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

Dynamics of an eco-epidemiological model with toxicity, treatment, time-varying incubation

  • Received: 03 March 2025 Revised: 24 April 2025 Accepted: 28 April 2025 Published: 20 May 2025
  • In this paper, we considered an eco-epidemiological model including toxicity, treatment, time-varying incubation, and Holling Ⅱ functional response. First, for the model without time delay, the positivity and boundedness of solutions were investigated, and some conditions for local asymptotic stability of all possible equilibriums and condition for global stability of the positive equilibrium were also given. Then, the local stability around the positive equilibrium and conditions for the existence of Hopf bifurcation of such model with time delay were explored. Furthermore, by using the method of multiple scales, a control strategy based on time delay was obtained to suppress oscillation. Moreover, the global stability of the positive equilibrium of the model with time-varying delay was investigated. Finally, the theoretical findings were validated through numerical simulations. Those results demonstrated that time-varying delays can induce more complex dynamics in systems. By combining the method of multiple scale with periodic perturbations, the oscillatory behavior induced by Hopf bifurcation can be effectively suppressed, providing a novel approach for stability control in time-varying delay systems.

    Citation: Rui Ma, Xin-You Meng. Dynamics of an eco-epidemiological model with toxicity, treatment, time-varying incubation[J]. Electronic Research Archive, 2025, 33(5): 3074-3110. doi: 10.3934/era.2025135

    Related Papers:

    [1] Hui Cao, Mengmeng Han, Yunxiao Bai, Suxia Zhang . Hopf bifurcation of the age-structured SIRS model with the varying population sizes. Electronic Research Archive, 2022, 30(10): 3811-3824. doi: 10.3934/era.2022194
    [2] Yuan Xue, Jinli Xu, Yuting Ding . Dynamics analysis of a diffusional immunosuppressive infection model with Beddington-DeAngelis functional response. Electronic Research Archive, 2023, 31(10): 6071-6088. doi: 10.3934/era.2023309
    [3] Wenbin Zhong, Yuting Ding . Spatiotemporal dynamics of a predator-prey model with a gestation delay and nonlocal competition. Electronic Research Archive, 2025, 33(4): 2601-2617. doi: 10.3934/era.2025116
    [4] Yanjiao Li, Yue Zhang . Dynamic behavior on a multi-time scale eco-epidemic model with stochastic disturbances. Electronic Research Archive, 2025, 33(3): 1667-1692. doi: 10.3934/era.2025078
    [5] Rina Su . Dynamic analysis for a class of hydrological model with time delay under fire disturbance. Electronic Research Archive, 2022, 30(9): 3290-3319. doi: 10.3934/era.2022167
    [6] Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215
    [7] Chen Wang, Ruizhi Yang . Hopf bifurcation analysis of a pine wilt disease model with both time delay and an alternative food source. Electronic Research Archive, 2025, 33(5): 2815-2839. doi: 10.3934/era.2025124
    [8] Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109
    [9] Xin Du, Quansheng Liu, Yuanhong Bi . Bifurcation analysis of a two–dimensional p53 gene regulatory network without and with time delay. Electronic Research Archive, 2024, 32(1): 293-316. doi: 10.3934/era.2024014
    [10] Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang . The dynamics of a delayed predator-prey model with square root functional response and stage structure. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150
  • In this paper, we considered an eco-epidemiological model including toxicity, treatment, time-varying incubation, and Holling Ⅱ functional response. First, for the model without time delay, the positivity and boundedness of solutions were investigated, and some conditions for local asymptotic stability of all possible equilibriums and condition for global stability of the positive equilibrium were also given. Then, the local stability around the positive equilibrium and conditions for the existence of Hopf bifurcation of such model with time delay were explored. Furthermore, by using the method of multiple scales, a control strategy based on time delay was obtained to suppress oscillation. Moreover, the global stability of the positive equilibrium of the model with time-varying delay was investigated. Finally, the theoretical findings were validated through numerical simulations. Those results demonstrated that time-varying delays can induce more complex dynamics in systems. By combining the method of multiple scale with periodic perturbations, the oscillatory behavior induced by Hopf bifurcation can be effectively suppressed, providing a novel approach for stability control in time-varying delay systems.



    Since the pioneering work studied by Lotka [1] and Volterra [2], predator-prey interaction models have become one of the most important and fascinating research areas in ecology. In recent years, researchers have expanded the classic predator-prey model into the eco-epidemiology. They focused on studying predator-prey systems with infectious diseases, further analyzing the dynamic relationship between predator and prey [3,4,5,6,7,8,9,10]. This interdisciplinary research approach helps us to better understand the impact of disease transmission on the structure and function of food webs. These research findings not only enrich ecological theory but also provide crucial insights for formulating more effective ecological management strategies. Beltrami and Carroll [3] established an eco-epidemiological model based on predator-prey, in which the prey population appears to be infected by a virus, forming an infected population. They discovered that such a system has been disrupted by a small amount of infectious factors, otherwise a stable structure would be observed. Meng et al. [5] showed that in the predator-prey system, the predator population can survive and spread the disease among the prey species. Recently, Sk and Pal [6] studied a predator-prey model with infected prey, and found that high levels of fear and low levels of refuge behavior can eliminate the disease from the system in deterministic and stochastic environments.

    Plankton, especially phytoplankton occupying the primary trophic level, forms the foundation of all aquatic food chains. These microscopic algae thrive in marine and freshwater environments, typically existing in low concentrations but capable of proliferating extensively on the water surface, resulting in dense cell aggregations known as algal blooms. While the majority of algae in aquatic environments are beneficial, a small number possess the ability to produce toxins. Viral infections within the plankton community can influence the changes of algal blooms, leading to alterations in the behavior and other aspects of aquatic and marine ecosystems. Based on the concept of plankton diseases, Kermack and McKendrick [11] proposed the classic SIR model to explore the impact of infections on populations. In recent years, many scholars have conducted extensive ecological epidemiological models to study the different ecological scenarios where populations are affected by external toxicity, disease transmission, prey refuge, Allee effect, fear effect, and other factors [12,13,14,15]. Among these studies, phytoplankton-zooplankton systems play a critical role in marine and freshwater ecosystems. Scholars have found that these toxic substances have a significant impact on the growth of zooplankton, and further affect the structure and function of the entire plankton community [15,16,17,18,19]. Chattopadhyay et al. [16] first proposed and studied the toxin-producing phytoplankton-zooplankton system. They suggested that toxin-producing phytoplankton (TPP) exert a biocontrol effect by reducing the grazing pressure exerted by zooplankton, thereby inhibiting zooplankton reproduction. Similarly, Gakkhar and Singh [18] proposed an eco-epidemiological time-delay model involving viral infection, TPP, and zooplankton system. Their results demonstrated that viral infections play a crucial role in controlling complex dynamic behaviors, including chaos. Additionally, variations in the toxin release rate of TPP significantly influence the emergence of chaos in the time-delay model. Huang et al. [19] studied aquatic toxicity through a simple mathematical model and concluded that high toxin levels can lead to population extinction. Srivastava and Thakur [15] proposed an eco-epidemiological model of an aquatic dynamical system:

    {St=D12S+rS(1S+Ik)cSIη1SPS+α+γ1I,It=D22I+cSIη2IPI+αd1Iγ1I,Pt=D32P+η3SPS+α+η4IPI+αd2Pθ(S+I)P, (1.1)

    where S, I, and P represent the densities of susceptible prey, infected prey, and predator at time t, respectively. The parameter θ represents the toxin release rate of susceptible and infected prey. The significance of other parameters in system (1.1) can be found in reference [15]. They analyzed the stability of both spatial and nonspatial models, and found that the density of predator is affected by virus-induced prey and toxicity.

    In the coexistence of disease and treatment, controlling the spread of disease is one of the primary objectives in the eco-epidemiological models [5,14,20,21,22,23]. Ghosh et al. [14] investigated an eco-epidemiological model incorporating the effects of fear, treatment, and cooperative hunting:

    {dSdt=(η+v(1η)v+y)rS(d+βI)S+ρIσ+I,dIdt=(βSδ(p+by)yρσ+I)I,dydt=(c(p+by)Im)y, (1.2)

    where ρIσ+I represents the treatment function, ρ>0 represents the maximum medical resources available for treatment, and 1σ>0 represents the saturation factor measuring the impact of delayed treatment on infected individuals. The meanings of other parameters can be found in reference [14]. The authors found that the fear of prey for predator, cooperative hunting, and treating infected prey with memory effect play a crucial role in preserving biodiversity.

    In an ecosystem with infectious disease, time delay is an important factor to describe the dynamic characteristics of the system. In recent years, many researchers have focused on delay-induced infection models to explore the combined effects of the infection process and time delay on population dynamic [18,24,25,26,27,28,29,30], where the delays are typically constant. However, in a real system, time delays are often not fixed and dynamically change over time, making time-varying delays more reflective of the complexity of a natural environment. Consequently, time-varying delays have been receiving increasing attention in predator-prey models, infectious disease transmission models, and broader ecological dynamics research. In recent years, researchers have explored how time-varying delays affect the stability and periodic behavior of systems through theoretical analysis and numerical simulations [31,32,33,34,35,36,37]. Research on this phenomenon not only helps to understand the fundamental mechanisms of ecosystem and epidemic dynamics, but also provides important theoretical support for devising effective management strategies. Wu et al. [35] developed a brucellosis transmission model incorporating seasonal alternation, density-dependent growth, stage structure, maturation delays, and time-varying latency periods. The numerical results showed that if the effects of time-varying latency or maturation delay are ignored, then the transmission of brucellosis would be overestimated (or underestimated). Liu and Meng [36] established an eco-epidemiological predator-prey system with time-varying delay, and obtained the permanence and global asymptotic stability of the time-varying delay system.

    Motivated by the above work, we will consider an eco-epidemiological model with toxicity, treatment and time-varying delay to analyze the interaction between toxicant prey and predators. This is a novel attempt to consider the effect of time-varying delay on prey and predators in a plankton system. Under the viral infection with and without time delay in a plankton system, investigating such complex dynamics are significant implications for both research on plankton population and ecological modeling.

    The organization of this paper is as follows. In Section 2, we perform the formulation of this model. In Section 3, we discuss the positivity and boundedness of the solution of the non-delayed system. Furthermore, we derive the conditions for the existence and local stability of five equilibriums of the system without delay. In Section 4, we focus on investigating the existence of Hopf bifurcation. Additionally, we apply the method of multiple scales (MMS) to the delay differential system. Building on this, we propose a control strategy that involves introducing time-varying perturbation to the delay, which aims to suppress oscillation. Moreover, the global stability of the positive equilibrium for model (2.1) is analyzed through the construction of a suitable Lyapunov function. In Section 5, we present some numerical simulations to validate the theoretical results obtained in the previous sections. We summarize our paper with some discussions in the last section.

    Motivated by the above papers, we take Noctiluca scintillan as the prey population and Paracalanu as the predator population. Results have shown that Noctiluca scintillan, a phytoplankton, produces toxins during its breeding season and can have adverse effect on zooplankton [17]. We give the following assumptions:

    (ⅰ) Assuming that only susceptible prey has the ability to reproduce, the disease only spreads within the prey population, and the disease is not inherited. When susceptible phytoplankton come into contact with infected phytoplankton, they also become infected. Therefore, the virus transmission is supposed to be Susceptible-Infected-Susceptible (SIS) type. However, the infected phytoplankton population still continues to grow at the same carrying capacity K.

    (ⅱ) The zooplankton consumes susceptible and infected phytoplankton in regard to Holling Ⅱ functional response. In addition, we hypothesize that the additional mortality of zooplankton is due to the phytoplankton liberating toxin instantaneously with rate θ. TPP populations do not release toxic chemicals unless there is a dense zooplankton population nearby [38].

    (ⅲ) In most eco-epidemiological models, researchers assume that infected prey recovers at a linear rate, and they do not consider the effect of treatment. In the real world, the process of treatment recovery can be complex. Thus, we consider a nonlinear treatment function, which may be more realistic.

    (ⅳ) Due to factors such as climate, environment, and individual differences, the incubation period of the virus may vary over time. Therefore, considering the time-varying incubation delay may be more in line with the actual situation.

    Thus, a time-varying delay eco-epidemiological model with toxicity and treatment is given as follows:

    {dSdt=rS(1S+IK)βSIμ1SPS+α+δIσ+I,dIdt=βS(tτ(t))I(tτ(t))μ2IPI+αρ1IδIσ+I,dPdt=μ3SPS+α+μ4IPI+αρ2Pθ(S+I)P, (2.1)

    where S(t), I(t), and P(t) represent the densities of the susceptible phytoplankton, infected phytoplankton, and the zooplankton at time t, respectively. τ(t) represents incubation period of disease and it is a continuously differentiable function. All parameters appeared in system (2.1) are positive. The biological significance of the remaining parameters of system (2.1) are given in Table 1. In addition, system (2.1) satisfies the nonnegative conditions S(m)=φ1(m)>0,I(m)=φ2(m)>0,P(m)=φ3(m)>0,m[τ,0], 0τ(t)τ,φi(m)C([τ,0),R3+0),i=1,2,3, where R3+0={(S(t),I(t),P(t):S(t)>0,I(t)>0,P(t)>0)}.

    Table 1.  Descriptions of all the parameters in (2.1).
    Symbols Descriptions Values Source
    r The intrinsic growth rate of susceptible phytoplankton population 22 Thakur[24]
    K Carrying capacity of phytoplankton population 200 Calibrate
    β Rate of disease transmission 0.06 Srivastava[15]
    α Represent the half saturation constant 15 Thakur[24]
    δ Maximum medical resource supplied for treatment 0.05 Calibrate
    1σ The saturation factor that measure the effect of the delay in treatment for the infected individuals 1.429 Calibrate
    μ1 The maximum predation rate of susceptible phytoplankton population 12.8 Calibrate
    μ2 The maximum predation rate of infected phytoplankton population 16.4 Calibrate
    μ3 Growth rate of zooplankton due to predation of susceptible phytoplankton population 7.5 Thakur[24]
    μ4 Growth rate of zooplankton due to predation of infected phytoplankton population 9.4 Thakur[24]
    ρ1 Death rate of infected phytoplankton population 3.4 Calibrate
    ρ2 Death rate of zooplankton 8.3 Thakur[24]
    θ The rate of toxin liberation by the toxin producing phytoplankton population 0.033 Calibrate

     | Show Table
    DownLoad: CSV

    When τ(t)=0, the non-delayed system is shown as follows.

    {dSdt=rS(1S+IK)βSIμ1SPS+α+δIσ+I,dIdt=βSIμ2IPI+αρ1IδIσ+I,dPdt=μ3SPS+α+μ4IPI+αρ2Pθ(S+I)P. (3.1)

    In this section, we will be devoted to discussing some biological properties of non-delayed system (3.1), such as positivity and boundedness of solutions. We also will analyze the local stability and global stability of system (3.1).

    Before analyzing the boundedness of solutions of system (3.1), we discuss the positivity of solutions.

    Lemma 3.1. The solution (S(t),I(t),P(t)) of system (3.1) with positive initial conditions (S(0),I(0),P(0))R3+ remains positive throughout the region for all t>0.

    Proof. From system (3.1), we get

    S(t)=S(0)exp{t0(r(1S(x)+I(x))K)βI(x)μ1P(x)S(x)+α)dx}>0,I(t)=I(0)exp{t0(βS(x)μ2P(x)I(x)+αρ1δσ+I(x))dx}>0,P(t)=P(0)exp{t0(μ3S(x)S(x)+α+μ4I(x)I(x)+αρ2θ(S(x)+I(x)))dx}>0.

    Hence, all the solutions of system (3.1) are positive for all t>0.

    Next, we will investigate that all positive solutions are uniformly bounded.

    Lemma 3.2. Assume that condition μ2μ3μ1μ4, then the set Ξ={(S,I,P)R3+:S+I+μ1μ3PK4rδ(r+δ)2} is a region of attraction for all solutions initiating in the positive quadrant, where δ=min{ρ1,ρ2}.

    Proof. To examine the boundedness of system (3.1), we define a function

    W(t)=S(t)+I(t)+μ1μ3P(t).

    Differentiating with respect to t, we have

    dWdt=dSdt+dIdt+μ1μ3dPdt.

    From system (3.1), we get

    dWdt=rS(1S+IK)ρ1Iμ2IPI+α+μ1μ3(μ4IPI+αρ2Pθ(S+I)P).

    Introducing the positive number δ and rewriting, we obtain

    dWdt+δWS(δ+r(1SK))(ρ1δ)Iμ1μ3(ρ2δ)P(μ2μ1μ4μ3)IPI+α,

    which implies

    dWdt+δWK4r(r+δ)2(ρ1δ)Iμ1μ3(ρ2δ)P(μ2μ1μ4μ3)IPI+α.

    Choosing δ=min{ρ1,ρ2} and μ2μ3μ1μ4, we have

    dWdt+δWK4r(r+δ)2. (3.2)

    Using the method of variation of constant for Eq (3.2), we get

    0<WK4rδ(r+δ)2(1eδt)+W0(eδt),

    where W0=W(S(0),I(0),P(0)).

    Therefore, it follows that

    lim supt+WK4rδ(r+δ)2.

    Then, all the solutions initiating in R3+ of system (3.1) are defined in the region

    Ξ={(S,I,P)R3+:S+I+μ1μ3PK4rδ(r+δ)2}.

    Thus, all solutions of system (3.1) are uniformly bounded for all t0 and Ξ is a positive invariant set.

    The eco-epidemiological system (3.1) has possible nonnegative feasible equilibria as follows: the trivial equilibrium E0(0,0,0) always exists. The boundary equilibrium E1(˜S,0,0) exists, where ˜S=K.

    Assume that E2(ˉS,ˉI,0) is the positive equilibrium of system (3.1), where ˉS and ˉI satisfy the following equations:

    {rˉS(1ˉS+ˉIK)βˉSˉI+δˉIσ+ˉI=0,βˉSˉIρ1ˉIδˉIσ+ˉI=0. (3.3)

    According to the second equation of (3.3), we can obtain

    ˉS=ρ1(σ+ˉI)+δβ(σ+ˉI). (3.4)

    Substituting (3.4) into the first equation of (3.3), we get

    b4ˉI4+b3ˉI3+b2ˉI2+b1ˉI+b0=0, (3.5)

    where

    b4=βρ1(βr),b3=rβKρ1+rρ21+β2δK3rβρ1σrβδρ21βσ2β2ρ1σβ2σ,b2=3rσρ1βK+rβδK+2ρ1δr+rρ21σ3rβσ2ρ12rσβδ2ρ21σ2βρ21βσδβ2σ2ρ1β2δσ+2β2δ2K,b1=3rρ1βσ2K+2rσβδKrρ21σ2+δ2rr4ρ1βσ2rβSσ3ρ21βρ21σ2βδ+δ3β2K,b0=σ3rρ1βK+σ2rβδK2δ2ρ1rδρ21σ3rδ2rσ.

    If (H1): β<r holds, then b4<0. By Descartes' rule of signs, Eq (3.5) has only one positive root ˉS, if and only if, one of the following conditions is met:

    (1) b3>0,b2>0,b1>0,b0>0;(2) b3<0,b2>0,b1>0,b0>0;(3) b3<0,b2<0,b1>0,b0>0;(4) b3<0,b2<0,b1<0,b0>0.

    We assume the assumption (H2): one of conditions (1)–(4) is true. Therefore, if (H1) and (H2) are satisfied, then system (3.1) has only one predator-free equilibrium E2(ˉS,ˉI,0), which is determined by Eqs (3.4) and (3.5).

    From the first equation of (3.1) under the case I=0, we have

    P=rμ1(S+α)(1SK). (3.6)

    Substituting (3.6) into the third equation of (3.1), we obtain

    S=μ3(θα+ρ2)+(θα+βμ3)24θρ2α2θ.

    If the assumption (H3):K>S, μ3>θα+ρ2 holds, then the equilibrium E3(S,0,P) exists.

    The positive equilibrium E(S,I,P) exists if S, I, and P are the positive solutions of the following equations:

    {rS(1S+IK)βSIμ1SPS+α+δIσ+I=0,βSIμ2IPI+αρ1IδIσ+I=0,μ3SPS+α+μ4IPI+αρ2Pθ(S+I)P=0. (3.7)

    By the second equation of (3.7), we get

    P=(I+α)(ρ1+δσ+IβS)μ2f(S,I). (3.8)

    The value of P is positive if the assumption (H4):ρ1+δσ+I<βS holds.

    Substituting (3.8) into the first and third equations of (3.7), then S and I are positive solution of following isoclines, such that

    {F1(S,I)rS(1S+IK)βSIμ1Sf(S,I)S+α+δIσ+I=0,F2(S,I)(μ3SS+α+μ4II+αρ2θ(S+I))f(S,I)=0. (3.9)

    We denote the two functions of Eq (3.9) as F1(S,I) and F2(S,I), respectively. Since F1 and F2 consist of polynomials and rational functions with nonzero denominators, then they are continuous. By selecting S[Smin,Smax] and I[Imin,Imax], F1(S,I) and F2(S,I) have opposite signs at the endpoints of the interval, satisfying the conditions of the intermediate value theorem (see Figure 1(a), (b)).

    Figure 1.  The intersection diagram of F1 and F2 with the zero plane. (a) Diagram of F1, (b) diagram of F2, (c) the intersection of F1 and F2 with the zero plane.

    Next, we theoretically prove that F1 and F2 have opposite signs at the endpoints of the interval. Taking the derivative with respect to S for F1, we get

    F1S=r(12S+IK)βIμ1(I+α)μ2(S+α)2[βS(2α+S)α(ρ1+δσ+I)].

    Obviously, F1S<0 when the assumption (A1) is true, where (A1):S is asymptotically close to K and βS(2α+S)>α(ρ1+δσ+I).

    Taking the derivative with respect to I for F1, we have

    F1I=(rKβ+μ1(ρ1βS)μ2(S+α))S+(μ1S(σα)μ2(S+α)+σ)δ(σ+I)2.

    If the assumption (A2) holds, where (A2):ρ1<βS and σ<α, then F1I<0. In conclusion, the above analysis shows that F1 is a decreasing function of S and I under the assumptions (A1) and (A2), respectively. Furthermore, we plot the monotonicity curves of F1 with respect to S and I at fixed values of I and S (see Figure 2(a), (b)), respectively. Consequently, F1 exhibits opposite signs at the endpoints of the interval.

    Figure 2.  The monotonicity curves of F1 and F2 with respect to S and I. (a) When I=40, F1 is monotonically decreasing with respect to S, (b) when S=119, F1 is monotonically decreasing with respect to I, (c) when I=40, F2 is monotonically decreasing with respect to S, (d) when S=119, F2 first monotonically increases and then monotonically decreases with respect to I.

    Next, taking the derivative with respect to S for F2, then we get

    F2S=I+αμ2[μ3(βSS+αα(ρ1+δσ+IβS)(S+α)2)+θ(ρ1+δσ+Iβ(2S+I))+β(μ4II+αρ2)].

    If the assumption (A3) is true, where (A3):βS<μ3α(ρ1+δσ+IβS)S+α, ρ1+δσ+I<β(2S+I) and μ4I<ρ2(I+α), then F2S<0. Without loss of generality, we treat S as a constant, whereupon F2 becomes a function of I. Substituting f into F2 yields

    F2(I)=1μ2(V3I3+V2I2+V1I+V0), (3.10)

    where V3=μ3δSσ, V2=(θβ)S, A1=βαS+μ4ρ1, V0=ρ1αμ3SS+α+ρ1ρ2α+ρ1αθS+βαμ3S2S+αβαρ2SβαθS2. It is direct from Eq (3.10) that

    F2(I)=1μ2(3V3I2+2V2I+V1).

    It is obvious that V3>0 and V1>0. If V20, then F2(I)>0 for I(0,). It means that F2(I) is an increasing function of I in (0,). If V2<0 and V22>3V3V1, let F2(I)=0, and we derive

    I=±2V223V3V1V23V3>0.

    When I(0,2V223V3V1V23V3), F2(I)>0. However, when I(2V223V3V1V23V3,2V223V3V1V23V3)F2(I)<0. That is, F2(I) is an increasing function as I(0,2V223V3V1V23V3) and a decreasing function as I(2V223V3V1V23V3,2V223V3V1V23V3). Similarly, we plot the monotonicity curves of F2 with respect to S and I at fixed values of I and S (see Figure 2(c), (d)), respectively. Therefore, F2 exhibits opposite signs at the endpoints of the interval.

    Additionally, the location of their intersection point is showed in Figure 1(c). From Figure 1(c), the figures of the two functions F1 and F2 intersect as S and I vary. As shown in the phase plane projection (see Figure 3), the intersection points of F1(S,I) and F2(S,I) are confined to the positive domain, confirming that the positive equilibrium E(S,I,P) exists.

    Figure 3.  Projection of the intersection curve for F1=F2=0. (a) Isolated plot of the intersection curve, (b)(c) When F1=F2=0, the intersection points between F1 and F2 are positive, (d) Projection of the intersection points.

    Furthermore, in Figure 4, we plot the isoclines of Eq (3.9) by choosing the parametric values mentioned in Table 1. It is clear from Figure 4 that the isoclines of Eq (3.9) intersect in the interior of the positive quadrant. Therefore, system (3.1) has at least a unique positive equilibrium.

    Figure 4.  Intersection point of isocline of system (3.1) at (S,I).

    In this subsection, the local stability of all equilibria of system (3.1) and the global stability of the positive equilibrium will be stated in the following theorems.

    Theorem 3.1. The trivial equilibrium E0 is always unstable.

    Proof. The Jacobian of system (3.1) is given by

    J=[j11j12j13j21j22j23j31j32j33], (3.11)

    where

    j11=r(12S+IK)βIμ1αP(S+α)2,j12=rSKβS+σδ(σ+I)2,j13=μ1SS+α,j21=βI,j22=βSμ2αP(I+α)2ρ1σδ(I+σ)2,j23=μ2II+α,j31=μ3αP(S+α)2θP,j32=μ4αP(I+α)2θP,j33=μ3SS+α+μ4II+αρ2θ(S+I)=0.

    The characteristic equation of system (3.1) at E0 is

    (λr)(λ+ρ1+δσ)(λ+ρ2)=0. (3.12)

    Since Eq (3.12) has a unique positive root λ=r, the trivial equilibrium E0 is always unstable.

    Theorem 3.2. If the assumption (H5):βK<ρ1+δσ, and μ3KK+α<ρ2+θK is satisfied, then E1 is stable, but it is unstable when one case of the reverse inequality holds.

    Proof. From (3.11), the characteristic equation of system (3.1) at E1 is

    (λ+r)(λ(βKρ1δσ))(λ(μ3KS+αρ2θK))=0. (3.13)

    Clearly, the characteristic values around E1 are λ1=r<0, λ2=βKρ1δσ, λ3=μ3KS+αρ2θK. If the assumption (H5) is true, then all characteristic values of Eq (3.13) have negative real parts, thus E1 is locally asymptotically stable. Conversely, if βK>ρ1+δσ or μ3KK+α>ρ2+θK, then Eq (3.13) has at least one characteristic value with the positive real part and E1 is unstable.

    We give some assumptions as follows:

    (H6):μ3ˉSˉS+α+μ4ˉIˉI+α<ρ2+θ(ˉS+ˉI),

    (H7):r(12ˉS+ˉIK)+β(ˉS+ˉI)<ρ1+σδ(ˉI+σ)2,

    (H8):r(12ˉS+ˉIK)(βˉSρ1σδ(ˉI+σ)2)+2β2ˉSˉI+rβˉSˉIK>βˉI(ρ1+2σδ(ˉI+σ)2).

    Theorem 3.3. If the assumptions (H1), (H2), (H6)(H8) are satisfied, then the equilibrium E2(ˉS,ˉI,0) is locally asymptotically stable.

    Proof. The characteristics equation of system (3.1) at the equilibrium E2 is given by

    [λ2+(ρ1+σδ(ˉI+σ)2βˉSβˉIr(12ˉS+ˉIK))λ+r(12ˉS+ˉIK)(βˉSρ1σδ(ˉI+σ)2)+2β2ˉIˉSβρ1ˉI2βˉIσδ(ˉI+σ)2+rβˉSˉIK](λμ3ˉSˉS+αμ4ˉIˉI+α+ρ2+θ(ˉS+ˉI))=0. (3.14)

    One eigenvalue of Eq (3.14) is μ3ˉSˉS+α+μ4ˉIˉI+αρ2θ(ˉS+ˉI), which is negative when the assumption (H6) holds.

    In Eq (3.14), the quadratic term gives two additional eigenvalues which satisfy the following conditions:

    λ1+λ2=r(12ˉS+ˉIK)+β(ˉS+ˉI)ρ1σδ(ˉI+σ)2,λ1λ2=r(12ˉS+ˉIK)(βˉSρ1σδ(ˉI+σ)2)+2β2ˉSˉIβρ1ˉIβˉI2σδ(ˉI+σ)2+rβˉSˉIK.

    Now, λ1+λ2<0 and λ1λ2>0 if the assumptions (H7) and (H8) hold.

    Therefore, the equilibrium E2 is locally asymptotically stable if the conditions (H6)(H8) hold.

    Theorem 3.4. If the assumptions (H3), (H9):βS<μ2Pα+ρ1+δσ, and (H10):r<2rSK+μ1αP(S+α)2,μ3αP(S+α)2>θP are true, then the equilibrium E3(S,0,P) is locally asymptotically stable.

    Proof. The Jacobian matrix corresponding to the disease-free equilibrium E3 is:

    J(E3)=[j11j12j13j21j22j23j31j32j33], (3.15)

    where

    j11=r(12SK)μ1αP(S+α)2,j12=rSKβS+δσ,j13=μ1SS+α,j21=0,j23=0,j22=βSμ2Pαρ1δσ,j31=μ3αP(S+α)2θP,j32=μ4PαθP,j33=0.

    One eigenvalue is βSμ2Pαρ1δσ>0 if the assumption (H9) holds. The others are the eigenvalue of sub-matrix

    J(E3)=[j11j13j31j33].

    The eigenvalues of sub-matrix J(E3) are negative if tr(J(E3))<0 and det(J(E3))>0, where

    tr(J(E3))=r(12SK)μ1αP(S+α)2,det(J(E3))=μ1SS+α(μ3αP(S+α)2θP).

    In fact, tr(J(E3))<0 and det(J(E3))>0 if the condition (H10) holds.

    Theorem 3.5. If the assumptions (H4) and (H11):V1>0,V2>0,V1V2V3>0 hold, then the positive equilibrium E(S,I,P) is locally asymptotically stable.

    Proof. The Jacobian matrix corresponding to the equilibrium E is given by

    J(E)=[e11e12e13e21e22e23e31e32e33].

    where eij=jij at E. Here, e13<0, e23<0 and e33=0. The characteristic equation of J(E) is given by

    λ3+V1λ2+V2λ+V3=0,

    where

    V1=(e11+e22),V2=e11e22e23e32e12e21e13e31,V3=e11e23e32+e13e31e22e12e23e31e13e21e32.

    From Routh-Hurwitz criteria, if the assumption (H11) holds, then system (3.1) is locally asymptotically stable at the positive equilibrium E.

    Next, we give the following lemma before proving the global stability of the positive equilibrium E of system (3.1).

    Lemma 3.3. If f is defined on [0,) and nonnegative such that f is integrable and uniformly continuous on [0,), then limtf(t)=0.

    Theorem 3.6. If the assumptions (H4) and (H12):o11>0,o11o22o212>0, o11o22o33+o12o23o31+o13o21o32o213o22o223o11o212o33>0 hold, then the positive equilibrium E of system (3.1) is globally asymptotically stable.

    Proof. Define the Lyapunov function for system (3.1) as

    V(S,I,P)=(SSSlnSS)+(IIIlnII)+k2(PP)2, (3.16)

    where k is an applicable positive constant.

    Differentiating Eq (3.16) with respect to time t, we have

    dVdt=SSSdSdt+IIIdIdt+k(PP)dPdt=SSS(rS(1S+IK)βSIμ1SPS+α+δIσ+I)+III(βSIμ2IPI+αρ1IδIσ+I)+k(PP)(μ3SPS+α+μ4IPI+αρ2Pθ(S+I)P)=(rK+δISS(σ+I)μ1Pψ1)(SS)2(μ2Pψ3δψ2)(II)2k(ρ2+θ(S+I)μ3SS+αμ4II+α)(PP)2+(σδSψ2rK)(SS)(II)+(kμ3αPψ1μ1S+αkθP)(SS)(PP)(μ2I+α+kθP)(II)(PP),

    where ψ1=(S+α)(S+α), ψ2=(I+σ)(I+σ), ψ3=(I+α)(I+α).

    The above expression can be reduced as

    dVdto11(SS)2o22(II)2o33(PP)2+o12(SS)(II)+o13(SS)(PP)o23(II)(PP)YTOY, (3.17)

    where YT=(|S(t)S|,|I(t)I|,|P(t)P|), and O=(oij)3×3,i,j=1,2,3. All entries of the matrix O3×3 are defined as

    o11=minS{rK+δISS(σ+I)μ1P},o22=μ2Pψ3δ,o33=k(ρ2+θ(S+I)μ3Sμ4I),o12=o21=σδψ2rK,o13=o31=12minS,P{kμ3αPμ1(S+α)kθP},o23=o32=12minI,P{μ2I+α+kθP}.

    The matrix O is positive if, and only if, all the principal minors of O are positive. If (H12) is held, then all the principal minors of O are positive.

    Therefore, from (3.17), we have that

    dVdtYTOYΛ[|S(u)S|2+|I(u)I|2+|P(u)P|2], (3.18)

    where Λ is the smallest eigenvalue. From (3.18), we have

    V+Λ[|S(u)S|2+|I(u)I|2+|P(u)P|2]V(S(0),I(0),P(0)). (3.19)

    By definition of V and (3.19), S(t), I(t), and P(t) are bounded uniformly on [0,+), and dSdt, dIdt, and dPdt are also bounded on [0,+). By Lemmas (3.3) and (3.19), we conclude that

    limt|S(t)S|2=0,limt|I(t)I|2=0,limt|P(t)P|2=0. (3.20)

    Thus, system (3.1) is globally asymptotically stable around the positive equilibrium E.

    In this section, we will investigate the stability and the existence of Hopf bifurcation of the delayed system.

    When the incubation period of the virus is constant, system (2.1) becomes as follows:

    {dSdt=rS(1S+IK)βSIμ1SPS+α+δIσ+I,dIdt=βS(tτ)I(tτ)μ2IPI+αρ1IδIσ+I,dPdt=μ3SPS+α+μ4IPI+αρ2Pθ(S+I)P. (4.1)

    Suppose ˜S(t)=S(t)S,˜I(t)=I(t)I,˜P(t)=P(t)P. For simplicity, ˜S, ˜I, and ˜P still are written as S, I, and P, respectively. Then, the linearized system at the positive equilibrium E(S,I,P) is given by

    {˙S(t)=a11S(t)+a12I(t)+a13P(t),˙I(t)=a21S(t)+a22I(t)+a23P(t)+b21S(tτ)+b22I(tτ),˙P(t)=a31S(t)+a32I(t)+a33P(t), (4.2)

    where

    a11=r(12S+IK)βIμ1αP(S+α)2,a12=rSKβS+σδ(I+σ)2,a13=μ1SS+α,a21=0,a22=μ2αP(I+α)2ρ1σδ(I+σ)2,a23=μ2II+α,a31=μ3αP(S+α)2θP,a32=μ4αP(I+α)2θP,a33=0,b21=βI,b22=βS.

    Let

    A=(a11a12a130a22a23a31a320),B=(000b21b220000).

    The characteristic equation of system (4.2) at E is

    λ3+P1λ2+P2λ+P3+(P4λ2+P5λ+P6)eλτ=0, (4.3)

    where

    P1=(a11+a22),P2=a11a22a23a32a13a31,P3=a11a23a32a12a23a31+a13a22a31,P4=b22,P5=a11b22a12b21,P6=a13a31b22a13a32b21.

    Suppose that Eq (4.3) admits a root λ=iω(ω>0). Then, substituting it in Eq (4.3) and separating real and imaginary parts, we have

    {P1ω2P3=(P6P4ω2)cos(ωτ)+P5ωsin(ωτ),ω3P2ω=(P4ω2P6)sin(ωτ)+P5ωcos(ωτ). (4.4)

    Squaring two equations of (4.4), respectively, and then summing them, we obtain

    ω6+Q1ω4+Q2ω2+Q3=0, (4.5)

    where Q1=P212P2P24,Q2=P222P1P3+2P4P6P25,Q3=P23P26.

    Choosing ω2=z, Eq (4.5) can be rewritten as

    f(z)=z3+Q1z2+Q2z+Q3=0. (4.6)

    Differentiating both sides of the equation with respect to z yields:

    f(z)=3z2+2Q1z+Q2=0,

    where the roots of f(z) can be expressed as,

    z1,2=Q1±Q213Q23.

    According to the reference [39] and the formula Fan in [40], we can get the discriminant of Eq (4.6)

    Δ=˜B4˜AˉC,

    where ˜A=Q213Q2,˜B=Q1Q29Q3,˜C=Q223Q1Q3.

    Lemma 4.1. From Eq (4.6), the following conclusions are true.

    (i) If (H13):Q30,Q213Q20 holds, then f(z) is monotonically increasing for z[0,+). Therefore, f(z) has no positive roots on [0,+), i.e., Eq (4.6) has no positive equilibria.

    (ii) If Δ>0, then Eq (4.6) only has one positive equilibrium.

    (iii) If Δ=0, then Eq (4.6) has two positive equilibria.

    (iv) If Δ<0, then Eq (4.6) has three positive equilibria.

    Without loss of generality, suppose that Eq (4.6) has three positive roots, denoted as z1z3, then Eq (4.5) has three roots: ω1=z1,ω2=z2,ω3=z3. According to Eq (4.4), we have

    τjk=1ωkarccos[(P3P1ω2k)(P4ω2kP6)+P5ω2k(ω2kP2)(P4ω2kP6)2+P25ω2k]+2πjωk,k=1,2,3,j=0,1,.

    Therefore, when τ=τjk, then ±iωk is a pair of pure imaginary roots of Eq (4.3), and the other roots have nonnegative real parts.

    Next, we will find the transversality condition for the occurrence of Hopf bifurcation. Let λ(τ)=α(τ)+iβ(τ) be the root of Eq (4.3) when τ=τjk, satisfying α(τjk)=0 and β(τjk)=ωk, k=1,2,3,j=0,1,2,. Denoting τ0=mink=1,2,3,j=0,1,{τjk}, we have the transversality condition.

    Theorem 4.1. If zk=ω2k and f(zk)0 hold, then dRe(λ(τ))dτλ=iω00.

    Proof. Differentiating Eq (4.3) with respect to τ, we have

    [dλdτ]1=(3λ2+2P1λ+P2)eλτ+2P4λ+P5λ(P4λ2+P5λ+P6)τλ. (4.7)

    Substituting λ=iω0 into Eq (4.7) and getting the real part, we obtain

    Re[dλdτ]λ=3ω40+2(P212P2P24)ω20+P222P1P3+2P4P6P25ω20((P6P4ω20)2+P25ω20)=f(zk)ω20((P6P4ω20)2+P25ω20),

    So, we have

    sign[ddτ(Re(λ))]λ=iω0=sign[Re(dλdτ)1]λ=iω0=sign{f(zk)}0,

    which is the required transversality condition.

    Based on the above analysis, the following theorem on stability of E and Hopf bifurcation can be obtained.

    Theorem 4.2. Consider the system (4.1) with τ>0.

    (i) If assumption (H13) holds, then the positive equilibrium E is locally asymptotically stable for all τ>0.

    (ii) If Δ0 or Δ<0 holds, then the positive equilibrium E is locally asymptotically stable when τ[0,τ0), but is unstable when τ>τ0.

    (iii) If all the conditions in (ii) hold and f(zk)0, then the system (4.1) undergoes a Hopf bifurcation at the positive equilibrium E when τ=τ0.

    The utilization of MMS is frequently employed for the examination of phenomena across varying temporal and spatial dimensions. In this subsection, we will apply this approach to system (2.1). The method is based on references [41,42,43,44,45,46].

    According to the previous analysis, the system has only one positive equilibrium X=E(S,I,P). To begin, it is imperative to linearize the system (4.1) at the equilibrium E, defined as ˆS(t)=S(t)S, ˆI(t)=I(t)I, ˆP(t)=P(t)P. For simplicity, ˆS, ˆI, and ˆP still are written as S, I, and P, respectively. So we obtain a differential equation about the variable X:

    ˙X(t)=F(X,Xτ), (4.8)

    where X=(S,I,P)T, Xτ=X(tτ).

    For MMS, we first define the new timescales and the time derivatives, respectively, as:

    Tk=ϵkt,ddt=k=0ϵkDk, (4.9)

    where Dk=Tk,kN{0}. To investigate Hopf bifurcation and use MMS, we need to perturb the bifurcation parameter in the system (4.8) near its critical value, that is, τ=τ0+ϵτϵ, where ϵ is a small positive parameter. According to reference [41], we assume that the solution for system (4.8) is of the following form:

    X(t)=X(T0,T1,T2,)=k=1ϵkXk(T0,T1,T2,). (4.10)

    Furthermore, adopting the different timescales, the delayed solutions of system (4.8) can be asymptotically expanded up and expressed in the following form:

    Xτ=k=1ϵkXk(T0τ0,T1ϵτ0,T2ϵ2τ0,). (4.11)

    Substituting Eqs (4.10) and (4.11) into system (4.8), using multivariate Taylor expansion, and collecting the terms by different powers of ϵ, we first consider the lowest order of ϵ,

    D0X1FXX1FXτX1τ=0, (4.12)

    where FX=FX, FXτ=FXτ. Further, we suppose that the solution of Eq (4.12) can be denoted as:

    X1=M1sin(ωT0)+N1cos(ωT0),X1τ=M1sin(ω(T0τ0))+N1cos(ω(T0τ0)), (4.13)

    where

    M1=(M11(T1,T2,),M12(T1,T2,),M13(T1,T2,))T,N1=(N11(T1,T2,),N12(T1,T2,),N13(T1,T2,))T.

    For simplicity, we will write the above equations in the following form:

    M1n(T1,T2,)=M1n,N1n(T1,T2,)=N1n,n=1,2,3.

    Substituting Eq (4.13) into Eq (4.12), we get

    ωM1cos(ωT0)ωN1sin(ωT0)FXM1sin(ωT0)FXN1cos(ωT0)FXτM1sin(ωT0ωτ0)FXτN1cos(ωT0ωτ0)=0.

    Then, we derive that M12, M13, N12, and N13 can be represented by M11 and N11, respectively. Furthermore, we get the following system of linear equations:

    M1j=α1j(M11,N11),N1j=β1j(M11,N11),j=2,3. (4.14)

    Similar to Eq (4.9), denote ddTk by Dk for k=0,1,. From Eq (4.14), we have

    D1M1i=α1i(D1M11,D1N11),D1N1i=β1i(D1M11,D1N11),i=2,3.

    Substituting Eqs (4.10) and (4.11) into system (4.8), and using Taylor expansion, we get the following equation at the second order of ϵ:

    D0X2(T0,T1,T2,)FXX2(T0,T1,T2,)FXτX2τ(T0τ0,T1,T2,)=g1(X1,X1τ),

    where

    g1(X1,X1τ)=12FXXX21+12FXτXτX21τ+FXXτX1X1τFXτ(τϵD0X1τ+τ0D1X1τ)D1X1=12FXX(M21(1cos(2ωT0))2+N21(1+cos(2ωT0))2+M1N1sin(2ωT0))+12FXτXτ(M1N1sin(2(ωT0ωτ0))+N21(1+cos(2(ωT0ωτ0)))2+M21(1cos(2(ωT0ωτ0)))2)+FXXτ(M1sin(ωT0)+N1cos(ωT0))×(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))FXτ(τϵω(M1cos(ω(T0τ0))N1sin(ω(T0τ0)))+τ0(D1M1sin(ω(T0τ0))+D1N1cos(ω(T0τ0))))D1M1sin(ωT0)D1N1cos(ωT0). (4.15)

    In order to ensure that the equation is solvable, the asymptotic oscillation terms including sin(ωT0) and cos(ωT0) must be eliminated, that is, suppose that the coefficients of sin(ωT0) and cos(ωT0) in Eq (4.15) are zero. From Eq (4.15), we can obtain the following equations:

    {D1M1=FXττϵωM1sin(ωτ0)+FXττϵωN1cos(ωτ0)FXττ0D1N1sin(ωτ0)FXττ0D1M1cos(ωτ0),D1N1=FXττϵωN1sin(ωτ0)FXττϵωM1cos(ωτ0)+FXττ0D1M1sin(ωτ0)FXττ0D1N1cos(ωτ0).

    After eliminating the asymptotic oscillation terms, the Eq (4.15) can be rewritten in the following form:

    D0X2FXX2FXτX2τ=12FXX(M21+N212+N21M212cos(2ωT0)+M1N1sin(2ωT0))+12FXτXτ(M21+N212+N21M212cos(2ω(T0τ0))+M1N1sin(2ω(T0τ0)))+FXXτM1sin(ωT0)(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))+FXXτN1cos(ωT0)(M1sin(ω(T0τ0))+N1cos(ω(T0τ0))). (4.16)

    In order to solve Eq (4.16) and satisfy the solvability conditions, the solution of Eq (4.16) is as follows:

    X2=M2sin(2ωT0)+N2cos(2ωT0)+O2,X2τ=M2sin(2ω(T0τ0))+N2cos(2ω(T0τ0))+O2,

    where X2=X2(T0,T1,T2,), M2=M2(T1,T2,T3,), N2=N2(T1,T2,T3,), O2=O2(T1,T2, T3,), and O2 can be denoted by M1 and N1, namely,

    O2=14FXX+FXτXτFX+FXτ(M21+N21).

    Therefore, O2 can be denoted by M11 and N11.

    Next, similar to ϵ and ϵ2, we repeat the above procedures and collect the terms of powers of ϵ3 from Eq (4.8),

    D0X3(T0,T1,)FXX3(T0,T1,)FXτX3τ(T0τ0,T1,)=g2(Xi,Xiτ),

    where i=1,2,

    g2=FXτ(12τ20D11X1τ+12τ2ϵD00X1τ+τϵτ0D01X1ττ0D2X1ττ0D1X2ττϵD1X1ττϵD0X2τ)+FXXX1X2+FXXτ(X1τX2+X1(X2ττ0D1X1ττϵD0X1τ))+X1τFXτXτ(X2ττ0D1X1ττϵD0X1τ)+12FXXXτX21X1τ+12FXXτXτX1X21τ+16FXτXτXτX31τ+16FXXXX31D2X1D1X2=FXτ(12τ20D11(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))+12τ2ϵ(ω2M1sin(ω(T0τ0))ω2N1cos(ω(T0τ0)))+τϵτ0(ωD1M1cos(ω(T0τ0))ωD1N1sin(ω(T0τ0)))τ0D2(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))τ0D1(M2sin(2ω(T0τ0))+N2cos(2ω(T0τ0))+O2)τϵD1(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))τϵ2ω(M2cos(2ω(T0τ0))N2sin(2ω(T0τ0)))+FXX(M1sin(ωT0)+N1cos(ωT0))(M2sin(2ωT0)+N2cos(2ωT0)+O2)+FXXτ(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))(M2sin(2ωT0)+N2cos(2ωT0)+O2)+FXXτ(M1sin(ωT0)+N1cos(ωT0))(M2sin(2ω(T0τ0))+N2cos(2ω(T0τ0))+O2)FXXτ(M1sin(ωT0))(τ0D1(M1sin(ω(T0τ0))+N1cos(ω(T0τ0))))FXXτN1cos(ωT0)(τ0D1(M1sin(ω(T0τ0))+N1cos(ω(T0τ0))))FXXτM1sin(ωT0)(τϵωM1cos(ω(T0τ0))τϵωN1sin(ω(T0τ0))))FXXτN1cos(ωT0)(τϵωM1cos(ω(T0τ0))τϵωN1sin(ω(T0τ0))))+FXτXτ(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))(M2sin(2ω(T0τ0))+N2cos(2ω(T0τ0))+O2τ0D1M1sin(ω(T0τ0))τ0D1N1cos(ω(T0τ0))τϵωM1cos(ω(T0τ0))+τϵωN1sin(ω(T0τ0)))+12FXXXτ(M1sin(ωT0)+N1cos(ωT0))2(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))+12FXXτXτ(M1sin(ωT0)+N1cos(ωT0))(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))2+16FXτXτXτ(M1sin(ω(T0τ0))+N1cos(ω(T0τ0)))3+16FXXX(M1sin(ωT0)+N1cos(ωT0))3D2(M1sin(ωT0)+N1cos(ωT0))D1(M2sin(2ωT0)+N2cos(2ωT0)+O2),

    where D00=2T20, D01=2T0T1, and D11=2T21. In general, g2 contains the Taylor expansion of the delay term Xτ, the interaction of the system states X1 and X2, and the higher order time derivative term.

    Similar to previous power, we require to eliminate the asymptotic oscillation terms to satisfy the solvability. After calculation, we obtain that the expressions of D2M11 and D2N11 are related to M11 and N11 at ϵ3, which are too complex to be showed here. Ultimately, restoring to the original timescale, we can obtain

    {dM11dt=ϵD1M11+ϵ2D2M11+,dN11dt=ϵD1N11+ϵ2D2N11+. (4.17)

    So as to conduct the normative analysis, the following polar coordinate transformations on Eq (4.17) are:

    M11=R(t)cos(φt),N11=R(t)sin(φt).

    As a consequence, when τ=τ0+ϵτϵ the amplitude frequency equation of system (4.17) can be written as

    {dRdt=p1(ϵ,τϵ)R(t)+p3(ϵ,τϵ)R3(t)+O(ϵ3),dφdt=p0(ϵ,τϵ)+p2(ϵ,τϵ)R2(t)+O(ϵ3), (4.18)

    where pi(ε,τε),i=0,1,2,3, depends on the parameters ε and τε. They describe the dynamic behavior of the system in polar coordinates, including the changes in amplitude R(t) and phase φ(t). By analyzing these coefficients, the stability and periodic behavior of the system can be understood.

    Bifurcation is a fundamental concept in dynamical systems. It can cause system instability, lead to periodic oscillations, or even chaotic behavior [24,47,48]. To enhance system performance, reliability, stability, and to address potential challenges in practical applications, many researchers have focused on bifurcation control. Generally speaking, the adopted methods of controlling bifurcations include: parameter tuning [47,49], feedback control [50,51], delayed feedback control [43,44,52], hybrid control [48,53], and dual control [54,55,56].

    As we know, the delay in system (4.1) can lead to periodic oscillations and Hopf bifurcation. Furthermore, we set τ=τ0+ϵτϵ, where ϵτϵ>0. In this subsection, we attempt to introduce a perturbation control. That is, applying time-varying perturbation to the delay achieves the goal of oscillation suppression [43,57]. Specifically, let

    τ(t)=τm+Lsin(Ωt), (4.19)

    where L and Ω represent magnitude and frequency of the perturbation, respectively, and τm=τ0+ϵτϵ. Assume that Ω and L are small, so that Lsin(Ωt) can be considered as a perturbation to τm. Consequently, the method of multiple timescales can be applied to analyze the time-varying delay system. Substituting Eq (4.19) into system (4.8), we have

    ˙X(t)=F(X,X(tτ(t))). (4.20)

    To derive the amplitude-frequency equation for system (4.17) when τ(t)=τ0+ϵτϵ+Lsin(Ωt), a re-scaling is necessary: ϵR(t)R(t). Following the procedure in Subsection 4.2, we obtain the amplitude-frequency equation as follows:

    {dRdt=p1(t)R(t)+p3(t)R3(t),dφdt=p0(t)+p2(t)R2(t), (4.21)

    where p1(t)=q0+q11sin(Ωt)+q12cos(Ωt), p0(t)=r0+r11sin(Ωt)+r12cos(Ωt), q0=q0(ϵτϵ,L,Ω), q11=q11(ϵτϵ,L,Ω), q12=q12(ϵτϵ,L,Ω), r0=r0(ϵτϵ,L,Ω), r11=r11(ϵτϵ,L,Ω), r12=r12(ϵτϵ,L,Ω). p2(t) and p3(t) are functions of ϵ.

    Because the first equation of (4.21) is only relative to R(t), the second equation depends on R(t). Therefore, we only pay attention to the first equation of (4.21):

    dRdt=p1(t)R(t)+p3(t)R3(t). (4.22)

    Note that Eq (4.22) is a Bernoulli equation with variable coefficients, and its solution can be derived by using the following lemma.

    Lemma 4.2. The solution to Eq (4.22) can be expressed analytically as follows:

    R(t)=Cet0p1(s)ds12C2t0p3(s2)e2s20p1(s1)ds1ds2, (4.23)

    where C is the initial value.

    It is worth noting that the decay of the solution of Eq (4.22) corresponds to the decay of a Bursting solution of Eq (4.8). In this subsection, a sufficient condition is derived for the solution to Eq (4.22) to decay. So, we have the following theorem.

    Theorem 4.3. A sufficient condition for the decay of solutions of system (4.21) is q0<0, where q0=q0(ϵτϵ,L,Ω).

    Proof. C>0 in Eq (4.22). Therefore, Eq (4.22) can be written as

    R(t)=1C2e2t0p1(s)ds2e2t0p1(s)dst0p3(s2)e2s20p1(s1)ds1ds2. (4.24)

    Utilizing the Fourier series expansion and combining with Eq (4.23), we get

    ˉCe2q0(t)Y1(t)+Y2(t)=2e2t0p1(s)dst0p3(s2)e2s20p1(s1)ds1ds2+ˉCe2t0p1(s)ds, (4.25)

    where ˉC=C2, Y1(t), and Y2(t) are periodic functions with a period of 2πΩ. Clearly, ˉC>0, and the long-term behavior of ˜v(t)=ˉCe2q0(t)Y1(t)+Y2(t) is governed by the sign of q0(t). When q0(t)0, ˉCe2q0(t)Y1(t) decays exponentially, thus the long-time behavior of ˜v(t) is determined by Y2(t), leading to periodic oscillation in R(t). On the other hand, when q0(t)<0, then ˜v(t) grows exponentially, which causes the oscillation of R(t)=1˜v(t) to die out exponentially over time. In summary, if the perturbation parameters of the delay satisfy certain conditions that ensure q0(t)<0, then the oscillation of system (2.1) can be controlled. Consequently, a critical boundary can be determined by solving q0(ϵτϵ,L,Ω)=0.

    Theorem 4.4. If (H4), (H14), and (H15) hold, then the positive equilibrium E of system (2.1) is globally asymptotically stable.

    Proof. Suppose a small disturbance occurs near the positive equilibrium E of the system (2.1); let w1(t)=S(t)S, w2(t)=I(t)I, w3(t)=P(t)P.

    By means of neglecting the higher-order nonlinear terms and applying a linear approximation, the following linearized system is obtained:

    {dw1dt=(12SKIKμ1PS+α+βI)w1(rSK+βS)w2μ1SS+αw3,dw2dt=βw1(tτ(t))w2(tτ(t))(ρ1+μ2PI+α)w2μ2II+αw3,dw3dt=(μ3PS+αθP)w1+(μ4PI+αθP)w2+(μ3SS+α+μ4II+αρ2θI)w3. (4.26)

    According to Lyapunov stability theory, the stability of the positive equilibrium of system (2.1) is equivalent to the stability of the zero solution of system (4.26).

    Define a Lyapunov-Krasouskii function near the positive equilibrium E as

    V(w1,w2,w3)=12(w21+w22+w23)+r12ttτ(t)w21(s)ds+r22ttτ(t)w22(s)ds, (4.27)

    where r1,r2 are appropriate positive constants.

    Taking the derivative of the above V function with respect to t, we have

    dVdt=w1dw1dt+w2dw2dt+w3dw3dt+r12(w21(1˙τ(t))w21(tτ(t)))+r22(w22(1˙τ(t))w22(tτ(t))), (4.28)

    where |˙τ(t)|δ, and δ is a positive constant. Substituting (4.26) into (4.28), we can obtain the following formulas

    dVdt=(12SKIKμ1PS+α+βI)w21(rSK+βS)w1w2μ1SS+αw1w3+βw2w1(tτ(t))w2(tτ(t))(ρ1+μ2PI+α)w22μ2II+αw2w3+(μ4PS+αθP)w1w3+(μ4PI+αθP)w2w3+(μ3SS+α+μ4II+αρ2θI)w23+r12(w21(1˙τ(t))w21(tτ(t)))+r22(w22(1˙τ(t))w22(tτ(t))).

    Using the Young inequality to deal with the cross terms in the above expression and perform scaling, we can obtain

    (4.29)

    where , , and is a symmetric matrix can be defined as:

    where

    The matrix is positive if, and only if, all the principal minors of are positive. If are positive, then all the principal minors of are positive. Further, if is true, then system (2.1) is globally asymptotically stable near the positive equilibrium .

    In this section, we have taken the same set of parameter values given in Table 1 to verify the theoretical results, and give some biological explanations. To start, we perform a sensitivity analysis on some parameters of system (3.1).

    To identify the parameters that significantly influence the output variables of system (3.1), we conduct the global sensitivity analysis on selected parameters. From system (3.1), we calculate the partial rank correlation coefficients (PRCCs) among parameters , , , , , and . Using the Latin hypercube sampling (LHS) method, we performed simulations of parameters in system (3.1). The baseline values of these parameters are presented in Table 2.

    Table 2.  The variation ranges of sensitive parameters in system (3.1).
    Parameters Basic values Minimum values Maximum values
    22 17.688 26.312
    200 160.8 239.2
    0.06 0.048 0.072
    0.05 0.041 0.06
    0.7 0.563 0.837
    0.033 0.027 0.039

     | Show Table
    DownLoad: CSV

    Based on the parameter values in Table 2, we analyze the impact of selected parameters on the correlation of the infected phytoplankton. Then, by generating scatter plots and conducting parameter samplings, we obtained the sampling results (see Figure 5) and dot plots (see Figure 6). The trend of monotonic increase (decrease) indicates the positive (negative) correlation between parameters and the model outputs. From Figure 6, these parameters show periodic correlations with system outputs, where , , and exhibit positive correlations with model outputs, demonstrates negative correlations with model outputs, and and show no significant correlation with model outputs.

    Figure 5.  Sampling results of samples of infected prey in system (3.1).
    Figure 6.  Scatter plots corresponding to different parameters of system (3.1).

    Using the parameter values in Table 1, we obtain the unique predator-free equilibrium and the positive equilibrium . For non-delayed system (3.1), the predator-free equilibrium and the positive equilibrium are locally asymptotically stable according to Theorems 3.3 and 3.5, respectively (see Figures 7 and 8). Thus, the susceptible phytoplankton, the infected phytoplankton, and zooplankton population are stable.

    Figure 7.  System (3.1) around the predator-free equilibrium is locally asymptotically stable. (a) Time series plot, (b) phase plot.
    Figure 8.  System (3.1) around the positive equilibrium is locally asymptotically stable. (a) Time series plot, (b) phase plot.

    Based on Theorem 3.6, is globally asymptotically stable (see Figure 9). This implies that regardless of variations in initial conditions, both the phytoplankton and zooplankton populations will converge to a stable equilibrium state.

    Figure 9.  System (3.1) at the positive equilibrium is globally asymptotically stable.

    For the time-varying delay system (2.1), the positive equilibrium of such a system remains unchanged by selecting the parameter values in Table 1. First, when we choose incubation period of disease , we find that system (2.1) is globally asymptotically stable at the positive equilibrium (see Figure 10), which shows that the conclusion of Theorem 4.1 is correct. However, the solution of system (2.1) is unstable when (see Figure 11). At the same time, when , the susceptible and infected phytoplankton population show periodicity (see Figure 11(a), (b)), and the zooplankton become extinct. Through numerical simulation, we find that time-varying delay is a key factor affecting the stability of system (2.1). When the value of is small enough, system (2.1) can maintain stability, which provides the possibility for taking some measures to control the widespread outbreak of the disease. On the contrary, when the time-varying delay is relatively large, the system may become unstable, and even chaos may occur. Therefore, early prevention and diagnosis are crucial for effectively preventing the spread of the disease and reducing it.

    Figure 10.  When , system (2.1) at the positive equilibrium is globally asymptotically stable.
    Figure 11.  When , system (5.1) at the positive equilibrium is unstable. (a) , (b) , (c) .

    For delayed system (4.1), the system has the same positive equilibrium . We can calculate that . Furthermore, we find that when , system (4.1) undergoes a Hopf bifurcation at , i.e., periodic solutions bifurcate from , as shown in Figure 12. According to the Theorem 4.2, we find that is locally asymptotically stable when (see Figure 13), but is unstable when (see Figure 14). That is, when the incubation period of disease transmission is less than the critical value, phytoplankton and zooplankton can coexist for a long time. When the delay exceeds the threshold, susceptible phytoplankton, infected phytoplankton, and zooplankton can still coexist at positive densities, but their densities oscillate periodically over time.

    Figure 12.  Bifurcation diagrams of system (4.1) by taking as the bifurcation parameter. (a) , (b) , (c) .
    Figure 13.  When , system (4.1) is locally asymptotically stable at the positive equilibrium . (a) , (b) , (c) .
    Figure 14.  When , the positive equilibrium of system (4.1) is unstable. (a) , (b) , (c) .

    Utilizing the method of multiple timescales discussed in Subsection 4.2, we can derive the following amplitude-frequency equation:

    (5.1)

    Multiplying both sides of the first equation of Eq (5.1) by simultaneously yields:

    (5.2)

    Let , and we can get a nontrivial steady-state solution for , that is,

    (5.3)

    By substituting (5.3) into the second equation of (5.1) and integrating it, we obtain

    where is initial phase.

    Next, we will primarily focus on the case of time-varying delay by utilizing periodic time-delay perturbation to suppress the oscillations of system. Furthermore, by employing the method of multiple scales to investigate the time-varying delay system, the following amplitude-frequency equation can be derived:

    (5.4)

    where

    and

    Assume that and . By solving , we can obtain the critical value of that induces attenuation of oscillation, which is . When exceeds the critical value (see Figure 13), we observe that the oscillation degree is significantly weakened, as shown in Figure 15. When , the oscillation gradually decays and the system eventually becomes stable (see Figure 16).

    Figure 15.  The oscillation suppression for . (a) , (b) , (c) .
    Figure 16.  The oscillation suppression for . (a) , (b) , (c) .

    The above analysis reveals that when time-varying perturbation is applied to the time delay, the oscillation of the system gradually decreases and even disappears. Thus, the oscillation suppression by periodic delay is effective in this paper. Moreover, when the disease latency delay is an approximate periodic disturbance function, the degree of instability of phytoplankton and zooplankton can gradually decrease, ultimately leading to a stable coexistence.

    To the best of our knowledge, only a few articles have incorporated time delays under different fundamental assumptions and analyzed their impact on system dynamics. For example, time delay in the viral infection process may destabilize phytoplankton density while zooplankton density remains unchanged [24], time delay in toxin liberation could destabilize an otherwise stable equilibrium [25], and a sufficiently large delay in fear-induced prey reproduction may trigger double Hopf bifurcation [30]. However, the influence of climate and environmental factors on time delays has not yet been considered. To address this gap, we propose treating delays as dynamic variables, which significantly improves the accuracy of population behavior characterization. To enhance the realism of the model, we considered the time delay as the incubation period of disease transmission, and replaced the constant delay with the time-varying delay. Thus, we developed a time-varying delay eco-epidemiological model incorporating toxicity, treatment, and Holling Ⅱ functional response in this paper.

    Then, we explored the dynamics of systems without delay, with constant delay, and with time-varying delay. First, we demonstrated the positivity and boundedness of the solution, and analyzed the local stability of five equilibriums of the non-delayed system (3.1), respectively. When there is no delay, we found that both the predator-free equilibrium and the positive equilibrium of system (3.1) are locally asymptotically stable (see Figures 7 and 8). Additionally, we investigated the global stability of the positive equilibrium by constructing an appropriate Lyapunov function (see Figure 9). Next, we analyzed the impact of time-varying delay on the stability of the system. When the time-varying delay is sufficiently small, the positive equilibrium is global asymptotically stable (see Figure 10). However, as the time-varying delay increases, the system may exhibit complex dynamic behaviors (see Figure 11). For the model with time delay, we studied the local stability of the positive equilibrium and Hopf bifurcation of system (4.1) by taking time delay as bifurcation parameter. We observed that Hopf bifurcation occurs when (see Figure 12). Our analysis revealed that when (less than the critical value 0.1093), the system remains locally asymptotically stable at the positive equilibrium (see Figure 13). However, when (greater than 0.1093), the system becomes unstable (see Figure 14). Moreover, numerical simulations were conducted using computational software to validate the theoretical findings. Our results demonstrated that time-varying delays generate rich and complex dynamics.

    We found that time delay can cause oscillations in system (4.1) through Hopf bifurcation. For our eco-epidemiological model, sustained oscillations between phytoplankton and zooplankton can undermine the stability of the system. Therefore, it is crucial to understand the factors that contribute to these oscillations and to identify strategies that can suppress the oscillation. Applying time-varying perturbation to the delay could serve as an effective control strategy to suppress oscillation, which is a delay-based control strategy. Using the method of multiple timescales, we derived the quantitative relationship between time delay and the periodic oscillation resulting from the Hopf bifurcation, as well as the critical value of the perturbation amplitude necessary to effectively control the oscillatory behavior of system. As shown in Figure 15, when , the oscillation is significantly reduced compared to those in Figure 13. Furthermore, when , the system eventually stabilizes (see Figure 16). The numerical results demonstrate that periodic perturbation of the time delay can successfully suppress the oscillations in systems. Thus, by combining the method of multiple timescales with periodic delay perturbations, we effectively suppress oscillatory behavior induced by Hopf bifurcation, providing a novel approach for stability control in time-delay systems.

    Moreover, several interesting topics could be explored in the future. For example, by considering that the release of toxins is not an instantaneous process and may be influenced by factors such as climate and temperature, we will consider another time-varying delay, denoting the delay in the release of phytoplankton toxin. Additionally, we will take into account the fear response of prey to predator and investigate its effect on prey population growth. Thus, we introduce the fear effect, where represents the degree of fear to reduce the growth of prey, and stands for reduced fear of disease transmission. Because fear of predator reduces the foraging activity of prey populations, which reduces disease transmission between prey, it is meaningful to consider the following time-varying delay eco-epidemiological model incorporating fear effect,

    The impact of the revisions on the the dynamics of model remains to be explored, and this will be discussed in the future.

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

    This work is supported by the National Natural Science Foundation of China (Grant Nos. 12161054 and 12161011), Funds for Innovative Fundamental Research Group Project of Gansu Province(Grant No. 24JRRA778), and the Doctoral Foundation of Lanzhou University of Technology.

    All authors declare no conflicts of interest in this paper.



    [1] A. J. Lotka, Analytical note on certain rythmic relations in organic systems, Proc. Natl. Acad. Sci. U.S.A., 6 (1920), 410–415. https://doi.org/10.1073/pnas.6.7.410 doi: 10.1073/pnas.6.7.410
    [2] V. Volterra, Variations and fluctuations of the number of individuals in animal species living together, in: Animal ecology, ICES J. Mar. Sci., 3 (1928), 3–51. https://doi.org/10.1093/icesjms/3.1.3
    [3] E. Beltrami, T. O. Carroll, Modelling the role of viral disease in recurrent phytoplankton blooms, J. Math. Biol., 32 (1994), 857–863.
    [4] J. Chattopadhyay, R. R. Sarkar, G. Ghosal, Removal of infected prey prevent limit cycle oscillations in an infected prey-predator system-a mathematical study, Ecol. Modell., 156 (2002), 113–121. https://doi.org/10.1016/S0304-3800(02)00133-3 doi: 10.1016/S0304-3800(02)00133-3
    [5] X. Y. Meng, N. N. Qin, H. F. Huo, Dynamics analysis of a predator-prey system with harvesting prey and disease in prey species, J. Biol. Dyn., 12 (2018), 342–374. https://doi.org/10.1080/17513758.2018.1454515 doi: 10.1080/17513758.2018.1454515
    [6] N. Sk, S. Pal, Dynamics of an infected prey-generalist predator system with the effects of fear, refuge and harvesting: Deterministic and stochastic approach, Eur. Phys. J. Plus, 137 (2022), 1–24. https://doi.org/10.1140/epjp/s13360-022-02348-9 doi: 10.1140/epjp/s13360-022-02348-9
    [7] M. X. Wang, S. W. Yao, The dynamics of an eco-epidemiological prey-predator model with infectious diseases in prey, Commun. Nonlinear Sci. Numer. Simul., 132 (2024). https://doi.org/10.1016/j.cnsns.2024.107902 doi: 10.1016/j.cnsns.2024.107902
    [8] X. Y. Meng, L. Xiao, Hopf bifurcation and Turing instability of a delayed diffusive zooplankton-phytoplankton model with hunting cooperation, Int. J. Bifurcation Chaos, 34 (2024), 2450090. https://doi.org/10.1142/S0218127424500901 doi: 10.1142/S0218127424500901
    [9] S. Akhtar, N. H. Gazi, S. Sarwardi, Mathematical modelling and bifurcation analysis of an eco-epidemiological system with multiple functional responses subjected to Allee effect and competition, Results Control Optim., 15 (2024), 100421. https://doi.org/10.1016/j.rico.2024.100421 doi: 10.1016/j.rico.2024.100421
    [10] S. X. Wu, X. Y. Meng, Dynamics of a delayed stage-structure predator-prey model with refuge and cooperation, Electron. Res. Arch., 33 (2025), 995–1036. https://doi.org/10.3934/era.2025045 doi: 10.3934/era.2025045
    [11] W. O. Kermack, A. G. McKendrick, A contribution to the mathematical theory of epidemics, Proc. R. Soc. London, Ser. A, Math. Phys. Sci., 115 (1927), 700–721. https://doi.org/10.1098/rspa.1927.0118 doi: 10.1098/rspa.1927.0118
    [12] S. Sharma, G. P. Samanta, A Leslie-Gower predator-prey model with disease in prey incorporating a prey refuge, Chaos, Solitons Fractals, 70 (2015), 69–84. https://doi.org/10.1016/j.chaos.2014.11.010 doi: 10.1016/j.chaos.2014.11.010
    [13] A. Kumar, P. K. Srivastava, A. Yadav, Delayed information induces oscillations in a dynamical model for infectious disease, Int. J. Biomath., 12 (2019), 163–196. https://doi.org/10.1142/S1793524519500207 doi: 10.1142/S1793524519500207
    [14] U. Ghosh, A. A. Thirthar, B. Mondal, P. Majumdar, Effect of fear, treatment, and hunting cooperation on an eco-epidemiological model: Memory effect in terms of fractional derivative, Iran. J. Sci. Technol. Trans. A Sci., 46 (2022), 1541–1554. https://doi.org/10.1007/s40995-022-01371-w doi: 10.1007/s40995-022-01371-w
    [15] S. C. Srivastava, N. K. Thakur, Modeling the spread and control of viral infection in damaged aquatic system: Emergence of patterns, Iran. J. Sci., 47 (2023), 467–487. https://doi.org/10.1007/s40995-023-01415-9 doi: 10.1007/s40995-023-01415-9
    [16] J. Chattopadhyay, R. R. Sarkar, A. El Abdllaoui, A delay differential equation model on harmful algal blooms in the presence of toxic substances, IMA J. Math. Appl. Med. Biol., 19 (2002), 137–161. https://doi.org/10.1093/imammb/19.2.137 doi: 10.1093/imammb/19.2.137
    [17] J. Chattopadhyay, R. R. Sarkar, S. Mandal, Toxin-producing plankton may act as a biological control for planktonic blooms-field study and mathematical modelling, J. Theor. Biol., 215 (2002), 333–344. https://doi.org/10.1006/jtbi.2001.2510 doi: 10.1006/jtbi.2001.2510
    [18] S. Gakkhar, A. Singh, A delay model for viral infection in toxin producing phytoplankton and zooplankton system, Commun. Nonlinear Sci. Numer. Simul., 15 (2010), 3607–3620. https://doi.org/10.1016/j.cnsns.2010.01.010 doi: 10.1016/j.cnsns.2010.01.010
    [19] Q. H. Huang, G. Seo, C. H. Shan, Bifurcations and global dynamics in a toxin-dependent aquatic population model, Math. Biosci., 296 (2018), 26–35. https://doi.org/10.1016/j.mbs.2017.11.013 doi: 10.1016/j.mbs.2017.11.013
    [20] C. Liu, Q. L. Zhang, J. N. Li, Global stability analysis and optimal control of a harvested eco-epidemiological prey predator model with vaccination and taxation, Abstr. Appl. Anal., 2013 (2013), 1–16. https://doi.org/10.1155/2013/950396 doi: 10.1155/2013/950396
    [21] G. Mortoja, P. Panja, K. Mondal, Dynamics of a predator-prey model with nonlinear incidence rate, Crowley-Martin type functional response and disease in prey population, Ecol. Genet. Genomics, 10 (2018), 100035. https://doi.org/10.1016/j.egg.2018.100035 doi: 10.1016/j.egg.2018.100035
    [22] M. Peng, Z. D. Zhang, C. W. Lim, X. D. Wang, Hopf bifurcation and hybrid control of a delayed ecoepidemiological model with nonlinear incidence rate and Holling type Ⅱ functional response, Math. Probl. Eng., 2018 (2018), 1–12. https://doi.org/10.1155/2018/6052503 doi: 10.1155/2018/6052503
    [23] B. N. Kahsay, O. D. Makinde, Eco-epidemiological model and optimal control analysis of tomato yellow leaf curl virus disease in tomato plant, Biotechnol. Bioeng., 2023 (2023). https://doi.org/10.1155/2023/4066236 doi: 10.1155/2023/4066236
    [24] N. K. Thakur, S. C. Srivastava, A. Ojha, Dynamical study of an eco-epidemiological delay model for plankton system with toxicity, Iran. J. Sci. Technol. Trans. A Sci., 45 (2021), 283–304. https://doi.org/10.1007/s40995-020-01042-8 doi: 10.1007/s40995-020-01042-8
    [25] K. Agnihotri, H. Kaur, The dynamics of viral infection in toxin producing phytoplankton and zooplankton system with time delay, Chaos, Solitons Fractals, 118 (2019), 122–133. https://doi.org/10.1016/j.chaos.2018.11.018 doi: 10.1016/j.chaos.2018.11.018
    [26] J. N. Lu, X. G. Zhang, R. Xu, Global stability and Hopf bifurcation of an eco-epidemiological model with time delay, Int. J. Biomath., 12 (2019), 1–21. https://doi.org/10.1142/S1793524519500621 doi: 10.1142/S1793524519500621
    [27] S. Biswas, P. K. Tiwari, S. Pal, Delay-induced chaos and its possible control in a seasonally forced eco-epidemiological model with fear effect and predator switching, Nonlinear Dyn., 104 (2021), 2901–2930. https://doi.org/10.1007/s11071-021-06396-1 doi: 10.1007/s11071-021-06396-1
    [28] S. Biswas, B. Ahmad, S. Khajanchi, Exploring dynamical complexity of a cannibalistic eco-epidemiological model with multiple time delay, Math. Methods Appl. Sci., 46 (2023), 4184–4211. https://doi.org/10.1002/mma.8749 doi: 10.1002/mma.8749
    [29] Z. W. Liang, X. Y. Meng, Stability and Hopf bifurcation of a multiple delayed predator-prey system with fear effect, prey refuge and Crowley-Martin function, Chaos, Solitons Fractals, 175 (2023), 113955. https://doi.org/10.1016/j.chaos.2023.113955 doi: 10.1016/j.chaos.2023.113955
    [30] S. C. Srivastava, N. K. Thakur, R. Singh, A. Ojha, Impact of fear and switching on a delay-induced eco-epidemiological model with Beverton-Holt functional response, Int. J. Dyn. Control, 12 (2024), 669–695. https://doi.org/10.1007/s40435-023-01216-3 doi: 10.1007/s40435-023-01216-3
    [31] S. A. Gourley, R. S. Liu, Y. J. Lou, Intra-specific competition and insect larval development: A model with time-dependent delay, Proc. R. Soc. Edinburgh Sect. A: Math., 147 (2017), 353–369. http://doi.org/10.1017/S0308210516000159 doi: 10.1017/S0308210516000159
    [32] F. X. Li, X. Q. Zhao, Dynamics of a periodic bluetongue model with a temperature-dependent incubation period, SIAM J. Appl. Math., 79 (2019), 2479–2505. https://doi.org/10.1137/18M1218364 doi: 10.1137/18M1218364
    [33] T. L. Zhang, X. Q. Zhao, Mathematical modeling for schistosomiasis with seasonal influence: A case study in Hubei, China, SIAM J. Appl. Math., 19 (2020), 1438–1471. https://doi.org/10.1137/19M1280259 doi: 10.1137/19M1280259
    [34] H. Y. Zhao, L. Shi, J. Wang, K. Wang, A stage structure HFMD model with temperature-dependent latent period, Appl. Math. Modell., 93 (2021), 745–761. https://doi.org/10.1016/j.apm.2021.01.010 doi: 10.1016/j.apm.2021.01.010
    [35] H. Wu, W. Chen, N. Wang, L. Zhang, H. Li, Z. Teng, A delayed stage-structure brucellosis model with interaction among seasonality, time-varying incubation and density-dependent growth, Int. J. Biomath., 16 (2023), 113–148. https://doi.org/10.1142/S1793524522501145 doi: 10.1142/S1793524522501145
    [36] X. Liu, X. Y. Meng, Dynamics of Bacterial white spot disease spreads in Litopenaeus Vannamei with time-varying delay, Math. Biosci. Eng., 20 (2023), 20748–20769. https://doi.org/10.3934/mbe.2023918 doi: 10.3934/mbe.2023918
    [37] C. Y. Yin, X. Y. Meng, J. M. Zuo, Modeling the effects of vaccinating strategies and periodic outbreaks on dengue in Singapore, J. Appl. Anal. Comput., 15 (2025), 1284–1309. https://doi.org/10.11948/20240205 doi: 10.11948/20240205
    [38] S. Gakkhar, K. Negi, A mathematical model for viral infection in toxin producing phytoplankton and zooplankton system, Appl. Math. Comput., 179 (2006), 301–313. https://doi.org/10.1016/j.amc.2005.11.166 doi: 10.1016/j.amc.2005.11.166
    [39] P. Yuan, L. L. Chen, M. S. You, H. Zhu, Dynamics complexity of generalist predatory mite and the leafhopper pest in tea plantations, J. Dyn. Differ. Equ., 35 (2023), 2833–2871. https://doi.org/10.1007/s10884-021-10079-1 doi: 10.1007/s10884-021-10079-1
    [40] S. Fan, A new extracting formula and a new distinguishing means on the one variable cubic equation, Natur. Sci. J. Hainan Teach. Coll., 2 (1989), 91–98.
    [41] S. L. Das, A. Chatterjee, Multiple scales without center manifold reductions for delay differential equations near Hopf bifurcations, Nonlinear Dyn., 30 (2002), 323–335. https://doi.org/10.1023/A:1021220117746 doi: 10.1023/A:1021220117746
    [42] A. H. Nayfeh, Order reduction of retarded nonlinear systems-the method of multiple scales versus center-manifold reduction, Nonlinear Dyn., 51 (2008), 483–500. https://doi.org/10.1007/s11071-007-9237-y doi: 10.1007/s11071-007-9237-y
    [43] S. Zhang, J. Xu, Oscillation control for n-dimensional congestion control model via time-varying delay, Sci. China Technol. Sci., 54 (2011), 2044–2053. https://doi.org/10.1007/s11431-011-4488-8 doi: 10.1007/s11431-011-4488-8
    [44] S. Zhang, J. Xu, Time-varying delayed feedback control for an internet congestion control model, Discrete Contin. Dyn. Syst. - Ser. B, 16 (2011), 653–668. https://doi.org/10.3934/dcdsb.2011.16.653 doi: 10.3934/dcdsb.2011.16.653
    [45] L. J. Pei, S. Wang, Double Hopf bifurcation of differential equation with linearly state-dependent delays via MMS, J. Appl. Math. Comput., 341 (2019), 256–276. https://doi.org/10.1016/j.amc.2018.08.040 doi: 10.1016/j.amc.2018.08.040
    [46] L. J. Pei, Y. M. Chen, S. Wang, Complicated oscillations and non-resonant double Hopf bifurcation of multiple feedback delayed control system of the gut microbiota, Nonlinear Anal. Real World Appl., 54 (2020), 103091. https://doi.org/10.1016/j.nonrwa.2020.103091 doi: 10.1016/j.nonrwa.2020.103091
    [47] S. Biswas, P. K. Tiwari, Delay-induced chaos and its possible control in a seasonally forced eco-epidemiological model with fear effect and predator switching, Nonlinear Dyn., 104 (2021), 2901–2930. https://doi.org/10.1007/s11071-021-06396-1 doi: 10.1007/s11071-021-06396-1
    [48] M. S. Shabbir, Q. Din, Understanding cannibalism dynamics in predator-prey interactions: Bifurcations and chaos control strategies, Qual. Theory Dyn. Syst., 23 (2024), 1–33. https://doi.org/10.1007/s12346-023-00908-7 doi: 10.1007/s12346-023-00908-7
    [49] D. Y. Li, H. Liu, H. T. Zhang, Bifurcation analysis in a predator-prey model with an Allee effect and a delayed mechanism, Acta Math. Sci., 43 (2023), 1415–1438. https://doi.org/10.1007/s10473-023-0324-z doi: 10.1007/s10473-023-0324-z
    [50] M. Z. Huang, X. Y. Song, Modeling and qualitative analysis of diabetes therapies with state feedback control, Int. J. Biomath., 7 (2014), 1450035. https://doi.org/10.1142/S1793524514500351 doi: 10.1142/S1793524514500351
    [51] V. S. Sharma, A. Singh, Strong resonance bifurcations and state feedback control in a discrete prey-predator model with harvesting effect, Qual. Theory Dyn. Syst., 22 (2023), 109. https://doi.org/10.1007/s12346-023-00805-z doi: 10.1007/s12346-023-00805-z
    [52] K. Pyragas, Delayed feedback control of chaos, Phil. Trans.: Math. Phys. Eng. Sci., 364 (2006), 2309–2334. https://doi.org/10.1098/rsta.2006.1827 doi: 10.1098/rsta.2006.1827
    [53] X. Zhang, Q. L. Zhang, Bifurcation analysis and control of a class of hybrid biological economic models, Nonlinear Anal. Hybrid Syst., 3 (2009), 578–587. https://doi.org/10.1016/j.nahs.2009.04.009 doi: 10.1016/j.nahs.2009.04.009
    [54] Z. G. Li, W. H. Chen, J. Yang, Concurrent active learning in autonomous airborne source search: Dual control for exploration and exploitation, IEEE Trans. Autom. Control, 68 (2023), 3123–3130. https://doi.org/10.1109/TAC.2022.3221907 doi: 10.1109/TAC.2022.3221907
    [55] G. Q. Tan, W. H. Chen, J. Yang, X. Tran, Z. Li, Dual control for autonomous airborne source search with Nesterov accelerated gradient descent: Algorithm and performance analysis, Neurocomputing, 630 (2025), 129729. https://doi.org/10.1016/j.neucom.2025.129729 doi: 10.1016/j.neucom.2025.129729
    [56] Z. G. Li, W. H. Chen, J. Yang, C. Liu, Cooperative active learning-based dual control for exploration and exploitation in autonomous search, IEEE Trans. Neural Networks Learn. Syst., 36 (2025), 2221–2233. https://doi.org/10.1109/TNNLS.2024.3349467 doi: 10.1109/TNNLS.2024.3349467
    [57] Y. G. Zheng, Z. H. Wang, Stability analysis of nolinear dynamics systems with slowly and periodically varying delay, Commun. Nonlinear Sci. Numer. Simul., 17 (2012), 3999–4009. https://doi.org/10.1016/j.cnsns.2012.02.026 doi: 10.1016/j.cnsns.2012.02.026
  • Reader Comments
  • © 2025 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(174) PDF downloads(24) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog