Research article Special Issues

An SIS epidemic model with individual variation

  • We study an extension of the stochastic SIS (Susceptible-Infectious-Susceptible) model in continuous time that accounts for variation amongst individuals. By examining its limiting behaviour as the population size grows we are able to exhibit conditions for the infection to become endemic.

    Citation: Philip K. Pollett. An SIS epidemic model with individual variation[J]. Mathematical Biosciences and Engineering, 2024, 21(4): 5446-5455. doi: 10.3934/mbe.2024240

    Related Papers:

    [1] Jianquan Li, Zhien Ma, Fred Brauer . Global analysis of discrete-time SI and SIS epidemic models. Mathematical Biosciences and Engineering, 2007, 4(4): 699-710. doi: 10.3934/mbe.2007.4.699
    [2] Meici Sun, Qiming Liu . An SIS epidemic model with time delay and stochastic perturbation on heterogeneous networks. Mathematical Biosciences and Engineering, 2021, 18(5): 6790-6805. doi: 10.3934/mbe.2021337
    [3] Peter Rashkov, Ezio Venturino, Maira Aguiar, Nico Stollenwerk, Bob W. Kooi . On the role of vector modeling in a minimalistic epidemic model. Mathematical Biosciences and Engineering, 2019, 16(5): 4314-4338. doi: 10.3934/mbe.2019215
    [4] Xin-You Meng, Tao Zhang . The impact of media on the spatiotemporal pattern dynamics of a reaction-diffusion epidemic model. Mathematical Biosciences and Engineering, 2020, 17(4): 4034-4047. doi: 10.3934/mbe.2020223
    [5] Jing Li, Zhen Jin, Yuan Yuan . Effect of adaptive rewiring delay in an SIS network epidemic model. Mathematical Biosciences and Engineering, 2019, 16(6): 8092-8108. doi: 10.3934/mbe.2019407
    [6] Carlos M. Hernández-Suárez, Carlos Castillo-Chavez, Osval Montesinos López, Karla Hernández-Cuevas . An application of queuing theory to SIS and SEIS epidemic models. Mathematical Biosciences and Engineering, 2010, 7(4): 809-823. doi: 10.3934/mbe.2010.7.809
    [7] Joan Ponce, Horst R. Thieme . Can infectious diseases eradicate host species? The effect of infection-age structure. Mathematical Biosciences and Engineering, 2023, 20(10): 18717-18760. doi: 10.3934/mbe.2023830
    [8] Marcin Choiński, Mariusz Bodzioch, Urszula Foryś . A non-standard discretized SIS model of epidemics. Mathematical Biosciences and Engineering, 2022, 19(1): 115-133. doi: 10.3934/mbe.2022006
    [9] Liqiong Pu, Zhigui Lin . A diffusive SIS epidemic model in a heterogeneous and periodically evolvingenvironment. Mathematical Biosciences and Engineering, 2019, 16(4): 3094-3110. doi: 10.3934/mbe.2019153
    [10] Ke Guo, Wanbiao Ma . Global dynamics of an SI epidemic model with nonlinear incidence rate, feedback controls and time delays. Mathematical Biosciences and Engineering, 2021, 18(1): 643-672. doi: 10.3934/mbe.2021035
  • We study an extension of the stochastic SIS (Susceptible-Infectious-Susceptible) model in continuous time that accounts for variation amongst individuals. By examining its limiting behaviour as the population size grows we are able to exhibit conditions for the infection to become endemic.



    The SIS (Susceptible-Infectious-Susceptible) Model was introduced by Weiss and Dishon [1] to study infections in a closed population of n individuals, infections that do not confer any long lasting immunity (gonorrhea, or the common cold, for example). If Y(t) is the number of infectives at time t, then (Y(t),t0) is a continuous-time Markov chain taking values in {0,1,,n} with transitions

    YY+1at rateλnY(nY)(infection)YY1at rateμY(recovery).

    The large-n behaviour of this model is well understood, because the proportion of infectives Y(t)/n obeys the following law of large numbers, which is an application of Theorem 3 of [2]; to the best of my knowledge, this result has not been stated explicitly elsewhere.

    Theorem 1. If Y(0)/ny0 as n, then (Y(t)/n) converges in probability, uniformly over finite time intervals, to the solution of the ODE

    dydt=λy(1ρy), (1.1)

    where ρ=μ/λ, namely

    y(t)=(1ρ)y0y0+(1ρy0)eλ(1ρ)t,y(0)=y0.

    We see immediately that if λ>μ then (1.1) has a globally stable equilibrium at y=1ρ, while if λ<μ the disease-free state 0 is globally stable. The two behaviours are illustrated in Figure 1: quasi stationarity (the infection becomes endemic) versus evanescence (the infection dies out). Notice how the sample paths of the process (Y(t)/n) (blue) "track" the limiting deterministic path y(t) (green).

    Figure 1.  Simulation of the SIS model with n=50 individuals, and with infection and recovery rates (a) λ=10, μ=2 (ρ<1, y=0.8), and (b) λ=2, μ=10 (ρ>1).

    Several elaborations of the basic SIS model have been proposed where the assumption of homogeneity among individuals is relaxed. Perhaps the most is extensive is the study of epidemics where transmission occurs within households [3,4,5,6,7,8]. Another [9] assumes that individuals encounter one another in a communal meeting place. We incorporate heterogeneity by assuming that individuals have their own, possibly different, infection and recovery rates. In particular, we assume that individual i (i=1,,n) has an exponentially distributed recovery period with mean μ1i and a resistance level λ1i (both finite). Let X(n)i be 1 or 0 according to whether individual i is infected or not, and let X(n)i=(X(n)1,,X(n)n) be the state of the population. Suppose that (X(n)i(t),t0) is a continuous-time Markov chain taking values in {0,1}n with transitions

    (,0,)(,1,)at rateλiˉX(n)i(,1,)(,0,)at rateμi,↑↑ Position i (i=1,,n)  (1.2)

    where ˉX(n)i=1nnj=1X(n)j is the proportion of the population that is infected. So we assume here that each infected individual makes the same contribution to the infection potential of the group of infectives. The model in described in [9] has μi=0 and λi to be interpreted as the proportion of time individual i spends in the communal meeting place. Daley et al. [13] also consider the case where μi=0 for all i and study the mean duration of the epidemic.

    Notice that the disease free state 0=(0,0,,0) is the sole absorbing state and, since (λi) and (μi) are strictly positive, the remaining states form a communicating class from which 0 is accessible, and indeed reached with probability 1. None-the-less, the process may exhibit quasi stationarity. By investigating the large-n behaviour of the model, we can determine conditions for such a quasi equilibrium to be achieved. We will see that accounting for variation amongst individuals can lead to different conclusions about equilibrium behaviour to those reached for the standard SIS model. Several examples are given to show a range of behaviour. For example, the model presented here may predict quasi stationarity in cases where the standard model predicts evanescence.

    Because the population is now heterogeneous, a quite different approach is needed. We will follow McVinish and Pollett [10], who considered a setup that also encapsulates metapopulation models, and other population models. Think of the individual characteristics θi:=(λi,μi) as (random) points in a subset S of R2+, and define sequences of random measures (σ(n)i) and random-measure-valued processes (m(n)t,t0) by

    σ(n)i(B)=#{θiB}/nandm(n)t(B)=#{θiB:X(n)i,t=1}/n,

    where B is any Borel subset of S. We will suppose that σ(n)idσ (weak convergence) for some non-random (probability) measure σ, and then hope to establish that m(n)tdm(n)t, so that properties of X(n)t can be approximated for large n. In particular, if the individual characteristics were chosen from some non-negative bivariate distribution, then σ(n)idσ would follow as a consequence of the law of large numbers.

    The sequences (σ(n)i) and (m(n)t) may be defined, equivalently, by

    h(θ)σ(n)i(dθ)=1nni=1h(θi)andh(θ)m(n)t(dθ)=1nni=1X(n)i,th(θi), (2.1)

    for h in Cb(S), the class of bounded continuous functions that map S to R. For example if (h1) then m(n)t(S)=m(n)t(dθ)=1nni=1X(n)i,t, the proportion of population that is infected at time t.

    Our main result is a consequence of Theorem 1 of [10].

    Theorem 2. Suppose that σ(n)idσ and m(n)0dm0 for some non-random measures σ and m0. Then, the sequence of measure-valued processes (m(n)t,t0) converges weakly to the unique solution (mt,t0) of

    (h,mt)=(h,m0)+t0L(h,ms)ds,hCb(S), (2.2)

    where (notation) (h,m)=h(θ)m(dθ), and

    L(h,mt):=mt(S)(λh(θ)σ(dθ)λh(θ)mt(dθ))μh(θ)mt(dθ).

    Lemma 5 of [10] implies that mt(B)σ(B), for all Borel sets B, and in particular that mt is absolutely continuous with respect to σ. It follows that mt has a (uniquely determined σ-a.e.) Radon-Nikodym derivative ϕt (0) with respect to σ: mt(B)=Bϕt(θ)σ(dθ). And further it implies that ϕt1.

    Now we may differentiate both sides of (2.2) with respect to σ to show that ϕt(λ,μ) must satisfy

    ϕt=ϕ0+t0(λ(1ϕs)ϕs(θ)σ(dθ)μϕs)ds,

    or equivalently,

    dϕdt=λ(1ϕt)ϕt(θ)σ(dθ)μϕt. (2.3)

    Note that ϕt(θ) can be interpreted as the limiting probability that an individual with characteristics θ=(λ,μ) is infected at time t.

    Equation (2.3) can be used to study the long-term (t) behaviour of our model. For example, any equilibrium ϕeq of (2.3) must satisfy 0=λ(1ϕeq)ϕeq(θ)σ(dθ)μϕeq. On setting ψ=ϕeq(θ)σ(dθ), we see that

    ϕeq(λ,μ) ( =ϕeq(θ) )=λψλψ+μ,

    and so, on integrating this over (λ,μ)S, we find that ψ must solve the equation

    ψ=R(ψ):=λψλψ+μσ(dλ,dμ). (3.1)

    The following result characterizes the large-n equilibrium behaviour of our model in terms of r0:=R(0+) (). It follows from Theorems 2, 3 and 4 of [10], noting that their Condition D holds in the present setting. Note also that since the integrand in (3.1) is less that 1, and σ is a probability measure, we can differentiate under the integral to obtain

    R(ψ)=λμ(λψ+μ)2σ(dλ,dμ),

    which implies that r0=(λ/μ)σ(dλ,dμ).

    Theorem 3. (a) If r01, then ψ=0 is the only fixed point of R, and ϕeq=0 is globally stable, that is, for all ϕ0, ϕt0 on S. The latter entails mt(B)0, for all B, and hence the disease free state is globally stable.

    (b) If r0>1, then R has two fixed points, 0 and a positive fixed point ψ, and if (m0(S)=) (ϕ0,σ)>0, then

    ϕtϕ:=λψλψ+μ.

    The latter entails mt(B)m(B), for all B, where

    m(B)=Bϕ(θ)σ(dθ)=Bλψλψ+μσ(dλ,dμ),

    in particular m(S)=ψ, thus implying quasi stationarity.

    Example 1. Suppose that the individual characteristics are chosen independently from Γ(aλ,rλ) and Γ(aμ,rμ) distributions, respectively. Notation: rλ and rμ are rate parameters. We have that

    r0=aλrλrμaμ1.

    Figure 2 illustrates (a) quasi-stationary behaviour, and (b) evanescence. In (a), the green line is at ψ=0.8035, very close to 0.8 (dashed red), which is the equilibrium of the homogeneous model using λ=10 and μ=2. Figure 3 also illustrates quasi-stationary behaviour, now with ψ=0.2435, but the corresponding homogeneous model (λ=1.4 and μ=1.5) predicts evanescence. This disparity between the two models happens when

    aμ1rμ<aλrλ<aμrμ.
    Figure 2.  Simulation of the SIS model with n=50 individuals, and with independent Gamma distributed infection and recovery rates.
    Figure 3.  Simulation of the SIS model with n=50 individuals, and with independent Gamma distributed infection and recovery rates.

    Note that ψ was evaluated by way of fixed point iteration of (3.1) using Matlab's integral2, which evaluates double integrals numerically.

    Example 2. Suppose that the individual characteristics are chosen from a bivariate Gamma distribution of the form given in Theorem 2 of [11]. The marginals are Gamma distributions with a common rate parameter, which we denote by r, and shape parameters aλ and aμ, respectively, and correlation ρ=aλaμ/c, where c>0. Using the representation Eq.3 on Page 768 of [11], we find that

    r0=aλrraμ1=aλaμ1.

    The simplest case has r=1, and bλ=bμ=1 (which necessitates aλ=aμ=c1, and c>1), and σ(dλ,dμ)=f(λ,μ)dλdμ, with

    f(x,y)={C(xy)c2Γ(2c,x)if yxC(xy)c2Γ(2c,y)if y>x,

    where C1=Γ(c)/(c1)2, and Γ(a,x)=xta1etdt is the complementary incomplete gamma function. Figure 4 illustrates quasi-stationary behaviour with ψ=0.1941, noting the corresponding homogeneous model (λ=μ=1.5) predicts evanescence.

    Figure 4.  Simulation of the SIS model with n=50 individuals, and with positively correlated (ρ=0.6) Gamma distributed infection and recovery rates.

    We can get a handle on various statistics associated with the process by way of the moment generating function (MGF) of mt:

    Mt(a,b)=eaλ+bμmt(dλ,dμ),

    which will be defined for (a,b) in a region containing the origin. On setting h(θ) (=h(λ,μ)) =exp(aλ+bμ) in (2.2) we obtain an integral equation for Mt:

    Mt(a,b)=M0(a,b)+t0(ms(S)(aΣ(a,b)aMs(a,b))bMs(a,b))ds,

    where Σ(a,b)=eaλ+bμσ(dλ,dμ) is the MGF of σ. This can be differentiated to obtain the PDE

    tMt(a,b)+Mt(0,0)aMt(a,b)+bMt(a,b)=Mt(0,0)Φ(a,b), (4.1)

    where

    Φ(a,b)=aΣ(a,b)=λeaλ+bμσ(dλ,dμ).

    On setting (/t)Mt(a,b)=0, we see that any equilibrium Meq of (4.1) must satisfy

    ψaMeq(a,b)+bMeq(a,b)=ψΦ(a,b), (4.2)

    where ψ=Meq(0,0). This PDE is easily solved using the method of characteristics. We find that

    Meq(a,b)=Meq(0,0)+ψλμ+ψλ(eaλ+bμ1)σ(dλ,dμ).

    Under the conditions of Theorem 3 (b), namely that r0>1 and (m0(S)=) (ϕ0,σ)>0, we will have that Mt(a,b)Meq(a,b), Meq being the MGF of m:

    Meq(a,b)=eaλ+bμm(dλ,dμ).

    We will necessarily have Meq(0,0)=ψ, and, since R(ψ)=ψ,

    Meq(a,b)=ψλμ+ψλeaλ+bμσ(dλ,dμ).

    Example 3. Suppose that the individual characteristics are chosen from independent exponential distributions with rate parameters l and m: σ(dλ,dμ)=lelλmemμdλdμ. With this choice we can perform some explicit calculations. For example,

    R(ψ)={ρψ(ρψ)2log(ρψ)ψρψif 0<ψρ12if 0<ψ=ρ0if ψ=0,

    where ρ=l/m. Therefore,

    R(ψ)={ρ(ρ+ψ)(ρψ)3log(ρψ)2ρ(ρψ)2if 0<ψρ16ρif 0<ψ=ρ.

    So, we have R(0+)=. This implies a stable equilibrium. In particular, R has two fixed points, 0 and a positive fixed point ψ. Again, a simple calculation yields

    Meq(a,b)=lmψ(laψ(mb))2log(laψ(mb))lmψ(la)(laψ(mb)),a<l, b<m.

    Notice that Meq(0,0)=R(ψ)=ψ. For example,

    μm(dλ,dμ)=Meqb(0,0)=ψ(l+mψ)(lmψ)22lmψ2(lmψ)3log(lmψ),

    a quantity which approximates the large-n long-term behaviour of (refer to (2.1))

    μm(n)t(dλ,dμ)=1nni=1μiX(n)i,t, (4.3)

    being the average recovery rate of infected individuals. This is illustrated in Figure 5, where the average recovery rate of infected individuals (cyan) is plotted along with the value of (4.3) (yellow).

    Figure 5.  Simulation of the SIS model with n=50 individuals, and with independent exponentially distributed infection and recovery rates.

    We have presented a model for an SIS infection that incorporates variation of individual characteristics. By imagining that these characteristics follow a point process on a subset of R2+, we are able to exploit existing results on random measures, and random measure-valued processes, to obtain large-population limits, and then predict long-term behaviour. These results provide a means of studying various statistics associated with the process, including the proportion of individuals infected. Several examples demonstrate a range of behaviour that cannot be gleaned from the standard stochastic SIS model based on mean values of the characteristics. Explicit results on the moment generating function of the (large-n) limiting process allow us to estimate additional quantities such as the average infection rates and average recovery rates.

    The methods described here do not provide explicit results on the quality of the approximations presented, and in particular how large n would need to be for our approximations to be faithful. It might be possible to adapt the results in Chapter 4 of [12], based on their discrete-time counterparts [13], to provide explicit rates of convergence by attempting to bound weighted sums of differences:

    1nni=1h(θi)|X(n)i,tpi,t|,

    where pt=(pi,t, i=1n) is some suitable deterministic law of motion; this is the subject of ongoing research. A simple extension to the model presented would be to replace the upward transition rate in (1.2) by a+λiˉX(n)i, where a>0, thus incorporating an external source of infection as in [14]. Now the disease-free state 0 would no longer be an absorbing state, but the methods employed here could be brought to bear, thus providing a means of estimating equilibrium behaviour.

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

    The authors declare there is no conflict of interest.



    [1] G.H. Weiss, M. Dishon, On the asymptotic behaviour of the stochastic and deterministic models of an epidemic, Math. Biosci., 11 (1971), 261–265. https://doi.org/10.1016/0025-5564(71)90087-3 doi: 10.1016/0025-5564(71)90087-3
    [2] T. G. Kurtz, Solutions of ordinary differential equations as limits of pure jump Markov processes, J. Appl. Probab., 7 (1970), 49–58. https://doi.org/10.2307/3212147 doi: 10.2307/3212147
    [3] F. Ball, Stochastic and deterministic models for SIS epidemics among a population partitioned into households, Math. Biosci., 158 (1999), 41–67. https://doi.org/10.1016/S0025-5564(98)10060-3 doi: 10.1016/S0025-5564(98)10060-3
    [4] F. G. Ball, O. D. Lyne, Optimal vaccination policies for stochastic epidemics among a population of households, Math. Biosci., 177–178 (2002), 333–354. https://doi.org/10.1016/S0025-5564(01)00095-5 doi: 10.1016/S0025-5564(01)00095-5
    [5] T. Britton, P. Neal, The time to extinction for an SIS-household-epidemic model, J. Math. Biol., 61 (2010), 763–779. https://doi.org/10.1007/s00285-009-0320-5 doi: 10.1007/s00285-009-0320-5
    [6] D. Clancy, P. O'Neill, Exact bayesian inference and model selection for stochastic models of epidemics among a community of households, Scand. J. Statist., 34 (2007), 259–274. https://doi.org/10.1111/j.1467-9469.2006.00522.x doi: 10.1111/j.1467-9469.2006.00522.x
    [7] D. Hiebler, Moment equations and dynamics of a household SIS epidemiological model, J. Math. Biol., 68 (2006), 1315–1333. https://doi.org/10.1007/s11538-006-9080-1 doi: 10.1007/s11538-006-9080-1
    [8] P. Neal, Efficient likelihood-free bayesian computation for household epidemics, Stat. Comput., 22 (2012), 1239–1256. https://doi.org/10.1007/s11222-010-9216-x doi: 10.1007/s11222-010-9216-x
    [9] V. Cane, R. McNamee, The spread of infection in a heterogeneous population, J. Appl. Probab., 19A (1982), 173–184. https://doi.org/10.1017/S0021900200034549 doi: 10.1017/S0021900200034549
    [10] R. McVinish, P. K. Pollett, The deterministic limit of a stochastic logistic model with individual variation, Math. Biosci., 241 (2013), 109–114. https://doi.org/10.1016/j.mbs.2012.10.001 doi: 10.1016/j.mbs.2012.10.001
    [11] S. Nadarajaha, A. K. Gupta, Some bivariate gamma distributions. Appl. Math. Letters, 19 (2006), 767–774. https://doi.org/10.1016/j.aml.2005.10.007
    [12] L. Hodgkinson, Approximations for finite spin systems and occupancy processes. PhD thesis, School of Mathematics and Physics, The University of Queensland, 2019.
    [13] L. Hodgkinson, R. McVinish, P. K. Pollett. Normal approximations for discrete-time occupancy processes, Stochastic Process. Appl., 130 (2020), 6414–6444. https://doi.org/10.1016/j.spa.2020.05.016 doi: 10.1016/j.spa.2020.05.016
    [14] P. Stone, H. Wilkinson-Herbots, V. Isham, A stochastic model for head lice infections, J. Math. Biol., 56 (2008), 743–763. https://doi.org/10.1007/s00285-007-0136-0 doi: 10.1007/s00285-007-0136-0
  • Reader Comments
  • © 2024 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(1410) PDF downloads(146) Cited by(0)

Figures and Tables

Figures(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog