Research article Special Issues

On the stability of the diffusive and non-diffusive predator-prey system with consuming resources and disease in prey species

  • Academic editor: Yang Kuang
  • This research deals with formulating a multi-species eco-epidemiological mathematical model when the interacting species compete for the same food sources and the prey species have some infection. It is assumed that infection does not spread vertically. Infectious diseases severely affect the population dynamics of prey and predator. One of the most important factors in population dynamics is the movement of species in the habitat in search of resources or protection. The ecological influences of diffusion on the population density of both species are studied. The study also deals with the analysis of the effects of diffusion on the fixed points of the proposed model. The fixed points of the model are sorted out. The Lyapunov function is constructed for the proposed model. The fixed points of the proposed model are analyzed through the use of the Lyapunov stability criterion. It is proved that coexisting fixed points remain stable under the effects of self-diffusion, whereas, in the case of cross-diffusion, Turing instability exists conditionally. Moreover, a two-stage explicit numerical scheme is constructed, and the stability of the said scheme is found by using von Neumann stability analysis. Simulations are performed by using the constructed scheme to discuss the model's phase portraits and time-series solution. Many scenarios are discussed to display the present study's significance. The impacts of the transmission parameter 𝛾 and food resource f on the population density of species are presented in plots. It is verified that the availability of common food resources greatly influences the dynamics of such models. It is shown that all three classes, i.e., the predator, susceptible prey and infected prey, can coexist in the habitat, and this coexistence has a stable nature. Hence, in the realistic scenarios of predator-prey ecology, the results of the study show the importance of food availability for the interacting species.

    Citation: Muhammad Shoaib Arif, Kamaleldin Abodayeh, Asad Ejaz. On the stability of the diffusive and non-diffusive predator-prey system with consuming resources and disease in prey species[J]. Mathematical Biosciences and Engineering, 2023, 20(3): 5066-5093. doi: 10.3934/mbe.2023235

    Related Papers:

    [1] Churni Gupta, Necibe Tuncer, Maia Martcheva . Immuno-epidemiological co-affection model of HIV infection and opioid addiction. Mathematical Biosciences and Engineering, 2022, 19(4): 3636-3672. doi: 10.3934/mbe.2022168
    [2] Shengqiang Liu, Lin Wang . Global stability of an HIV-1 model with distributed intracellular delays and a combination therapy. Mathematical Biosciences and Engineering, 2010, 7(3): 675-685. doi: 10.3934/mbe.2010.7.675
    [3] Georgi Kapitanov, Christina Alvey, Katia Vogt-Geisse, Zhilan Feng . An age-structured model for the coupled dynamics of HIV and HSV-2. Mathematical Biosciences and Engineering, 2015, 12(4): 803-840. doi: 10.3934/mbe.2015.12.803
    [4] Nawei Chen, Shenglong Chen, Xiaoyu Li, Zhiming Li . Modelling and analysis of the HIV/AIDS epidemic with fast and slow asymptomatic infections in China from 2008 to 2021. Mathematical Biosciences and Engineering, 2023, 20(12): 20770-20794. doi: 10.3934/mbe.2023919
    [5] Nicolas Bacaër, Xamxinur Abdurahman, Jianli Ye, Pierre Auger . On the basic reproduction number R0 in sexual activity models for HIV/AIDS epidemics: Example from Yunnan, China. Mathematical Biosciences and Engineering, 2007, 4(4): 595-607. doi: 10.3934/mbe.2007.4.595
    [6] Kazunori Sato . Basic reproduction number of SEIRS model on regular lattice. Mathematical Biosciences and Engineering, 2019, 16(6): 6708-6727. doi: 10.3934/mbe.2019335
    [7] Georgi Kapitanov . A double age-structured model of the co-infection of tuberculosis and HIV. Mathematical Biosciences and Engineering, 2015, 12(1): 23-40. doi: 10.3934/mbe.2015.12.23
    [8] Xue-Zhi Li, Ji-Xuan Liu, Maia Martcheva . An age-structured two-strain epidemic model with super-infection. Mathematical Biosciences and Engineering, 2010, 7(1): 123-147. doi: 10.3934/mbe.2010.7.123
    [9] Andrew Omame, Sarafa A. Iyaniwura, Qing Han, Adeniyi Ebenezer, Nicola L. Bragazzi, Xiaoying Wang, Woldegebriel A. Woldegerima, Jude D. Kong . Dynamics of Mpox in an HIV endemic community: A mathematical modelling approach. Mathematical Biosciences and Engineering, 2025, 22(2): 225-259. doi: 10.3934/mbe.2025010
    [10] Wenshuang Li, Shaojian Cai, Xuanpei Zhai, Jianming Ou, Kuicheng Zheng, Fengying Wei, Xuerong Mao . Transmission dynamics of symptom-dependent HIV/AIDS models. Mathematical Biosciences and Engineering, 2024, 21(2): 1819-1843. doi: 10.3934/mbe.2024079
  • This research deals with formulating a multi-species eco-epidemiological mathematical model when the interacting species compete for the same food sources and the prey species have some infection. It is assumed that infection does not spread vertically. Infectious diseases severely affect the population dynamics of prey and predator. One of the most important factors in population dynamics is the movement of species in the habitat in search of resources or protection. The ecological influences of diffusion on the population density of both species are studied. The study also deals with the analysis of the effects of diffusion on the fixed points of the proposed model. The fixed points of the model are sorted out. The Lyapunov function is constructed for the proposed model. The fixed points of the proposed model are analyzed through the use of the Lyapunov stability criterion. It is proved that coexisting fixed points remain stable under the effects of self-diffusion, whereas, in the case of cross-diffusion, Turing instability exists conditionally. Moreover, a two-stage explicit numerical scheme is constructed, and the stability of the said scheme is found by using von Neumann stability analysis. Simulations are performed by using the constructed scheme to discuss the model's phase portraits and time-series solution. Many scenarios are discussed to display the present study's significance. The impacts of the transmission parameter 𝛾 and food resource f on the population density of species are presented in plots. It is verified that the availability of common food resources greatly influences the dynamics of such models. It is shown that all three classes, i.e., the predator, susceptible prey and infected prey, can coexist in the habitat, and this coexistence has a stable nature. Hence, in the realistic scenarios of predator-prey ecology, the results of the study show the importance of food availability for the interacting species.



    In the last 15 years, US deaths due to the opioid epidemic have quadrupled from nearly 12,000 in 2002 to more than 47,000 in 2017 [1]. In October 2017, the US Department of Health and Human Services declared the opioid crisis a national public health emergency [2]. The increase in injection drug use and reduction of behavioral inhibition have also contributed to the spread of infectious diseases, particularly HIV [3]. The two epidemics – opioid and HIV – are intertwined and modeling them in tandem will lead to understanding their interdependence [4].

    The HIV epidemic, which begun in 1981, has been modeled extensively both on immunological level and on epidemiological level. On within-host level typically models involve healthy and infected CD4 cells and the virus [5,6,7,8,9]. On the between-host scale, the very basic models include susceptible, infected and AIDS classes [10,11]. Multi-scale models of HIV, mostly of the nested type [12], have also received significant attention in the recent years [13,14,15,16].

    Modeling of the opioid epidemic is more recent and picked up with the increasing importance of the substance abuse disorder. Early models [17] assume opioid use can be modeled similarly to infectious disease spread and that assumption has been used in current models of heroin use [18,19,20,21]. The transition to opioid use typically starts from prescription drug misuse and modeling of that has also been drawing attention lately [22].

    Despite the importance of the HIV and the opioid epidemics and the clear interdependence between the two, relatively little modeling has been done at the interface of the two epidemics. The within-host modeling of the interplay between opioid and HIV has first drawn attention. Based on experiments in monkeys, Vaidya et al. [23,24,25,26] model the within-host dynamics of HIV and an opioid. The between host dynamics has been addressed even less. Duan et al. [27] models the two epidemics on population level and finds that the best control strategy should reduce the probability of opioid affected individuals getting HIV and should target the drug abuse epidemic. In this paper, we investigate the interplay of the two epidemics with a multi-scale model that incorporates both a within-host component and a between host component.

    The sexual contacts leading to HIV are not homogeneous across the population. Typically, few individuals in the population partake in a lot of contacts, while most of the members of the population partake in few contacts. This type of heterogeneity of contacts in HIV transmission is modeled by scale-free networks. Modeling infectious disease dynamics on networks has been also drawing attention in the recent years, resulting in a book devoted to this topic [28]. We use here a network modeling approach introduced in [29] which gives a closed form model equations [30]. In this modeling framework of networks, ODE and age structured models have been investigated but the only multi-scale model on networks in closed form seems to be [16]. Here we expand the modeling framework developed in [16] to include the opioid epidemic thus considering for the first time a multi-scale network model of two diseases. This model is in closed form and we are able to perform analysis on it.

    In Section 2 we introduce the multi-scale network model of HIV and opioid which consists of within-host component, between-host component and linking functions used to connect the two scales. In Section 3 we discuss the existence and stability of the disease-free equilibrium and compute an explicit form of the reproduction numbers. In Section 4 we focus on the existence and stability of the semi-trivial equilibria. In Section 4 we also compute the invasion numbers of the two epidemics. In Section 5 we perform simulations and consider different scenarios with differing parameter values. In Section 6 we summarize our results.

    We modify a well-known within-host model of HIV by explicitly including the opioid drug concentration C(τ) and its impact on the average susceptibility of target cells. The model takes the following form.

    {dTdτ=sdTTk(C)ViT, dTidτ=k(C)ViTδiTi,dVidτ=NvδiTicVi,dCdτ=ΛdcC, (2.1)

    Here T are the target cells, Ti are the infected target cells and Vi is the virus (HIV). Target cells are produced at rate s and cleared at rate dT. Infected cells die at a rate δi, releasing Nv viral particles at death. The clearance rate of the virus is denoted by c. Opioid is taken at doses Λ is degraded at rate dc. Infection rate of target cells by HIV in the presence of opioid is given by

    k(C)=k0+k1C(τ)C0+C(τ),

    where C0 is the half saturation constant, k0 is the transmission coefficient in the absence of opioid and k1 is a maximal increase in infection rate due to opioids.

    A shortcoming of this model is that it does not consider multiple infection routes and drug resistant viral strains. This problem could have been remedied if a model such as the one formulated in [31] was used instead, as the base model. But because of the network structure, the system is already complicated, and addition of strains and infection route would further complicate it. We believe addition of strains and infection route would not give any different insights with the network structures.

    To introduce the model, we define a complex network with size (number of nodes) N where each node is either occupied by an individual or vacant. The states for the epidemic transmission process on the network are divided into vacant state E, susceptible state S, opioid state U, infected state V, co-affected state i(τ,t) and AIDS state A. Nodes change states at rates to be introduced, and HIV transmission is governed by the network connections. A vacant state becomes susceptible state at the recruitment rate. Susceptible, opioid, infected, co-affected and AIDS states can change their state into a vacant state at natural death rate μ or at disease-induced death rates du, dv, di(τ), da, respectively. A susceptible state can be infected with HIV and change into an infected state, or can become opioid-dependent and change into opioid state. HIV and opioid states can get co-affected by adding the other disease. An opioid state or co-affected state can move to a susceptible or HIV-infected states respectively due to treatment denoted as δ. HIV-infected or co-affected states can move to the AIDS state at rates γv and γi(τ).

    For an epidemic network, degree of a node is the number of contacts the node has with other nodes. We assume that the network contacts are HIV type contacts. That is an edge between any two nodes represents a potential for transmission of HIV, either through sexual contact or intravenous drug usage. For a network with maximal degree n, the average network degree is given by

    <k>=nk=1kp(k),

    where p(k) is the probability that a randomly chosen node has degree k. That is, <k> is a constant, when n is pre-specified. This is a standard notation for average degree. For further background, [32] is a good source for network science. Empirical studies suggest that many real-life HIV networks have scale-free degree distribution p(k)kη, where 2<η<3 [33,34]. The conditional probability p(j|k) that a node with degree k is connected to a node with degree j, is given by

    p(j|k)=jp(j)<k>.

    Basically, p(j|k) gives probability that an individual with k contacts is connected to an individual with j contacts. We assume contacts that lead to opioid addiction are homogeneous (transmission of opioid addiction can be between two nodes with the same probability). With the structure of a complex network and infection age, let Sk(t), Uk(t), Vk(t), Ak(t), be the number of susceptible, opioid-dependent, HIV infected and AIDS nodes respectively with degree k at time t for k{1,2,,n}. Let ik(τ,t) be the density of co-affected nodes of degree k at time t and with infection age τ. Then we can formulate the following network multi-scale model of HIV and opioid epidemics.

    {dSk(t)dt=Λkλu(t)Sk(t)kλv(t)Sk(t)μSk(t)+δUk(t),dUk(t)dt=λu(t)Sk(t)kqvλv(t)Uk(t)(μ+du+δ)Uk(t),dVk(t)dt=kλv(t)Sk(t)quλu(t)Vk(t)(μ+dv+γv)Vk(t)+δ0σ(τ)ik(t,τ)dτ,ik(t,τ)dt+ksik(t,τ)dτ=(μ+di(τ)+γi(τ)+δσ(τ))ik(t,τ),ksik(t,0)=kqvλv(t)Uk(t)+quλu(t)Vk(t)dAk(t)dt=γvVk(t)+0γi(τ)ik(t,τ)dτ(μ+da)Ak(t). (2.2)

    The total population size of degree k is Nk(t)=Sk(t)+Uk(t)+Vk(t)+Ik(t) where Ik(t)=0ik(t,τ)dτ. The force of HIV infection, λv(t) takes into account the heterogeneous mixing in the network:

    λv(t)=1<k>nk=1kp(k)λkv(t) where λkv(t)=βkv1Vk(t)+0βkv2(τ)ik(t,τ)dτNk(t). (2.3)

    Thus λkv(t) denotes the force of infection from a node with degree k, where βkv1 is the infection rate from Vk(t) (HIV-infected node with degree k) per effective contact, and similarly, βkv2(τ) is the time-since-infection dependent infection rate from ik(τ,t) (co-affected node with degree k) per effective contact. Then the force of infection λv(t) for the heterogeneous network model is obtained by summing over all the degrees of the network λkv(t) times the probability that the node with degree k is linked to the node with degree j. This is the average force of infection from each contact, and is therefore multiplied by k in the equations (2.2) for HIV transmission to nodes of degree k. The force of opioid addiction, λu(t) is then given by,

    λu(t)=nk=1λku(t) where λku(t)=βuUk(t)+0ik(t,τ)dτNk(t). (2.4)

    Since the opioid contacts are homogeneous within the network, the force of addiction is obtained by summing over all the force of addictions of degree k, λku(t). The within-host model remains the same as in (2.1) and the linking functions are given below.

    To link the within-host and between host models, we use data [35] to determine the form of the linking of the transmission coefficient βkv2(τ) [36] to the viral load. Fitting to the data [36], we obtain the following function for βkv2(τ):

    βkv2(τ)=βk0Vri(τ)B+Vri(τ),

    where r1. Further, we use the suggested functions in [37] to link the remaining τ-dependent rates:

    di(τ)=d0(T(0)T)+d1,γi(τ)=γ0(T(0)T),σ(τ)=σ0,

    where βk0,B,d0,d1,γ0,σ0, are constants. Disease-induced death rate di and transition to AIDS rate γi do not depend explicitly on the viral load because the viral load is high during the acute HIV phase but these rates are low during this same stage.

    Equilibria are time-independent solutions of the system and often determine its long-term behavior. We find the equilibria of this system by setting the derivatives with respect to t equal to zero.

    {Λkλu(t)Sk(t)kλv(t)Sk(t)μSk(t)+δUk(t)=0,λu(t)Sk(t)kqvλv(t)Uk(t)(μ+du+δ)Uk(t)=0,kλv(t)Sk(t)quλu(t)Vk(t)(μ+dv+γv)Vk(t)+δ0σ(τ)ik(t,τ)dτ=0,ksik(t,τ)dτ=(μ+di(τ)+γi(τ)+δσ(τ))ik(t,τ),ksik(0)=kqvλv(t)Uk(t)+quλu(t)Vk(t). (3.1)

    At the disease-free equilibrium Uk(t), Vk(t) and ik(t,τ) are zero for all k. So, the disease-free equilibrium is given by

    ε0=(Λ1μ,0,0,0,,Λnμ,0,0,0).

    To determine the stability of the disease free equilibrium and to find the basic reproduction numbers of HIV infection and opioid addiction, denoted by Ru and Rv respectively, we linearize the system around the disease-free equilibrium. We take Sk(t)=S0k+xk(t), Uk(t)=uk(t), Vk(t)=vk(t), Nk(t)=N0k+nk(t) and ik(t,τ)=yk(t,τ). Then linearizing the system (2.2) takes the following form

    {dxk(t)dt=λu(t)S0kkλv(t)S0kμxk(t)+δuk(t),duk(t)dt=λu(t)S0k(μ+du+δ)uk(t),dvk(t)dt=kλv(t)S0k(μ+dv+γv)vk(t)+δ0σ(τ)yk(t,τ)dτ,yk(t,τ)dt+ksyk(t,τ)dτ=(μ+di(τ)+γi(τ)+δσ(τ))yk(t,τ),ksyk(t,0)=0 (3.2)

    where,

    λu(t)=nk=1βuuk(t)+0yk(t,τ)dτN0k,
    λv(t)=1<k>nk=1kp(k)βkv1vk(t)+0βkv2(τ)yk(t,τ)dτN0k,andN0k=S0k=Λkμ.

    We look for solutions of the form xk(t)=xk0eλt, uk(t)=uk0eλt, vk(t)=vk0eλt, yk(t,τ)=yk(τ)eλt and obtain the following eigenvalue problem,

    {(λ+μ)xk0+λu(t)S0k+kλv(t)S0kδuk0=0,(λ+μ+du+δ)uk0λu(t)S0k=0,(λ+μ+dv+γv)vk0kλv(t)S0kδ0σ(τ)yk(τ)dτ=0,ksyk(t,τ)dτ+λyk(τ)=(μ+di(τ)+γi(τ)+δσ(τ))yk(τ),ksyk(0)=0 (3.3)

    Solving the fourth equation of (3.3) we get

    yk(τ)=yk(0)π(τ)eλτks=0 (3.4)

    where

    π(τ)=e1ksτ0μ+di(ξ)+γi(ξ)+δσ(ξ)dξ.

    Substituting (3.4) in the second equation of (3.3) we have,

    uk0=S0k(λ+μ+du+δ)βunj=1uj0N0j. (3.5)

    We notice that all uk0 have the same sign. Summing both sides from 1 to n after dividing with N0k, we get

    nk=1uk0N0k=βu(λ+μ+du+δ)nk=1S0kN0knj=1uj0N0j (3.6)

    If nk=1uk0N0k=0 then uk0=0 for all k which is not the case. So cancelling nk=1uk0N0k from both sides of the equation we obtain,

    1=βu(λ+μ+du+δ)nk=1S0kN0k=nβu(λ+μ+du+δ) (3.7)

    since N0k=S0k=Λkμ. We define

    Ru=nβuμ+du+δ. (3.8)

    Substituting (3.4) in the third equation of (3.3) we have,

    vk0=kS0k(λ+μ+dv+γv)1<k>nj=1jp(j)βjv1vj0N0j. (3.9)

    We note that all vk0 have the same sign. Multiplying both sides by 1<k>nk=1kp(k)βkv1N0k and summing from 1 to n we obtain,

    1<k>nk=1kp(k)βkv1vk0N0k=1<k>nk=1k2p(k)βkv1S0kN0k(λ+μ+dv+γv)1<k>nj=1jp(j)βjv1vj0N0j. (3.10)

    If 1<k>nk=1kp(k)βkv1vk0N0k=0, then vk0=0 for all k which is not the case. So cancelling 1<k>nk=1kp(k)βkv1vk0N0k from both sides we get,

    1=1<k>nk=1k2p(k)βkv1S0kN0k(λ+μ+dv+γv)=1<k>nk=1k2p(k)βkv1(λ+μ+dv+γv). (3.11)

    So we define

    Rv=1<k>nk=1k2p(k)βkv1(μ+dv+γv). (3.12)

    Now we can prove the following theorem,

    Theorem 1. If max{Ru,Rv}<1 then the disease free equilibrium is locally asymptotically stable. If max{Ru,Rv}>1, then the disease-free equilibrium is unstable.

    Proof. Suppose

    G1(λ)=nβuλ+μ+du+δ,G2(λ)=1<k>nk=1k2p(k)βkv1(λ+μ+dv+γv). (3.13)

    Then we notice that G1(0)=Ru, G2(0)=Rv, limλG1(λ)=0, limλG2(λ)=0.

    We claim that if max{Ru,Rv}<1 then the disease free equilibrium is locally asymptotically stable, that is all the solutions of G1(λ)=1 and G2(λ)=1, have negative real parts. To show this, we proceed by way of contradiction. Suppose one of the equations G1(λ)=1 and G2(λ)=1 has a solution λ0 with (λ0)0. Then,

    1=|G1(λ0)||G1(λ0)||G1(0)|=Ru,

    or

    1=|G2(λ0)||G2(λ0)||G2(0)|=Rv.

    This is a contradiction. Hence ε0 is locally asymptotically stable when max{Ru,Rv}<1. If max{Ru,Rv}>1, let us assume, without loss of generality, Ru>1. Then G1(λ)=1 has a positive solution, λ. Thus, the disease free equilbrium is unstable.

    In this section, we prove the existence and stability of the two boundary equilibria E1 and E2 corresponding to opioid addiction and HIV transmission in a single population respectively. To obtain E1 we let Sk(t)=Sk1, Uk(t)=Uk1 and Vk(t)=0 and ik(t,τ)=0 for all k, i.e., E1=(S11,U11,0,0,,Sn1,Un1,0,0). We get the following equations

     ΛkμSk1Sk1λu(U1)+δUk1=0 Sk1λu(U1)(μ+du+δ)Uk1=0 λu(U1)=βunk=1Uk1Nk1 (4.1)

    From the second equation of (4.1) we get,

    Sk1=μ+du+δβuUk1nj=1Uj1Nj1.

    Summing both sides of the equation from k=1 to k=n, and dividing by Nk1 we get,

    nk=1Sk1Nk1=μ+du+δβu=nRu (4.2)

    We know Sk1+Uk1Nk1=1, and so

    nk=1Sk1+Uk1Nk1=n (4.3)

    i.e.,

    nk=1Uk1Nk1=n(11Ru)

    Plugging in the values in the first equation of (4.1) and solving for Uk1 we get

    Uk1=ΛkμRu1+μ+du=Λk(μ+du)(1μ(1Ru)(μ+du)) (4.4)

    Now when Ru<1, nk=1Uk1Nk1<0, which implies Uk1<0 for at least one k. When Ru>1, Uk1>0. Thus, E1 exists if and only if Ru>1.

    To obtain E2 we let Sk(t)=Sk2, Vk(t)=Vk2 and Uk(t)=0 and ik(t,τ)=0 for all k, i.e., E2=(S12,0,V12,0,,Sn2,0,Vn2,0). We get the following equations

     ΛkμSk2kSk2λv(V2)=0, kSk2λv(V2)(μ+dv+γv)Vk2=0, λv(V2)=1<k>nj=1jp(j)βjv1Vj2Nj2. (4.5)

    From the second equation of (4.5) we get

    kSk2=μ+dv+γvλv(V2)Vk2.

    Dividing both sides by Nk2 and multiplying with 1<k>nk=1kp(k)βkv1 and adding from 1 to n we get,

    1<k>nk=1k2p(k)βkv1Sk2Nk2=(μ+dv+γv). (4.6)

    We know Sk2+Vk2Nk2=1, and so

    1<k>nk=1k2p(k)βkv1(Vk2+Sk2)Nk2=(μ+dv+γv)Rv, (4.7)

    which gives us,

    1<k>nk=1k2p(k)βkv1Vk2Nk2=(μ+dv+γv)(Rv1).

    So when Rv<1, E2 does not exist. Now

    Nk2=Vk2+Sk2=Vk2+Vk2(μ+dv+γvkλv(V2)),

    and so we obtain,

    Vk2Nk2=11+μ+dv+γvkλv(V2), (4.8)
    λv(V2)=1<k>nk=1kp(k)βkv111+μ+dv+γvkλv(V2)=1<k>nk=1kp(k)βkv1kλv(V2)kλv(V2)+(μ+dv+γv). (4.9)

    λv(V2)=0 is a solution to (4.5) which gives the disease free equilibrium. Then for a boundary equilibrium, λv(V2)>0 is a root of f(λv), where

    f(λv)=1<k>nk=1k2p(k)βkv1kλv(V2)+(μ+dv+γv)1.

    As λv increases f decreases. limλvf(λv)=1. But f(0)=Rv1>0. Then if Rv>1, f(λv) has a unique zero, giving us a unique boundary equilibrium for the system.

    To find the invasion number of HIV and stability of E1 we first linearize the system (2.2) around E1. We set Sk(t)=xk(t)+Sk1, Uk(t)=uk(t)+Uk1, Vk(t)=vk(t), ik(t,τ)=yk(t,τ) and Nk(t)=nk(t)+Nk1, the system for the perturbations becomes,

    {dxk(t)dt=Sk1λu(u,y)xk(t)βunj=1Uj1Nj1+Sk1βunj=1Uj1nkN2j1μxk(t)+δuk(t)kSk1λv(v,y),duk(t)dt=Sk1λu(u,y)+xk(t)βunj=1Uj1Nj1Sk1βunj=1Uj1nkN2j1kqvUk1λv(v,y)(μ+du+δ)uk(t),dvk(t)dt=kSk1λv(v,y)quvk(t)βunj=1Uj1Nj1(μ+dv+γv)vk(t)+δ0σ(τ)yk(t,τ)dτ,yk(t,τ)dt+ksyk(t,τ)dτ=(μ+di(τ)+γi(τ)+δσ(τ))yk(t,τ),ksyk(t,0)=kqvUk1λv(v,y)+quvk(t)βunj=1Uj1Nj1, (4.10)

    where,

    λu(u,y)=βunj=1uj+0yj(t,τ)Nj1,
    λv(v,y)=1<k>nj=1jp(j)βjv1vj(t)+0βjv2(τ)yj(t,τ)dτNj1.

    We look for solutions of the form xk(t)=xkeλt, uk(t)=ukeλt, vk(t)=vkeλt, yk(τ,t)=yk(τ)eλt and obtain the following eigenvalue problem,

    {λxk=Sk1λu(u,y)xkβunj=1Uj1Nj1+Sk1βunj=1Uj1nkN2j1μxk+δukkSk1λv(v,y),λuk=Sk1λu(u,y)+xkβunj=1Uj1Nj1Sk1βunj=1Uj1nkN2j1kqvUk1λv(v,y)(μ+du+δ)uk,λvk=kSk1λv(v,y)quvkβunj=1Uj1Nj1(μ+dv+γv)vk+δ0σ(τ)yk(τ)dτ,ksyk(τ)dτ+λyk=(μ+di(τ)+γi(τ)+δσ(τ))yk(τ),ksyk(0)=kqvUk1λv(v,y)+quvkβunk=1Uk1Nk1, (4.11)

    where,

    λu(u,y)=βunj=1uj+0yj(τ)dτNj1,

    and

    λv(v,y)=1<k>nj=1jp(j)βjv1vj+0βjv2(τ)yj(τ)dτNj1. (4.12)

    Now, using the third, fourth and fifth equation of (4.11) we will compute the invasion number of HIV.

    {(λ+μ+dv+γv)vk=kSk1λv(v,y)quvkβunj=1Uj1Nj1+δ0σ(τ)yk(τ)dτ,ksyk(τ)dτ+λyk=(μ+di(τ)+γi(τ)+δσ(τ))yk(τ),ksyk(0)=kqvUk1λv(v,y)+quvkβunj=1Uj1Nj1.

    From the second equation of (4.2) we get yk(τ)=yk(0)π(τ)eλτks, where π(τ) is as defined before. Suppose K=βuqunk=1Uk1Nk1 and Q(λ)=δ[0σ(τ)π(τ)eλτksdτ]. Then from the first and third equations of (4.2) we get,

    {(λ+μ+dv+γv+K)vkQ(λ)yk(0)=kSk1λv(v,y),Kvk+ksyk(0)=kqvUk1λv(v,y). (4.13)

    Solving for vk and yk(0) we obtain,

     vk=kskSk1+kqvUk1Q(λ)ks(λ+μ+dv+γv+K)KQ(λ)λv(v,y), yk(0)=kqv(λ+μ+dv+γv+K)Uk1+KkSk1ks(λ+μ+dv+γv+K)KQ(λ)λv(v,y).

    Supplying these values in Eq (4.12), and cancelling λv(v,y) from both sides of the equation, we obtain,

    1=1<k>nj=1j2p(j)[βjv1Nj1ksjSj1+jqvUj1Q(λ)ks(λ+μ+dv+γv+K)KQ(λ)+0βjv2(τ)π(τ)dτNj1jqv(λ+μ+dv+γv+K)Uj1+KjSj1ks(λ+μ+dv+γv+K)KQ(λ)]. (4.14)

    We define

    R1vi=1<k>nj=1j2p(j)[βjv1Nj1ksjSj1+jqvUj1δ[0σ(τ)π(τ)dτ]ks(μ+dv+γv+K)Kδ[0σ(τ)π(τ)dτ]+0βjv2(τ)π(τ)dτNj1jqv(μ+dv+γv+K)Uj1+KjSj1ks(μ+dv+γv+K)Kδ[0σ(τ)π(τ)dτ]]. (4.15)

    We call R1vi the invasion reproduction number of HIV infection. Now suppose,

    Gvi(λ)=1<k>nj=1j2p(j)[βjv1Nj1ksSj1+qvUj1Q(λ)ks(λ+μ+dv+γv+K)KQ(λ)+0βjv2(τ)eλτπ(τ)dτNj1qv(λ+μ+dv+γv+K)Uj1+KSj1ks(λ+μ+dv+γv+K)KQ(λ)]

    0βjv2(τ)eλτπ(τ)dτ=βj(λ). βj(λ) is bounded above by βj(0) and Q(λ) is bounded above by Q(0). Then Gvi(0)=R1vi and limλGvi(λ)=0. Suppose (4.14) has a solution λ=x+iy with (λ)=x0 and R1vi<1. First we prove the following result.

    |qv(λ+μ+dv+γv+K)Uj1Nj1|+|KSj1Nj1||ks(λ+μ+dv+γv+K)||KQ(0)||qv(μ+dv+γv+K)Uj1Nj1|+|KSj1Nj1||ks(μ+dv+γv+K)||KQ(0)| (4.16)

    Proof. To prove (4.16) we write down the left hand side of the inequality,

    |qv(λ+μ+dv+γv+K)Uj1Nj1|+|KSj1Nj1||ks(λ+μ+dv+γv+K)||KQ(0)|=qvC1z+KC2kszKQ(0)=f(z),

    where, C1=Uj1Nj1, C2=Sj1Nj1 and z=(x+μ+dv+γv+K)2+y2. Since f(z)<0, f(z) is a decreasing function. That is when z(0,0)z(x,y), f(z(0,0))f(z(x,y)). But f(z(0,0)) is just the right hand side of (4.16). Using (4.16) we can now state the following,

    1=|Gvi(λ)|=|1<k>nj=1j2p(j)[βjv1Nj1ksSj1+qvUj1Q(λ)ks(λ+μ+dv+γv+K)KQ(λ)+βj(λ)Nj1qv(λ+μ+dv+γv+K)Uj1+KSj1ks(λ+μ+dv+γv+K)KQ(λ)]|1<k>nj=1j2p(j)[|βjv1Nj1ksSj1+qvUj1Q(0)ks(μ+dv+γv+K)KQ(0)|+|βj(0)qv(λ+μ+dv+γv+K)Uj1Nj1+KSj1Nj1ks(λ+μ+dv+γv+K)KQ(0)|]1<k>nj=1j2p(j)[|βjv1Nj1ksSj1+qvUj1Q(0)ks(μ+dv+γv+K)KQ(0)|+βj(0)|qv(λ+μ+dv+γv+K)Uj1Nj1|+|KSj1Nj1||ks(λ+μ+dv+γv+K)||KQ(0)|]1<k>nj=1j2p(j)[|βjv1Nj1ksSj1+qvUj1Q(0)ks(μ+dv+γv+K)KQ(0)|+|qv(μ+dv+γv+K)Uj1Nj1|+|KSj1Nj1||ks(μ+dv+γv+K)||KQ(0)|]=|Gvi(0)|=R1vi<1. (4.17)

    This is a contradiction. So (4.14) only has solutions with non-negative real parts when R1vi<1.

    Now let us suppose, R1vi>1. It can be shown that Gvi(λ) is decreasing. Then since Gvi(0)=R1vi>1 and limλGvi(λ)=0 for real and positive λ, (4.14) must have at least one positive root when Rvi>1. Now, from (4.1)–(4.3) we get,

    Sk1=nRuUk1n(11Ru)=Uk1Ru1,

    i.e.,

    Uk1Ru1+Uk1Nk1=Uk1Nk1(1Ru1+1)=1.

    Solving the equation we get,

     Uk1Nk1=11Ru, Sk1Nk1=1Ru. (4.18)

    To find the remaining eigenvalues, satisfying the third, fourth and fifth equation of (4.11), yj(τ)=0 and vj=0 for all j=1,2,,n. The first two equations then just reduce to

     λxk=Sk1βunj=1ujNjxkβunj=1Uj1Nj1+Sk1βunj=1Uj1nkN2j1μxk+δuk, λuk=Sk1βunj=1ujNj+xkβunj=1Uj1Nj1Sk1βunj=1Uj1nkN2j1(μ+du+δ)uk, (4.19)

    Adding the two equations and solving for nk we get

    nk=duuk(λ+μ),

    i.e.,

    xk=λ+μ+duλ+μuk.

    Replacing xk and nk in the second equation of (4.19) we get,

     (λ+μ+du+δ)uk=Sk1βunj=1ujNj1λ+μ+duλ+μukβunj=1Uj1Nj1+Sk1βunj=1Uj1N2j1duuj(λ+μ) (λ+μ+du+δ+βun(11Ru)λ+μ+duλ+μ)uk=Sk1βunj=1ujNj1(1+duUj1Nj1λ+μ). (4.20)

    Multiplying both sides of the equation 1Nk1 and summing over 1 to n,

    (λ+μ+du+δ+βun(11Ru)λ+μ+duλ+μ)nk=1ukNk1=nk=1Sk1Nk1βunj=1ujNj1(1+duUj1Nj1λ+μ). (4.21)

    nj=1ujNj1=0 implies from (4.20) all uk would be zero, which would not be of interest. nj=1ujNj10 for non-equilibrium points, and cancelling the expression on both sides, then the characteristic equation becomes,

    (λ+μ+du+δ)(λ+μ)+βu(λ+μ+du)n(11Ru)=βu(λ+μ+du(11Ru))nRu. (4.22)

    Rewriting this equation as a quadratic equation, we get

    λ2+(2μ+du+δ+βunβunRu)λ+(μ+du+δ)μ+βun(11Ru)(μ+du)βun1Ru(μ+du(11Ru))=0 (4.23)

    Simplifying the equation, we have λ2+bλ+c=0 where b=μ+βun>0 and c=(11Ru)(βun(μ+du)(μ+du+δ)du)>0 since, βun>(μ+du+δ) when Ru>1. Hence this quadratic equation has only roots with negative real parts. Combining the work above we can conclude,

    Theorem 2. The unique boundary equilibrium E1 is locally asymptotically stable if R1vi<1, and it is unstable if R1vi>1.

    To find the invasion number of opioid addiction and stability of E2 we first linearize the system (2.2) around E2. We set Sk(t)=xk(t)+Sk2, Uk(t)=uk(t), Vk(t)=vk(t)+Vk2, ik(t,τ)=yk(t,τ) and Nk(t)=nk(t)+Nk2. The system for the perturbations becomes,

    {dxk(t)dt=Sk2λ2u(u,y)μxk(t)+δuk(t)kSk2λ2v(v,y)C1kxk+kSk21<k>nj=1jp(j)βjv1Vj2  njN j22,duk(t)dt=Sk2λ2u(u,y)kqvukC1(μ+du+δ)uk(t),dvk(t)dt=kxkC1+kSk2λ2v(v,y)kSk21<k>nj=1jp(j)βjv1Vj2njN2j2quVk2λ2u(u,y)(μ+dv+γv)vk(t)+δ0σ(τ)yk(t,τ)dτ,yk(t,τ)dt+ksyk(t,τ)dτ=(μ+di(τ)+γi(τ)+δσ(τ))yk(t,τ),ksyk(t,0)=kqvC1uk+quVk2λ2u(u,y), (4.24)

    where,

    λ2u(u,y)=βunj=1uj+0yj(t,τ)Nj2
    λ2v(v,y)=1<k>nj=1jp(j)βjv1vj(t)+0βjv2(τ)yj(t,τ)dτNj2

    and

    C1=1<k>nj=1jp(j)βjv1Vj2Nj2.

    We look for solutions of the form xk(t)=xkeλt, uk(t)=ukeλt, vk(t)=vkeλt, yk(t,τ)=yk(τ)eλt and obtain the following eigenvalue problem,

    {λxk=Sk2λ2u(u,y)μxk+δukkSk2λ2v(v,y)C1kxk+kSk21<k>nj=1jp(j)βkv1Vj2njN2j2,λuk=Sk2λ2u(u,y)kqvukC1(μ+du+δ)uk,λvk=kxkC1+kSk2λ2v(v,y)kSk21<k>nj=1jp(j)βkv1Vj2njN2j2quVk2λ2u(u,y)(μ+dv+γv)vk+δ0σ(τ)yk(τ)dτ,ksyk(τ)dτ+λyk=(μ+di(τ)+γi(τ)+δσ(τ))yk(τ),ksyk(0)=kqvC1uk+quVk2λ2u(u,y). (4.25)

    From the fourth equation of (4.25) we get

    yk(τ)=yk(0)π(τ)eλksτ. (4.26)

    Let Q(λ)=0π(τ)eλksτdτ. From the second equation of (4.25) we get

    uk=Sk2λ+μ+du+δ+kqvC1βunj=1uj+yj(0)Q(λ)Nj2.

    Multiplying both sides of this equation with 1Nk2 we get,

    ukNk2=βuSk2(λ+μ+du+δ+kqvC1)Nk2nj=1uj+yj(0)Q(λ)Nj2. (4.27)

    From the fifth equation of (4.25) we get,

    ksyk(0)=kqvC1uk+quVk2βunj=1uj+yj(0)Q(λ)Nj2.

    Multiplying both sides of this equation with Q(λ)Nk2 we get,

    yk(0)Q(λ)Nk2=kqvC1Q(λ)βuSk2ks(λ+μ+du+δ+kqvC1)Nk2nj=1uj+yj(0)Q(λ)Nj2+Q(λ)quVk2βuNk2ksnj=1uj+yj(0)Q(λ)Nj2. (4.28)

    Summing both side of (4.27) and (4.28) from 1 to n and adding together we get,

    S(λ)=S(λ)[nk=1(1+kqvC1Q(λ))βuSk2ks(λ+μ+du+δ+kqvC1)Nk2+Q(λ)quβuksnk=1Vk2Nk2] (4.29)

    where, S(λ)=nj=1uj+yj(0)Q(λ)Nj2. Since S(λ)=0 implies from (4.27) uk=0 for all k, S(λ)0. We cancel S(λ) from both sides and get,

    1=nk=1(1+kqvC1Q(λ))βuSk2ks(λ+μ+du+δ+kqvC1)Nk2+Q(λ)quβuksnk=1Vk2Nk2. (4.30)

    Let Π=0π(τ)dτ. We define

    R2ui=nk=1(1+kqvC1Π)βuSk2ks(μ+du+δ+kqvC1)Nk2+Πquβuksnk=1Vk2Nk2. (4.31)

    We call R2ui the invasion reproduction number of opioid addiction. We claim that when R2ui<1 the boundary equilibrium E2 is locally asymptotically stable, that is all the roots of (4.30) have negative real parts. Suppose

    Gui(λ)=nk=1(1+kqvC1Q(λ))βuSk2ks(λ+μ+du+δ+kqvC1)Nk2+Q(λ)quβuksnk=1Vk2Nk2.

    Then Gui(0)=R2ui and limλGui(λ)=0. Assume the Eq (4.30) has roots with non-negative real part (λ)>0. The Eq (4.30) satisfies,

     1=|nk=1(1+kqvC1Q(λ))βuSk2ks(λ+μ+du+δ+kqvC1)Nk2+Q(λ)quβuksnk=1Vk2Nk2| |nk=1(1+kqvC1Q(λ))βuSk2ks(λ+μ+du+δ+kqvC1)Nk2|+|Q(λ)quβuksnk=1Vk2Nk2| |nk=1(1+kqvC1Q((λ)))βuSk2ks(λ+μ+du+δ+kqvC1)Nk2|+|Q((λ))quβuksnk=1Vk2Nk2| |nk=1(1+kqvC1Π)βuSk2ks(μ+du+δ+kqvC1)Nk2|+|Πquβuksnk=1Vk2Nk2| R2ui<1 (4.32)

    This is a contradiction. Hence all roots of (4.30) have negative real parts when R2ui<1. Now let us suppose, R2ui>1. Then since Gui(λ)<0 when λ>0, Gui(λ) is decreasing when λ>0. But we have, Gui(0)=R2ui>1 and limλGui(λ)=0. Then (4.30) has at least one positive root when R2ui>1. If λ is not a solution of characteristic equation (4.30), we have uj=0, yj(0)=0, the remaining two equations of (4.25) then just reduce to

     λxk=μxkkSk21<k>nj=1jp(j)βjv1vjNj2C1kxk+kSk21<k>nj=1jp(j)βjv1Vj2njN2j2,λvk=kxkC1+kSk21<k>nj=1jp(j)βjv1vjNj2kSk21<k>nj=1jp(j)βjv1Vj2njN2j2(μ+dv+γv)vk. (4.33)

    Adding the two equations and solving for nk we get

    nk=(dv+γv)vk(λ+μ),

    i.e.,

    xk=λ+μ+dv+γvλ+μvk.

    Using these values for nk and xk in the second equation of (4.33) we get

    (λ+μ+dv+γv+kC1(λ+μ+dv+γv)λ+μ)vk=kSk21<k>nj=1jp(j)βjv1Vj2Nj2(λ+μ+(dv+γv)Vj2Nj2λ+μ). (4.34)

    Dividing both sides by Nk2 and readjusting we obtain,

    vkNk2=kSk2Nk21(λ+μ+dv+γv)(λ+μ+kC1)1<k>nj=1jp(j)βjv1vjNj2(λ+μ+(dv+γv)Vj2Nj2). (4.35)

    Multiplying both sides of this equation with 1<k>nk=1kp(k)βkv1(λ+μ+(dv+γv)Vk2Nk2), we get,

    T(λ)=1<k>nk=1k2p(k)βkv1Sk2Nk2(λ+μ+(dv+γv)Vk2Nk2)(λ+μ+dv+γv)(λ+μ+kC1)T(λ), (4.36)

    where T(λ)=1<k>nj=1jp(j)βjv1vjNj2(λ+μ+(dv+γv)Vj2Nj2). Since T(λ)=0 implies from (4.35), vk=0 for all k, T(λ)0 and we get the following characteristic equation,

    1=1<k>nk=1k2p(k)βkv1Sk2Nk2(λ+μ+(dv+γv)Vk2Nk2)(λ+μ+dv+γv)(λ+μ+kC1). (4.37)

    From (4.8) we obtain,

    Vk2Nk2=kC1kC1+μ+dv+γv,
    Sk2Nk2=μ+dv+γvkC1+μ+dv+γv.

    Assume the Eq (4.37) has roots with non-negative real part. Using (4.9), λ0 the Eq (4.37) satisfies,

     1=|1<k>nk=1k2p(k)βkv1Sk2Nk2(λ+μ+(dv+γv)Vk2Nk2)(λ+μ+dv+γv)(λ+μ+kC1)| =|1<k>nk=1k2p(k)βkv1μ+dv+γvkC1+μ+dv+γv(λ+μ+(dv+γv)kC1kC1+μ+dv+γv)(λ+μ+dv+γv)(λ+μ+kC1)| <|1<k>nk=1k2p(k)βkv1μ+dv+γvkC1+μ+dv+γv1(λ+μ+dv+γv)| 1<k>nk=1k2p(k)βkv1μ+dv+γvkC1+μ+dv+γv1(μ+dv+γv)=1 (4.38)

    This is a contradiction. So we can state the following theorem,

    Theorem 3. The unique boundary equilibrium E2 is locally asymptotically stable if R2ui<1, and is unstable if R2ui>1.

    We present a numerical scheme for the immuno-epidemiological models (2.1) and (2.2). The within-host model, consisting of ordinary differential equations can be solved by a stiff ODE solver in MATLAB.

    For the between-host model we introduce a finite-difference method. We discretize the domain

    D={(t,τ): 0tT, 0τA}

    where A is a maximal infection age and time T<, a maximal time. We take Δt=Δτ, with ks=1, and so the points in age and line direction can be computed as,

    τm=mΔttj=jΔt.

    Setting M=[AΔt] and N=[TΔt], we obtain A=MΔt, T=NΔt. The numerical method computes approximations to the solution at the mesh points. We assume Sk(tj)=Sjk, Uk(tj)=Ujk, Vk(tj)=Vjk and ik(tj,τm)=ijm,k. We summarize the numerical method below,

    { λjv=1knk=1βv1Vjk+Mm=1Δtβv2mij+1m,kNjk,j=0,,N1; λju=nk=1βuUjk+Mm=1Δtij+1m,kNjk,j=0,,N1; Sj+1k=Sjk+ΛΔt+δUjkΔt1+kλjvΔt+λjuΔt+μΔt,j=0,,N1; Uj+1k=Ujk+λjuSj+1kΔt1+kqvλjvΔt+(μ+du+δ)Δt,j=0,,N1; Vj+1k=Vjk+kλjvSj+1kΔt+δΔtMm=1σmij+1m,kΔt1+quλjuΔt+(μ+dv+γv)Δt,j=0,,N1; ij+1m+1,k=ijm,k1+μΔt+(dm+γm+δσm)Δt,j=0,,N1; m=0,,M1, ij+10,k=kqvUj+1kλjv+quVj+1kλju,j=0,,N1; Nj+1k=Njk+Δt(ΛduUj+1k(dv+γv)Vj+1kMm=1(dm+γm)ijm,k)1+μΔt,j=0,,N1. (5.1)

    To study the coexistence equilibrium analytically is not feasible for this model. So we take the help of simulations to predict the existence of coexistence equilibrium in a scale free network scenario. We consider specific parameter values for which R2ui and R1vi are greater than 1. The simulations suggest that the coexistence equilibrium exists and is stable. Given parameter values are constant, the invasion number of HIV, R1vi, seem to show dependence on the size of the network used. With the same parameter values given in Table 3, when the network contains 200 nodes, R1vi is close to 1.4, while with 300 nodes, R1vi increases to 2.9. The invasion number of opioid addiction R1vi remains stable near the same value 1.2 when size of the network is increased from 200 nodes to 300 nodes. This is to be expected since the spread of opioid has been considered homogeneous over the network. Simulations suggesting a coexistence equilibrium are shown in Figure 1.

    Table 1.  List of parameters of the within-host model.
    Notation Meaning
    s Production rate of healthy T-cells
    dT Clearance rate of healthy T-cells
    T Number of healthy cells
    Ti Number of infected cells
    Vi Number of virions
    δi Death rate of infected cells
    c Clearance rate of virions
    k0 Transmission coeffiecient in absence of opioid
    k1 Maximal increase in infection rate due to opioid
    C0 Half saturation constant
    dc Rate of degradation of opioid

     | Show Table
    DownLoad: CSV
    Table 2.  Definitions of parameters and dependent variables of the between-host model.
    Parameter/Variable Description
    Sk(t) Number of susceptible individuals at time t with degree k 
    Vk(t) Number of HIV infected individuals at time t with degree k 
    Uk(t) Number of opioid addicted individuals at time t with degree k 
    ik(t,τ) Density of co-affected individuals with coinfection age τ at time t with degree k 
    Ak(t) Number of individuals with AIDS at time t with degree k 
    Λk Constant recruitment rate for nodes with degree k 
    βu Transmission rate of opioid addiction
    βv1 Transmission rate of HIV infection of HIV infected only individuals
    βv2(τ) Transmission rate of HIV infection of coinfected individuals
    μ Natural death rate
    du Death rate induced by opioid addiction
    dv Death rate induced by HIV infection
    di(τ) Death rate induced by coaffection at coaffection age τ 
    da Death rate induced by AIDS
    δ Recovery rate from opioid addiction
    σ(τ) Recovery rate from coinfection to HIV only
    qu Increase coefficient of HIV infection due to opioid usage
    qv Increase coefficient of opioid usage due to HIV infection
    γv Rate of transition from HIV to AIDS
    γi(τ) Rate of transition from coinfection to AIDS

     | Show Table
    DownLoad: CSV
    Table 3.  Parameter estimation results from [38].
    Parameter Estimated Value Units
    βu 0.385676 1/time
    βv1 0.0551 1/time
    k0 0.00011046 1/time
    B 15318.9 vRNA/ml
    δ 0.118227 1/time
    qu 0.867138 Unitless
    qv 30.6189 Unitless
    du 0.00817752 1/time
    dv 0.0144092 1/time
    da 1.2766e+11 1/time
    d0 2.72895e-07 ml/(time×cells)
    d1 3.4671e-06 1/time
    γv 0.0223488 1/time
    γ0 1.63927e-12 1/time
    σ 0.000270006 Unitless
    s 22843.6 CD4 count/(time×ml)
    d 0.0766824 1/time
    k1 2.02785e-05 vRNA/(CD4 count×time)
    δi 0.725266 1/time
    Nvδi 8465.63 vRNA/(CD4 count×time)

     | Show Table
    DownLoad: CSV
    Figure 1.  Simulations with the network model. In this simulation the average degree is 7.63 and R1vi=1.2173, R2ui=1.1607. The number of nodes is 100. The maximal degree is 28 but there are no occupied nodes with degree 1, 2, 3.

    We define U(t) = nk=1Uk(t) as the total opioid addicted population. Similarly we define V(t), I(t) and S(t) as the total HIV infected, total co-infected and total susceptible population respectively. The network utilized had 200 nodes. The parameters βu and βkv1 are estimated by the following formulas, βu=Ru(μ+du+δ); and βkv1=Rv(μ+dv+γv). Since the model we consider does not include treatment for HIV, we consider Rv=5.5. This estimate is an average value collected from [39], which gives an estimation of basic reproduction number of HIV in Rural South West Uganda. Estimated value of Ru according to [27] would be close to 1.1. Given the fact that a high percentage of US citizens mentioned it would be easy for someone to access opioids for illicit purposes, according to a poll conducted in 2018 (46 percent) and people who misuse opioids often get them from a family member or friend who has a prescription [40], we considered the Ru to be higher in range, around 3.25. The maximal degree n for the following simulations (except Figure 1) is 43, with number of nodes being 200.

    The two parameters in (2.2) that are of particular interest are qu and qv, which determine how much one epidemics impacts the other. That is qv determines how likely opioid users are to get infected by HIV compared to non-users, and qu determines how likely HIV infected people are to become opioid addicted compared to non-infected people. In [27] the estimated value of the qv term equivalent was 94.5, and in [38], the estimated value of qv was 30.62. Both estimates suggest a high dependence effect of opioid usage on HIV infection. In our simulations we take the lesser estimate of qv=30.62. The total co-affected population due to varying values of qv is simulated, such as 0.5qv,qv,1.5qv,2qv,2.5qv,3qv. The estimated value for qu in [38] was below 1, but HIV-infected persons are more likely to have chronic pain, receive opioid analgesic treatment, receive higher doses of opioids, and to have substance use disorders and mental illness compared with the general population, putting them at increased risk for opioid use disorder [41]. So the simulations were done with the fitted value for qu along with double, five times and ten times the fitted value. While the total number of co-affected varies according to the network, the trend seems to be similar, with qv increasing, the total number of co-affected increases (Figure 2). A definite situation of interest is when qu=4.3, the maximum value seems to be achieved when qv is close to 60, not at the highest value of approximately 90. The simulation was repeated with these values for different network sizes and provided the same result.

    Figure 2.  Figure shows total co-affected individuals for six different values of qv. Top Left: qu=0.86, Top Right: qu=1.72, Bottom Left: qu=4.3, Bottom Right: qu=8.6. The other parameters used are given in Table 3.

    A second set of simulations were performed, taking the base value of qu=1.72. The four individual cases have all other parameters and network values same, only qv is varied, from 15,31,62 and 93 respectively. While some of the cases do have permutations, over all the trend is similar, and opposite to the simulations in Figure 2. That is qu increasing causes the total number of co-affected to decrease (Figure 3). Again we notice a situation of interest, when qv=60, the maximum value seems to be achieved when qu is close to 1.72, not at the lowest value of approximately 0.86. Repeated simulation with differing sized networks did not show changes in the results.

    Figure 3.  Figure shows total co-affected individuals for six different values of qu, Top Left: qv=15, Top Right: qv=30, Bottom Left: qv=60, Bottom Right: qv=90. The other parameters used are given in Table 3.

    Another parameter of interest is δ, which denotes the rate of recovery from addiction in the epidemiological model. The estimate for that in [27] is close to 0.033, while in [38] the estimate is approximately 0.11. If successful treatment is considered without subtracting the relapses, the rate would probably be close to 0.05 [42]. To simulate for differing values of δ, we consider βu and βkv1 as constants, directly taking the values from Table 3. All the other parameter values are as mentioned in Table 3. We consider four differing situations for δ, with the value being 0.02, the fitted value 0.11, and target high values of 0.5 and δ=1. We also investigated different scenarios with differing network sizes, with number of nodes being 100, 300 and 500 respectively. Interestingly, irrespective of network size, the total number of co-affected people appeared to be higher with the value of δ increasing (Figure 4). One quite plausible explanation would be the higher recovery would decrease the number of opioid overdose deaths significantly, thereby increasing the co-affected prevalence.

    Figure 4.  Figure shows total co-affected individuals for four different values of δ, Top Left: Network Size 100, Top Right: Network Size 300, Bottom: Network size 500. The other parameters used are given in Table 3.

    We formulate a within-host model linked with a dynamic network HIV/opioid coinfection epidemiological model with demography, through epidemiological parameters. The system is described by ordinary differential equations coupled with partial differential equations in a nested fashion. The network multi-scale model here is an extension of the multi-scale model considered in [38]. The disease free equilibrium of the system always exists and is locally asymptotically stable when both the basic reproduction numbers of opioid and HIV, Ru and Rv are less than 1.

    The boundary equilibrium E1 exists when Ru is more than 1 and E2 exist when Rv is more than 1. We define the invasion reproduction numbers R1vi and R2ui. The invasion reproduction number R2ui gives the reproduction of the opioid users when the population is at the equilibrium E2, that is, when HIV infection alone is at equilibrium in the single population. The invasion reproduction number R1vi gives the reproduction of the HIV infection at the equilibrium E1, that is when the opioid transmission alone is at equilibrium in the single population. When R1vi<1, E1 is locally stable and when R1vi>1, E1 is unstable. When R2ui<1, E2 is locally stable and when R2ui>1, E2 is unstable. The model is too complicated to compute or consider the stability of an endemic equilibrium, analytically, but simulations suggest that there is an interior equilibrium potentially under the condition that both invasion numbers are larger than one.

    We use fitted parameters from [38], to perform simulations, to explore the effect of the change of the parameters qu, qv and δ. The parameters qu and qv represent the effect of one epidemic on the other, and we simulated for plausible values of qu and qv, to get estimates of co-affected prevalence. The estimate of qv, the likelihood of a heroin user to be infected with HIV, in [27] was approximately 94, and since a large number of opioid users progress to becoming heroin users, (Data from 2011 showed that an estimated 4 to 6 percent who misuse prescription opioids switch to heroin and about 80 percent of people who used heroin first misused prescription opioids) [43], chances of the qv estimate for all illicit opioid usage being close to the "heroin only" estimate is high. We notice that increase in qv causes the co-affected population to rise significantly, which indicates that control strategies focusing on reducing HIV infection among the opioid addicted population would be effective in decoupling the epidemics, corroborating with the conclusions in [27].

    To the best of our knowledge, there have not been previous models with a parameter similar to qu, that provides an estimate for the effect of HIV infection on opioid use. The simulations from our model show that increase of qu in general causes the number of co-affected to decline. That is the prevalence of co-affected people declines sharply with the increase to more HIV infected people being addicted to opioids. This does corroborate real life data, since deaths due to overdose in the US per year were aproximately 120,000 in the year 2020, and the number increased 15 percent by 2021 [44], compared to the number of deaths due to HIV being around 18,500 in 2020 [45]. We noticed significant increase in the prevalence of co-affected individuals, when the parameter δ, representing the recovery from opioid usage was increased significantly. The sharp increase points to the idea that control measures should focus more on treatment of opioid use disorder, and that uncoupling the two epidemics is a priority to prevent loss of human lives. Counseling about the dangers of opioid addiction for HIV infected people must be provided, and similarly counseling about the dangers of getting infected with HIV should be provided to reported opioid addicts.

    In summary, we have developed a novel multi-scale network model of HIV and opioid epidemics. We have analized the model and obtained conditions for HIV-only to persist or opioid-only to persist. Simulations suggest that the two epidemics can co-exist for some parameter values. Simulations further suggest that decreasing qv decreases the number of co-affected and may lead to decopuling the epidemics. Thus control measures targeted at reducing qv should be coupled with treatment of opioid affected individuals which is consistent with our previois findings.

    Maia Martcheva was partially supported by NSF DMS-1951595. Necibe Tuncer was partially supported by NSF DMS-1951626.

    The authors declare there is no conflict of interest.

    The code was written by one of the authors and is available on request by contacting Churni Gupta.



    [1] Y. Huang, F. Chen, L. Zhong, Stability analysis of a prey predator model with Holling type III response function incorporating a prey refuge, Appl. Math. Comput., 182 (2006), 672–683. https://doi.org/10.1016/j.amc.2006.04.030 doi: 10.1016/j.amc.2006.04.030
    [2] J. D. Reeve, Environmental variability, migration, and persistence in host-parasitoid systems, Am. Nat., 132 (1988), 810–836. https://doi.org/10.1086/284891 doi: 10.1086/284891
    [3] W. W. Murdoch, C. J. Briggs, R. M. Nisbet, W. S. C. Gurney, A. Stewart-Oaten, Aggregation and stability in met population models, Am. Nat., 140 (1992), 41–58. https://doi.org/10.1086/285402 doi: 10.1086/285402
    [4] M. M. Myerscough, M. Darwen, W. Hogarth, Stability, persistence and structural stability in a classical predator-prey model, Ecol. Model., 89 (1996), 31–42. https://doi.org/10.1016/0304-3800(95)00117-4 doi: 10.1016/0304-3800(95)00117-4
    [5] Q. L. Peng, L. S. Chen, Asymptotic behavior of the nonautonomous two-species Lotka-Volterra competition models, Comput. Math. Appl., 27 (1994), 53–60. https://doi.org/10.1016/0898-1221(94)90085-X doi: 10.1016/0898-1221(94)90085-X
    [6] F. Chen, Positive periodic solutions of neutral Lotka Volterra system with feedback control, Appl. Math. Comput., 162 (2005), 1279–1302. https://doi.org/10.1016/j.amc.2004.03.009 doi: 10.1016/j.amc.2004.03.009
    [7] B. Dubey, A prey-predator model with a reserved area, Nonlinear Anal. Model. Control, 12 (2007), 479–494. https://doi.org/10.15388/NA.2007.12.4.14679 doi: 10.15388/NA.2007.12.4.14679
    [8] B. Dubey, P. Chandra, P. Sinha, A model for fishery resource with reserve area, Nonlinear Anal. Real World Appl., 4 (2003), 625–637. https://doi.org/10.1016/S1468-1218(02)00082-2 doi: 10.1016/S1468-1218(02)00082-2
    [9] C. M. Roberts, N. V. Polunin, Marine reserves: simple solutions to managing complex fisheries?, Ambio, 1993 (1993), 363–368.
    [10] Y. Takeuchi, N. Adachi, Existence and bifurcation of stable equilibrium in two-prey, one-predator communities, Bull. Math. Biol., 45 (1983), 877–900. https://doi.org/10.1016/S0092-8240(83)80067-6 doi: 10.1016/S0092-8240(83)80067-6
    [11] N. Wang, M. Zhao, H. Yu, C. Dai, B. Wang, P. Wang, Bifurcation behavior analysis in a predator-prey model, Discrete Dyn. Nat. Soc., 2016 (2016), 3565316. https://doi.org/10.1155/2016/3565316 doi: 10.1155/2016/3565316
    [12] J. Wang, J. Shi, J. Wei, Predator prey system with strong Allee effect in prey, J. Math. Biol., 62 (2011), 291–331. https://doi.org/10.1007/s00285-010-0332-1 doi: 10.1007/s00285-010-0332-1
    [13] S. R. Zhou, Y. F. Liu, G. Wang, The stability of predator prey systems subject to the Allee effects, Theor. Popul. Biol., 67 (2005), 23–31. https://doi.org/10.1016/j.tpb.2004.06.007 doi: 10.1016/j.tpb.2004.06.007
    [14] M. Yavuz, N. Sene, Stability analysis and numerical computation of the fractional predator prey model with the harvesting rate, Fractal Fractional, 4 (2020), 35. https://doi.org/10.3390/fractalfract4030035 doi: 10.3390/fractalfract4030035
    [15] M. Haque, A predator prey model with disease in the predator species only, Nonlinear Anal. Real World Appl., 11 (2010), 2224–2236. https://doi.org/10.1016/j.nonrwa.2009.06.012 doi: 10.1016/j.nonrwa.2009.06.012
    [16] L. Fang, J. Wang, The global stability and pattern formations of a predator prey system with consuming resource, Appl. Math. Lett., 58 (2016), 49–55. https://doi.org/10.1016/j.aml.2016.01.020 doi: 10.1016/j.aml.2016.01.020
    [17] B. E. Ainseba, M. Bendahmane, A. Noussair, A reaction diffusion system modeling predator prey with prey-taxis, Nonlinear Anal. Real World Appl., 9 (2008), 2086–2105. https://doi.org/10.1016/j.nonrwa.2007.06.017 doi: 10.1016/j.nonrwa.2007.06.017
    [18] S. Chen, J. Yu, Stability and bifurcation on predator-prey systems with nonlocal prey competition, Discrete Contin. Dyn. Syst., 38 (2018), 43. https://doi.org/10.3934/dcds.2018002 doi: 10.3934/dcds.2018002
    [19] S. Djilali, Pattern formation of a diffusive predator prey model with herd behavior and nonlocal prey competition, Math. Methods Appl. Sci., 43 (2020), 2233–2250. https://doi.org/10.1002/mma.6036 doi: 10.1002/mma.6036
    [20] J. D. Ferreira, S. H. Da Silva, V. S. H. Rao, Stability analysis of predator prey models involving cross-diffusion, Phys. D Nonlinear Phenom., 400 (2019), 132141. https://doi.org/10.1016/j.physd.2019.06.007 doi: 10.1016/j.physd.2019.06.007
    [21] S. Kant, V. Kumar, Stability analysis of predator prey system with migrating prey and disease infection in both species, Appl. Math. Model., 42 (2017), 509–539. https://doi.org/10.1016/j.apm.2016.10.003 doi: 10.1016/j.apm.2016.10.003
    [22] K. M. Owolabi, Numerical approach to chaotic pattern formation in diffusive predator prey system with Caputo fractional operator, Numer. Methods Partial Differ. Equations, 37 (2021), 131–151. https://doi.org/10.1002/num.22522 doi: 10.1002/num.22522
    [23] D. Song, C. Li, Y. Song, Stability and cross-diffusion-driven instability in a diffusive predator prey system with hunting cooperation functional response, Nonlinear Anal. Real World Appl., 54 (2020), 103106. https://doi.org/10.1016/j.nonrwa.2020.103106 doi: 10.1016/j.nonrwa.2020.103106
    [24] T. Zhang, Y. Xing, H. Zong, M. Han, Spatio-temporal dynamics of a reaction-diffusion system for a predator prey model with hyperbolic mortality, Nonlinear Dyn., 78 (2014), 265–277. https://doi.org/10.1007/s11071-014-1438-6 doi: 10.1007/s11071-014-1438-6
    [25] U. Ghosh, S. Pal, M. Banerjee, Memory effect on Bazykin prey-predator model: Stability and bifurcation analysis, Chaos Solitons Fractals, 143 (2021), 110531. https://doi.org/10.1016/j.chaos.2020.110531 doi: 10.1016/j.chaos.2020.110531
    [26] K. M. OWolabi, Computational dynamics of predator-prey model with the power-law kernel, Results Phys., 21 (2021), 103810. https://doi.org/10.1016/j.rinp.2020.103810 doi: 10.1016/j.rinp.2020.103810
    [27] W. Shatanawi, A. Raza, M. S. Arif, M. Rafiq, M. Bibi, M. Mohsin, Essential features preserving dynamics of stochastic Dengue model, Comput. Model. Eng. Sci., 126 (2021), 201–215. https://doi.org/10.32604/cmes.2021.012111 doi: 10.32604/cmes.2021.012111
    [28] W. Shatanawi, A. Raza, M. S. Arif, K. Abodayeh, M. Rafiq, M. Bibi, Design of nonstandard computational method for stochastic susceptible–infected–treated–recovered dynamics of coronavirus model, Adv. Differ. Equations, 2020 (2020), 1–15. https://doi.org/10.1186/s13662-019-2438-0 doi: 10.1186/s13662-019-2438-0
    [29] M. S. Arif, A. Raza, M. Rafiq, M. Bibi, J. N. Abbasi, A. Nazeer, U. Javed, Numerical simulations for stochastic computer virus propagation model, Comput. Mater. Contin., 62 (2020), 6177. https://doi.org/10.32604/cmc.2020.08595 doi: 10.32604/cmc.2020.08595
    [30] S. A. Pasha, Y. Nawaz, M. S. Arif, The modified homotopy perturbation method with an auxiliary term for the nonlinear oscillator with discontinuity, J. Low Freq. Noise Vib. Active Control, 38 (2019), 13631373. https://doi.org/10.1177/0962144X18820454 doi: 10.1177/0962144X18820454
    [31] S. X. Wu, X. Y. Meng, Dynamics of a delayed predator-prey system with fear effect, herd behavior and disease in the susceptible prey, AIMS Math., 6 (2021), 3654–3685. https://doi.org/10.3934/math.2021218 doi: 10.3934/math.2021218
    [32] B. Ghanbari, On approximate solutions for a fractional prey–predator model involving the Atangana–Baleanu derivative, Adv. Differ. Equations, 2020 (2020), 1–24. https://doi.org/10.1186/s13662-019-2438-0 doi: 10.1186/s13662-019-2438-0
    [33] B. Ghanbari, On the modeling of the interaction between tumor growth and the immune system using some new fractional and fractional-fractal operators, Adv. Differ. Equations, 2020 (2020), 1–32. https://doi.org/10.1186/s13662-019-2438-0 doi: 10.1186/s13662-019-2438-0
    [34] B. Ghanbari, A fractional system of delay differential equation with nonsingular kernels in modeling hand-foot-mouth disease, Adv. Differ. Equation, 2020 (2020), 536. https://doi.org/10.1186/s13662-020-02993-3 doi: 10.1186/s13662-020-02993-3
    [35] B. Ghanbari, On novel non differentiable exact solutions to local fractional Gardner's equation using an effective technique, Math. Methods Appl. Sci., 44 (2021), 4673–4685. https://doi.org/10.1002/mma.7060 doi: 10.1002/mma.7060
    [36] B. Ghanbari, A new model for investigating the transmission of infectious diseases in a prey-predator system using a non-singular fractional derivative, Math. Methods Appl. Sci., 2021 (2021), forthcoming. https://doi.org/10.1002/mma.7412 doi: 10.1002/mma.7412
    [37] B. Ghanbari, A. Atangana, Some new edge detecting techniques based on fractional derivatives with non-local and non-singular kernels, Adv. Differ. Equation, 435 (2020). https://doi.org/10.1186/s13662-020-02890-9. doi: 10.1186/s13662-020-02890-9
    [38] S. Saha, A. Maiti, G. P. Samanta, A Michaelis–Menten predator–prey model with strong Allee effect and disease in prey incorporating prey refuge, Int. J. Bifurcation Chaos, 28 (2018), 1850073. https://doi.org/10.1142/S0218127418500736 doi: 10.1142/S0218127418500736
    [39] S. Saha, G. P. Samanta, Analysis of a predator–prey model with herd behavior and disease in prey incorporating prey refuge, Int. J. Biomath., 12 (2019), 1950007. https://doi.org/10.1142/S1793524519500074 doi: 10.1142/S1793524519500074
    [40] S. Saha, G. P. Samanta, A prey–predator system with disease in prey and cooperative hunting strategy in predator, J. Phys. A Math. Theor., 53 (2020), 485601. https://doi.org/10.1088/1751-8121/abbc7b doi: 10.1088/1751-8121/abbc7b
    [41] A. Mondal, A. K. Pal, G. P. Samanta, On the dynamics of evolutionary Leslie-Gower predator-prey eco-epidemiological model with disease in predator, Ecol. Genet. Genomics, 10 (2019), 100034. https://doi.org/10.1016/j.egg.2018.11.002 doi: 10.1016/j.egg.2018.11.002
    [42] S. Sharma, G. P. Samanta, Analysis of a two prey one predator system with disease in the first prey population, Int. J. Dyn. Control, 3 (2015), 210–224. https://doi.org/10.1007/s40435-014-0107-4 doi: 10.1007/s40435-014-0107-4
    [43] 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
    [44] B. Ghanbari, Chaotic behaviors of the prevalence of an infectious disease in a prey and predator system using fractional derivatives, Math. Methods Appl. Sci., 44 (2021), 9998–10013. https://doi.org/10.1002/mma.7386 doi: 10.1002/mma.7386
    [45] A. Ejaz, Y. Nawaz, M. S. Arif, D. S. Mashat, K. Abodayeh, Stability analysis of predator-prey system with consuming resource and disease in predator species, CMES Comput. Model. Eng. Sci., 132 (2022), 489–506. https://doi.org/10.32604/cmes.2022.019440 doi: 10.32604/cmes.2022.019440
    [46] M. S. Arif, K. Abodayeh, A. Ejaz, Computational modeling of reaction-diffusion COVID-19 model having isolated compartment, CMES Comput. Model. Eng. Sci., 2022 (2022), 1–25.
    [47] Y. Peng, G. Zhang, Dynamics analysis of a predator prey model with herd behavior and nonlocal prey competition, Math. Comput. Simul., 170 (2020), 366–378. https://doi.org/10.1016/j.matcom.2019.11.012 doi: 10.1016/j.matcom.2019.11.012
  • This article has been cited by:

    1. Chelsea Spence, Mary E. Kurz, Thomas C. Sharkey, Bryan Lee Miller, Scoping Literature Review of Disease Modeling of the Opioid Crisis, 2024, 0279-1072, 1, 10.1080/02791072.2024.2367617
    2. Eric Numfor, Necibe Tuncer, Maia Martcheva, Optimal control of a multi-scale HIV-opioid model, 2024, 18, 1751-3758, 10.1080/17513758.2024.2317245
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(1894) PDF downloads(98) Cited by(3)

Figures and Tables

Figures(12)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog