τ0=min{τ+0,τ0}, and {τl}+l=0={τ1,2,3l}+l=0 such that τ0<τ1<τ2<<τl<, where τ0=min{τ10,τ20,τ30}.

    Let λ(τ)=α(τ)+iω3(τ) be a root of Eq. (.24) near τ=τj and α(τj)=0, ω3(τj)=ω3, j=0,1,2,. Further, differentiating Eq. (.24) with respect to τ and substituting λ=iω3, τ=τj, j=0,1,2,, we obtain

    Re(dλdτ)1λ=iω3,τ=τj=M31M33+M32M34M233+M234, (.31)
    M31=P16ω33sin(τjω3)+ω3(P15sin(τjω3)+P18sin(τjω3))ω32(P14cos(τjω3)+P17cos(τjω3)),M32=ω32(P14sin(τjω3)+P17sin(τjω3))P16ω33cos(τjω3)+ω3(P15cos(τjω3)+P18cos(τjω3)),M33=ω3(P14τjsin(τjω3)+2P16sin(τjω3)P17τjsin(τjω3))+ω32(P16τjcos(τjω3)3)+P14cos(τjω3)P15τjcos(τjω3)+P17cos(τjω3)P18τcos(τjω3)+P12,M34=P16τjω32sin(τjω3)P14sin(τjω3)+P15τjsin(τjω3)P17sin(τjω3)+P18τjsin(τjω3)+ω3(P14τjcos(τjω3)+2P16cos(τjω3)P17τjcos(τjω3)+2P11).

    Proof of Lemma 4.9 and Theorem 4.10: If τ1>0,τ2(0,τ2j). In this case, we take τ1 as a bifurcation parameter and τ2 is in stable interval. Let λ= iω4,(ω4>0) be a root of Eq. (4.3). Putting λ=iω4 in Eq. (4.3) and separating imaginary and real parts, we have

    P15cos(ω4τ1)+P14ω4sin(ω4τ1)=P11ω24P13(P18P16ω24)cos(ω4τ2)P17ω4sin(ω4τ2),P14ω4cos(ω4τ1)P15sin(ω4τ1)=P12ω4+ω34P17ω4cos(ω4τ2)+(P18P16ω24)sin(ω4τ2). (.32)

    From above equation, we obtain the following equation with respect to ω4

    h11(ω4)+h12(ω4)cos(ω4τ2)+h13(ω4)sin(ω4τ2)=0, (.33)

    with

    h11(ω4)=ω64+(P2112P12+P216)ω44+(P212+P172P11P132P16P18P214)ω24+P213P215+P218,h12(ω4)=2ω44(P17P11P16)+2(P12P17+P13P16P11P18)ω24+2P13P18,h13(ω4)=2ω54P16+2ω4(P13P17P18P12)2ω34(P11P17P16P12P18).

    Assume that H1: Eq. (.33) has a positive root ω4. Eliminating sin(ω4τ1) from Eq. (.32), we obtain

    τ1j=1ω4cos1[P14ω44+(P11P15P12P14)ω42P13P15(P15P18+(P17P14P15P16)ω42)cos(τ2ω4)+((P18P14P15P17)ω4P14P16ω43)sin(τ2ω4)P215+(P14ω4)2+2jπ], (.34)

    j=0,1,2,. Now differentiating Eq. (4.3) with respect to τ1 and substituting λ=iω4, τ1=τ1j, j=0,1,2,, we obtain

    Re(dλdτ1)1λ=iω4,τ1=τ1j=M41M43+M42M44M243+M244, (.35)

    where

    M41=P15ω4sin(τ1jω4)P11ω42cos(τ1jω4),M42=P11ω42sin(τ1jω4)+P15ω4cos(τ1jω4),M43=ω4(2P16sin(τ2ω4)P11τ1jsin(τ1jω4)+P17τ2sin(τ2ω4))ω42(P16τ2cos(τ2ω4)3)+P11cos(τ1jω4)P17cos(τ2ω4)P15τ1jcos(τ1jω4)+P18τ2cos(τ2ω4)+P12,M44=P11sin(τ1jω4)+P17sin(τ2ω4)+P15τ1jsin(τ1jω4)P18τ2sin(τ2ω4)ω4(2P16cos(τ2ω4)P11τ1cos(τ1jω4)+P17τ2cos(τ2ω4)+2P11)+P16τ2ω42sin(τ2ω4).

    Let p1=SS,p2=EE,p3=II and τ1=ζ+τ1, where ζR. Normalize the delay τ1 by the time-scaling tt/τ1. Then, the reduced system (4.1) becomes in the following functional differential form:

    ˙p(t)=Lζ(pt)+F(ζ,pt), (.36)

    where Lζ:CR3 and F:R×CR3 are defined as

    Lζφ=(ζ+τ1)(Δ1maxφ(0)+Δ2maxφ(τ2τ1)+Δ3maxφ(1)),

    and

    F(ζ,φ)=(α11φ1(0)φ3(0)+α12φ1(0)φ3(1)+α13φ3(0)φ3(1)α21φ1(0)φ3(0)+α22φ1(0)φ3(1)+α23φ3(0)φ3(1)α31φ23(τ2τ1)),

    with

    Δ1max=(J110J12J13J14J150J16J17),Δ3max=(00J1800J19000),Δ2max=(00000000J20),

    and

    α11=βf(I)=α21,α12=βIf(I)f(I)2=α22,α13=βSf(I)f(I)2=α23,α31=γIg(I)[g(I)2g(I)g(I)g(I)2g

    By the Riesz representation theorem [58], there exists a function \vartheta (\theta, \zeta) , whose components are of bounded variation for \theta \in [-1, 0] such that

    \begin{equation*} L_{\zeta}\varphi = \int_{-1}^{0}d\vartheta (\theta, \zeta)\varphi (\theta). \end{equation*}

    We can choose

    \begin{equation*} \begin{aligned} \vartheta (\theta, \zeta) = \left\lbrace \begin{array}{cc} (\tau_{1}^{*} +\zeta)(\Delta_{1 \max}+\Delta_{2 \max}+\Delta_{3 \max}),& \theta = 0, \\ (\tau_{1}^{*} +\zeta)(\Delta_{2 \max}+\Delta_{3 \max}),& \theta \in \left[-\frac{\tau_{2}^{*}}{\tau_{1}^{*}},0 \right), \\ (\tau_{1}^{*} +\zeta)(\Delta_{3 \max}),& \theta \in \left(-1, -\frac{\tau_{2}^{*}}{\tau_{1}^{*}} \right) , \\ 0, & \theta = -1, \end{array} \right. \end{aligned} \end{equation*}

    with \delta (\theta) = \left\lbrace \begin{array}{cc} 0, & \theta \neq 0, \\ 1, & \theta = 1. \end{array}\right.

    For \varphi \in C([-1, 0], \mathbb{R}^{3}) , define

    \begin{equation*} A(\zeta)\varphi = \left\lbrace \begin{array}{cc} \frac{d \varphi (\theta)}{d\theta}, & -1 \leq \theta < 0, \\ \int_{-1}^{0}d\vartheta (\theta, \zeta)\varphi (\theta),& \theta = 0, \end{array} \right. \end{equation*}
    \begin{equation*} R(\zeta) \varphi = \left\lbrace \begin{array}{cc} 0, & -1 \leq \theta < 0, \\ F(\zeta, \theta), & \theta = 0. \end{array} \right. \end{equation*}

    Then, the system (.36) is equivalent to the following system

    \begin{equation} \dot{p}(t) = A(\zeta)p(t)+R(\zeta)p(t). \end{equation} (.37)

    For \kappa \in C'([0, 1], (\mathbb{R}^{3})^{*}) , the adjoint operator A^* of A(0) is defined by

    \begin{equation*} A^*(\kappa) = \left\lbrace \begin{array}{cc} -\frac{d \kappa (s)}{ds}, & 0 < s \leq 1, \\ \int_{-1}^{0}d\vartheta^{T} (s, 0)\kappa (-s),& s = 0. \end{array} \right. \end{equation*}

    For \varphi \in C([-1, 0], \mathbb{R}^{3}) and \kappa \in C'([0, 1], (\mathbb{R}^{3})^{*}) , we define the following bilinear inner form:

    \begin{equation} \langle \kappa (s), \varphi (s) \rangle = \bar{\kappa} (0)\varphi (0) - \int_{\theta = -1}^{0} \int_{\nu = 0}^{\theta} \bar{\kappa}(\nu -\theta)d\vartheta(\theta)\varphi (\nu)d\nu, \end{equation} (.38)

    where \vartheta(\theta) = \vartheta(\theta, 0) .

    From subsection 4.3, it can be seen that \pm i \tau_{1}^{*}\omega_{4}^{*} are the eigenvalues of A(0) , therefore, \pm i \tau_{1}^{*}\omega_{4}^{*} are also the eigenvalues of A^{*}(0) . Let p(\theta) = (1, p_2, p_3)^{T}e^{i \tau_{1}^{*}\omega_{4}^{*}\theta} and p^*(s) = D(1, p_2^*, p_3^*)^{T}e^{i \tau_{1}^{*}\omega_{4}^{*}s} be the eigenvectors of A(0) and A^{*}(0) corresponding to the eigenvalues + i \tau_{1}^{*}\omega_{4}^{*} and - i \tau_{1}^{*}\omega_{4}^{*} , respectively. Hence, we obtain

    \begin{align*} p_2 = &\frac{{\omega_{4}^{*}}^{2}-(J_{11}+J_{17})i\omega_{4}^{*}+J_{11}J_{17} + J_{20} (J_{11}-i\omega_{4}^{*})e^{-i\tau_{2}^{*}\omega_{4}^{*}}}{J_{16}(J_{12}+J_{18}e^{-i\tau_{1}^{*}\omega_{4}^{*}})}, \quad p_3 = \frac{i\omega_{4}^{*} - J_{11}}{J_{12}+J_{18}e^{-i\tau_{1}^{*}\omega_{4}^{*}}} \\ p_2^* = & \frac{-i\omega_{4}^{*} - J_{11}}{J_{13}} , \quad p_3^* = \frac{J_{11}(J_{13}-J_{15})-J_{15}i\omega_{4}^{*}+(J_{13}J_{18}-J_{19}(J_{11}+i\omega_{4}^{*}))e^{i\tau_{1}^{*}\omega_{4}^{*}}}{J_{13}(-i\omega_{4}^{*}-J_{17}-J_{20}e^{i\tau_{2}^{*}\omega_{4}^{*}})} \end{align*}

    In order to show that \langle p^*, \bar{p} \rangle = 0 and \langle p^*, p \rangle = 1 , we need to compute the value of \bar{D} . In view of Eq. (.38), we may choose

    \begin{equation*} \begin{aligned} \bar{D} = \left[1+p_2 \bar{p}_2^*+p_3 \bar{p}_3^* +\tau_{1}^{*}e^{-i\tau_{1}^{*}\omega_{4}^{*}}(J_{18}p_3 + J_{19}p_3 \bar{p}_3^*)+\tau_{2}^{*}e^{-i\tau_{2}^{*}\omega_{4}^{*}}J_{20}p_2 \bar{p}_2^* \right]^{-1}. \end{aligned} \end{equation*}

    Using the same notation as in Hassard et al. [57], we compute the coordinates to describe the center manifold C_0 at \zeta = 0 . Let p_t be the solution of Eq. (.36), when \zeta = 0 . Define

    \begin{equation} \begin{aligned} z(t) & = \langle p^*, p_t \rangle, \\ W(t, \theta) & = p_t(\theta)-2 Re\left\lbrace z(t)p (\theta)\right\rbrace. \end{aligned} \end{equation} (.39)

    On the center manifold C_0 , we have

    \begin{equation*} W(t, \theta) = W(z(t), \bar{z}(t), \theta), \end{equation*}

    where

    \begin{equation*} \begin{aligned} W(z(t), \bar{z}(t), \theta) = & W(z, \bar{z}),\\ = & W_{20}(\theta)\frac{z^2}{2}+W_{11}(\theta)z\bar{z}+W_{02}(\theta)\frac{\bar{z}^2}{2} +W_{30}(\theta)\frac{z^3}{2}+ \ldots , \end{aligned} \end{equation*}

    and z and \bar{z} are local coordinates for center manifold C_0 in the direction of p^* and \bar{p}^* . Note that W is also real if p_t is real, we consider only real solutions. For solutions p_t \in C_0 of Eq. (.36), since \zeta = 0 ,

    \begin{equation*} \begin{aligned} \dot{z}(t) = & i\tau_{1}^{*}\omega_{4}^{*} z +\bar{p}^*(0)F(0, W(z, \bar{z}, \theta))+2 Re\left\lbrace z(t) p(\theta) \right\rbrace ,\\ = & i\tau_{1}^{*}\omega_{4}^{*} z +g(z, \bar{z}), \end{aligned} \end{equation*}

    where

    \begin{equation} g(z, \bar{z}) = g_{20}\frac{z^2}{2}+g_{11}z\bar{z}+g_{02}\frac{\bar{z}^2}{2} +g_{21}\frac{z^2 \bar{z}}{2}+ \ldots , \end{equation} (.40)

    then it follows from Eq. (.36), that

    \begin{equation*} \begin{aligned} p_t = & W(t, \theta) +2 Re \left\lbrace z(t) p(\theta) \right\rbrace ,\\ = & W_{20}(\theta)\frac{z^2}{2}+W_{11}(\theta)z\bar{z}+W_{02}(\theta)\frac{\bar{z}^2}{2}+(1, p_2)e^{i\tau_{1}^{*}\omega_{4}^{*} \theta}z+(1, \bar{p}_2)e^{-i\tau_{1}^{*}\omega_{4}^{*} \theta}\bar{z}+ \ldots . \end{aligned} \end{equation*}

    Hence, we have

    \begin{equation*} \begin{aligned} g(z, \bar{z}) = \bar{p}^*(0)F(0,p_t). \end{aligned} \end{equation*}

    After some algebraic calculation and comparing the coefficients with (.40), we obtain

    \begin{align*} g_{20} = &2 \bar{D}\tau_{1}^{*}[ (\alpha_{11}p_3 + \alpha_{12}p_3 +\alpha_{13}p_3^2)+p_2^*(\alpha_{21}p_3 +\alpha_{22}p_3+\alpha_{23}p_3^2)+p_3^* (\alpha_{31}p_3^2)],\\ g_{11} = & \bar{D}\tau_{1}^{*}[(\alpha_{11}\bar{p}_3+\alpha_{11}p_3+\alpha_{12}p_3 + \alpha_{12} \bar{p}_3 + 2 \alpha_{13}p_3 \bar{p}_3)+p_2^*(\alpha_{21}\bar{p}_3+\alpha_{21}p_3\\ &+\alpha_{22}\bar{p}_3+\alpha_{22}p_3+2\alpha_{23}p_3 \bar{p}_3)+p_3^*(2 \alpha_{31}p_3 \bar{p}_3)],\\ g_{02} = &2 \bar{D}\tau_{1}^{*}[ (\alpha_{11}\bar{p}_3 +\alpha_{12}\bar{p}_3 + \alpha_{13} \bar{p}_3^2)+p_2^*(\alpha_{21}\bar{p}_3 +\alpha_{22}\bar{p}_3+\alpha_{23}\bar{p}_3^2)+p_3^* (\alpha_{31} \bar{p}_3^2)],\\ g_{21} = & \bar{D}\tau_{1}^{*}[(\alpha_{11}\bar{p}_3 W_{20}^{(1)}(0)+2\alpha_{11}p_3 W_{11}^{(1)}(0)+2\alpha_{11}W_{11}^{(3)}(0)+\alpha_{11}W_{20}^{(3)}(0)\\ &+\alpha_{12}W_{20}^{(1)}(0)\bar{p}_{3}+2\alpha_{12}p_3W_{11}^{(1)}(0)+2 \alpha_{12}W_{11}^{(3)}(-1)+\alpha_{12}W_{20}^{(3)}(-1)\\ &+\alpha_{13}\bar{p}_3 W_{20}^{(3)}(0)+2\alpha_{13} p_3 W_{11}^{(3)}(0) + 2 p_3 W_{11}^{(3)}(-1) +W_{20}^{(3)}(-1) )\\ &+p_2^*(\alpha_{21}\bar{p}_3 W_{20}^{(1)}(0)+2\alpha_{21}p_3 W_{11}^{(1)}(0)+2\alpha_{21}W_{11}^{(3)}(0)+\alpha_{21}W_{20}^{(3)}(0)\\ &+\alpha_{22}W_{20}^{(1)}(0)\bar{p}_{3}+2\alpha_{22}p_3W_{11}^{(1)}(0)+2 \alpha_{22}W_{11}^{(3)}(-1)+\alpha_{22}W_{20}^{(3)}(-1)\\ &+\alpha_{23}\bar{p}_3 W_{20}^{(3)}(0)+2\alpha_{23} p_3 W_{11}^{(3)}(0) + 2 p_3 W_{11}^{(3)}(-1) +W_{20}^{(3)}(-1))\\ &+ p_3^* (2 \alpha_{31} W_{20}^{(3)}\left( -\frac{\tau_{2}^{*}}{\tau_{1}^{*}}\right) \bar{p}_{3} +4 \alpha_{31} W_{11}^{(3)}\left( -\frac{\tau_{2}^{*}}{\tau_{1}^{*}}\right))]. \end{align*}

    with

    \begin{align*} W_{20}(\theta) = & \frac{ig_{20}p(0)}{\tau_{1}^{*}\omega_{4}^{*}}e^{i\tau_{1}^{*}\omega_{4}^{*} \theta}+\frac{i\bar{g}_{02}\bar{p}(0)}{3\tau_{1}^{*}\omega_{4}^{*}}e^{-i\tau_{1}^{*}\omega_{4}^{*} \theta}+E_1 e^{2i\tau_{1}^{*}\omega_{4}^{*} \theta},\\ W_{11}(\theta) = &-\frac{ig_{11}p(0)}{\tau_{1}^{*}\omega_{4}^{*}}e^{i\tau_{1}^{*}\omega_{4}^{*} \theta}+\frac{i\bar{g}_{11}\bar{p}(0)}{\tau_{1}^{*}\omega_{4}^{*}}e^{-i\tau_{1}^{*}\omega_{4}^{*} \theta}+E_2. \end{align*}

    E_1 and E_2 can be determined by the following equations

    \begin{equation*} \begin{aligned} E_1 = 2\left( \begin{array}{ccc} J_{11}^{*}& 0 & -J_{12} +J_{18}^{*} \\ -J_{13}& J_{14}^{*} & -J_{15} +J_{19}^{*} \\ 0 & -J_{16} & -J_{17} +J_{20}^{*} \end{array} \right)^{-1} \times \left(\begin{array}{c} E_{11}'\\ E_{12}'\\ E_{13}' \end{array} \right), \end{aligned} \end{equation*}
    \begin{equation*} \begin{aligned} E_2 = -\left( \begin{array}{ccc} J_{11}& 0 & J_{12}+J_{18}\\ J_{13}& J_{14} & J_{15}+J_{19} \\ 0 & J_{16} & J_{17}+J_{20} \end{array} \right)^{-1} \times \left(\begin{array}{c} E_{14}'\\ E_{15}'\\ E_{16}' \end{array} \right), \end{aligned} \end{equation*}

    where

    \begin{align*} J_{11}^{*} = &2i\omega_{4}^{*} -J_{11},\; J_{18}^{*} = -J_{18}e^{-2i\tau_{1}^{*}\omega_{4}^{*} },\; J_{14}^{*} = 2i\omega_{4}^{*} -J_{14},\\ J_{19}^{*} = &-J_{19}e^{-2i\tau_{1}^{*}\omega_{4}^{*} },\; J_{20}^{*} = 2i\omega_{4}^{*}-J_{20}e^{-2i\tau_{2}^{*}\omega_{4}^{*}},\\ E_{11}' = &\alpha_{11}p_3 + \alpha_{12}p_3 +\alpha_{13}p_3^2,\; E_{12}' = \alpha_{21}p_3 +\alpha_{22}p_3+\alpha_{23}p_3^2,\; E_{13}' = \alpha_{31}p_3^2,\\ E_{14}' = &\alpha_{11}\bar{p}_3+\alpha_{11}p_3+\alpha_{12}p_3 + \alpha_{12} \bar{p}_3 + 2 \alpha_{13}p_3 \bar{p}_3,\; E_{16}' = 2 \alpha_{31}p_3 \bar{p}_3, \\ E_{15}' = &\alpha_{21}\bar{p}_3+\alpha_{21}p_3+\alpha_{22}\bar{p}_3+\alpha_{22}p_3+2\alpha_{23}p_3 \bar{p}_3. \end{align*}

    Hence, we can determine the following values

    \begin{equation*} \begin{aligned} c_1(0) & = \frac{i}{2\tau_{1}^{*}\omega_{4}^{*}}\left( g_{20}g_{11}-2|g_{11}|^2-\frac{|g_{02}|^2}{3}\right) +\frac{g_{21}}{2},\\ \beta_2 & = 2Re\left\lbrace c_1(0)\right\rbrace,\\ \mu_2 & = -\frac{Re \left\lbrace c_1(0)\right\rbrace }{Re \left\lbrace \lambda'(\tau_{1}^{*})\right\rbrace },\\ T_2 & = -\frac{Im\left\lbrace c_1(0)\right\rbrace +\mu_2 Im\left\lbrace \lambda'(\tau_{1}^{*}) \right\rbrace}{\tau_{1}^{*}\omega_{4}^{*}}. \end{aligned} \end{equation*}

    The above expressions regulate the behavior of bifurcating periodic solution at the critical value \tau_{1}^{*} . Specifically, \mu_2 regulates the direction of Hopf bifurcation: if \mu_2 > 0\; (\mu_2 < 0) then the Hopf bifurcation is supercritical (subcritical) and bifurcating periodic solution exists for \tau_{1} > \tau_{1}^{*} \; (\tau_{1} < \tau_{1}^{*}) . \beta_ 2 regulates the stability of the bifurcating periodic solution: the bifurcating periodic solution is stable (unstable) if \beta_2 < 0 \; (\beta_2 > 0) . T_2 regulates the period of bifurcating periodic solution: if T_2 > 0 , the period increases, otherwise decreases.

    Proof of Theorem 5.1: To prove this theorem, we use Theorem 4.3 of [60]. We note that disease-free equilibrium (D_0) is unstable whenever R_0 > 1 for both systems (2.2) and (2.3). To show that (2.2) and (2.3) satisfy all the conditions of Theorem 4.3, when R_0 > 1 , we choose X = \mathbb{R}^{4} and E = \mho. The maximal invariant set M on the boundary \partial \mho is the singleton {D_0} and is isolated. Thus the hypothesis of Theorem 4.3 of [60] holds for both systems (2.2) and (2.3). We obtained that our systems have a unique endemic equilibrium whenever R_0 > 1. Hence the necessary and sufficient condition for uniform persistence in Theorem 4.3 of [60] is equivalent to D_0 being unstable.



    [1] F. Brauer, C. Castillo-Chavez, Mathematical models in population biology and epidemiology, New York: Springer, 2 (2012).
    [2] K. Dietz, J. A. Heesterbeek, Daniel Bernoulli's epidemiological model revisited, Math. Biosci., 180 (2002), 1–21. doi: 10.1016/S0025-5564(02)00122-0
    [3] W. O. Kermack, A. G. McKendrick, A contribution to the mathematical theory of epidemics, Proc. R. Soc. Lond. A Math. Phys. Sci., 115 (1927), 700–721.
    [4] H. W. Hethcote, The mathematics of infectious diseases, SIAM Rev. Soc. Ind. Appl. Math., 42 (2000), 599–653.
    [5] R. M. Anderson, B. Anderson, R. M. May, Infectious diseases of humans: dynamics and control, Oxford University Press, 1992.
    [6] S. Ruan, W. Wang, Dynamical behavior of an epidemic model with a nonlinear incidence rate, J. Differ. Equ., 188 (2003), 135–163. doi: 10.1016/S0022-0396(02)00089-X
    [7] J. P. Tripathi, S. Abbas, Global dynamics of autonomous and nonautonomous SI epidemic models with nonlinear incidence rate and feedback controls, Nonlinear Dyn, 86 (2016), 337–351. doi: 10.1007/s11071-016-2892-0
    [8] L. J. Gross, Mathematical models in plant biology: An overview, Applied Mathematical Ecology, (1989), 385–407.
    [9] W. M. Liu, H. W. Hethcote, S. A. Levin, Dynamical behavior of epidemiological models with nonlinear incidence rates, J. Math. Biol., 25 (1987), 359–380. doi: 10.1007/BF00277162
    [10] Y. Cai, Y. Kang, M. Banerjee, W. Wang, A stochastic SIRS epidemic model with infectious force under intervention strategies, J. Differ. Equ., 259 (2015), 7463–7502. doi: 10.1016/j.jde.2015.08.024
    [11] M. Zhou, T. Zhang, Global Analysis of an SEIR Epidemic Model with Infectious Force under Intervention Strategies, J. Appl. Math. Phys., 7 (2019), 1706. doi: 10.4236/jamp.2019.78117
    [12] J. Cui, Y. Sun, H. Zhu, The impact of media on the control of infectious diseases, J. Dyn. Differ. Equ., 20 (2008), 31–53. doi: 10.1007/s10884-007-9075-0
    [13] J.A. Cui, X. Tao, H. Zhu, An SIS infection model incorporating media coverage, Rocky Mt. J. Math., (2008), 1323–1334.
    [14] D. Xiao, S. Ruan, Global analysis of an epidemic model with nonmonotone incidence rate, Math. Biosci., 208 (2007), 419–429. doi: 10.1016/j.mbs.2006.09.025
    [15] Y. Xiao, T. Zhao, S. Tang, Dynamics of an infectious diseases with media/psychology induced non-smooth incidence, Math. Biosci. Eng., 10 (2013), 445. doi: 10.3934/mbe.2013.10.445
    [16] S. Tang, Y. Xiao, L. Yuan, R. A. Cheke, J. Wu, Campus quarantine (Fengxiao) for curbing emergent infectious diseases: Lessons from mitigating A/H1N1 in Xi'an, China, J. Theor. Biol., 295 (2012), 47–58. doi: 10.1016/j.jtbi.2011.10.035
    [17] Y. Xiao, S. Tang, J. Wu, Media impact switching surface during an infectious disease outbreak, Sci. Rep., 5 (2015), 7838. doi: 10.1038/srep07838
    [18] A. B. Gumel, S. Ruan, T. Day, J. Watmough, F. Brauer, P. Van den Driessche, et al., Modelling strategies for controlling SARS outbreaks, Proc. R. Soc. Lond. B Biol. Sci., 271 (2004), 2223–2232. doi: 10.1098/rspb.2004.2800
    [19] J. Zhang, J. Lou, Z. Ma, J. Wu, A compartmental model for the analysis of SARS transmission patterns and outbreak control measures in China, Appl. Math. Comput., 162 (2005), 909–924.
    [20] S. Bugalia, V. P. Bajiya, J. P. Tripathi, M. T. Li, G. Q. Sun, Mathematical modeling of COVID-19 transmission: The roles of intervention strategies and lockdown, Math. Biosci. Eng., 17 (2020), 5961–5986. doi: 10.3934/mbe.2020318
    [21] V. P. Bajiya, S. Bugalia, J. P. Tripathi, Mathematical modeling of COVID-19: Impact of non-pharmaceutical interventions in India, Chaos Interdiscipl. J. Nonlinear Sci., 30 (2020), 113143. doi: 10.1063/5.0021353
    [22] M. T. Li, G. Q. Sun, J. Zhang, Y. Zhao, X. Pei, L. Li, et al., Analysis of COVID-19 transmission in Shanxi Province with discrete time imported cases, Math. Biosci. Eng., 17 (2020), 3710. doi: 10.3934/mbe.2020208
    [23] C. N. Ngonghala, E. Iboi, S. Eikenberry, M. Scotch, C. R. MacIntyre, M. H. Bonds, et al., Mathematical assessment of the impact of non-pharmaceutical interventions on curtailing the 2019 novel Coronavirus, Math. Biosci., 325 (2020), 108364. doi: 10.1016/j.mbs.2020.108364
    [24] S. Batabyal, A. Batabyal, Mathematical computations on epidemiology: A case study of the novel coronavirus (SARS-CoV-2), Theory Biosci., (2021), 1–16.
    [25] S. Batabyal, COVID-19: Perturbation dynamics resulting chaos to stable with seasonality transmission, Chaos Solitons Fractals, 145 (2021), 110772. doi: 10.1016/j.chaos.2021.110772
    [26] W. Wang, S. Ruan, Bifurcations in an epidemic model with constant removal rate of the infectives, J. Math. Anal. Appl., 291 (2004), 775–793. doi: 10.1016/j.jmaa.2003.11.043
    [27] W. Wang, Epidemic models with nonlinear infection forces, Math. Biosci. Eng., 3 (2006), 267. doi: 10.3934/mbe.2006.3.267
    [28] C. Shan, H. Zhu, Bifurcations and complex dynamics of an SIR model with the impact of the number of hospital beds, J. Differ. Equ., 257 (2014), 1662–1688. doi: 10.1016/j.jde.2014.05.030
    [29] G. H. Li, Y. X. Zhang, Dynamic behaviors of a modified SIR model in epidemic diseases using nonlinear incidence and recovery rates, PLoS One, 12 (2017), e0175789. doi: 10.1371/journal.pone.0175789
    [30] R. Mu, A. Wei, Y. Yang, Global dynamics and sliding motion in A (H7N9) epidemic models with limited resources and Filippov control, J. Math. Anal. Appl., 477 (2019), 1296–1317. doi: 10.1016/j.jmaa.2019.05.013
    [31] H. Zhao, L. Wang, S. M. Oliva, H. Zhu, Modeling and Dynamics Analysis of Zika Transmission with Limited Medical Resources, Bull. Math. Biol., 82 (2020), 1–50. doi: 10.1007/s11538-019-00680-3
    [32] A. Sirijampa, S. Chinviriyasit, W. Chinviriyasit, Hopf bifurcation analysis of a delayed SEIR epidemic model with infectious force in latent and infected period, Adv. Differ. Equ., 1 (2018), 348.
    [33] E. Beretta, D. Breda, An SEIR epidemic model with constant latency time and infectious period, Math. Biosci. Eng., 8 (2011), 931. doi: 10.3934/mbe.2011.8.931
    [34] Z. Zhang, S. Kundu, J. P. Tripathi, S. Bugalia, Stability and Hopf bifurcation analysis of an SVEIR epidemic model with vaccination and multiple time delays, Chaos Solitons Fractals, 131 (2020), 109483. doi: 10.1016/j.chaos.2019.109483
    [35] 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. doi: 10.1007/s00285-017-1173-y
    [36] A. K. Misra, A. Sharma, V. Singh, Effect of awareness programs in controlling the prevalence of an epidemic with time delay, J. Biol. Sys., 19 (2011), 389–402. doi: 10.1142/S0218339011004020
    [37] F. Al Basir, S. Adhurya, M. Banerjee, E. Venturino, S. Ray, Modelling the Effect of Incubation and Latent Periods on the Dynamics of Vector-Borne Plant Viral Diseases, Bull. Math. Biol., 82 (2020), 1–22. doi: 10.1007/s11538-019-00680-3
    [38] S. Liao, W. Yang, Cholera model incorporating media coverage with multiple delays, Math. Methods Appl. Sci., 42 (2019), 419–439. doi: 10.1002/mma.5175
    [39] K. A. Pawelek, S. Liu, F. Pahlevani, L. Rong, A model of HIV-1 infection with two time delays: mathematical analysis and comparison with patient data, Math. Biosci., 235 (2012), 98–109. doi: 10.1016/j.mbs.2011.11.002
    [40] D. Greenhalgh, S. Rana, S. Samanta, T. Sardar, S. Bhattacharya, J. Chattopadhyay, Awareness programs control infectious disease–multiple delay induced mathematical model, Appl. Math. Comput., 251 (2015), 539–563.
    [41] H. 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. doi: 10.1080/17513758.2016.1229050
    [42] F. Al Basir, S. Ray, E. Venturino, Role of media coverage and delay in controlling infectious diseases: A mathematical model, Appl. Math. Comput., 337 (2018), 372–385.
    [43] J. Liu, Bifurcation analysis for a delayed SEIR epidemic model with saturated incidence and saturated treatment function, J. Biol. Dyn., 13 (2019), 461–480. doi: 10.1080/17513758.2019.1631965
    [44] G. H. Li, Y. X. Zhang, Dynamic behaviors of a modified SIR model in epidemic diseases using nonlinear incidence and recovery rates, PLoS One, 12 (2017), e0175789. doi: 10.1371/journal.pone.0175789
    [45] D. P. Gao, N. J. Huang, S. M. Kang, C. Zhang, Global stability analysis of an SVEIR epidemic model with general incidence rate, Bound. Value Probl., (2018), 42.
    [46] G. Huang, Y. Takeuchi, W. Ma, D. Wei, Global stability for delay SIR and SEIR epidemic models with nonlinear incidence rate, Bull. Math. Biol., 72 (2010), 1192–1207. doi: 10.1007/s11538-009-9487-6
    [47] E. Beretta, Y. Kuang, Geometric stability switch criteria in delay differential systems with delay dependent parameters, SIAM J. Math. Anal., 33 (2002), 1144–1165. doi: 10.1137/S0036141000376086
    [48] Q. An, E. Beretta, Y. Kuang, C. Wang, H. Wang, Geometric stability switch criteria in delay differential equations with two delays and delay dependent parameters, J. Differ. Equ., 266 (2019), 7073–7100. doi: 10.1016/j.jde.2018.11.025
    [49] Y. Kuang, editor, Delay differential equations: with applications in population dynamics, Academic Press, 1993.
    [50] X. Yang, Generalized form of Hurwitz-Routh criterion and Hopf bifurcation of higher order, Appl. Math. Lett., 15 (2002), 615–621. doi: 10.1016/S0893-9659(02)80014-3
    [51] C. Castillo-Chavez, B. Song, Dynamical models of tuberculosis and their applications, Math. Biosci. Eng., 1 (2004), 361. doi: 10.3934/mbe.2004.1.361
    [52] C. Castillo-Chavez, Z. Feng, W. Huang, On the computation of R_0 and its role on global stability, IMA Volumes in Mathematics and Its Applications, 1 (2002), 229.
    [53] Y. Y. Min, Y. G. Liu, Barbalat lemma and its application in analysis of system stability, Journal of Shandong University (engineering science), 37 (2007), 51–55.
    [54] S. Ruan, J. Wei, On the zeros of transcendental functions with applications to stability of delay differential equations with two delays, Dynamics of Continuous Discrete and Impulsive Systems Series A, 10 (2003), 863–874.
    [55] H. L. Freedman, V. S. Rao, The trade-off between mutual interference and time lags in predator-prey systems, Bull. Math. Biol., 45 (1983), 991–1004. doi: 10.1016/S0092-8240(83)80073-1
    [56] L. H. Erbe, H. I. Freedman, V. S. Rao, Three-species food-chain models with mutual interference and time delays, Math. Biosci., 80 (1986), 57–80. doi: 10.1016/0025-5564(86)90067-2
    [57] B. D. Hassard, N. D. Kazarinoff, Y. H. Wan, Y. W. Wan, Theory and applications of Hopf bifurcation, CUP Archive, 41 (1981).
    [58] J. J. Benedetto, W. Czaja, Riesz Representation Theorem, In Integration and Modern Analysis, Birkhäuser Boston, (2009), 321–357.
    [59] J. Wu, Symmetric functional differential equations and neural networks with memory, Trans. Am. Math. Soc., 350 (1998), 4799–4838. doi: 10.1090/S0002-9947-98-02083-2
    [60] H. I. Freedman, S. Ruan, M. Tang, Uniform persistence and flows near a closed positively invariant set, J. Dyn. Differ. Equ., 6 (1994), 583–600. doi: 10.1007/BF02218848
    [61] Worldometer, coronavirus, available from: https://www.worldometers.info/coronavirus/.
    [62] The print news, available from: https://theprint.in/world/this-is-how-france\-italy-and-spain-are-easing-their-lockdowns-one-step-at-a-time/413060/.
    [63] World Health Organization, COVID-19. Available from: https://covid19.who.int/.
    [64] X. Lin, H. Wang, Stability analysis of delay differential equations with two discrete delays, Canadian Appl. Math. Quart., 20 (2012), 519–533.
    [65] L. Wang, J. Wang, H. Zhao, Y. Shi, K. Wang, P. Wu, et al., Modelling and assessing the effects of medical resources on transmission of novel coronavirus (COVID-19) in Wuhan, China, Math. Biosci. Eng., 17 (2020), 2936–2949. doi: 10.3934/mbe.2020165
  • This article has been cited by:

    1. Hongli Zhu, Shiyong Liu, Wenwen Zheng, Haimanote Belay, Weiwei Zhang, Ying Qian, Yirong Wu, Tadesse Guadu Delele, Peng Jia, Vinícius Silva Belo, Assessing the dynamic impacts of non-pharmaceutical and pharmaceutical intervention measures on the containment results against COVID-19 in Ethiopia, 2022, 17, 1932-6203, e0271231, 10.1371/journal.pone.0271231
    2. Ya Chen, Juping Zhang, Zhen Jin, Optimal control of an influenza model with mixed cross-infection by age group, 2023, 206, 03784754, 410, 10.1016/j.matcom.2022.11.019
    3. Udai Kumar, Partha Sarathi Mandal, Jai Prakash Tripathi, Vijay Pal Bajiya, Sarita Bugalia, SIRS epidemiological model with ratio‐dependent incidence: Influence of preventive vaccination and treatment control strategies on disease dynamics, 2021, 44, 0170-4214, 14703, 10.1002/mma.7737
    4. Harendra Pal Singh, Sumit Kaur Bhatia, Yashika Bahri, Riya Jain, Optimal control strategies to combat COVID-19 transmission: A mathematical model with incubation time delay, 2022, 9, 26667207, 100176, 10.1016/j.rico.2022.100176
    5. Qing Yang, Xinhong Zhang, Daqing Jiang, Asymptotic behavior of a stochastic SIR model with general incidence rate and nonlinear Lévy jumps, 2022, 107, 0924-090X, 2975, 10.1007/s11071-021-07095-7
    6. Vijay Pal Bajiya, Jai Prakash Tripathi, Vipul Kakkar, Yun Kang, Modeling the impacts of awareness and limited medical resources on the epidemic size of a multi-group SIR epidemic model, 2022, 15, 1793-5245, 10.1142/S1793524522500450
    7. Weiwei Zhang, Thomas Huggins, Wenwen Zheng, Shiyong Liu, Zhanwei Du, Hongli Zhu, Ahmad Raza, Ahmad Hussen Tareq, Assessing the Dynamic Outcomes of Containment Strategies against COVID-19 under Different Public Health Governance Structures: A Comparison between Pakistan and Bangladesh, 2022, 19, 1660-4601, 9239, 10.3390/ijerph19159239
    8. Yihan Xing, Oleg Gaidai, Multi-regional COVID-19 epidemic forecast in Sweden, 2023, 9, 2055-2076, 205520762311629, 10.1177/20552076231162984
    9. François M. Castonguay, Brian F. Borah, Seonghye Jeon, Gabriel Rainisch, Patsy Kelso, Bishwa B. Adhikari, Daniel J. Daltry, Leah S. Fischer, Bradford Greening, Emily B. Kahn, Gloria J. Kang, Martin I. Meltzer, The public health impact of COVID-19 variants of concern on the effectiveness of contact tracing in Vermont, United States, 2024, 14, 2045-2322, 10.1038/s41598-024-68634-x
    10. Akanksha Rajpal, Sumit Kaur Bhatia, Shashank Goel, Praveen Kumar, Time delays in skill development and vacancy creation: Effects on unemployment through mathematical modelling, 2024, 130, 10075704, 107758, 10.1016/j.cnsns.2023.107758
    11. S. Swetha, S. Sindu Devi, K. Kannan, Optimal Control Strategies for COVID-19 Using SEIQR Mathematical Model, 2024, 0369-8203, 10.1007/s40010-024-00898-4
    12. Sarita Bugalia, Jai Prakash Tripathi, Assessing potential insights of an imperfect testing strategy: Parameter estimation and practical identifiability using early COVID-19 data in India, 2023, 123, 10075704, 107280, 10.1016/j.cnsns.2023.107280
    13. Shashank Goel, Sumit Kaur Bhatia, Jai Prakash Tripathi, Sarita Bugalia, Mansi Rana, Vijay Pal Bajiya, SIRC epidemic model with cross-immunity and multiple time delays, 2023, 87, 0303-6812, 10.1007/s00285-023-01974-w
    14. Zhijun Wu, Vincent Antonio Traag, Beyond six feet: The collective behavior of social distancing, 2024, 19, 1932-6203, e0293489, 10.1371/journal.pone.0293489
    15. Toshiaki Takayanagi, Estimating parameter values and initial states of variables in a mathematical model of coronavirus disease 2019 epidemic wave using the least squares method, Visual Basic for Applications, and Solver in Microsoft Excel, 2023, 4, 26669900, 100111, 10.1016/j.cmpbup.2023.100111
    16. Zuo‐jing Yin, Han Xiao, Stuart McDonald, Vladimir Brusic, Tian‐yi Qiu, Dynamically adjustable SVEIR(MH) model of multiwave epidemics: Estimating the effects of public health measures against COVID‐19, 2023, 95, 0146-6615, 10.1002/jmv.29301
    17. Aditya Pandey, Archana Singh Bhadauria, Vijai Shanker Verma, Impact of awareness and time-delayed saturated treatment on the transmission of infectious diseases, 2024, 17, 26667207, 100463, 10.1016/j.rico.2024.100463
    18. S. Dickson, S. Padmasekaran, K. Lakshmanan, Stability of delayed fractional order SEIQI_cRVW mathematical model for Omicron variant, 2024, 12, 2195-268X, 1392, 10.1007/s40435-023-01287-2
    19. Sarita Bugalia, Jai Prakash Tripathi, Hao Wang, Mutations make pandemics worse or better: modeling SARS-CoV-2 variants and imperfect vaccination, 2024, 88, 0303-6812, 10.1007/s00285-024-02068-x
    20. Akanksha Rajpal, Sumit Kaur Bhatia, Shashank Goel, Sanyam Tyagi, Praveen Kumar, Epidemic and unemployment interplay through bi-level multi delayed mathematical model, 2025, 229, 03784754, 758, 10.1016/j.matcom.2024.10.027
    21. A. Venkatesh, M. Prakash Raj, B. Baranidharan, Mohammad Khalid Imam Rahmani, Khawaja Tauseef Tasneem, Mudassir Khan, Jayant Giri, Analyzing steady-state equilibria and bifurcations in a time-delayed SIR epidemic model featuring Crowley-Martin incidence and Holling type II treatment rates, 2024, 10, 24058440, e39520, 10.1016/j.heliyon.2024.e39520
    22. May Anne E. Mata, Rey Audie S. Escosio, El Veena Grace A. Rosero, Jhunas Paul T. Viernes, Loreniel E. Anonuevo, Bryan S. Hernandez, Joel M. Addawe, Rizavel C. Addawe, Carlene P.C. Pilar-Arceo, Victoria May P. Mendoza, Aurelio A. de los Reyes, Analyzing the dynamics of COVID-19 transmission in select regions of the Philippines: A modeling approach to assess the impact of various tiers of community quarantines, 2024, 10, 24058440, e39330, 10.1016/j.heliyon.2024.e39330
    23. Vikram Singh, Shikha Kapoor, Sandeep kumar Gupta, Sandeep Sharma, Modelling the leadership role of police in controlling COVID-19, 2024, 12, 2544-7297, 10.1515/cmb-2024-0010
    24. Abraham J. Arenas, Gilberto González-Parra, Miguel Saenz Saenz, Qualitative Analysis of a COVID-19 Mathematical Model with a Discrete Time Delay, 2024, 13, 2227-7390, 120, 10.3390/math13010120
    25. Jiayin Zhang, Zhengyi Li, Tao Feng, Dynamics of a SIS Epidemic Model Under Environmental Stochasticity and Resource Availability, 2025, 48, 0126-6705, 10.1007/s40840-025-01865-x
  • Reader Comments
  • © 2021 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(5418) PDF downloads(459) Cited by(25)

Figures and Tables

Figures(29)  /  Tables(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog