This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\jyear

2023

[1]\orgdivDepartment of Education, \orgnameTokyo Gakugei University, \orgaddress\street4-1-1 Nukuikita-machi, \cityKoganei-shi, \postcode184-8501, \stateTokyo, \countryJapan

Basic concepts for the Kermack and McKendrick model with static heterogeneity

\fnmHisashi \surInaba [email protected] *
Abstract

In this paper, we consider the infection-age-dependent Kermack–McKendrick model in which host individuals are distributed in a continuous state space. To provide a mathematical foundation for the heterogeneous model, we develop a L1L^{1}-framework to formulate basic epidemiological concepts. First, we show the mathematical well-posedness of the basic model under appropriate conditions allowing the unbounded parameters with non-compact domain. Next we define the basic reproduction number and prove the pandemic threshold theorem. We then present a systematic procedure to compute the effective reproduction number and the herd immunity threshold. Finally we give some illustrative examples and concrete results by using the separable mixing assumption.

keywords:
Kermack-McKendrick model, basic reproduction number, effective reproduction number, herd immunity threshold, heterogeneity
pacs:
[

MSC Classification]92D30, 92D25, 45G10

1 Introduction

Among the mathematical studies inspired by COVID-19, the Kermack-McKendrick epidemic model of 1927 Kermack1927 has played an important role. In particular, it was recognized that to understand the complex dynamics of recurrent waves of epidemics, we need to extend the basic model to account for individual heterogeneity described by the continuous trait variable that reflects biological or social heterogeneity of individuals Britton2020 ; Diekmann2023 ; Gomes2022 ; Montalban2022 ; Neipel2020 ; Tkachenko2021a ; Tkachenko2021b .

As a special case of individual heterogeneity described by continuous trait variables, spatial extensions of the Kermack–McKendrick model have a long history. Kendall Kendall1957 proposed a spatial extension of the Kermack–McKendrick model and stated his Pandemic Threshold Theorem. Baily Bailey1975 defines a pandemic when the proportion of individuals contracting the disease, whatever the distance from the initial focus of infection is, is greater or equal to the root of the final size equation, provided that the basic reproduction number exceeds unity. Diekmann Diekmann1978 , Thieme Thieme1977a ; Thieme1977b , Webb Webb1980 ; Webb1981 , Rass and Radcliffe Rass2003 and Inaba Inaba2014 extended Kendall’s pandemic threshold result to the infection-age structured Kermack–McKendrick model with continuous individual heterogeneity.

On the other hand, the complex behavior of COVID-19 has suggested us that the social and biological individual heterogeneity would play a most important role in the global dynamics of an outbreak. Therefore, we need to develop a mathematical model that can take into account the social and biological individual heterogeneity, which could be qualitatively different from the geographic distribution. It should be noted that before COVID-19, several authors had already studied such kind of general heterogeneous models (Katriel2012 , Novozhilov2008 , Novozhilov2012 , Thieme2009 ). As is mentioned above, the practical importance of the heterogeneity model has been widely discussed by many authors in recent years, but its rigorous mathematical foundation is still unclear.

In this paper, we consider the infection-age-dependent Kermack–McKendrick model, where host individuals are distributed in a continuous state space Ωn\Omega\subset\mathbb{R}^{n}. To provide mathematical foundation for the heterogeneous model with unbounded susceptibility and non-compact domain, we develop a L1L^{1}-framework to formulate basic epidemiological concepts. As is shown in Diekmann and Inaba Diekmann2023 , the basic model is written as a simple renewal equation, which essentially has only one model component (the force of infection), and can be reduced to a variety of compartmental models under certain assumptions. First, we prove the well-posedness of the basic system. Next we define the basic reproduction number and prove the pandemic threshold theorem. We then present a systematic procedure to compute the effective reproduction number and the herd immunity threshold. Finally we give some illustrative examples and concrete results by using the separable mixing assumption. In particular, we discuss the critical condition for epidemic resurgence.

2 The general Kermack–McKendrick model with static heterogeneity

We consider a closed population with static individual heterogeneity. Let S(t,x)S(t,x) be the density of completely susceptible, never infected population with individual trait xx at time tt. From our biological interpretation, it is reasonable to assume that S(t,)L+1(Ω)S(t,\cdot)\in L^{1}_{+}(\Omega). The Kermack and McKendrick epidemic model with individual static heterogeneity xΩx\in\Omega for a closed population is formulated as follows Bootsma2023a ; Bootsma2023b ; Diekmann2023 :

S(t,x)t=F(t,x)S(t,x),\displaystyle\frac{\partial S(t,x)}{\partial t}=-F(t,x)S(t,x), (1)
F(t,x)=Ω0A(τ,x,σ)F(tτ,σ)S(tτ,σ)𝑑τ𝑑σ,\displaystyle F(t,x)=\int_{\Omega}\int_{0}^{\infty}A(\tau,x,\sigma)F(t-\tau,\sigma)S(t-\tau,\sigma)d\tau d\sigma,

where xΩx\in\Omega is the individual heterogeneity (trait) parameter, Ωn\Omega\subset\mathbb{R}^{n} is its state space and A(τ,x,σ)A(\tau,x,\sigma) denotes the expected contribution from infecteds with trait σ\sigma to the force of infection applied to susceptibles of trait xx as a function of the time τ\tau elapsed since infection took place.

For simplicity, we consider only the case of continuous heterogeneity distribution. To deal with a mixture of continuous and discrete heterogeneity, we can use the measure-theoretic formulation Diekmann2023 . For basic observations about the model (1), the reader can refer to chapter 6 of Diekmann2000 or chapter 8 of Diekmann2013 . It is also noted that this kind of the renewal equation formulation is also useful to consider the Kermack–McKendrick endemic model Breda2012 . The reader may find McKendrick equation (PDE) approach to the Kermack–McKendrick model in Inaba2001 ; Inaba2014 ; Inaba2016 ; Inaba2017 .

Under additional assumptions, the basic system (1) can be reduced to a variety of compartment models (see Diekmann2013 , Chap. 8). Here a simple example is given to show the reduction of the basic integral system to the traditional compartmental SIR model.

Example 1

Let γ\gamma be the recovery rate and let I(t,x)I(t,x) be the density of infected individuals. Then we have

I(t,x)=teγ(tτ)F(τ,x)S(τ,x)𝑑τ,I(t,x)=\int_{-\infty}^{t}e^{-\gamma(t-\tau)}F(\tau,x)S(\tau,x)d\tau, (2)

and it follows that

I(t,x)t=γI(t,x)+F(t,x)S(t,x),\frac{\partial I(t,x)}{\partial t}=-\gamma I(t,x)+F(t,x)S(t,x), (3)

which can be supplemented with

R(t,x)t=γI(t,x),\frac{\partial R(t,x)}{\partial t}=\gamma I(t,x), (4)

where R(t,x)R(t,x) denotes the density of recovered individuals. Suppose that A(τ,x,σ)=eγτβ(x,σ)A(\tau,x,\sigma)=e^{-\gamma\tau}\beta(x,\sigma). Then we have

F(t,x)\displaystyle F(t,x) =Ω0A(τ,x,σ)F(tτ,σ)S(tτ,σ)𝑑τ𝑑σ\displaystyle=\int_{\Omega}\int_{0}^{\infty}A(\tau,x,\sigma)F(t-\tau,\sigma)S(t-\tau,\sigma)d\tau d\sigma (5)
=Ωβ(x,σ)I(t,σ)𝑑σ.\displaystyle=\int_{\Omega}\beta(x,\sigma)I(t,\sigma)d\sigma.

Then the Kermack and McKendrick model (1) is reduced to the “SIR” (compartmental) epidemic model with trait varibale xx:

S(t,x)t=S(t,x)Ωβ(x,σ)I(t,σ)𝑑σ,\displaystyle\frac{\partial S(t,x)}{\partial t}=-S(t,x)\int_{\Omega}\beta(x,\sigma)I(t,\sigma)d\sigma, (6)
I(t,x)t=S(t,x)Ωβ(x,σ)I(t,σ)𝑑σγI(t,x),\displaystyle\frac{\partial I(t,x)}{\partial t}=S(t,x)\int_{\Omega}\beta(x,\sigma)I(t,\sigma)d\sigma-\gamma I(t,x),
R(t,x)t=γI(t,x).\displaystyle\frac{\partial R(t,x)}{\partial t}=\gamma I(t,x).

Kendall Kendall1957 used the above SIR model (6) with heterogeneity to study the geographic spread of the epidemic, and formulated the Pandemic Threshold Theorem (PTT), where xx is the spatial location of the individual. Hadeler Hadeler2017 considered this type of model by assuming that xx is the chronological age of individuals. In fact, as far as we consider fast epidemics compared to the demographic timescale, aging and demographic turnover have no remarkable effect on the epidemics, so the chronological age of individuals can be considered as a static heterogeneity. Moreover, as is shown in Tkachenko2021a , Montalban2022 and Diekmann2023 , the heterogeneous SIR model (6) can be reduced to a ODE system with nonlinear incidence rate if the host heterogeneity distribution is given by the gamma distribution and β(x,σ)\beta(x,\sigma) is proportional to xx. As was pointed out in Thieme Thieme2009 , the infinite-dimensional ODE model is mathaematically quite difficult if the parameters are not bounded in the non-compact domain, so we start from the renewal integral equation.

3 The initial value problem

For the heterogeneous model (1), the existence and uniqueness for the total orbit starting from the disease-free steady state S(,x)=N(x)S(-\infty,x)=N(x) is still an open problem, although it has been solved for the homogeneous case Diekmann1977 . Instead, following Rass and Radcliff Rass2003 , we consider here the initial value problem for the basic system (1).

Let N()X+:=L+1(Ω)N(\cdot)\in X_{+}:=L^{1}_{+}(\Omega) be the distribution of host population heterogeneity (traits) in a steady state and NX\|N\|_{X} is the total host size: NX:=Ω|N(x)|𝑑x\|N\|_{X}:=\int_{\Omega}\lvert N(x)\rvert dx, where X\|\cdot\|_{X} denotes the L1L^{1} norm of the space X:=L1(Ω)X:=L^{1}(\Omega). We assume that at time t=0t=0, the totally susceptible host population N(x)N(x) is exposed to infection from the outside. So S(0,x)=N(x)S(0,x)=N(x) and it holds that

S(t,x)t=F(t,x)S(t,x),\displaystyle\frac{\partial S(t,x)}{\partial t}=-F(t,x)S(t,x), (7)
F(t,x)=Ω0tA(τ,x,σ)F(tτ,σ)S(tτ,σ)𝑑τ𝑑σ+G(t,x),\displaystyle F(t,x)=\int_{\Omega}\int_{0}^{t}A(\tau,x,\sigma)F(t-\tau,\sigma)S(t-\tau,\sigma)d\tau d\sigma+G(t,x),

where t>0t>0 and G(t,x)G(t,x) is the given initial data for the force of infection, which is assumed to be generated by an external infection.

For the initial value problem, its mathematical well-posedness can be established by reducing it to the problem of the integral equation for the cumulative force of infection. Let W(t,x)W(t,x) be the cumulative force of infection:

W(t,x):=0tF(τ,x)𝑑τ.W(t,x):=\int_{0}^{t}F(\tau,x)d\tau. (8)

Then we have

S(t,x)=N(x)eW(t,x),S(t,x)=N(x)e^{-W(t,x)}, (9)

and

W(t,x)\displaystyle W(t,x) =H(t,x)+0t𝑑ηΩ0ηA(τ,x,σ)(Sη(ητ,σ))𝑑τ𝑑σ\displaystyle=H(t,x)+\int_{0}^{t}d\eta\int_{\Omega}\int_{0}^{\eta}A(\tau,x,\sigma)\left(-\frac{\partial S}{\partial\eta}(\eta-\tau,\sigma)\right)d\tau d\sigma (10)
=H(t,x)+Ω0tA(τ,x,σ)(N(σ)S(tτ,σ))𝑑τ𝑑σ\displaystyle=H(t,x)+\int_{\Omega}\int_{0}^{t}A(\tau,x,\sigma)(N(\sigma)-S(t-\tau,\sigma))d\tau d\sigma
=H(t,x)+Ω0tA(τ,x,σ)N(σ)(1eW(tτ,σ))𝑑τ𝑑σ\displaystyle=H(t,x)+\int_{\Omega}\int_{0}^{t}A(\tau,x,\sigma)N(\sigma)(1-e^{-W(t-\tau,\sigma)})d\tau d\sigma

where

H(t,x):=0tG(τ,x)𝑑τ.H(t,x):=\int_{0}^{t}G(\tau,x)d\tau. (11)

The integral equation (10) was traditionally studied as the Volterra-Hammerstein integral equation Thieme1977a ; Thieme1977b ; Thieme1980 . Diekmann Diekmann1978 used the renewal equation (10) to analyze the Kermack–McKendrick model with spatial heterogeneity. For traditional space-dependent models, much attention has been paid to the spatial interaction kernel like as A(τ,x,σ)=u(τ)v(xσ)A(\tau,x,\sigma)=u(\tau)v(x-\sigma) and its traveling wave solutions in a noncompact domain Ω\Omega. On the other hand, recent COVID-19 inspired studies assumes that the feature is not necessarily the spatial heterogeneity, but the biological or social individual heterogeneity, so the separable kernel as A(τ,x,σ)=a(x)b(τ)c(σ)A(\tau,x,\sigma)=a(x)b(\tau)c(\sigma) played a much more important rule Diekmann2023 .

To treat unbounded susceptibility and its non-compact state space, we first give existence and uniqueness results for the solution of (10) under new assumptions different from Diekmann1978 and Inaba Inaba2014 . Since F(t,x)S(t,x)F(t,x)S(t,x) should be integrable with respect to xx, it is reasonable to expect that WBC+(+;Y)W\in BC_{+}(\mathbb{R}_{+}\mathchar 24635\relax\;Y), where YY be a Banach space defined by Y:={N1ψ:ψX}Y:=\left\{N^{-1}\psi:\psi\in X\right\} with norm ϕY=Ω|ϕ(x)|N(x)𝑑x\|\phi\|_{Y}=\int_{\Omega}\lvert\phi(x)\rvert N(x)dx. Then we have ϕY=ϕNX\|\phi\|_{Y}=\|\phi N\|_{X}. Hence, we assume that a map P:ϕNϕP:\phi\to N\phi is an isometry from a Banach space YY to a Banach space XX. For any T>0T>0, let CT:=C([0,T];Y)C_{T}:=C([0,T]\mathchar 24635\relax\;Y) be the set of continuous functions on [0,T][0,T] equipped with the norm fCT:=sup0tTf(t,)Y\|f\|_{C_{T}}:=\sup_{0\leq t\leq T}\|f(t,\cdot)\|_{Y}. We also make the following technical assumptions:

Assumption 1.
  1. 1.

    There exist nonnegative measurable functions aa, bb, cc and a number α>1\alpha>1 such that

    a(x)b(τ)c(σ)A(τ,x,σ)αa(x)b(τ)c(σ).a(x)b(\tau)c(\sigma)\leq A(\tau,x,\sigma)\leq\alpha a(x)b(\tau)c(\sigma). (12)

    where aY+a\in Y_{+}, bL+1(+)b\in L^{1}_{+}(\mathbb{R}_{+}) and cnY+c^{n}\in Y_{+} for n=1,2n=1,2. Furthermore, there exists a bounded continuous function QBC+(+)Q\in BC_{+}(\mathbb{R}_{+}) such that

    a(x)Q(t)H(t,x)αa(x)Q(t).a(x)Q(t)\leq H(t,x)\leq\alpha a(x)Q(t). (13)
  2. 2.

    It holds that

    limh0ΩΩ0N(x)|A(τ+h,x,σ)A(τ,x,σ)|N(σ)𝑑τ𝑑σ𝑑x=0,\lim_{h\to 0}\int_{\Omega}\int_{\Omega}\int_{0}^{\infty}N(x)\lvert A(\tau+h,x,\sigma)-A(\tau,x,\sigma)\rvert N(\sigma)d\tau d\sigma dx=0, (14)

    where we assume that A(τ+h,x,σ)=0A(\tau+h,x,\sigma)=0 if τ+h+\tau+h\notin\mathbb{R}_{+} and

    limh0ΩΩ0|N(x+h)A(τ,x+h,σ)N(x)A(τ,x,σ)|N(σ)𝑑τ𝑑σ𝑑x=0,\lim_{h\to 0}\int_{\Omega}\int_{\Omega}\int_{0}^{\infty}\lvert N(x+h)A(\tau,x+h,\sigma)-N(x)A(\tau,x,\sigma)\rvert N(\sigma)d\tau d\sigma dx=0, (15)

    where we assume that N(x+h)A(τ,x+h,σ)=0N(x+h)A(\tau,x+h,\sigma)=0 if x+hΩx+h\notin\Omega.

Let CT+C_{T+} be the positive cone of CTC_{T}. Define an operator \mathcal{F} for ϕCT+\phi\in C_{T+} by

(ϕ)(t,x)=H(t,x)+Ω0tA(τ,x,σ)N(σ)(1eϕ(tτ,σ))𝑑τ𝑑σ,\mathcal{F}(\phi)(t,x)=H(t,x)+\int_{\Omega}\int_{0}^{t}A(\tau,x,\sigma)N(\sigma)(1-e^{-\phi(t-\tau,\sigma)})d\tau d\sigma, (16)

where we assume HCT+H\in C_{T+}.

Proposition 1.

\mathcal{F} is a continuous map from CT+C_{T+} into itself.

Proof: For ϕCT+\phi\in C_{T+}, we have

(ϕ)(t,x)H(x,t)+Ω0tA(τ,x,σ)N(σ)𝑑τ𝑑σαa(x)Q(t)+αLa(x)cY,\mathcal{F}(\phi)(t,x)\leq H(x,t)+\int_{\Omega}\int_{0}^{t}A(\tau,x,\sigma)N(\sigma)d\tau d\sigma\leq\alpha a(x)Q(t)+\alpha La(x)\|c\|_{Y}, (17)

where L:=0b(τ)𝑑τL:=\int_{0}^{\infty}b(\tau)d\tau. Then for every t[0,T]t\in[0,T], we have (ϕ)(t,)Y+\mathcal{F}(\phi)(t,\cdot)\in Y_{+} and F(ϕ)(t,)Y\|F(\phi)(t,\cdot)\|_{Y} is uniformly bounded. Let

J(t,x;ϕ):=Ω0tA(tτ,x,σ)N(σ)(1eϕ(τ,σ))𝑑τ𝑑σ.J(t,x\mathchar 24635\relax\;\phi):=\int_{\Omega}\int_{0}^{t}A(t-\tau,x,\sigma)N(\sigma)(1-e^{-\phi(\tau,\sigma)})d\tau d\sigma.

Then we have

Ω|J(t1,x;ϕ)J(t2,x;ϕ)|N(x)𝑑xQ1(t1,t2)+Q2(t1,t2),\int_{\Omega}\lvert J(t_{1},x\mathchar 24635\relax\;\phi)-J(t_{2},x\mathchar 24635\relax\;\phi)\rvert N(x)dx\leq Q_{1}(t_{1},t_{2})+Q_{2}(t_{1},t_{2}), (18)

where

Q1(t1,t2):=Ω𝑑xΩ𝑑σ0t1N(x)|A(t1τ,x,σ)A(t2τ,x,σ)|N(σ)𝑑τ,Q_{1}(t_{1},t_{2}):=\int_{\Omega}dx\int_{\Omega}d\sigma\int_{0}^{t_{1}}N(x)\lvert A(t_{1}-\tau,x,\sigma)-A(t_{2}-\tau,x,\sigma)\rvert N(\sigma)d\tau,
Q2(t1,t2):=Ω𝑑xΩ𝑑σt1t2N(x)|A(t2τ,x,σ)|N(σ)𝑑τ.Q_{2}(t_{1},t_{2}):=\int_{\Omega}dx\int_{\Omega}d\sigma\int_{t_{1}}^{t_{2}}N(x)\lvert A(t_{2}-\tau,x,\sigma)\rvert N(\sigma)d\tau.

From our assumption (12)-(14), for any ϵ>0\epsilon>0 there exists δ>0\delta>0 such that

(ϕ)(t1,)(ϕ)(t2,)YH(t1,)H(t2,)Y+Q1(t1,t2)+Q2(t1,t2)<ϵ,\|\mathcal{F}(\phi)(t_{1},\cdot)-\mathcal{F}(\phi)(t_{2},\cdot)\|_{Y}\leq\|H(t_{1},\cdot)-H(t_{2},\cdot)\|_{Y}+Q_{1}(t_{1},t_{2})+Q_{2}(t_{1},t_{2})<\epsilon, (19)

if |t1t2|<δ\lvert t_{1}-t_{2}\rvert<\delta. Then (ϕ)CT+\mathcal{F}(\phi)\in C_{T+}. Next observe that for ϕjCT+\phi_{j}\in C_{T+},

(ϕ1)(t,)(ϕ2)(t,)YαLaYΩc(σ)N(σ)|eϕ1(tτ,σ)eϕ2(tτ,σ)|𝑑σ.\|\mathcal{F}(\phi_{1})(t,\cdot)-\mathcal{F}(\phi_{2})(t,\cdot)\|_{Y}\leq\alpha L\|a\|_{Y}\int_{\Omega}c(\sigma)N(\sigma)\lvert e^{-\phi_{1}(t-\tau,\sigma)}-e^{-\phi_{2}(t-\tau,\sigma)}\rvert d\sigma. (20)

From the Schwarz inequality, it follows that

Ωc(σ)N(σ)|eϕ1(tτ,σ)eϕ2(tτ,σ)|𝑑σ\displaystyle\int_{\Omega}c(\sigma)N(\sigma)\lvert e^{-\phi_{1}(t-\tau,\sigma)}-e^{-\phi_{2}(t-\tau,\sigma)}\rvert d\sigma
(Ωc2(σ)N(σ)|eϕ1(tτ,σ)eϕ2(tτ,σ)|𝑑σ)12\displaystyle\leq\left(\int_{\Omega}c^{2}(\sigma)N(\sigma)\lvert e^{-\phi_{1}(t-\tau,\sigma)}-e^{-\phi_{2}(t-\tau,\sigma)}\rvert d\sigma\right)^{\frac{1}{2}}
×(ΩN(σ)|eϕ1(tτ,σ)eϕ2(tτ,σ)|𝑑σ)12,\displaystyle~{}~{}~{}~{}~{}~{}\times\left(\int_{\Omega}N(\sigma)\lvert e^{-\phi_{1}(t-\tau,\sigma)}-e^{-\phi_{2}(t-\tau,\sigma)}\rvert d\sigma\right)^{\frac{1}{2}},

where

Ωc2(σ)N(σ)|eϕ1(tτ,σ)eϕ2(tτ,σ)|𝑑σ2c2Y,\int_{\Omega}c^{2}(\sigma)N(\sigma)\lvert e^{-\phi_{1}(t-\tau,\sigma)}-e^{-\phi_{2}(t-\tau,\sigma)}\rvert d\sigma\leq 2\|c^{2}\|_{Y},

and

ΩN(σ)|eϕ1(tτ,σ)eϕ2(tτ,σ)|𝑑σ\displaystyle\int_{\Omega}N(\sigma)\lvert e^{-\phi_{1}(t-\tau,\sigma)}-e^{-\phi_{2}(t-\tau,\sigma)}\rvert d\sigma
ΩN(σ)|ϕ1(tτ,σ)ϕ2(tτ,σ)|𝑑σϕ1ϕ2CT.\displaystyle\leq\int_{\Omega}N(\sigma)\lvert\phi_{1}(t-\tau,\sigma)-\phi_{2}(t-\tau,\sigma)\rvert d\sigma\leq\|\phi_{1}-\phi_{2}\|_{C_{T}}.

So we have

(ϕ1)(ϕ2)CTαLaY2c2Yϕ1ϕ2CT.\|\mathcal{F}(\phi_{1})-\mathcal{F}(\phi_{2})\|_{C_{T}}\leq\alpha L\|a\|_{Y}\sqrt{2\|c_{2}\|_{Y}\|\phi_{1}-\phi_{2}\|_{C_{T}}}. (21)

Then \mathcal{F} is a continuous positive operator.

Proposition 2.

\mathcal{F} has at least one fixed point.

Proof: Define a subset Φ:={ϕn:n=0,1,2,}CT+\Phi:=\{\phi_{n}:n=0,1,2,\cdots\}\subset C_{T+}, where

ϕ0=H,ϕn=(ϕn1),(n1).\phi_{0}=H,\quad\phi_{n}=\mathcal{F}(\phi_{n-1}),~{}(n\geq 1). (22)

From (19) in the proof of Proposition 1, Φ\Phi is equicontinuous on [0,T][0,T]. Next, for each t[0,T]t\in[0,T], define a set Φ(t):={ϕn(t,):n=0,1,2,}Y+\Phi(t):=\{\phi_{n}(t,\cdot):n=0,1,2,\cdots\}\subset Y_{+}. From the assumption 1, it follows that supn=0,1,2,..ϕn(t,)Y<\sup_{n=0,1,2,..}\|\phi_{n}(t,\cdot)\|_{Y}<\infty, and it follows from (17) that

limrr|ϕn(t,x)|N(x)𝑑x=0,limh+00h|ϕn(t,x)|N(x)𝑑x=0,\lim_{r\to\infty}\int_{r}^{\infty}\lvert\phi_{n}(t,x)\rvert N(x)dx=0,\quad\lim_{h\to+0}\int_{0}^{h}\lvert\phi_{n}(t,x)\rvert N(x)dx=0, (23)

uniformly in ϕn(t,)Φ(t)\phi_{n}(t,\cdot)\in\Phi(t). From (15), we have

limh0Ω|N(x+h)ϕn(t,x+h)N(x)ϕn(t,x)|𝑑x=0,\lim_{h\to 0}\int_{\Omega}\lvert N(x+h)\phi_{n}(t,x+h)-N(x)\phi_{n}(t,x)\rvert dx=0, (24)

uniformly in ϕn(t,)Φ(t)\phi_{n}(t,\cdot)\in\Phi(t). Then it follows from the Fréchet-Kolmogorov criterion (Theorem B.2. in Smith2011 ) that the set {ϕn(t,)N}n=1,2,L+1(Ω)\{\phi_{n}(t,\cdot)N\}_{n=1,2,...}\subset L^{1}_{+}(\Omega) has compact closure. Then we can choose a convergent subsequence from Φ(t)\Phi(t), so Φ(t)\Phi(t) is relatively compact in Y+Y_{+}. Thanks to Ascoli’s theorem Lang1993 , Φ\Phi is also relatively compact in CT+C_{T+}. Then we can choose a convergent subsequence {ϕn(k)}k=1,2,Φ\{\phi_{n(k)}\}_{k=1,2,...}\subset\Phi, and we conclude that limkϕn(k)=ϕCT+\lim_{k\to\infty}\phi_{n(k)}=\phi_{\infty}\in C_{T+} exists and ϕ=limk(ϕn(k1))=(limkϕn(k1))=(ϕ)\phi_{\infty}=\lim_{k\to\infty}\mathcal{F}(\phi_{n(k-1)})=\mathcal{F}(\lim_{k\to\infty}\phi_{n(k-1)})=\mathcal{F}(\phi_{\infty}). Thus ϕ\phi_{\infty} is a positive fixed point of \mathcal{F}.

Proposition 3.

\mathcal{F} has at most one positive fixed point.

Proof: At first, note that the operator \mathcal{F} is monotonically non-decreasing positive operator. From our assumption 1, we have

h(ϕ)(t)a(x)(ϕ)(t,x)αh(ϕ)(t)a(x),h(\phi)(t)a(x)\leq\mathcal{F}(\phi)(t,x)\leq\alpha h(\phi)(t)a(x), (25)

where h:CT+BC+([0,T]:+)h:C_{T+}\to BC_{+}([0,T]:\mathbb{R}_{+}) is given by

h(ϕ)(t):=Q(t)+Ω0tb(τ)c(σ)N(σ)(1eϕ(tτ,σ))𝑑τ𝑑σ.h(\phi)(t):=Q(t)+\int_{\Omega}\int_{0}^{t}b(\tau)c(\sigma)N(\sigma)(1-e^{-\phi(t-\tau,\sigma)})d\tau d\sigma. (26)

It also holds that for s(0,1)s\in(0,1),

(sϕ)(t,x)s(ϕ)(t,x)+η(ϕ,s)(t)a(x),\mathcal{F}(s\phi)(t,x)\geq s\mathcal{F}(\phi)(t,x)+\eta(\phi,s)(t)a(x), (27)

where

η(ϕ,s)(t):=Ω0tb(τ)c(σ)N(σ)(1esϕ(tτ,σ)s(1eϕ(tτ,σ))dτdσ.\eta(\phi,s)(t):=\int_{\Omega}\int_{0}^{t}b(\tau)c(\sigma)N(\sigma)(1-e^{-s\phi(t-\tau,\sigma)}-s(1-e^{-\phi(t-\tau,\sigma)})d\tau d\sigma. (28)

Then it is easy to see that η(ϕ,s)(t)>0\eta(\phi,s)(t)>0 for any ϕCT+{0}\phi\in C_{T+}\setminus\{0\} and any s(0,1)s\in(0,1). Suppose that \mathcal{F} has two non-zero fixed points ϕ1\phi_{1} and ϕ2\phi_{2}. Then for every t>0t>0, h(ϕj)(t,x)>0h(\phi_{j})(t,x)>0 and it follows that

ϕ1(t,x)=(ϕ1)(t,x)h(ϕ1)(t)a(x)h(ϕ1)(t)αh(ϕ2)(t)(ϕ2)(t,x)=h(ϕ1)(t)αh(ϕ2)(t)ϕ2(t,x).\phi_{1}(t,x)=\mathcal{F}(\phi_{1})(t,x)\geq h(\phi_{1})(t)a(x)\geq\frac{h(\phi_{1})(t)}{\alpha h(\phi_{2})(t)}\mathcal{F}(\phi_{2})(t,x)=\frac{h(\phi_{1})(t)}{\alpha h(\phi_{2})(t)}\phi_{2}(t,x). (29)

For a fixed t>0t>0, define a number k:=sup{μ:ϕ1(t,x)μϕ2(t,x)}k:=\sup\{\mu:\phi_{1}(t,x)\geq\mu\phi_{2}(t,x)\}. It follows from (29) that k>0k>0. Suppose that 0<k<10<k<1. Then we can observe that

ϕ1(t,x)\displaystyle\phi_{1}(t,x) =(ϕ1)(t,x)(kϕ2)(t,x)k(ϕ2)(t,x)+η(ϕ2,k)(t)a(x)\displaystyle=\mathcal{F}(\phi_{1})(t,x)\geq\mathcal{F}(k\phi_{2})(t,x)\geq k\mathcal{F}(\phi_{2})(t,x)+\eta(\phi_{2},k)(t)a(x) (30)
=kϕ2(t,x)+η(ϕ2,k)(t)a(x)(k+η(ϕ2,k)αh(ϕ2)(t))ϕ2(t,x),\displaystyle=k\phi_{2}(t,x)+\eta(\phi_{2},k)(t)a(x)\geq\left(k+\frac{\eta(\phi_{2},k)}{\alpha h(\phi_{2})(t)}\right)\phi_{2}(t,x),

which contradicts the definition of kk. Then we conclude that k1k\geq 1 and ϕ1(t,x)ϕ2(t,x)\phi_{1}(t,x)\geq\phi_{2}(t,x). Switching the roles of ϕ1\phi_{1} and ϕ2\phi_{2}, we can repeat the same argument to prove that ϕ2(t,x)ϕ1(t,x)\phi_{2}(t,x)\geq\phi_{1}(t,x), so we conclude that ϕ1=ϕ2\phi_{1}=\phi_{2} for every t[0,T]t\in[0,T]. Then \mathcal{F} has at most one positive fixed point

From Propositions 2 and 3, the equation (10) has a unique positive solution for any T>0T>0, and it is uniformly bounded, so the global positive solution of (10) exists uniquely for t+t\in\mathbb{R}_{+}.

Proposition 4.

For the global solution of (10), there exists W(,)Y+W(\infty,\cdot)\in Y_{+} such that limtW(t,)W(,)Y=0\lim_{t\to\infty}\|W(t,\cdot)-W(\infty,\cdot)\|_{Y}=0 and W(,x)W(\infty,x) satisfies the limit equation

W(,x)=H(,x)+ΩΘ(x,σ)N(σ)(1eW(,σ))𝑑σ,W(\infty,x)=H(\infty,x)+\int_{\Omega}\Theta(x,\sigma)N(\sigma)(1-e^{-W(\infty,\sigma)})d\sigma, (31)

where

Θ(x,σ):=0A(τ,x,σ)𝑑τ.\Theta(x,\sigma):=\int_{0}^{\infty}A(\tau,x,\sigma)d\tau. (32)

Proof: Since W(t,)NX=L+1(Ω)W(t,\cdot)N\in X=L^{1}_{+}(\Omega) is monotonically non-decreasing with respect to tt and supt>0W(t,)Y<\sup_{t>0}\|W(t,\cdot)\|_{Y}<\infty, it follows from B. Levi’s theorem that there exists a MXM\in X such that limtW(t,)N=MX\lim_{t\to\infty}W(t,\cdot)N=M\in X. Then limtW(t,)W(,)Y=0\lim_{t\to\infty}\|W(t,\cdot)-W(\infty,\cdot)\|_{Y}=0 where W(,):=M/NW(\infty,\cdot):=M/N. Observe that

|Ω𝑑σ0tA(τ,x,σ)N(σ)(1eW(tτ,σ))𝑑τΩ𝑑σ0A(τ,x,σ)N(σ)(1eW(,σ))𝑑τ|\displaystyle\Big{\lvert}\int_{\Omega}d\sigma\int_{0}^{t}A(\tau,x,\sigma)N(\sigma)(1-e^{-W(t-\tau,\sigma)})d\tau-\int_{\Omega}d\sigma\int_{0}^{\infty}A(\tau,x,\sigma)N(\sigma)(1-e^{-W(\infty,\sigma)})d\tau\Big{\rvert}
αa(x)0tb(τ)𝑑τΩc(σ)N(σ)|eW(tτ,σ)eW(,σ)|𝑑σ\displaystyle\leq\alpha a(x)\int_{0}^{t}b(\tau)d\tau\int_{\Omega}c(\sigma)N(\sigma)\lvert e^{-W(t-\tau,\sigma)}-e^{-W(\infty,\sigma)}\rvert d\sigma
+αa(x)tb(τ)𝑑τΩc(σ)N(σ)eW(,σ)𝑑σ,\displaystyle~{}~{}~{}+\alpha a(x)\int_{t}^{\infty}b(\tau)d\tau\int_{\Omega}c(\sigma)N(\sigma)e^{-W(\infty,\sigma)}d\sigma,

where it is clear that the second term on the right-hand side goes to zero when tt\to\infty. For the first term, we observe that

0t𝑑τb(τ)Ωc(σ)N(σ)|eW(tτ,σ)eW(,σ)|𝑑σ\displaystyle\int_{0}^{t}d\tau b(\tau)\int_{\Omega}c(\sigma)N(\sigma)\lvert e^{-W(t-\tau,\sigma)}-e^{-W(\infty,\sigma)}\rvert d\sigma
0tb(τ)𝑑τΩc(σ)N(σ)|W(tτ,σ)W(,σ)|𝑑σ\displaystyle\leq\int_{0}^{t}b(\tau)d\tau\int_{\Omega}c(\sigma)N(\sigma)\lvert W(t-\tau,\sigma)-W(\infty,\sigma)\rvert d\sigma
0t𝑑τb(τ)(2c2NXW(tτ,)W(,)Y)12,\displaystyle\leq\int_{0}^{t}d\tau b(\tau)\left(2\|c^{2}N\|_{X}\|W(t-\tau,\cdot)-W(\infty,\cdot)\|_{Y}\right)^{\frac{1}{2}},

where we use the Schwarz inequality as in the proof of Proposition 1. Since

b(τ)(2c2NXW(tτ,)W(,)Y)12\displaystyle b(\tau)\left(2\|c^{2}N\|_{X}\|W(t-\tau,\cdot)-W(\infty,\cdot)\|_{Y}\right)^{\frac{1}{2}}
b(τ)(2c2NXsupt0W(t,σ)W(,σ)Y)12L1(+),\displaystyle~{}~{}\leq b(\tau)\left(2\|c^{2}N\|_{X}\sup_{t\geq 0}\|W(t,\sigma)-W(\infty,\sigma)\|_{Y}\right)^{\frac{1}{2}}\in L^{1}(\mathbb{R}_{+}),

then it follows from the dominated convergence theorem that

limt0t𝑑τb(τ)(2c2NXW(tτ,)W(,)Y)12=0,\lim_{t\to\infty}\int_{0}^{t}d\tau b(\tau)\left(2\|c^{2}N\|_{X}\|W(t-\tau,\cdot)-W(\infty,\cdot)\|_{Y}\right)^{\frac{1}{2}}=0,

Si if we let tt\to\infty in the equation (10), we have (31).

4 The basic reproduction number

Let B(t,x):=F(t,x)S(t,x)B(t,x):=F(t,x)S(t,x) be the incidence at time t>0t>0. Then we can rewrite (7) as an initial value problem for SS and BB:

S(t,x)t=B(t,x),\displaystyle\frac{\partial S(t,x)}{\partial t}=-B(t,x), (33)
B(t,x)=S(t,x)[Ω0tA(τ,x,σ)B(tτ,σ)𝑑τ𝑑σ+G(t,x)].\displaystyle B(t,x)=S(t,x)\left[\int_{\Omega}\int_{0}^{t}A(\tau,x,\sigma)B(t-\tau,\sigma)d\tau d\sigma+G(t,x)\right].

In the disease invasion phase, the incidence is described by the linearized equation as

B(t,x)=N(x)Ω0tA(τ,x,σ)B(tτ,σ)𝑑τ𝑑σ+N(x)G(t,x).B(t,x)=N(x)\int_{\Omega}\int_{0}^{t}A(\tau,x,\sigma)B(t-\tau,\sigma)d\tau d\sigma+N(x)G(t,x). (34)

where BB denotes the incidence rate in the invasion phase for the totally susceptible population. Define the net reproduction operator acting on the birth state space X=L+1(Ω)X=L^{1}_{+}(\Omega) as

(K(τ)ϕ)(x):=N(x)ΩA(τ,x,σ)ϕ(σ)𝑑σ,ϕX.(K(\tau)\phi)(x):=N(x)\int_{\Omega}A(\tau,x,\sigma)\phi(\sigma)d\sigma,~{}\phi\in X. (35)

Then (34) is formulated as an abstract renewal equation

B(t)=NG(t)+0tK(τ)B(tτ)𝑑τ,B(t)=NG(t)+\int_{0}^{t}K(\tau)B(t-\tau)d\tau, (36)

where B(t)=B(t,)B(t)=B(t,\cdot) and NG(t)=N()G(t,)NG(t)=N(\cdot)G(t,\cdot) are vector-valued functions from +\mathbb{R}_{+} to X+X_{+}.

The next generation operator (NGO)111The NGO was first introduced in Diekmann1990 , see also Inaba2017 . acting on XX associated with the renewal process (36) is defined by

T:=0K(τ)𝑑τ,T:=\int_{0}^{\infty}K(\tau)d\tau, (37)

which is the linear integral operator defined by

(Tϕ)(x)=N(x)ΩΘ(x,σ)ϕ(σ)𝑑σ,ϕX.(T\phi)(x)=N(x)\int_{\Omega}\Theta(x,\sigma)\phi(\sigma)d\sigma,~{}\phi\in X. (38)

Now we introduce the additional condition:

Assumption 2.
  1. 1.

    For any fL+{0}f\in L^{\infty}_{+}\setminus\{0\}, it follows that

    f,aN=Ωf(x)a(x)N(x)𝑑x>0.\langle f,aN\rangle=\int_{\Omega}f(x)a(x)N(x)dx>0. (39)
  2. 2.

    cL+(Ω)c\in L^{\infty}_{+}(\Omega) and

    c,ϕ=Ωc(x)ϕ(x)𝑑x>0,\langle c,\phi\rangle=\int_{\Omega}c(x)\phi(x)dx>0, (40)

    for all ϕX+{0}\phi\in X_{+}\setminus\{0\}.

  3. 3.

    The following holds uniformly for σΩ\sigma\in\Omega,

    limh0Ω|N(x+h)Θ(x+h,σ)N(x)Θ(x,σ)|𝑑x=0,\lim_{h\to 0}\int_{\Omega}\lvert N(x+h)\Theta(x+h,\sigma)-N(x)\Theta(x,\sigma)\rvert dx=0, (41)

    where we assume that Θ(x,σ)=0\Theta(x,\sigma)=0 for xΩx\notin\Omega.

Under the assumptions 1-2, the next generation operator TT is a positive, bounded linear operator, and we can show that

Proposition 5.

Under the assumptions 1-2, TT is compact and nonsupporting, so r(T)r(T) is the dominant eigenvalue of TT.

Proof: From

La(x)N(x)c,ϕ(Tϕ)(x)αLa(x)N(x)c,ϕ.La(x)N(x)\langle c,\phi\rangle\leq(T\phi)(x)\leq\alpha La(x)N(x)\langle c,\phi\rangle. (42)

We know that TT is a bounded linear operator on XX and it follows that for ϕX+\phi\in X_{+}, and for every integer n1n\geq 1, it holds that

(Tnϕ)(x)Lnc,ϕc,Nan1N(x)a(x).(T^{n}\phi)(x)\geq L^{n}\langle c,\phi\rangle\langle c,Na\rangle^{n-1}N(x)a(x). (43)

By the assumption 2, TnϕT^{n}\phi is a quasi-interior point in the cone X+X_{+}. Then TT is a strictly nonsupporting operator (Definition 10.3, Inaba2017 ). It follows from the positive operator theory that the spectral radius r(T)r(T) is the dominant eigenvalue of TT. Next we show the compactness of TT for the case that Ω=[0,)\Omega=[0,\infty). Let ϕU:={ϕX:ϕXM}\phi\in U:=\{\phi\in X:\|\phi\|_{X}\leq M\}. Then we have TϕXαLcaYM\|T\phi\|_{X}\leq\alpha L\|c\|_{\infty}\|a\|_{Y}M, where c:=supxΩ|c(x)|\|c\|_{\infty}:=\sup_{x\in\Omega}\lvert c(x)\rvert. Next we can see that

limrr|(Tϕ)(x)|𝑑xlimrαLMcrN(x)a(x)𝑑x=0,\lim_{r\to\infty}\int_{r}^{\infty}\lvert(T\phi)(x)\rvert dx\leq\lim_{r\to\infty}\alpha LM\|c\|_{\infty}\int_{r}^{\infty}N(x)a(x)dx=0,
limh+00h|(Tϕ)(x)|𝑑xlimh+0αMLc0hN(x)a(x)𝑑x=0,\lim_{h\to+0}\int_{0}^{h}\lvert(T\phi)(x)\rvert dx\leq\lim_{h\to+0}\alpha ML\|c\|_{\infty}\int_{0}^{h}N(x)a(x)dx=0,

where the convergence is uniform in ϕU\phi\in U. Finally it follows from the condition (41) that

limh+00|(Tϕ)(x+h)(Tϕ)(x)|𝑑x=0,\lim_{h\to+0}\int_{0}^{\infty}\lvert(T\phi)(x+h)-(T\phi)(x)\rvert dx=0,

uniformly in ϕU\phi\in U. Then we can use the Fréchet-Kolomogorov criterion (see Smith2011 , Theorem B.2.) to conclude that T(U)T(U) has compact closure.

Under the assumptions 1-2, the basic reproduction number R0R_{0} is defined by its spectral radius r(T)r(T). In fact, it follows from the well-known Renewal Theorem that B(t)B(t) is asymptotically proportional to er0tϕ(x)e^{r_{0}t}\phi(x), where the growth rate r0r_{0}\in\mathbb{R} and the density ϕX+\phi\in X_{+} satisfy the eigenvalue problem ϕ=K^(r0)ϕ\phi=\hat{K}(r_{0})\phi, where K^(λ)\hat{K}(\lambda), λ\lambda\in\mathbb{C} is the Laplace transform of KK defined by K^(λ):=0eλτK(τ)𝑑τ\hat{K}(\lambda):=\int_{0}^{\infty}e^{-\lambda\tau}K(\tau)d\tau. Then the spectral radius r(K^(λ))r(\hat{K}(\lambda)), λ\lambda\in\mathbb{R} is the positive eigenvalue of K^(λ)\hat{K}(\lambda). Then the intrinsic growth rate (asymptotic Malthusian parameter for BB) r0r_{0}, is given as the real root such that r(K^(r0))=1r(\hat{K}(r_{0}))=1 and ϕ\phi is the positive eigenvector of K^(r0)\hat{K}(r_{0}) associated with its eigenvalue unity. Since the spectral radius r(K^(λ))r(\hat{K}(\lambda)), λ\lambda\in\mathbb{R} is monotonically decreasing with respect to λ\lambda, so the sign relation sign(r0)=sign(R01){\rm sign}(r_{0})={\rm sign}(R_{0}-1) holds, where R0=r(K^(0))=r(T)R_{0}=r(\hat{K}(0))=r(T). (Inaba2017 , chapter 10).

5 Pandemic Threshold Theorem

As is mentioned in Section 3, the generalized pandemic threshold theorem has been proved by Thieme Thieme1977a , Diekmann Diekmann1978 , and Inaba Inaba2014 under different formulations, respectively. Here we give a simple proof under the new assumption.

From (9), limtS(t,x)\lim_{t\to\infty}S(t,x) exists and it holds that

W(,x)=log(S(x)N(x)).W(\infty,x)=-\log\left(\frac{S_{\infty}(x)}{N(x)}\right). (44)

Let tt\to\infty in (10), we have

W(,x)=H(,x)+Ψ(W(,))(x).W(\infty,x)=H(\infty,x)+\Psi(W(\infty,\cdot))(x). (45)

where Ψ\Psi is a nonlinear operator from Y+Y_{+} to Y+Y_{+} defined by

(Ψψ)(x)=ΩΘ(x,σ)N(σ)(1eψ(σ))𝑑σ,ψY+\displaystyle(\Psi\psi)(x)=\int_{\Omega}\Theta(x,\sigma)N(\sigma)(1-e^{-\psi(\sigma)})d\sigma,\quad\psi\in Y_{+} (46)

where H(,)Y+H(\infty,\cdot)\in Y_{+}. To estimate W(,)W(\infty,\cdot), the fixed point of Ψ\Psi plays a key role.

Proposition 6.

If R01R_{0}\leq 1, Ψ\Psi has no positive fixed point.

Proof: Let ψY+\psi\in Y_{+} be a positive fixed point in Y+Y_{+}. Then we have NψX+N\psi\in X_{+} and

N(x)ψ(x)\displaystyle N(x)\psi(x) =N(x)ΩΘ(x,σ)N(σ)(1eψ(σ))𝑑σ\displaystyle=N(x)\int_{\Omega}\Theta(x,\sigma)N(\sigma)(1-e^{-\psi(\sigma)})d\sigma (47)
N(x)ΩΘ(x,σ)N(σ)ψ(σ)𝑑σ=(TNψ)(x),\displaystyle\leq N(x)\int_{\Omega}\Theta(x,\sigma)N(\sigma)\psi(\sigma)d\sigma=(TN\psi)(x),

Let fX+f^{*}\in X_{+}^{*} be the adjoint eigenfunctional of TT associated with the eigenvalue r(T)=R0r(T)=R_{0}. Taking the duality pairing in (47), we have

f,Nψf,TNψ=R0f,Nψ,\langle f^{*},N\psi\rangle\leq\langle f^{*},TN\psi\rangle=R_{0}\langle f^{*},N\psi\rangle,

where f,ϕ\langle f^{*},\phi\rangle denotes the value of ff^{*} at ϕX\phi\in X, and the equality holds only if ψ=0\psi=0. Then we have R0>1R_{0}>1. So if R01R_{0}\leq 1, there is no positive fixed point.

Although we skip the proof, it is intuitively clear that W(,)W(\infty,\cdot) is monotonically non-decreasing with respect to H(,)H(\infty,\cdot). Then Proposition 6 implies that if R01R_{0}\leq 1, W(,)0W(\infty,\cdot)\to 0 when H(,)0H(\infty,\cdot)\to 0.

Let P:YXP:Y\to X be the isomorphism between XX and YY such that Pψ=NψP\psi=N\psi. Define an operator Φ\Phi from XX into itself by Φ=PΨP1\Phi=P\Psi P^{-1}. Then we have

Φ(ϕ)(x)=N(x)ΩΘ(x,σ)N(σ)(1eN1(σ)ϕ(σ))𝑑σ,ϕX+.\Phi(\phi)(x)=N(x)\int_{\Omega}\Theta(x,\sigma)N(\sigma)(1-e^{-N^{-1}(\sigma)\phi(\sigma)})d\sigma,\quad\phi\in X_{+}. (48)
Proposition 7.

Under the assumption 1-2, Φ\Phi is a compact and concave operator222The definition of the concave operator is given in Krasnoselskii Krasnoselskii1964 . The reader may consult Inaba1990 and Inaba2017 . in XX.

Proof: First we show that Φ\Phi is a compact operator. For simplicity, we consider the case that Ω=[0,)\Omega=[0,\infty). As is seen in Proposition 5, it is sufficient to check the conditions for the compactness of Φ(U)X=L1(Ω)\Phi(U)\subset X=L^{1}(\Omega) for any bounded set UX+U\subset X_{+} (Theorem B.2. in Smith2011 ). In fact, it is easy to see Φ(ϕ)XαLNaXc,N\|\Phi(\phi)\|_{X}\leq\alpha L\|Na\|_{X}\langle c,N\rangle for any ϕX+\phi\in X_{+}. Then the condition supϕX+Φ(ϕ)X<\sup_{\phi\in X_{+}}\|\Phi(\phi)\|_{X}<\infty holds. Next it holds that

limrr|Φ(ϕ)(x)|𝑑xlimrαLc,NrN(x)a(x)𝑑x=0,\lim_{r\to\infty}\int_{r}^{\infty}\lvert\Phi(\phi)(x)\rvert dx\leq\lim_{r\to\infty}\alpha L\langle c,N\rangle\int_{r}^{\infty}N(x)a(x)dx=0,
limh+00h|Φ(ϕ)(x)|𝑑xlimrαLc,N0hN(x)a(x)𝑑x=0.\lim_{h\to+0}\int_{0}^{h}\lvert\Phi(\phi)(x)\rvert dx\leq\lim_{r\to\infty}\alpha L\langle c,N\rangle\int_{0}^{h}N(x)a(x)dx=0.

Finally let

J:=0|Φ(ϕ)(x+h)Φ(ϕ)(x)|𝑑x.J:=\int_{0}^{\infty}\lvert\Phi(\phi)(x+h)-\Phi(\phi)(x)\rvert dx.

Then we can observe that for ϕX+\phi\in X_{+},

Ω|Φ(ϕ)(x+h)Φ(ϕ)(x)|𝑑xΩΩ|N(x+h)Θ(x+h,σ)N(x)Θ(x,σ)|𝑑xN(σ)𝑑σ.\int_{\Omega}\lvert\Phi(\phi)(x+h)-\Phi(\phi)(x)\rvert dx\leq\int_{\Omega}\int_{\Omega}\lvert N(x+h)\Theta(x+h,\sigma)-N(x)\Theta(x,\sigma)\rvert dxN(\sigma)d\sigma. (49)

where it follows from the assumption 2 that the right-hand side of (49) goes to zero uniformly for ϕX+\phi\in X_{+} when h0h\to 0. Then we conclude that Φ(X+)\Phi(X_{+}) has compact closure. Next, we prove the concavity of Φ\Phi. First it is clear that Φ\Phi is monotone non-decreasing. Next, from the fact that 1etxt(1ex)1-e^{-tx}\geq t(1-e^{-x}) for 0t10\leq t\leq 1 and x0x\geq 0, it follows that for 0t10\leq t\leq 1 and ϕX+\phi\in X_{+},

Φ(tϕ)tΦ(ϕ),0t1,\Phi(t\phi)\geq t\Phi(\phi),\quad 0\leq t\leq 1, (50)

Now we can observe that Φ\Phi is u0u_{0}-positive, that is, there exists a nonzero element u0X+u_{0}\in X_{+} such that for a nonzero ϕX+\phi\in X_{+} the inequality

ϵ(ϕ)u0Φ(ϕ)αϵ(ϕ)u0,\epsilon(\phi)u_{0}\leq\Phi(\phi)\leq\alpha\epsilon(\phi)u_{0}, (51)

holds, where ϵ\epsilon denotes a positive functional on X+X_{+}. In fact, under the assumption 1, it holds that for ϕX+\phi\in X_{+},

La(x)N(x)ϵ(ϕ)Φ(ϕ)(x)αLa(x)N(x)ϵ(ϕ),La(x)N(x)\epsilon(\phi)\leq\Phi(\phi)(x)\leq\alpha La(x)N(x)\epsilon(\phi), (52)

where ϵ\epsilon is a strictly positive functional on X+X_{+} given by

ϵ(ϕ):=Ωc(σ)N(σ)(1eN1(σ)ϕ(σ))𝑑σ.\epsilon(\phi):=\int_{\Omega}c(\sigma)N(\sigma)(1-e^{-N^{-1}(\sigma)\phi(\sigma)})d\sigma. (53)

Therefore if we choose u0(x)=La(x)N(x)u_{0}(x)=La(x)N(x), we obtain (51), and (50) and (51) are sufficient conditions for concavity of Φ\Phi.

Proposition 8.

Φ\Phi has at most one positive fixed point.

Proof: Observe that for k(0,1)k\in(0,1),

Φ(kϕ)kΦ(ϕ)\displaystyle\Phi(k\phi)-k\Phi(\phi) =N(x)ΩΘ(x,σ)N(σ)(1ekN1(σ)k(1eN1(σ)ϕ(σ)))𝑑σ\displaystyle=N(x)\int_{\Omega}\Theta(x,\sigma)N(\sigma)(1-e^{-kN^{-1}(\sigma)}-k(1-e^{-N^{-1}(\sigma)\phi(\sigma)}))d\sigma (54)
a(x)N(x)η(ϕ;k),\displaystyle\geq a(x)N(x)\eta(\phi\mathchar 24635\relax\;k),

where

η(ϕ;k):=LΩc(σ)N(σ)(1ekN1(σ)ϕ(σ)k(1eN1(σ)ϕ(σ)))𝑑σ.\eta(\phi\mathchar 24635\relax\;k):=L\int_{\Omega}c(\sigma)N(\sigma)(1-e^{-kN^{-1}(\sigma)\phi(\sigma)}-k(1-e^{-N^{-1}(\sigma)\phi(\sigma)}))d\sigma.

Then it is easy to see that η(ϕ,k)>0\eta(\phi,k)>0 for any ϕX+{0}\phi\in X_{+}\setminus\{0\} and k(0,1)k\in(0,1). Therefore, Φ\Phi is monotone and concave in X+X_{+} and for any k(0,1)k\in(0,1), it follows that

Φ(kϕ)kΦ(ϕ)+η(ϕ,k)a(x)N(x).\Phi(k\phi)\geq k\Phi(\phi)+\eta(\phi,k)a(x)N(x). (55)

where aNX+aN\in X_{+} is a quasi-interior point. Then we can apply Lemma 4.8 in Inaba1990 , Φ\Phi has at most one positive fixed point.

Proposition 9.

If R0>1R_{0}>1, Ψ\Psi has a unique positive fixed point.

Proof: It is easy to see that Φ\Phi maps the positive cone X+X_{+} into the bounded convex subset {ϕX+:ϕXαLNaX}\{\phi\in X_{+}:\|\phi\|_{X}\leq\alpha L\|Na\|_{X}\}, and its Fréchet derivative at the origin, denoted by Φ[0]\Phi^{\prime}[0], is the next generation operator TT. Then it follows from Krasnoselskii’s theorem (Theorem 4.11 in Krasnoselskii1964 , Inaba1990 , Inaba2017 ) that Φ\Phi has at least one positive fixed point if r(Φ[0])=r(T)=R0>1r(\Phi^{\prime}[0])=r(T)=R_{0}>1, and it is unique from Proposition 8. Therefore, Ψ=P1ΦP\Psi=P^{-1}\Phi P has a unique fixed point in Y+Y_{+}

Proposition 10.

Suppose that Ψ\Psi has a unique positive fixed point VY+{0}V\in Y_{+}\setminus\{0\}. Then it holds that

W(,x)V(x).W(\infty,x)\geq V(x). (56)

Proof: Define a positive operator \mathcal{H} in Y+Y_{+} by (ϕ)(x):=H(,x)+Ψ(ϕ)(x)(\mathcal{H}\phi)(x):=H(\infty,x)+\Psi(\phi)(x). Then W(,)W(\infty,\cdot) is its fixed point and \mathcal{H} is a monotone operator. Define a sequence ϕn\phi_{n} by ϕn=(ϕn1)\phi_{n}=\mathcal{H}(\phi_{n-1}) (n1n\geq 1) and ϕ0=V\phi_{0}=V. Then we have ϕ1Ψ(ϕ0)=ϕ0\phi_{1}\geq\Psi(\phi_{0})=\phi_{0}, so it follows iteratively that ϕnϕn1\phi_{n}\geq\phi_{n-1}. Then we conclude that limnϕn(x)=W(,x)ϕ0(x)=V(x)\lim_{n\to\infty}\phi_{n}(x)=W(\infty,x)\geq\phi_{0}(x)=V(x).

From propositions 8, 9 and 10, we know that is, if R0>1R_{0}>1, the epidemic occurs for all traits xΩx\in\Omega no matter how small the initial size of the infecteds is, and its final size distribution is estimated from below by the final size distribution VV. On the other hand, the final size goes to zero if the initial size of infecteds goes to zero under the condition R01R_{0}\leq 1.

6 The effective reproduction number, control and HIT

In order to consider the herd immunity threshold and to estimate the effect of preventive control, we introduce here the idea of the effective reproduction number. Preventive control that works by reducing the availability of susceptibles in the host population is called S-control, while the I-control means that the preventive control acts primarily on the infectivity of the host individuals Heesterbeek2006 .

Let Z:=L(Ω)YZ:=L^{\infty}(\Omega)\subset Y. For ϕZ+\phi\in Z_{+}, we define the effective next generation operator at state ϕZ+\phi\in Z_{+} for S-control by TS[ϕ]:X+X+T_{S}[\phi]:X_{+}\to X_{+}:

(TS[ϕ]ψ)(x):=N(x)eϕ(x)ΩΘ(x,σ)ψ(σ)𝑑σ,ψX,(T_{S}[\phi]\psi)(x):=N(x)e^{-\phi(x)}\int_{\Omega}\Theta(x,\sigma)\psi(\sigma)d\sigma,~{}\psi\in X, (57)

and for I-control by TI[ϕ]:X+X+T_{I}[\phi]:X_{+}\to X_{+}:

(TI[ϕ]ψ)(x):=N(x)ΩΘ(x,σ)eϕ(σ)ψ(σ)𝑑σ,ψX,(T_{I}[\phi]\psi)(x):=N(x)\int_{\Omega}\Theta(x,\sigma)e^{-\phi(\sigma)}\psi(\sigma)d\sigma,~{}\psi\in X, (58)

which are both bounded linear operators in XX. From our assumption 2, both operators are compact and nonsupporting, and TS[0]=TI[0]T_{S}[0]=T_{I}[0] is none other than the next generation operator.

The biological meaning of the state ϕ\phi in the S-control is that eϕ(x)e^{-\phi(x)} is the reduction of susceptibility, or the fraction 1eϕ(x)1-e^{-\phi(x)} is immunized by vaccination or by natural infection, so the susceptible density N(x)N(x) is replaced by N(x)eϕ(x)N(x)e^{-\phi(x)} in the next generation operator TT to get TST_{S}. On the other hand, in the I-control, eϕ(x)e^{-\phi(x)} is the reduction of infectivity, or 1eϕ(x)1-e^{-\phi(x)} is removed from the infectious state by vaccination or by non-pharmaceutical intervention.

The spectral radius of TS[ϕ]T_{S}[\phi] (TI[ϕ]T_{I}[\phi]) is defined as the effective reproduction number in the state ϕ\phi for S-control (I-control), denoted by RS[ϕ]R_{S}[\phi] (RI[ϕ]R_{I}[\phi]), that is,

RS[ϕ]=r(TS[ϕ]),RI[ϕ]=r(TI[ϕ]),R_{S}[\phi]=r(T_{S}[\phi]),\quad R_{I}[\phi]=r(T_{I}[\phi]), (59)

Note that RS[0]=RI[0]R_{S}[0]=R_{I}[0] is none other than the basic reproduction number R0R_{0}.

Proposition 11.

If ϕZ+\phi\in Z_{+}, it holds that RS[ϕ]=RI[ϕ]R_{S}[\phi]=R_{I}[\phi].

Proof: Let U[ϕ]:X+X+U[\phi]:X_{+}\to X_{+} be a positive operator defined by (U[ϕ]ψ)(x)=eϕ(x)ψ(x)(U[\phi]\psi)(x)=e^{-\phi(x)}\psi(x). Then if ϕZ+\phi\in Z_{+}, then U1U^{-1} is a bounded linear operator, and it follows that TS[ϕ]=UTI[ϕ]U1T_{S}[\phi]=UT_{I}[\phi]U^{-1}, that is, TS[ϕ]T_{S}[\phi] and TI[ϕ]T_{I}[\phi] are similar operators to each other. Then RS[ϕ]=RI[ϕ]R_{S}[\phi]=R_{I}[\phi] holds.

If the vaccination coverage is given by ϵ(x)\epsilon(x), then the susceptible population is (1ϵ(x))N(x)(1-\epsilon(x))N(x). Let the state ϕ\phi be given by ϕ(x)=log(1ϵ(x))\phi(x)=-\log(1-\epsilon(x)), then ϵ\epsilon is the vaccination coverage, and the partially immunized host population has the effective reproduction number given by RS[ϕ]R_{S}[\phi].

Let Ψ[ϕ]\Psi^{\prime}[\phi] be the Fréchet derivative of Ψ\Psi at ϕZ+\phi\in Z_{+}. Then it holds that

(Ψ[ϕ]y)(x)=ΩΘ(x,σ)N(σ)eϕ(σ)y(σ)𝑑σ,yZ+.(\Psi^{\prime}[\phi]y)(x)=\int_{\Omega}\Theta(x,\sigma)N(\sigma)e^{-\phi(\sigma)}y(\sigma)d\sigma,\quad y\in Z_{+}. (60)
Proposition 12.

For the effective reproduction number at state ϕZ+\phi\in Z_{+}, it follows that

RS[ϕ]=RI[ϕ]=r(Ψ[ϕ]).R_{S}[\phi]=R_{I}[\phi]=r(\Psi^{\prime}[\phi]). (61)

Proof: From proposition 11, we have RS[ϕ]=RI[ϕ]R_{S}[\phi]=R_{I}[\phi]. Let PϕP_{\phi} be a bounded linear operator from YY to XX given by (Pϕy)(x)=N(x)eϕ(x)y(x)(P_{\phi}y)(x)=N(x)e^{-\phi(x)}y(x). Then Ψ[ϕ]\Psi^{\prime}[\phi] is a similar operator of TS[ϕ]T_{S}[\phi], that is,

Pϕ1TS[ϕ]Pϕ=Ψ[ϕ].P_{\phi}^{-1}T_{S}[\phi]P_{\phi}=\Psi^{\prime}[\phi]. (62)

Therefore we have RS[ϕ]=r(TS[ϕ])=r(Ψ[ϕ])R_{S}[\phi]=r(T_{S}[\phi])=r(\Psi^{\prime}[\phi]).

If W(t,x)W(t,x) is the cumulative FOI of the renewal equation (10), then

(TS[W(t,)]ψ)(x)\displaystyle(T_{S}[W(t,\cdot)]\psi)(x) =N(x)exp(W(t,x))ΩΘ(x,σ)ψ(σ)𝑑σ\displaystyle=N(x)\exp(-W(t,x))\int_{\Omega}\Theta(x,\sigma)\psi(\sigma)d\sigma (63)
=S(t,x)ΩΘ(x,σ)ψ(σ)𝑑σ.\displaystyle=S(t,x)\int_{\Omega}\Theta(x,\sigma)\psi(\sigma)d\sigma.

Then RS[W(t,)]R_{S}[W(t,\cdot)] gives the effective reproduction number at time tt in the natural epidemic process described by (10). In particular, we call RS[W(,)]R_{S}[W(\infty,\cdot)] the final reproduction number, which gives the reproduction number for the remaining susceptible population after the end of the epidemics.

Proposition 13.

Let W(t,x)W(t,x) be the cumulative FOI for (10). Then RS[W(t,)]R_{S}[W(t,\cdot)] is monotonically decreasing with respect to tt and RS[W(,)]<1R_{S}[W(\infty,\cdot)]<1, and there exists a unique t>0t^{*}>0 such that RS[W(t,)]=1R_{S}[W(t^{*},\cdot)]=1 if R0>1R_{0}>1.

Proof: Note that RS[W(t,)]R_{S}[W(t,\cdot)] is monotonically decreasing with respect to tt, since W(t,)W(t,\cdot) is monotonically increasing. It follows from 1exexx1-e^{-x}\geq e^{-x}x for x0x\geq 0 that

W(,x)\displaystyle W(\infty,x) =ΩΘ(x,σ)N(σ)(1eW(,σ))𝑑σ\displaystyle=\int_{\Omega}\Theta(x,\sigma)N(\sigma)(1-e^{-W(\infty,\sigma)})d\sigma (64)
ΩΘ(x,σ)N(σ)eW(,σ)W(,σ)𝑑σ\displaystyle\geq\int_{\Omega}\Theta(x,\sigma)N(\sigma)e^{-W(\infty,\sigma)}W(\infty,\sigma)d\sigma
=(Ψ[W(,)]W(,))(x).\displaystyle=(\Psi^{\prime}[W(\infty,\cdot)]W(\infty,\cdot))(x).

From (64), we obtain an inequality in XX:

(PϕW(,))(x)(TS[ϕ]PϕW(,))(x).(P_{\phi}W(\infty,\cdot))(x)\geq(T_{S}[\phi]P_{\phi}W(\infty,\cdot))(x). (65)

Let fϕX+f^{*}_{\phi}\in X^{*}_{+} be the adjoint eigenfunctional of TS[ϕ]T_{S}[\phi] associated with the eigenvalue RS[W(,)]=r(Ψ[W(,)])=r(TS[ϕ])R_{S}[W(\infty,\cdot)]=r(\Psi^{\prime}[W(\infty,\cdot)])=r(T_{S}[\phi]). Taking the duality pairing in (65), it follows that fϕ,PϕW(,)RS[W(,)]fϕ,PϕW(,)\langle f^{*}_{\phi},P_{\phi}W(\infty,\cdot)\rangle\geq R_{S}[W(\infty,\cdot)]\langle f^{*}_{\phi},P_{\phi}W(\infty,\cdot)\rangle, where the equality does not hold because W(,)0W(\infty,\cdot)\neq 0. Then we know that RS[W(,)]<1R_{S}[W(\infty,\cdot)]<1. If R0=RS[0]>1R_{0}=R_{S}[0]>1, there exists a unique t>0t^{*}>0 such that RS[W(t,)]=1R_{S}[W(t^{*},\cdot)]=1, because r(Ψ[W(t,)])r(\Psi^{\prime}[W(t,\cdot)]) is continuous, monotone decreasing with respect to tt, and it moves from R0>1R_{0}>1 to RS[W(,)]<1R_{S}[W(\infty,\cdot)]<1 as tt goes from 0 to \infty.

Traditionally, the herd immunity threshold (HIT) is defined as the proportion of the host population that must be immune to prevent an outbreak. When the HIT is reached by mass vaccination, it can be called the critical coverage of immunization (CCI).

Define a subset 𝒲Z+\mathcal{W}\subset Z_{+} by

𝒲:={ϕZ+:RS[ϕ]=RI[ϕ]=1}.\mathcal{W}:=\left\{\phi\in Z_{+}:R_{S}[\phi]=R_{I}[\phi]=1\right\}. (66)

and let ϵ(x)=1eϕ(x)\epsilon(x)=1-e^{-\phi(x)} for ϕ𝒲\phi\in\mathcal{W}, then ϵ\epsilon gives the critical coverage of intervention for S-control, or for I-control. For the S-control, the HIT for ϕ𝒲\phi\in\mathcal{W} is calculated as

HIT[ϕ]=11NΩN(x)eϕ(x)𝑑x=Ωϵ(x)ω(x)𝑑x,{\rm HIT[\phi]}=1-\frac{1}{N}\int_{\Omega}N(x)e^{-\phi(x)}dx=\int_{\Omega}\epsilon(x)\omega(x)dx, (67)

where ω(x)=N(x)/ΩN(σ)𝑑σ\omega(x)=N(x)/\int_{\Omega}N(\sigma)d\sigma denotes the trait profile of the host population.

That is, the HIT is not uniquely determined for the heterogeneous population. As is seen in Proposition 13, for the natural epidemic process described by (10), if R0>1R_{0}>1, there exists a unique time tt^{*} such that W(t,)𝒲W(t^{*},\cdot)\in\mathcal{W}. Then the HIT is reached at time tt^{*}, but the time tt^{*} and the HIT for the natural infection generally depend on the initial data. For the I-control, for ϕ𝒲\phi\in\mathcal{W}, eϕ(x)e^{-\phi(x)} gives the critical reduction level of infectivity.

7 Results of the separable mixing assumption

Here we consider the results of the separable mixing assumption:

A(τ,x,σ)=a(x)b(τ)c(σ),H(t,x)=a(x)Q(t),A(\tau,x,\sigma)=a(x)b(\tau)c(\sigma),\quad H(t,x)=a(x)Q(t), (68)

where we assume that QQ is a bounded continuous positive function on +\mathbb{R}_{+} and that limtQ(t)=Q()>0\lim_{t\to\infty}Q(t)=Q(\infty)>0 exists. Then we can concretely compute the basic indices defined so far.

7.1 RS[ϕ]R_{S}[\phi] and HIT

From (60) and (68), we easily obtain expressions:

(Ψ[ϕ]y)(x)=La(x)Ωc(σ)N(σ)eϕ(σ)y(σ)𝑑σ,yY+.(\Psi^{\prime}[\phi]y)(x)=La(x)\int_{\Omega}c(\sigma)N(\sigma)e^{-\phi(\sigma)}y(\sigma)d\sigma,~{}y\in Y_{+}. (69)
RS[ϕ]\displaystyle R_{S}[\phi] =r(Ψ[ϕ])=LΩc(σ)N(σ)eϕ(σ)a(σ)𝑑σ\displaystyle=r(\Psi^{\prime}[\phi])=L\int_{\Omega}c(\sigma)N(\sigma)e^{-\phi(\sigma)}a(\sigma)d\sigma (70)
=R0Ωκ(x)eϕ(x)𝑑x,\displaystyle=R_{0}\int_{\Omega}\kappa(x)e^{-\phi(x)}dx,

where

κ(x):=a(x)c(x)N(x)Ωa(σ)c(σ)N(σ)𝑑σ,\kappa(x):=\frac{a(x)c(x)N(x)}{\int_{\Omega}a(\sigma)c(\sigma)N(\sigma)d\sigma}, (71)

Then it follows that

𝒲\displaystyle\mathcal{W} ={ϕZ+:LΩc(σ)N(σ)eϕ(σ)a(σ)𝑑σ=1}.\displaystyle=\left\{\phi\in Z_{+}:L\int_{\Omega}c(\sigma)N(\sigma)e^{-\phi(\sigma)}a(\sigma)d\sigma=1\right\}. (72)
={ϕZ+:Ωκ(x)eϕ(x)𝑑x=1R0}.\displaystyle=\left\{\phi\in Z_{+}:\int_{\Omega}\kappa(x)e^{-\phi(x)}dx=\frac{1}{R_{0}}\right\}.

Note that 𝒲\mathcal{W}\neq\emptyset, because a constant function

ϕ(x)=log(LΩc(σ)N(σ)a(σ)𝑑σ)=logR0,\phi(x)=\log\left(L\int_{\Omega}c(\sigma)N(\sigma)a(\sigma)d\sigma\right)=\log R_{0}, (73)

belongs to 𝒲\mathcal{W}. In this constant case, the HIT (or CCI) is given by

1eϕ=11R0,1-e^{-\phi}=1-\frac{1}{R_{0}}, (74)

which traditional result implies that the epidemic will not occur if the fraction greater than 11/R01-1/R_{0} is immunized in the susceptible host population.

Example 2

It is easy to see that increasing the immunization fraction in populations with larger κ(x)\kappa(x) is more effective to lower the effective reproduction number. Let us show an example. Again let us assume that Ω=[0,)\Omega=[0,\infty), a(x)=xa(x)=x and cc is constant. Then we have

κ(x)=xω(x)0xω(x)𝑑x.\kappa(x)=\frac{x\omega(x)}{\int_{0}^{\infty}x\omega(x)dx}. (75)

Suppose that

0(ϵ(x)ϵ,ω)(xx,ω)ω(x)𝑑x0,\int_{0}^{\infty}(\epsilon(x)-\langle\epsilon,\omega\rangle)(x-\langle x,\omega\rangle)\omega(x)dx\geq 0, (76)

where f,ω:=0f(x)ω(x)𝑑x\langle f,\omega\rangle:=\int_{0}^{\infty}f(x)\omega(x)dx denotes the mean of ff. Then we obtain

ϵ,κϵ,ω,\langle\epsilon,\kappa\rangle\geq\langle\epsilon,\omega\rangle, (77)

because if we expand (76), we get ϵx,ωϵ,ωx,ω0\langle\epsilon x,\omega\rangle-\langle\epsilon,\omega\rangle\langle x,\omega\rangle\geq 0. It follows from ϵx,ω/x,ω=ϵ,κ\langle\epsilon x,\omega\rangle/\langle x,\omega\rangle=\langle\epsilon,\kappa\rangle that ϵ,κϵ,ω\langle\epsilon,\kappa\rangle\geq\langle\epsilon,\omega\rangle. Therefore, if ϕ(x)=log(1ϵ(x))𝒲\phi^{*}(x)=-\log(1-\epsilon^{*}(x))\in\mathcal{W}, we have

ϵ,κ=11R0ϵ,ω.\langle\epsilon^{*},\kappa\rangle=1-\frac{1}{R_{0}}\geq\langle\epsilon^{*},\omega\rangle. (78)

The condition (76) and its result (78) imply that if the group with above-average susceptibility is more immunized than the average immunization rate, the critical immunization fraction of the population ϵ,ω\langle\epsilon^{*},\omega\rangle is less than 11/R0=ϵ,κ1-1/R_{0}=\langle\epsilon^{*},\kappa\rangle, which is the HIT for the homogeneous case. It is intuitively obvious that selective immunization of susceptible groups will achieve herd immunity by immunizing a smaller percentage of the population.

7.2 HIT in the natural epidemics

In order to calculate the HIT in the natiral epidemic, observe that (10) is written as

W(t,x)=a(x)Q(t)+a(x)Ω0tb(τ)c(σ)N(σ)(1eW(tτ,σ))𝑑τ𝑑σ.W(t,x)=a(x)Q(t)+a(x)\int_{\Omega}\int_{0}^{t}b(\tau)c(\sigma)N(\sigma)(1-e^{-W(t-\tau,\sigma)})d\tau d\sigma. (79)

Therefore, there exists a function w(t)w(t) such that W(t,a)=a(x)w(t)W(t,a)=a(x)w(t) and

w(t)=Q(t)+Ω0tb(τ)c(σ)N(σ)(1ea(σ)w(tτ))𝑑τ𝑑σ.w(t)=Q(t)+\int_{\Omega}\int_{0}^{t}b(\tau)c(\sigma)N(\sigma)(1-e^{-a(\sigma)w(t-\tau)})d\tau d\sigma. (80)

From Proposition 9, we have

RS[W(t,)]=r(Ψ[a()w(t)])=LΩc(σ)a(σ)ea(σ)w(t)N(σ)𝑑σ.R_{S}[W(t,\cdot)]=r(\Psi^{\prime}[a(\cdot)w(t)])=L\int_{\Omega}c(\sigma)a(\sigma)e^{-a(\sigma)w(t)}N(\sigma)d\sigma. (81)

In this case, w(t)w(t^{*}) satisfies

RS[W(t,)]=RS[a()w(t)]=1,R_{S}[W(t^{*},\cdot)]=R_{S}[a(\cdot)w(t^{*})]=1, (82)

is uniquely determined as the unique positive root λ\lambda of the characteristic equation

LΩc(σ)a(σ)ea(σ)λN(σ)𝑑σ=1.L\int_{\Omega}c(\sigma)a(\sigma)e^{-a(\sigma)\lambda}N(\sigma)d\sigma=1. (83)

Since λ\lambda is independent of the initial data, the HIT in the natural epidemics is calculated as

HIT[aλ]=1Ωω(x)ea(x)λ𝑑x,{\rm HIT}[a\lambda]=1-\int_{\Omega}\omega(x)e^{-a(x)\lambda}dx, (84)

and it is also independent from the initial data, although the time tt^{*} such that λ=w(t)\lambda=w(t^{*}) depends on the initial data QQ.

Example 3

Suppose that Ω=+\Omega=\mathbb{R}_{+}, a(x)=xa(x)=x, c(σ)=cc(\sigma)=c is constant and the trait profile ω\omega is given by the gamma function with mean 11 and variance 1/p1/p:

ω(x)=ppΓ(p)xp1exp.\omega(x)=\frac{p^{p}}{\Gamma(p)}x^{p-1}e^{-xp}. (85)

Using the formula (81), we can compute the effective reproduction number as follows

RS[W(t,)]\displaystyle R_{S}[W(t,\cdot)] =NLc0ω(σ)σeσw(t)𝑑σ\displaystyle=NLc\int_{0}^{\infty}\omega(\sigma)\sigma e^{-\sigma w(t)}d\sigma (86)
=NLcpp+1(p+w(t))p+1=R0s(t)1+1p,\displaystyle=\frac{NLcp^{p+1}}{(p+w(t))^{p+1}}=R_{0}s(t)^{1+\frac{1}{p}},

where

s(t):=1N0S(t,x)𝑑x,s(t):=\frac{1}{N}\int_{0}^{\infty}S(t,x)dx,

denotes the susceptible fraction at time tt, and it is calculated as

s(t)=(pp+w(t))p.s(t)=\left(\frac{p}{p+w(t)}\right)^{p}. (87)

Then the HIT in the natural epidemics is given by

1s(t)=1R0pp+1,1-s(t^{*})=1-R_{0}^{-\frac{p}{p+1}}, (88)

which may be far from the homogeneous case of p=p=\infty. These power-law results have been used in many studies related to COVID-19 Bootsma2023b ; Diekmann2023 ; Montalban2022 ; Tkachenko2021a .

7.3 Final size

In the renewal equation (80), if we let tt\to\infty, we obtain an equation for w()w(\infty):

w()=Q()+LΩc(σ)N(σ)(1ea(σ)w())𝑑σ.w(\infty)=Q(\infty)+L\int_{\Omega}c(\sigma)N(\sigma)(1-e^{-a(\sigma)w(\infty)})d\sigma. (89)

Then the final susceptible distribution is given by

S(x)=N(x)ea(x)w(),S_{\infty}(x)=N(x)e^{-a(x)w(\infty)}, (90)

and the attack rate (final size distribution) is calculated as

1S(x)N(x)=1ea(x)w()1-\frac{S_{\infty}(x)}{N(x)}=1-e^{-a(x)w(\infty)} (91)

If R0>1R_{0}>1,

F(x):=LΩc(σ)N(σ)(1ea(σ)x)𝑑σ.F(x):=L\int_{\Omega}c(\sigma)N(\sigma)(1-e^{-a(\sigma)x})d\sigma. (92)

has a unique fixed point v>0v>0 and w()vw(\infty)\geq v. Then the final size is estimated from below as

1S(x)N(x)1ea(x)v.1-\frac{S_{\infty}(x)}{N(x)}\geq 1-e^{-a(x)v}. (93)

7.4 I-control and epidemic resurgence

As was seen above, the final reproduction number Re()R_{e}(\infty) is less than unity, the epidemic resurgence does not occur after the natural eradication of the disease. However, if the infectivity is reduced by non-pharmaceutical intervention (such as universal masking and social distance), after the end of the controlled outbreak, there could remain susceptible population whose effective reproduction number is greater than one. Then if the intervention is lifted after the end of the first epidemic wave, the resurgence of the epidemic could occur. To formulate the condition for the resurgence, we calculate the critical level of the control for the infectivity.

Again we adopt the separable mixing assumption. Let S[ϕ](x)=N(x)ewϕ()a(x)S_{\infty}[\phi](x)=N(x)e^{-w_{\phi}(\infty)a(x)} be the final susceptible distribution under the I-control with state ϕ\phi, where wϕ()w_{\phi}(\infty) is the positive root of the final size equation

wϕ()=Q()+LΩc(x)eϕ(x)N(x)(1ewϕ()a(x))𝑑x.w_{\phi}(\infty)=Q(\infty)+L\int_{\Omega}c(x)e^{-\phi(x)}N(x)(1-e^{-w_{\phi}(\infty)a(x)})dx. (94)

Then the root wϕ()w_{\phi}(\infty) is non-increasing with respect to ϕ\phi.

The effective next generation operator after lifting the prevention policy is given by

(Twϕ()aϕ)(x)=La(x)N(x)ewϕ()a(x)c,ϕ.(T_{w_{\phi}(\infty)a}\phi)(x)=La(x)N(x)e^{-w_{\phi}(\infty)a(x)}\langle c,\phi\rangle. (95)

and the effective reproduction number is

r(Twϕ()a)=LΩc(σ)N(σ)ewϕ()a(σ)a(σ)𝑑σ.r(T_{w_{\phi}(\infty)a})=L\int_{\Omega}c(\sigma)N(\sigma)e^{-w_{\phi}(\infty)a(\sigma)}a(\sigma)d\sigma. (96)

Then we can define the set of critical I-control by

Σ:={ϕZ+:LΩc(σ)N(σ)ewϕ()a(σ)a(σ)𝑑σ=1}.\Sigma:=\left\{\phi\in Z_{+}:L\int_{\Omega}c(\sigma)N(\sigma)e^{-w_{\phi}(\infty)a(\sigma)}a(\sigma)d\sigma=1\right\}. (97)

Let λ\lambda^{*} be the unique positive root of the characteristic equation:

LΩc(σ)N(σ)eλa(σ)a(σ)𝑑σ=1,L\int_{\Omega}c(\sigma)N(\sigma)e^{-\lambda^{*}a(\sigma)}a(\sigma)d\sigma=1, (98)

whose existence is guaranteed by the assumption R0=LΩc(x)N(x)a(x)𝑑x>1R_{0}=L\int_{\Omega}c(x)N(x)a(x)dx>1. So, we know that

Σ\displaystyle\Sigma ={ϕZ+:wϕ()=λ}\displaystyle=\{\phi\in Z_{+}:w_{\phi}(\infty)=\lambda^{*}\} (99)
={ϕ:λ=Q()+LΩc(x)eϕ(x)N(x)(1eλa(x))𝑑x}.\displaystyle=\left\{\phi:\lambda^{*}=Q(\infty)+L\int_{\Omega}c(x)e^{-\phi(x)}N(x)(1-e^{-\lambda^{*}a(x)})dx\right\}.

If λ>Q()\lambda^{*}>Q(\infty), Σ\Sigma is not empty. In fact, if assume that ϕ\phi is a constant function ϕ(x)=ϕ\phi(x)=\phi^{*}, we can easily see that

ϕ=log(λQ()LΩc(x)N(x)(1eλa(x))𝑑x)Σ.\phi^{*}=-\log\left(\frac{\lambda^{*}-Q(\infty)}{L\int_{\Omega}c(x)N(x)(1-e^{-\lambda^{*}a(x)})dx}\right)\in\Sigma. (100)

If the I-control is so strong as ϕ(x)>ϕ\phi(x)>\phi^{*} for all xΩx\in\Omega, the resurgence occurs if the prevention policy is lifted after the previous epidemics have ended.

Example 4

In general, it is difficult to calculate the threshold value λ\lambda^{*} explicitly. However, this is not the case if we consider the limiting epidemics described by the homogeneous model with constant parameters. Assume that the host population is homogeneous, all parameters are constant, so assume that a(x)=βa(x)=\beta, b(τ)=eγτb(\tau)=e^{-\gamma\tau} and c=1c=1. From (98), we have

LΩc(σ)N(σ)eλa(σ)a(σ)𝑑σ=R0eλβ=1,L\int_{\Omega}c(\sigma)N(\sigma)e^{-\lambda^{*}a(\sigma)}a(\sigma)d\sigma=R_{0}e^{-\lambda^{*}\beta}=1, (101)

hence we have λ=(logR0)/β\lambda^{*}=(\log R_{0})/\beta. From (100) and assuming that Q()=0Q(\infty)=0, it follows that

ϕ=log(λNγΩω(x)(1eλβ)𝑑x)=log(logR0R01).\phi^{*}=-\log\left(\frac{\lambda^{*}}{\frac{N}{\gamma}\int_{\Omega}\omega(x)(1-e^{-\lambda^{*}\beta})dx}\right)=-\log\left(\frac{\log R_{0}}{R_{0}-1}\right). (102)

In the homogeneous model with constant parameters, if the transmission control is so strong that ϕ>ϕ\phi>\phi^{*}, and QQ is negligibly small (that is, the initial size of the infected is so small), the supercritical size of the susceptible population remains after the end of the previous epidemic wave, and a resurgence can occur.

8 Final remarks

In the classical theory of the Kermack MacKendrick model, the basic epidemiological concepts as the basic reproduction number, the effective reproduction number, the herd immunity threshold and the final size can be mathematically formulated by using the renewal integral equation. The threshold principle tells us that an outbreak occurs in a totally susceptible population if and only if R0>1R_{0}>1 and the lower bound of the final size is given as the unique root of the final size equation. In this paper, we have extended these basic epidemiological concepts and results to allow the theory to account for individual static heterogeneity. Our main results provide a mathematical basis for the recent COVID-19-inspired arguments and calculations based on the heterogeneous Kermack-McKendrick model.

Among the COVID-19-related studies, the most notable ones were that individual variation in susceptibility lowers the herd immunity threshold, and it was shown that the infinite-dimensional compartment model with individual heterogeneity described by the gamma distribution can be transformed into a finite-dimensional ordinary differential equation system with power-law interaction term.

To justify these arguments, we have to deal with unbounded susceptibility and infectivity with non-compact domain. Our assumption 1 is sufficient to prove the well-posedness of the heterogeneous model with unbounded parameters. On the other hand, our assumption 2 still excludes the unbounded infectivity, so it is a future challenge to construct a theory to deal with more general parameters. Moreover, it would also be necessary to consider the dynamic heterogeneity to understand the real epidemic waves.

\bmhead

Acknowledgments This work was supported by JSPS KAKENHI Grant Number 22K03433 and Japan Agency for Medical Research and Development (JP23fk0108685).

Declarations

Not applicable

References

  • (1) N. T. J. Bailey (1975), The Mathematical Theory of Infectious Diseases and its Applications, 2nd Edition, Charles Griffin, London.
  • (2) M. C. J. Bootsma, K. M. D. Chan, O. Diekmann and H. Inaba (2023a), Separable mixing: the general formulation and a particuar example focusing on mask efficiency, Math. Biosci. Eng. 20(10), 17661-17671.
  • (3) M. C. J. Bootsma, K. M. D. Chan, O. Diekmann and H. Inaba (2023b), The effect of host population heterogeneity on epidemic outbreaks, Math. Appl. Sci. Eng., submitted
  • (4) D. Breda, O. Diekmann, W. F. de Graaf, A. Pugliese and R. Vermiglio (2012): On the formulation of epidemic models (an appraisal of Kermack and McKendrick), J. Biol. Dyn., 6, sup2: 103117.
  • (5) T. Britton, F. Ball and P. Trapman (2020), A mathematical model reveals the influence of population heterogeneity on herd immunity to SARS-CoV2, Science 369, 846-849.
  • (6) O. Diekmann (1977), Limiting behaviour in an epidemic model, Nonl. Anal. Theor. Meth. Appls. 1(5), 459-470.
  • (7) O. Diekmann (1978), Threshold and travelling waves for the geographical spread of infection, J. Math. Biol. 6, 109-130.
  • (8) O. Diekmann, J. A. P. Heesterbeak and J. A. J. Metz (1990), On the definition and the computation of the basic reproduction ratio R0R_{0} in models for infectious diseases in heterogeneous populations, J. Math. Biol. 28, 365-382.
  • (9) O. Diekmann and J. A. P. Heesterbeak (2000), Mathematical Epidemiology of Infectious Diseases: Model Building, Analysis and Interpretation, John Wiley and Sons, LTD, Chichester.
  • (10) O. Diekmann, J. A. P. Heesterbeek, T. Britton (2013), Mathematical Tools for Understanding Infectious Disease Dynamics, Princeton University Press.
  • (11) O. Diekmann and H. Inaba (2023), A systematic procedure for incorporating separable static heterogeneity into compartmental epidemic models, J. Math. Biol. 86, Article Number 29.
  • (12) M. G. M. Gomes, M. U. Ferreira, R. M. Corder, J. G. King, C. Souto-Maior, C. Penha-Gonçalves, G. Gonçalves, M. Chikina, W. Pegden and R. Aguas (2022). Individual variation in susceptibility or exposure to SARS-CoV-2 lowers the herd immunity threshold. J. Theor. Biol. 540, 111063.
  • (13) K. P. Hadeler (2017), Topics in Mathematical Biology, Springer.
  • (14) J. A. P. Heesterbeek and M. G. Roberts (2006), The type-reproduction number TT in models for infectious disease control. Math. Biosci. 206, 3-10.
  • (15) H. Inaba (1990), Threshold and stability results for an age-structured epidemic model, J. Math. Biol. 28(4): 411-434.
  • (16) H. Inaba (2001), Kermack and McKendrick revisited: The variable susceptibility model for infectious diseases, Japan J. Indust. Appl. Math. 18(2): 273-292.
  • (17) H. Inaba (2013), On the definition and the computation of the type-reproduction number TT for structured populations in heterogeneous environments. J. Math. Biol. 66, 1065-1097.
  • (18) H. Inaba (2014), On a pandemic threshold theorem of the early Kermack–McKendrick model with individual heterogeneity, Math. Popul. Stud. 21(2), 95-111.
  • (19) H. Inaba (2016), Endemic threshold analysis for the Kermack–McKendrick reinfection model, Josai Mathematical Monographs vol. 9, 105-133.
  • (20) H. Inaba (2017), Age-Structured Population Dynamics in Demography and Epidemiology, Springer.
  • (21) G. Katriel (2012), The size of epidemics in populations with heterogeneous susceptibility. J. Math. Biol. 65, 237–262.
  • (22) W. O. Kermack and A. G. McKendrick (1927), A contribution to the mathematical theory of epidemics. Proc. Roy. Soc. London. Series A, Containing papers of a mathematical and physical character, 115(772), 700-721.
  • (23) D. G. Kendall (1957), Discussion of “Measles periodicity and community size" by M. S. Bartlett, J. Roy. Statist. Soc. A120, 48-70.
  • (24) M. A. Krasnoselskii (1964). Positive Solutions of Operator Equations. Groningen, Noordhoff.
  • (25) S. Lang (1993), Real and Functional Analysis, Third Edition, Springer.
  • (26) A. Montalbán, R. M. Corder and M. G. M. Gomes (2022). Herd immunity under individual variation and reinfection, J. Math. Biol. 85, Article number:2.
  • (27) J. Neipel, J. Bauermann, S. Bo, T. Harmon, and F. Jülicher (2020), Power-law population heterogeneity governs epidemic waves, PLoS ONE 15(10), e0239678, doi: 10.1371/journal.pone.0239678
  • (28) A. S. Novozhilov (2008). On the spread of epidemics in a closed heterogeneous population. Math. Biosci. 215, 177–185.
  • (29) A. S. Novozhilov (2012). Epidemiological models with parametric heterogeneity: Deterministic theory for closed populations. Math. Model. Nat. Phenom. Vol. 7, No. 3, 147-167.
  • (30) L. Rass and J. Radcliffe (2003), Spatial Deterministic Epidemics, Mathematical Surveys and Monographs Vol. 102, Ame. Math. Soc.
  • (31) H. L. Smith and H. R. Thieme (2011), Dynamical Systems and Population Persistence, Amer. Math. Soc.
  • (32) H. R. Thieme (1977a). A model for the spatial spread of an epidemic. J. Math. Biol. 4, 337-351.
  • (33) H. R. Thieme (1977b). The asymptotic behaviour of solutions of nonlinear integral equations. Math. Zeit. 157, 141-154.
  • (34) H. R. Thieme (1980). On the boundedness and the asymptotic behaviour of the non-negative solutions of Volterra-Hammerstein integral equations, Manuscripta Mathematica 31, 379-412.
  • (35) H. R. Thieme (2009). Distributed susceptibility: a challenge to persistence theory in infectious disease models, Disc. Conti. Dyn. Sys. Ser B, Vol 12, No. 4, 865-882..
  • (36) A. V. Tkachenko, S. Maslov, A. Elbanna, G. N. Wong, Z. J. Weiner and N. Goldenfeld (2021a), Time-dependent heterogeneity leads to transient suppression of the COVID-19 epidemic, not herd immunity, PNAS Apr 2021, 118 (17) e2015972118; DOI: 10.1073/pnas.2015972118
  • (37) A. V. Tkachenko, S. Maslov, T. Wang, A. Elbanna, G. N. Wong and N. Goldenfeld (2021b). Stochastic social behavior coupled to COVID-19 dynamics leads to waves, plateaus, and an endemic state, eLife, https://doi.org/10.7554/eLife.68341
  • (38) G. F. Webb (1980). An age-dependent epidemic model with spatial diffusion. Arch. Rat. Mech. Anal. 75(1), 91-102.
  • (39) G. F. Webb (1981). A reaction-diffusion model for a deterministic diffusive epidemic. J. Math. Anal. Appl. 84(1), 150-161.