
Citation: Kento Okuwa, Hisashi Inaba, Toshikazu Kuniya. Mathematical analysis for an age-structured SIRS epidemic model[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 6071-6102. doi: 10.3934/mbe.2019304
[1] | Kento Okuwa, Hisashi Inaba, Toshikazu Kuniya . An age-structured epidemic model with boosting and waning of immune status. Mathematical Biosciences and Engineering, 2021, 18(5): 5707-5736. doi: 10.3934/mbe.2021289 |
[2] | Toshikazu Kuniya, Hisashi Inaba . Hopf bifurcation in a chronological age-structured SIR epidemic model with age-dependent infectivity. Mathematical Biosciences and Engineering, 2023, 20(7): 13036-13060. doi: 10.3934/mbe.2023581 |
[3] | Mostafa Adimy, Abdennasser Chekroun, Claudia Pio Ferreira . Global dynamics of a differential-difference system: a case of Kermack-McKendrick SIR model with age-structured protection phase. Mathematical Biosciences and Engineering, 2020, 17(2): 1329-1354. doi: 10.3934/mbe.2020067 |
[4] | Andrea Franceschetti, Andrea Pugliese, Dimitri Breda . Multiple endemic states in age-structured $SIR$ epidemic models. Mathematical Biosciences and Engineering, 2012, 9(3): 577-599. doi: 10.3934/mbe.2012.9.577 |
[5] | Abdennasser Chekroun, Mohammed Nor Frioui, Toshikazu Kuniya, Tarik Mohammed Touaoula . Global stability of an age-structured epidemic model with general Lyapunov functional. Mathematical Biosciences and Engineering, 2019, 16(3): 1525-1553. doi: 10.3934/mbe.2019073 |
[6] | Zhong-Kai Guo, Hai-Feng Huo, Hong Xiang . Global dynamics of an age-structured malaria model with prevention. Mathematical Biosciences and Engineering, 2019, 16(3): 1625-1653. doi: 10.3934/mbe.2019078 |
[7] | Abdelheq Mezouaghi, Salih Djillali, Anwar Zeb, Kottakkaran Sooppy Nisar . Global proprieties of a delayed epidemic model with partial susceptible protection. Mathematical Biosciences and Engineering, 2022, 19(1): 209-224. doi: 10.3934/mbe.2022011 |
[8] | Jinliang Wang, Xiu Dong . Analysis of an HIV infection model incorporating latency age and infection age. Mathematical Biosciences and Engineering, 2018, 15(3): 569-594. doi: 10.3934/mbe.2018026 |
[9] | Junyuan Yang, Rui Xu, Xiaofeng Luo . Dynamical analysis of an age-structured multi-group SIVS epidemic model. Mathematical Biosciences and Engineering, 2019, 16(2): 636-666. doi: 10.3934/mbe.2019031 |
[10] | Toshikazu Kuniya, Yoshiaki Muroya, Yoichi Enatsu . Threshold dynamics of an SIR epidemic model with hybrid of multigroup and patch structures. Mathematical Biosciences and Engineering, 2014, 11(6): 1375-1393. doi: 10.3934/mbe.2014.11.1375 |
In a seminal series of papers published during the 1920s and the 1930s, Kermack and McKendrick proposed infection–age structured epidemic models that take into account demography of the host population, the waning immunity and reinfection of recovered individuals ([1,2]). In their models, the total population is decomposed into three compartments, the susceptibles, the infectious and the recovered populations (SIR model), and it is assumed that reinfection occurs for the recovered population depending on the time since recovery (recovery-age). Recently the concept of reinfection is recognized as more and more important in understanding emerging and reemerging infectious diseases, since it makes the control of infectious diseases difficult, and a waning immunity is widely observed among many infectious diseases. In fact, the recovered individuals or vaccinated individuals could be reinfected as time passes owing to the natural decay of host immunity, or a genetic change in the circulating virus. The Kermack–McKendrick infection-age dependent reinfection model has been reinvestigated by several authors ([3,4,5,6,7]), and it was shown that a backward bifurcation of endemic steady states is possible to occur.
On the other hand, we can formulate another type of reinfection model such that recovered individuals can return to the full susceptible class due to the loss of immunity, which model is simply called a SIRS model. So far SIRS epidemic models have been studied by several authors. Hethcote [8] first examined an ODE model for SIRS epidemic and shown that there exists a unique globally stable endemic steady state if and only if R0>1, where R0 denotes the basic reproduction number (although Hethcote did not use this notation at 1976). Aron ([9,10,11]) developed an SIRS model with recovery-age for malaria epidemic by which we can take into account the boosting of immunity. If we consider the recovery-age independent case, Aron's SIRS model has a globally stable endemic steady state [12]. Although in many cases, the host population is assumed to be in a demographic steady state, Busenberg and Hadeler [13] and Busenberg and van den Driessche [14] considered a homogeneous SIRS epidemic model in a changing host population. Recently Nakata et al. [15] developed an infection-age structured SIRS epidemic model and studied global stability of the endemic equilibrium by using the Lyapunov method. It is noted that the infection-age structured SIRS model can be formulated by a scalar nonlinear renewal integral equation ([16,6]).
Different from the above-mentioned existing SIRS models, we here investigate a SIRS epidemic model with chronological age structure, which kind of model was proposed by Tudor [17] in 1985. As the age-structured SIRS model is an extension of the well-known age-structured SIR epidemic model [18], we can make use of many ideas developed for the SIR model, but we have to develop new technique to deal with problems arising from the reversion of susceptibility for recovered individuals. For application purpose, it is important to clear the implication of reinfection on R0 and the critical coverage of immunization, because the reinfection phenomena would make disease control more difficult and complex. In fact, quantitative threshold results of the SIRS epidemic are similar to those of the SIR epidemic, but its controllability is very much different from the SIR epidemic. An important effect of vaccination policy is reduction of the effective size of the susceptible population, however in the reinfection model, there is a possibility that a disease can invade a fully vaccinated population, and we are naturally led to the idea of the reinfection threshold ([6,7]). In other words, for the SIRS reinfection model, mass-vaccination policy is not necessarily almighty.
For simplicity, in this paper we only treat the case that the host population is assumed to be in a demographic steady state, so the force of infection is given by the pseudo mass-action type [19]. The reader may refer to [20] and [21] for more complex model formulation to take into account subclinical infection. Moreover, it is noted that our explicit bifurcation and persistence calculations are based on the separable mixing assumption for the transmission coefficient (Assumption 3.8). The separable mixing assumption implies that there is no correlation between the age of the infecteds and the age of susceptibles when their contacts occur. Further, we neglect vertical transmission because our analysis is sufficiently complex even without vertical transmission and also its essential features of reversion phenomena could be well understood by considering horizontal transmission.
In the following, we first give a standard proof for the well-posedness of the normalized age-structured SIRS model. Next we examine existence of endemic steady states by fixed point arguments and bifurcation method, where we introduce the next generation operator and the basic reproduction number R0 to formulate endemic threshold results. Thirdly we investigate stability of steady states by bifurcation calculation and comparative method, and we show existence of a compact attractor and discuss the global behavior based on the population persistence theory [22]. Finally we give some numerical examples and discuss the effect of mass-vaccination.
We here consider an infectious disease in a closed age-structured host population and assume that the disease confers temporary immunity on the recovered individuals. For simplicity, we neglect the disease-induced death rate, the effect of infection on fertility, the latent period and the infection-age dependency of the parameters. Let S(t,a) be the age density of susceptible individuals at time t, I(t,a) be the age density of infected individuals at time t and R(t,a) be the age density of recovered individuals at time t. The basic age-structured SIRS model is then formulated by the following system of McKendrick equations:
(∂∂t+∂∂a)S(t,a)=−μ(a)S(t,a)−λ(t,a)S(t,a)+δ(a)R(t,a),(∂∂t+∂∂a)I(t,a)=λ(t,a)S(t,a)−(μ(a)+γ(a))I(t,a),(∂∂t+∂∂a)R(t,a)=γ(a)I(t,a)−(μ(a)+δ(a))R(t,a),S(t,0)=∫ω0m(a)N(t,a)da,I(t,0)=R(t,0)=0, | (2.1) |
where 0<ω<∞ is the maximum attainable chronological age of the host population, m(a) and μ(a) are the age-specific birth rate and death rate respectively, γ(a) and δ(a) are the age-dependent recovery rate and the loss-of-immunity rate respectively. Let β(a,σ) be the transmission coefficient between susceptibles with age a and infecteds with age σ. The age-density function of the host population is N(t,a)=S(t,a)+I(t,a)+R(t,a) and the force of infection is given by
λ(t,a)=∫ω0β(a,σ)I(t,σ)dσ. | (2.2) |
Then the host population satisfies the stable population model [7]:
∂N(t,a)∂t+∂N(t,a)∂a=−μ(a)N(t,a),N(t,0)=∫ω0m(a)N(t,a)da. | (2.3) |
Define the survival function by
ℓ(a):=exp(−∫a0μ(σ)dσ). | (2.4) |
We assume that the net reproduction rate (demographic basic reproduction number) of the host population is unity:
∫ω0m(a)ℓ(a)da=1. | (2.5) |
Without loss of generality, we can assume that the host population has already reached the demographic steady state:
S(t,a)+I(t,a)+R(t,a)=N(a):=Bℓ(a), | (2.6) |
where N(a) is the demographic stationary host population and B>0 denotes its number of birth per unit time.
From technical reasons, we assume that m,γ,δ∈L∞+(0,ω), β∈L∞+((0,ω)×(0,ω)) and μ∈L1loc,+(0,ω) with ∫ω0μ(σ)dσ=∞, which implies ℓ(ω)=0. Let β∞, γ∞ and δ∞ be the essential supremum of β, γ and δ respectively.
Let
s(t,a)=S(t,a)N(a),i(t,a)=I(t,a)N(a),r(t,a)=R(t,a)N(a). |
Then the basic system (2.1) can be written as the normalized system:
(∂∂t+∂∂a)s(t,a)=−λ[a|i(t,⋅)]s(t,a)+δ(a)r(t,a),(∂∂t+∂∂a)i(t,a)=λ[a|i(t,⋅)]s(t,a)−γ(a)i(t,a),(∂∂t+∂∂a)r(t,a)=γ(a)i(t,a)−δ(a)r(t,a),s(t,0)=1,i(t,0)=r(t,0)=0, | (2.7) |
where λ[a∣ψ],ψ∈L1(0,ω) is the mapping on E:=L1(0,ω) defined by
λ[a∣ψ]=∫ω0β(a,σ)N(σ)ψ(σ)dσ, | (2.8) |
and
s(t,a)+i(t,a)+r(t,a)=1,∀(t,a)∈R+×[0,ω]. |
In what follows, we mainly investigate the normalized SIRS epidemic model (2.7).
Since s=1−i−r, the state space for (i,r)-system is a convex closed set in E2:=L1(0,ω)×L1(0,ω) given as
C={(i,r)∈L1+(0,ω)×L1+(0,ω)∣0≤i+r≤1}. | (2.9) |
Let ϕ=(ϕ1(a),ϕ2(a))T∈E2 and let us introduce operators A and F on E2 as
(Aϕ)(a)=(−dda00−dda)(ϕ1(a)ϕ2(a)), | (2.10) |
D(A)={ϕ∈E2∣ϕ∈AC[0,ω],ϕ(0)=0}, | (2.11) |
F(ϕ)(a)=(λ[a∣ϕ1](1−ϕ1(a)−ϕ2(a))−γ(a)ϕ1(a)γ(a)ϕ1(a)−δ(a)ϕ2(a)), | (2.12) |
where AC[0,ω] is the set of real-valued absolutely continuous functions on [0,ω]. Then (i,r)-system can be formulated as a semilinear Cauchy problem on E2:
ddtu(t)=Au(t)+F(u(t)),u(0)=u0. | (2.13) |
The linear operator A, which is called population operator, generates the C0-semigroup {T(t)}t≥0 on E=L2(0,ω):
(T(t)ϕ)(a)={(ϕ1(a−t)ϕ2(a−t))for a>t0for a<t,ϕ=(ϕ1ϕ2)∈D(A). | (2.14) |
Then the state space C is positively invariant with respect to the semiflow defined by {etA}t≥0, that is,
etA(C)⊂C for all t≥0. | (2.15) |
Lemma 2.1. The operator F is Lipschitz continuous. Moreover, there exists a constant α∈(0,1) such that
(I+αF)(C)⊂C. | (2.16) |
Proof. Lipschitz continuity is obvious. Observe that
u(a)+αF(u)(a)=(u1(a)+αλ[a∣u1](1−u1(a)−u2(a))−αγ(a)u1(a)(1−αδ(a))u2(a)+αγ(a)u1(a)). |
Thus it is easy to see that u+αF(u)≥0 if u∈C, 1−αδ∞>0 and 1−αγ∞>0. Furthermore, it follows that if u∈C, then
(u1(a)+αλ[a∣u1](1−u1(a)−u2(a))−αγ(a)u1(a))+((1−αδ(a))u2(a)+αγ(a)u1(a)) |
≤(u1(a)+u2(a))(1−αλ[a∣u1])+αλ[a∣u1]≤1. |
Hence we have proved that (I+αF)(C)⊂C.
By using the method in [23], we obtain the following proposition.
Proposition 2.2. Let u0∈C. Then the Cauchy problem (2.13) has a unique mild solution in C. The mild solution u(t) is given by the following variation of constants formula:
u(t)=e−1αtetAu0+1α∫t0e−1α(t−s)e(t−s)A[u(s)+αF(u(s))]ds. | (2.17) |
Proof. First we choose α such that (2.16) holds. Define the series {un}n≥1 iteratively as
u0(t)=u0,un+1(t)=e−1αtetAu0+1α∫t0e−1α(t−s)e(t−s)A[un(s)+αF(un(s))]ds. |
Since (2.15) and (2.16) hold, if un∈C, then un+1∈C. In fact, un+1 is a convex linear combination of etAu0 and [un(s)+αF(u(s))] with e−1αt+1α∫t0e−1α(t−s)ds = 1. Because of the Lipschitz continuity of F, un(t) converges to the mild solution u(t)∈C uniformly as n→∞.
We now consider the existence of endemic steady states. First note that the endemic steady state (s∗(a),i∗(a),r∗(a))T satisfies the following ODE system:
ddas∗(a)=−λ∗(a)s∗(a)+δ(a)r∗(a),ddai∗(a)=λ∗(a)s∗(a)−γ(a)i∗(a),ddar∗(a)=γ(a)i∗(a)−δ(a)r∗(a),s∗(0)=1,i∗(0)=r∗(0)=0, | (3.1) |
where
λ∗(a):=∫ω0β(a,σ)N(σ)i∗(σ)dσ, | (3.2) |
Δ(a):=exp(−∫a0δ(σ)dσ),Γ(a):=exp(−∫a0γ(σ)dσ). |
Formally solving the above ODEs, we have the following expressions:
s∗(a)=e−∫a0λ∗(σ)dσ+∫a0e−∫aσλ∗(σ)dσδ(σ)r∗(σ)dσ, | (3.3) |
i∗(a)=∫a0Γ(a)Γ(σ)λ∗(σ)s∗(σ)dσ, | (3.4) |
r∗(a)=∫a0Δ(a)Δ(σ)γ∗(σ)i∗(σ)dσ. | (3.5) |
Inserting the above expression into (3.3), we obtain
s∗(a)=e−∫a0λ∗(σ)dσ+∫a0e−∫aσλ∗(ξ)dξδ(σ)∫σ0Δ(σ)Δ(η)γ(η)∫η0Γ(η)Γ(ζ)λ∗(ζ)s∗(ζ)dζdηdσ. | (3.6) |
Let b∗(a):=λ∗(a)s∗(a) be the density of newly infecteds at steady state and define a nonlinear operator f given by
f[ϕ](a,σ):=ϕ(a)e−∫aσϕ(ξ)dξ,ϕ∈E. | (3.7) |
Moreover we define
Π(σ,ζ):=∫σζf[δ](σ,η)f[γ](η,ζ)dη, | (3.8) |
which denotes the transition probability that individuals recovered at age ζ become susceptible again at age σ. Then for all ζ∈[0,ω] it holds that
∫ωζΠ(σ,ζ)dσ≤(1−e−‖δ‖E)(1−e−‖γ‖E)<1. | (3.9) |
In fact, we can observe that
∫ωζΠ(σ,ζ)dσ=∫ωζdη∫ωηf[δ](σ,η)dσf[γ](η,ζ)=∫ωζdη(1−e−∫ωηδ(ξ)dξ)f[γ](η,ζ)≤(1−e−‖δ‖E)∫ωζf[γ](η,ζ)dη, |
which shows (3.9).
From (3.6), we have
b∗(a)=f[λ∗](a,0)+(T[λ∗]b∗)(a), | (3.10) |
where T[λ∗] is a linear operator in E=L1(0,ω) defined as:
(T[λ∗]ϕ)(a):=∫a0∫aζf[λ∗](a,σ)Π(σ,ζ)dσϕ(ζ)dζ. | (3.11) |
Lemma 3.1. There exists a number k∈(0,1) such that ‖T[λ∗]‖≤k uniformly for λ∗∈E+.
Proof. For a given λ∗∈E+, it follows from (3.9) that
∫ω0(T[λ∗]ϕ)(a)da=∫ω0da∫a0∫aζλ∗(a)e−∫aσλ∗(x)dxΠ(σ,ζ)dσϕ(ζ)dζ=∫ω0dσ∫ωσλ∗(a)e−∫aσλ∗(x)dxda∫σ0Π(σ,ζ)ϕ(ζ)dζ≤(1−e−‖λ∗‖E)∫ω0dσ∫σ0Π(σ,ζ)ϕ(ζ)dζ≤(1−e−‖λ∗‖E)(1−e−‖δ‖E)(1−e−‖γ‖E)‖ϕ‖E, |
which shows that ‖T[λ∗]‖≤k<1 with k:=(1−e−‖δ‖E)(1−e−‖γ‖E). Thus we have our conclusion.
If λ∗∈E+=L1+(0,ω) is given, (3.10) is a Volterra integral equation with respect to b∗. As the Volterra operator has the spectral radius zero, (3.10) is solved as following:
b∗=(I−T[λ∗])−1f[λ∗](⋅,0). | (3.12) |
Therefore, we obtain a fixed point equation for the force of infection λ∗:
λ∗(a)=(Ψλ∗)(a):=∫ω0β(a,σ)N(σ)i∗(σ)dσ=∫ω0β(a,σ)N(σ)∫σ0Γ(σ)Γ(η)b∗(η)dηdσ=∫ω0∫ωηβ(a,σ)N(σ)Γ(σ)Γ(η)dσ((I−T[λ∗])−1f[λ∗](⋅,0))(η)dη, | (3.13) |
where Ψ is a nonlinear operator from E+ into itself defined by the right hand side of (3.13).
Then the Fréchet derivative of Ψ at zero is given by
(Ψ′[0]ϕ)(a)=∫ω0∫ωηβ(a,σ)N(σ)Γ(σ)Γ(η)dσϕ(η)dη. | (3.14) |
For simplicity, we call K=Ψ′[0] the next generation operator (NGO) and its spectral radius R0:=r(K) the basic reproduction number for the normalized system, although K is a similar operator of the next generation operator for the original system (2.1) (Chapter 9, [7]). If K and R0 are defined for the original system (2.1), for given age distribution ϕ of primary cases, Kϕ represents the age distribution of secondary cases, and the number R0 means the expected number of secondary cases produced by an infected individual during its entire period of infectiousness in a completely susceptible population. The reader may refer to [24,25,26,7] for the original implications of R0. See also [27] for a practical approach to the computation of R0.
In order to solve the fixed point problem λ∗=Ψλ∗ in E:=L1(0,ω), we use a corollary of the well-known theorem by Krasnoselskii ([28,7]):
Theorem 3.2. Suppose that E is a real Banach space and E+ is a positive cone of E. Let Ψ is a positive operator on E+ which has a strong Fréchet derivative at the origin K=Ψ′[0], satisfies Ψ(0)=0 and Ψ(E+) is bounded. Moreover, K has a positive eigenvector v0∈E+ associated with eigenvalue λ0>1, but has no eigenvector in E+ with unity. Then, Ψ has at least one nonzero fixed point in Ψ(E+).
According to [18], we adopt the following technical assumptions for the transmission coefficient β(a,σ), which is a natural assumption to make the next generation operator becomes nonsupporting and compact.
Assumption 3.3. 1. There exist numbers δ0∈(0,ω) and β_>0 such that
β(a,η)≥β_for almost all (a,η)∈(0,ω)×(ω−δ0,ω). | (3.15) |
2. β∈L∞+((0,ω)×(0,ω)) is extended into L∞+(R2) by β(a,σ)=0 for (a,σ)∉(0,ω)×(0,ω) and satisfies
limh→0∫ω0|β(a+h,η)−β(a,η)|da=0uniformly for η∈R. | (3.16) |
Here we summarize basic definitions from positive operator theory [7]. Let E be a real or complex Banach space and let E∗ be its dual space. Then, E∗ is a space of all linear functionals on E. In the following, we write the value of f∈E∗ at ψ∈E as ⟨f,ψ⟩. A closed subset C⊂E is called the cone (or positive cone) if the following conditions hold: (1) C+C⊂C, (2) λ≥0⇒λC⊂C, (3) C∩(−C)={0} and (4) C≠{0}. With respect to the cone C, we write x≤y if y−x∈C and x<y if y−x∈C∖{0}. If the set {ψ−ϕ∣ψ,ϕ∈C} is dense in E, the cone C is said to be total. If E=C−C, C is called a reproducing cone. Let B(E) be a set of bounded linear operators from E into itself. Let r(T) be the spectral radius of T∈B(E) and let Pσ(T) be the point spectrum of T. The dual cone C∗ is a subset of E∗ composed of all positive linear functionals. f∈C∗ is called a positive linear functional if ⟨f,ψ⟩≥0 for all ψ∈C. ψ∈C is called a quasi-interior point or nonsupporting point if ⟨f,ψ⟩>0 for all f∈C∗∖{0}. A positive linear functional f∈C∗ is called strictly positive if ⟨f,ψ⟩>0 for all ψ∈C+. A nonzero operator T∈B(E) is called positive if T(C)⊂C. If (T−S)(C)⊂C for T,S∈B(E), we write S≤T. A positive operator T∈B(E) is called semi-nonsupporting if, for any ψ∈C+ and f∈C∗∖{0}, there exists an integer p=p(ψ,f) such that ⟨f,Tpψ⟩>0. A positive operator T∈B(E) is called nonsupporting if, for any ψ∈C+ and f∈C∗∖{0}, there exists an integer p=p(ψ,f) such that ⟨f,Tnψ⟩>0 for all n≥p. A positive operator T∈B(E) is called strictly nonsupporting if, for any ψ∈C+, there exists a positive integer p=p(ψ) such that Tnψ is a quasi-interior point of C for all n≥p.
Lemma 3.3. The next generation operator K is nonsupporting and compact.
Proof. Define the positive linear functional f0∈E∗+ by
⟨f0,ψ⟩:=∫ω0∫ωηβ0(σ)N(σ)Γ(σ)Γ(η)dσψ(η)dη, |
where
β0(σ)={β_for σ∈(ω−δ0,ω)0otherwise. | (3.17) |
Then Kψ≥⟨f0,ψ⟩e for all ψ∈E+, where e=1∈E+, which implies
Kn+1ψ≥⟨f0,ψ⟩⟨f0,e⟩ne,∀n∈N. |
Thus for arbitrary F∈E∗+∖{0}, ψ∈E+∖{0} and n≥1,
⟨F,Knψ⟩≥⟨f0,ψ⟩⟨f0,e⟩n−1⟨F,e⟩>0. |
This shows K is nonsupporting. Next we show the compactness of K. Let C be an arbitrary bounded subset in L1+(0,ω), and take M>0 such that supϕ∈C‖ϕ‖E≤M. For all ϕ∈C, using Assumption 1.1,
limh→0∫R|Kϕ(a+h)−Kϕ(a)|da≤limh→0∫R∫ω0∫ωη|β(a+h,σ)−β(a,σ)|N(σ)Γ(σ)Γ(η)dσϕ(η)dηda≤limh→0∫R∫ω0ϕ(η)dη∫ω0|β(a+h,σ)−β(a,σ)|Bdσda≤BMlimh→0∫R∫ω0|β(a+h,σ)−β(a,σ)|dσda=0. |
By the Fréchet-Kolmogorov criterion for the compactness of sets in Lp(R) ([29,22]), Ψ(C) is relatively compact. This shows that K is compact.
Lemma 3.5. Let E+=L1+(0,ω) and ΩM:={ϕ∈E+:‖ϕ‖E≤M}. There exists a number M>0 such that Ψ(E+)⊂ΩM.
Proof. Define the nonlinear operator G:λ∗→b∗ in E+ by Gϕ=(I−T[ϕ])−1f[ϕ](⋅,0). From Lemma 3.1, it follows that
‖ϕ‖E≤‖(I−T[ϕ])−1‖‖f[ϕ](⋅,0)‖E≤11−k. |
Let
M:=11−k∫ω0da∫ω0β(a,σ)N(σ)dσ. |
Then it is easy to see that ‖Ψλ∗‖E≤M, from which we have Ψ(E+)⊂ΩM.
From the well-known Krein-Rutman's Theorem, we know that r(K) is a positive eigenvalue if r(K)>0, and it is a pole of the resolvent because K is compact. Then we can apply Sawashima's results for nonsupporting operator to obtain the following properties (see [30,31]):
Proposition 3.6. Suppose that the cone E+ is total, K is compact, nonsupporting with respect to E+ and r(K)>0. Then the following holds:
1. r(K)∈Pσ(K)∖{0} and r(K) is a simple pole of the resolvent (λI−K)−1.
2. The eigenspace corresponding to r(K) is one-dimensional and its eigenvector v0∈E+ is a quasi-interior point. Any eigenvector in E+ is proportional to v0.
3. The adjoint eigenspace corresponding to r(K) is one-dimensional and its eigenfunctional f∈E∗∖{0} is strictly positive.
As is seen above, the idea of being nonsupporting for positive operator is an infinite-dimensional extension of the primitivity of nonnegative matrices in the finite-dimensional case.
Using the above facts, we can show the main theorem in this section:
Theorem 3.7. If R0>1, there exists at least one endemic steady state, while there is no endemic steady state if R0≤1.
Proof. If R0>1, thanks to the Lemma 3.4 and above Proposition 3.6, there is no eigenvector of K which is corresponding to unity in E+. Then by Lemma 3.5 and Theorem 3.2, we can show the first half of statement. Next suppose that R0≤1 and there exists an endemic steady state and the force of infection at the endemic steady state is given by λ∗>0. From (3.10), it follows that
(I−T[λ∗])−1f[λ∗](⋅,0)=b∗≤λ∗s∗<λ∗. | (3.18) |
Then we know that λ∗=Ψλ∗<Ψ′[0]λ∗, which implies that R0=r(Ψ′[0])>1. This contradicts our assumption. Then there is no endemic steady state if R0≤1.
Next we show that an endemic steady state bifurcates forwardly at R0=1. For this purpose, we adopt the following assumption called separable mixing assumption which means that there is no correlation between the age of the infected individuals and that of the susceptible individuals.
Assumption 3.8. There exist β1,β2∈L∞+(0,ω) such that β(a,σ)=β1(a)β2(σ).
Then the next generation operator K is represented as follows:
(Kϕ)(a)=β1(a)∫ω0∫ωηβ2(σ)N(σ)Γ(σ)Γ(η)dσϕ(η)dη. | (3.19) |
So the range of K is one-dimensional and β1 becomes a positive eigenvector as
(Kβ1)(a)=β1(a)∫ω0∫ωηβ2(σ)N(σ)Γ(σ)Γ(η)dσβ1(η)dη, | (3.20) |
which shows that the basic reproduction number is given by
R0=r(K)=∫ω0∫ωηβ2(σ)N(σ)Γ(σ)Γ(η)dσβ1(η)dη. | (3.21) |
Let us introduce a bifurcation parameter ϵ>0 and suppose that β1=ϵβ10, where the standard susceptibility β10 is chosen such as R0=1 if ϵ=1. Then the basic reproduction number is equal to ϵ and it holds that
∫ω0∫ωηβ2(σ)N(σ)Γ(σ)Γ(η)dσβ10(η)dη=1. | (3.22) |
Then the infection force at the endemic steady state is λ∗(a)=cϵβ10(a) for some c>0. Hence the fixed point problem (3.13) is rewritten as
Θ(c,ϵ):=ϵ∫ω0∫ωηβ2(σ)N(σ)Γ(σ)Γ(η)((I−T[ϵcβ10])−1κ[c;ϵ])(η)dσdη−1=0, | (3.23) |
where
κ[c,ϵ](η)=β10(η)e−ϵc∫η0β10(z)dz. | (3.24) |
Then it follows from (3.22) that Θ(0,1)=0. Observe that
(I−T[ϵcβ10])−1κ[c;ϵ]=∞∑n=0Tn[ϵcβ10]κ[c;ϵ], | (3.25) |
where
(T[ϵcβ10]κ[c;ϵ])(η)=∫η0∫ηζf[ϵcβ10](η,σ)Π(σ,ζ)dσκ[c;ϵ](ζ)dζ. | (3.26) |
Then it follows that
∂∂c∞∑n=0(Tn[ϵcβ10]κ[c;ϵ])(η)|(c,ϵ)=(0,1)=−β10(η)∫η0β10(ζ)dζ+∫η0∫ηζβ10(η)Π(σ,ζ)dσβ10(ζ)dζ=β10(η)∫η0β10(ζ)[−1+∫ηζΠ(σ,ζ)dσ]dζ<0, | (3.27) |
from which we can conclude that
∂Θ∂c(0,1)<0. | (3.28) |
From the Implicit Function Theorem, Θ(c,ϵ)=0 can be solved as c=c(ϵ) with c(1)=0 at the neighborhood of (c,ϵ)=(0,1). Corresponding to a positive root c>0 of Θ(c,ϵ)=0, there exists an endemic steady state. Moreover it follows from (3.22) that
∂Θ∂ϵ(0,1)=1, | (3.29) |
and so
c′(1)=−(∂Θ∂c(0,1))−1∂Θ∂ϵ(0,1)=−(∂Θ∂c(0,1))−1>0, | (3.30) |
which implies that c(ϵ)>0 if ϵ>0 and |ϵ−1| is small enough. Then a positive steady state forwardly bifurcates at ϵ=R0=1. Then we have the following bifurcation result:
Proposition 3.9. For the separable mixing case, an endemic steady state forwardly bifurcates from the disease-free steady state when R0 crosses unity.
Next we consider the stability of steady states of the system (2.7).
The system (2.7) has a unique disease-free steady state: (s0(a),i0(a),r0(a))T=(1,0,0)T. In order to consider the dynamics around the disease-free steady state, we introduce the small perturbation terms:
x(t,a)=s(t,a)−s0(a),y(t,a)=i(t,a)−i0(a),z(t,a)=r(t,a)−r0(a), |
where x(t,0)=y(t,0)=z(t,0)=0. Then the second equation in (2.7) is rewritten as
(∂∂t+∂∂a)y(t,a)=λ[a∣y(t,⋅)](x(t,a)+1)−γ(a)y(t,a). | (4.1) |
Neglecting the second order term of small perturbation, we obtain the linearized equation of (4.1):
(∂∂t+∂∂a)y(t,a)=λ[a|y(t,⋅)]−γ(a)y(t,a),y(t,0)=0. | (4.2) |
This equation describes the dynamics of the initial invasion phase of the infected population. Define two linear operators A0 and F0 on E as the following:
A0=−dda−γ(a),D(A0)={x∈E∣y∈AC[0,ω],y(0)=0}, |
F0y(a)=λ[a∣y]=∫ω0β(a,σ)N(σ)y(σ)dσ,y∈E. |
Then the equation (4.2) is transformed into the following linear Cauchy problem
ddtu(t)=(A0+F0)u(t). | (4.3) |
Note that from Assumption 3.3, F0∈B(E) is a compact operator. Define the linear operator Sζ for given ζ∈ρ(A0) as follows:
Sζv(a):=F0R(ζ,A0)v(a)=∫ω0∫ωηβ(a,σ)N(σ)e−ζ(σ−η)Γ(σ)Γ(η)dσv(η)dη,v∈E, | (4.4) |
where R(ζ,A0)=(ζ−A0)−1. Equation (4.3) has been well studied and the following statement can be proved as Lemma 4.7 of this paper ([18], [32]):
Lemma 4.1. A0+F0 has a compact resolvent and it holds that
σ(A0+F0)=Pσ(A0+F0)=Σ:={ζ∈C∣1∈Pσ(Sζ)}. | (4.5) |
So we are going to investigate the properties of the operator Sζ instead of A0+F0, which determines the location of eigenvalues of the linearized generator A0+F0.
Lemma 4.2. The operator Sζ is compact and nonsupporting for all ζ∈R.
Proof. The operator Sζ is the composition of the compact operator F0 and the bounded operator R(ζ;A0) on E, so Sζ is compact.
Define the strictly positive linear functional fζ∈E∗+ by
⟨fζ,ψ⟩:=∫ω0∫ωηβ0(σ)N(σ)e−(σ−η)ζΓ(σ)Γ(η)dσψ(η)dη, |
then Sζψ≥⟨fζ,ψ⟩e for all ψ∈E+ and ζ∈R. Therefore, we obtain
Sn+1ζψ≥⟨fζ,ψ⟩⟨fζ,e⟩ne∀n∈N. |
Thus for arbitrary F∈E∗+∖{0}, ψ∈E+∖{0} and n≥1,
⟨F,Snζψ⟩≥⟨fζ,ψ⟩⟨fζ,e⟩n−1⟨F,e⟩>0, |
which implies Sζ is nonsupporting.
Lemma 4.3. There is a unique ζ0∈R such that r(Sζ0)=1 and the sign relation holds:
sign(ζ0)=sign(R0−1) | (4.6) |
Moreover, ζ0 is the dominant characteristic root, that is, ℜζ<ζ0 for any ζ∈Σ∖{ζ0}.
To prove Lemma 4.3 we need the following theorem on the monotone property of spectral radius.
Theorem 4.4 ([31]). Let E be a Banach lattice. Suppose that S,T∈B(E) are positive operators. Then, the following holds:
1. If S≤T, then r(S)≤r(T).
2. If S,T are semi-nonsupporting and compact, S≤T,S≠T and r(T)≠0, then r(S)<r(T).
Proof of Lemma 4.3. To prove the first half we will show that
limζ→−∞r(Sζ)=∞,limζ→+∞r(Sζ)=0, | (4.7) |
and that the mapping R∋ζ→r(Sζ) is strictly decreasing and continuous. Applying Proposition 3.5, r(Sζ) is an eigenvalue of S∗ζ and its corresponding eigenfunctional Fζ∈E∗+ is strictly positive. Then we have
r(Sζ)⟨Fζ,e⟩=⟨r(Sζ)Fζ,e⟩=⟨S∗ζFζ,e⟩=⟨Fζ,Sζe⟩≥⟨fζ,e⟩⟨Fζ,e⟩. |
Because of the strict positivity of Fζ, we can divide the both sides of the above inequality by ⟨Fζ,e⟩>0 to obtain r(Sζ)≥⟨fζ,e⟩. Since
⟨fζ,e⟩=∫ω0∫ωηβ0(σ)N(σ)e−(σ−η)ζΓ(σ)Γ(η)dσdη≥∫ωω−δ0N(σ)∫σ0β_e−(σ−η)ζΓ(σ)Γ(η)dηdσ=∫ωω−δ0N(σ)∫σ0β_e−(σ−η)ζΓ(σ)Γ(η)dηdσ≥∫ωω−δ0N(σ)∫σ0β_e−(σ−η)(ζ+γ∞)dηdσ=β_∫ωω−δ0N(σ)1−e−σ(ζ+γ∞)ζ+γ∞dσ |
holds, it follows that lim infζ→−∞r(Sζ)=∞ by the Fatou's lemma. Furthermore, we define a strictly positive functional gζ∈E∗+ by
⟨gζ,ψ⟩:=¯β∫ω0∫ωηN(σ)e−(σ−η)ζΓ(σ)Γ(η)dσψ(η)dη. |
Using the same argument as above, we obtain
r(Sζ)≤⟨gζ,e⟩≤β∞∫ω0N(σ)1−e−σζζdσ. |
By the reverse Fatou's lemma, lim supζ→+∞r(Sζ)=0. Consequently, (4.7) holds. Note that if ζ>μ, then we have Sζ≤Sμ and Sζ≠Sμ, so it follows from Theorem 4.4 that r(Sζ)>r(Sμ). As Sζ is a compact operator for any ζ∈R and r(Sζ)>0, it follows from the Krein-Rutman's theorem that its spectral radius r(Sζ) is a positive eigenvalue, so it is a continuous function of ζ. Then r(Sζ)=1 has a unique real root ζ0 such that sign(ζ0)=sign(r(S0)−1). Since r(S0)=r(K)=R0, we have the sign relation (4.6).
Next we show the latter half. If we take ζ∈Σ, there exists an eigenfunction yζ∈E∖{0} such that Sζyζ=yζ. If we use the notation |yζ|(a):=|yζ(a)|, it follows that
|yζ|(a)=|Sζyζ(a)|=|∫ω0∫ωηβ(a,σ)N(σ)e−ζ(σ−η)Γ(σ)Γ(η)dσyζ(η)dη|≤∫ω0∫ωηβ(a,σ)N(σ)|e−ζ(σ−η)|Γ(σ)Γ(η)dσ|yζ|(η)dη=Sℜζ|yζ|(η). |
In short, |yζ|≤|Sζyζ|≤Sℜζ|yζ|. Let Fℜζ be an eigenfunctional of S∗ℜζ corresponding to the eigenvalue r(Sℜζ). By Proposition 3.5, it is strictly positive. Then we have
r(Sℜζ)⟨Fℜζ,|yζ|⟩=⟨Fℜζ,Sℜζ|yζ|⟩≥⟨Fℜζ,|yζ|⟩ |
and dividing the both sides by ⟨Fℜζ,|yζ|⟩>0, we obtain r(Sℜζ)≥1. This implies ℜζ≤ζ0 for all ζ∈Σ by the sign relation. Let us show that ζ=ζ0 if ℜζ=ζ0. In this case, we have |yζ|≤Sℜζ|yζ|=Sζ0|yζ|. In particular |yζ|=Sζ0|yζ| holds. In fact, if we assume |yζ|<Sζ0|yζ| then
⟨Fζ0,|yζ|⟩<⟨Fζ0,Sζ0|yζ|⟩=⟨S∗ζ0Fζ0,|yζ|⟩=r(Sζ0)⟨Fζ0,|yζ|⟩=⟨Fζ0,|yζ|⟩, |
which is a contradiction. Let y0 be the eigenfunction of Sζ0 corresponding to 1=r(Sζ0). Since |yζ| is a positive eigenfunction of the nonsupporting operator Sζ0, it follows from Proposition 3.6 that there exists c>0 such that |yζ|=cy0 and, without loss of generality, we can assume that c=1. Hence the function yζ is represented as yζ(a)=eiv(a)y0(a) for some real valued function v:(0,ω)→R. From the relation |Sζyζ|=Sζ0y0 with ζ=ζ0+ℑζ, we have
|∫ω0∫ωηβ(a,σ)N(σ)e−ζ(σ−η)Γ(σ)Γ(η)dσeiv(η)y0(η)dη|=∫ω0∫ωηβ(a,σ)N(σ)e−ζ0(σ−η)Γ(σ)Γ(η)y0(η)dσdη. | (4.8) |
Applying Lemma 6.12 in [33], it follows that −ℑζ(σ−η)+v(η)=κ for some constant κ. Inserting yζ=eiv(a)y0 into the relation Sζyζ=yζ, we have eiκSζ0y0=eiv(a)y0, so κ=v(a) and ℑζ=0. Then ζ0 is the strictly dominant eigenvalue.
Finally we can conclude the following stability theorem.
Theorem 4.5. If R0<1, the disease-free steady state is locally asymptotically stable whereas unstable if R0>1.
Proof. The C0-semigroup {etA0} is zero for t≥ω, so it is eventually norm continuous. Since F0 is compact, {et(A0+F0)}t≥0 is also eventually norm continuous [34]. Then applying the spectral mapping theorem, etω0(A0+F0)=ets(A0+F0) for all t≥0. Then it holds that
ζ0=maxζ∈Σℜζ=maxζ∈σ(A0+F0)ℜζ=s(A0+F0)=ω0(A0+F0), |
where ω(A) denotes the growth bound of the semigroup etA. Then it follows that
sign(ζ0)=sign(R0−1)=sign(ω(A0+F0)). |
Thus by the principle of linearized stability, the disease-free steady state is locally asymptotically stable if R0<1, while it is unstable if R0>1.
Theorem 4.6. If R0<1, the disease-free steady state is globally asymptotically stable.
Proof. Let U(t) and V(t) be the semiflows induced by the mild solution of (2.13) and the mild solution v(t) of
v(t)=e−1αtetAu0+1α∫t0e−1α(t−s)e(t−s)A[v(s)+αF′[0](v(s))]ds. |
Since F(u)≤F′[0]u for u∈C, by using iterative argument, it is easily seen that U(t)≤V(t) in C. For the asymptotic behavior of the linearized equation, it is shown above that if R0=r(K)<1, then
limt→∞V(t)u0=0,u0∈C, | (4.9) |
which implies the global stability of the disease-free steady state.
Next we consider the local stability of endemic steady states. Throughout this subsection, we again adopt the separable mixing assumption. As is shown in the previous section, the endemic steady state exists if and only if R0>1, so we assume this supercriticality condition to consider the stability of the endemic steady state. Let (s∗(a),i∗(a),r∗(a))T∈E3+=(L1+(0,ω))3 be an endemic steady state. Again let us introduce the small perturbation terms:
x(t,a)=s(t,a)−s∗(a),y(t,a)=i(t,a)−i∗(a),z(t,a)=r(t,a)−r∗(a). |
Then the system (2.7) is rewritten as
(∂∂t+∂∂a)x(t,a)=−λ[a|y(t,⋅)]s∗(a)−λ∗(a)x(t,a)+δ(a)z(t,a),(∂∂t+∂∂a)y(t,a)=λ[a|y(t,⋅)]s∗(a)+λ∗(a)x(t,a)−γ(a)y(t,a),(∂∂t+∂∂a)z(t,a)=γ(a)y(t,a)−δ(a)z(t,a), | (4.10) |
x(t,0)=y(t,0)=z(t,0)=0. |
Since x(t,a)+y(t,a)+z(t,a)=0, (4.10) is reduced to (y,z)-system, so it can be formulated as an abstract linear problem on the Banach space E2:
ddtv(t)=Av(t)+Fv(t),v(0)=v0, | (4.11) |
where A is defined as the same as A in Section 2, and
(Fv)(a)=(λ[a|v1]s∗(a)−λ∗(a)(v1(a)+v2(a))−γ(a)v2(a)γ(a)v1(a)−δ(a)v2(a)),v=(v1v2)∈E2. | (4.12) |
Now let us consider the resolvent equation for A+F:
(ζ−(A+F))v=u,u∈E2,v=(v1v2)∈D(A). | (4.13) |
By the definition of the operators A and F, v1 satisfies
v1(a)=∫a0e−∫aσ2(ζ+γ(z)+λ∗(z))dz×(−λ∗(σ2)∫σ20e−∫σ2σ3(ζ+δ(z))dz(γ(σ3)v1(σ3)+u2(σ3))dσ3+s∗(σ2)λ[σ2|v1]+u1(σ2))dσ2. | (4.14) |
Once v1 is determined by (4.14), v2 is calculated as
v2(a)=∫a0e−ζ(a−σ)Δ(a)Δ(σ)[γ(σ)v1(σ)+u2(σ)]dσ. | (4.15) |
Define an operator Vζ on E by
(Vζϕ)(a):=−∫a0e−∫aσ2(ζ+γ(z)+λ∗(z))dzλ∗(σ2)∫σ20e−∫σ2σ3(ζ+δ(z))dzγ(σ3)ϕ(σ3)dσ3dσ2. | (4.16) |
Moreover, define an operator g and a given function h as
g[ϕ,ζ](a):=∫a0e−∫aσ2(ζ+γ(z)+λ∗(z))dzs∗(σ2)ϕ(σ2)dσ2, | (4.17) |
h[u,ζ](a):=∫a0e−∫aσ2(ζ+γ(z)+λ∗(z))dz[−λ∗(σ2)∫σ20e−∫σ2σ3(ζ+δ(z))dzu2(σ3)dσ3+u1(σ2)]dσ2. | (4.18) |
Thus if we suppose λ=λ[⋅∣v1] is given, the equation (4.14) with respect to v1 is rewritten as
v1(a)=(Vζv1)(a)+g[λ,ζ](a)+h[u,ζ](a). | (4.19) |
Since Vζ is Volterra type operator, R(1;Vζ):=(I−Vζ)−1 exists and (4.19) is solved as:
v1=R(1;Vζ)(g[λ,ζ]+h[u,ζ]). | (4.20) |
Then we have
λ(a)=∫ω0β(a,σ)N(σ)v1(σ)dσ=∫ω0β(a,σ)N(σ)R(1;Vζ)(g[λ,ζ]+h[u,ζ])(σ)dσ=∫ω0β(a,σ)N(σ)R(1;Vζ)g[λ,ζ](σ)dσ+∫ω0β(a,σ)N(σ)R(1;Vζ)h[u,ζ](σ)dσ=:(Wζλ)(a)+ξ(a;u,ζ), | (4.21) |
where Wζ is an integral operator from L1(0,ω) into itself defined by
(Wζϕ)(a):=∫ω0β(a,σ)N(σ)R(1;Vζ)g[ϕ,ζ](σ)dσ. | (4.22) |
Roughly speaking, if (I−Wζ)−1 exists, λ can be calculated as (I−Wζ)−1ξ and the resolvent (ζ−(A+F))−1 exists.
Lemma 4.7. For the linearized generator A+F at the steady state, it holds that
σ(A+F)=Pσ(A+F)={ζ∈C:1∈Pσ(Wζ)}, | (4.23) |
where σ(A) denotes the spectrum of A.
Proof. From the expression (4.20) and (4.15), we know that A+F has a compact resolvent, so it holds that σ(A+F)=Pσ(A+F) (Theorem 6.29, [35]). Let S:={ζ∈C:1∈σ(Wζ)}. From the above argument, it follows that C∖S⊂ρ(A+F), where ρ(A) denotes the resolvent set of A. Thus S⊃σ(A+F)=Pσ(A+F). Since Wζ is a compact operator, its spectrum different from zero is an eigenvalue (Theorem 6.26, [35]), so there exists an eigenfunction ϕζ such that Wζϕζ=ϕζ if ζ∈S. In this case,
(v1v2)=(((I−Vζ)−1g[ϕζ,ζ])(a)∫a0e−ζ(a−σ)Δ(a)Δ(σ)γ(σ)v1(σ)dσ), |
becomes an eigenfunction of A+F associated with eigenvalue ζ. Hence we have S⊂Pσ(A+F). Then we have (4.23).
In the following, we again adopt the separable mixing assumption that
β(a,σ)=β1(a)β2(σ), | (4.24) |
in order to simply make use of bifurcation arguments explicitly, although our argument could be applied to the model with the general transmission coefficient as is observed in [18].
In the separable mixing case, we can observe that
((I−Wζ)β1)(a)=β1(a)(1−∫ω0β2(σ)N(σ)R(1;Vζ)g[β1,ζ](σ)dσ), | (4.25) |
which shows that β1 is a positive eigenvector of Wζ.
As is shown in section 3, if we introduce a bifurcation parameter ϵ>0 such that β1=ϵβ10 and R0=1 when ϵ=1, we have λ∗=c(ϵ)ϵβ10 for a right neighborhood at ϵ=1, c(1)=0 and c′(1)>0. Then the parametrized model has the basic reproduction number ϵ. Now we define a two parameter function Λ as
Λ(ζ,ϵ):=1−∫ω0β2(σ)N(σ)R(1;Vζ,ϵ)g[ϵβ10,ζ](σ)dσ, | (4.26) |
where
(Vζ,ϵϕ)(a):=−c(ϵ)ϵ∫a0e−∫aσ2(ζ+γ(z)+c(ϵ)ϵβ10(z))dzβ10(σ2)∫σ20e−∫σ2σ3(ζ+δ(z))dzγ(σ3)ϕ(σ3)dσ3dσ2. | (4.27) |
Then the operator I−Wζ is not invertible if and only if Λ(ζ,ϵ)=0.
Since Vζ=0 for ϵ=1, we have Λ(0,1)=1−R0=0. Observe that
∂Λ∂ζ(0;1)=−∫ω0β2(σ)N(σ)∫σ0(σ2−σ)Γ(σ)Γ(σ2)β10(σ2)dσ2dσ>0. | (4.28) |
Therefore it follows from the Implicit Function Theorem that Λ(ζ,ϵ)=0 can be solved as ζ=ζ(ϵ) with ζ(1)=0 in the neighborhood of (ζ,ϵ)=(1,0). From Lemma 4.1, we know that ζ(1)=0 is the dominant real eigenvalue of the linearized generator A+F when ϵ=1.
Theorem 4.8. For the separable mixing case, the endemic steady state bifurcated from the disease-free steady state is locally asymptotically stable, if R0>1 and |R0−1| is sufficiently small.
Proof. Observe that
∂Λ∂ϵ(0,1)=−∂∂ϵ∫ω0β2(σ)N(σ)∞∑k=0(Vkζ,ϵg[ϵβ10,ζ])(σ)dσ|(ζ,ϵ)=(0,1). | (4.29) |
On the first term of the summation in (4.29),
−∂∂ϵ∫ω0β2(σ)N(σ)g[ϵβ10,ζ](σ)dσ|(ζ,ϵ)=(0,1)=−∂∂ϵ∫ω0β2(σ)N(σ)∫σ0e−∫σσ2(ζ+γ(z)+ϵc(ϵ)β10(z))dzs∗(σ2;ϵ)ϵβ10(σ2)dσ2dσ|(ζ,ϵ)=(0,1), | (4.30) |
where s∗(⋅;ϵ) is the prevalence of susceptibles in the bifurcated endemic steady state with the bifurcation parameter ϵ. Since λ∗(a)s∗(a)=b∗(a) and (3.12), we have
∂∂ϵs∗(σ2;1)=∂∂ϵ1ϵc(ϵ)β10(σ2)∞∑k=0T[ϵc(ϵ)β10]kf[ϵc(ϵ)β10](σ2,0)|ϵ=1=∂∂ϵ1ϵc(ϵ)β10(σ2){f[ϵc(ϵ)β10](σ2,0)+T[ϵc(ϵ)β10]f[ϵc(ϵ)β10](σ2,0)}|ϵ=1=−c′(1)∫σ20β10(z)[1−∫σ2zΠ(σ,z)dσ]dz<0. | (4.31) |
Therefore we obtain
−∂∂ϵ∫ω0β2(σ)N(σ)g[ϵβ10,ζ](σ)dσ|ϵ=1,ζ=0=J1+J2+J3, | (4.32) |
where
J1=c′(1)∫ω0β2(σ)N(σ)∫σ0∫σσ2β10(ζ)dζΓ(σ)Γ(σ2)β10(σ2)dσ2dσ,J2=∫ω0β2(σ)N(σ)∫σ0Γ(σ)Γ(σ2)β10(σ2)∫σ20β10(ζ)[1−∫σ2ζΠ(σ3,ζ)dσ3]dζdσ2dσ,J3:=−∫ω0β2(σ)N(σ)∫σ0Γ(σ)Γ(σ2)β10(σ2)dσ2dσ. | (4.33) |
From our assumption, we have J3=−R0=−1. Moreover, it follows from (3.27) and (3.30) that
J2=−c′(1)∂Θ∂c(0,1)=1. | (4.34) |
On the second term of the summation in (4.29),
−∂∂ϵ∫ω0β2(σ)N(σ)Vζ,ϵg[ϵβ10,ζ](σ)dσ|ϵ=1,ζ=0=∫ω0β2(σ)N(σ)∫σ0e−∫σσ2γ(z)dzβ10(σ2)c′(1)∫σ20e−∫σ2σ3δ(z)dzγ(σ3)×∫σ30e−∫σ3σ4γ(z)dzβ10(σ4)dσ4dσ3dσ2dσ=:J4 | (4.35) |
The third and the subsequent term in (4.29) is equal to zero, hence we obtain
∂Λ∂ϵ(0;1)=J1+J2+J3+J4>0. | (4.36) |
Therefore we conclude that
ζ′(1)=−(∂Λ∂ζ(0;1))−1∂Λ∂ϵ(0;1)<0, | (4.37) |
which shows that the dominant real eigenvalue moves to the left when ϵ crosses the unity. Using the Rouché theorem ([36,37]), we can show that another eigenvalues stays in the left half plane if |ϵ−1| is small enough. Then we conclude that the forwardly bifurcated small endemic steady state is locally asymptotically stable.
In this section, we again adopt the separable mixing assumption 3.8 to discuss the persistence threshold result of the SIRS model. The reader may find elementary ideas and techniques in [22,38,39]. A key idea is that the weak persistence could imply the strong persistence if there exists a compact attractor for the semiflow defined by the basic dynamical system.
First we consider the weak persistence of the semiflow induced from our SIRS epidemic model. The persistence can be seen as a mathematical formulation of the disease endemicity.
Definition 5.1. Let X be an arbitrary nonempty set and ρ:X→R+. A semiflow Φ:[0,∞)×X→X is called uniformly weakly ρ-persistent, if there exists some ϵ>0 such that
lim supt→∞ρ(Φ(t,x))>ϵfor all x∈X with ρ(x)>0, | (5.1) |
and is called uniformly strongly ρ-persistent, if there exists some ϵ>0 such that
lim inft→∞ρ(Φ(t,x))>ϵfor all x∈X with ρ(x)>0. | (5.2) |
Now we set the state space X={(x1,x2,x3)T∈E3+|x1+x2+x3=1} and the continuous semiflow Φ(t,x0)=u(t;x0), which is the solution of (2.7) with the initial value x0. Let us consider the persistence of the system (2.7) under the separable mixing assumption 3.8. Then the force of infection λ is represented as the form of separation of variables:
λ[a∣i(t,⋅)]=β1(a)∫ω0β2(σ)N(σ)i(t,σ)dσ=β1(a)ϕ(t), | (5.3) |
where
ϕ(t):=∫ω0β2(σ)N(σ)i(t,σ)dσ. | (5.4) |
Integrating along the characteristic line, for t>a, we have an expression
i(t,a)=∫a0Γ(a)Γ(η)λ[η∣i(t−a+η,⋅)]s(t−a+η,η)dη=∫a0Γ(a)Γ(η)β1(η)ϕ(t−a+η)s(t−a+η,η)dη. | (5.5) |
Lemma 5.2. If ‖i0‖E>0, then ‖i(t)‖E>0 for all t>0, where i(t):=i(t,⋅)∈L1(0,ω).
Proof. Let t0=inf{t∈R+:‖i(t)‖E=0} and suppose that t0<∞. As ‖i(t)‖E is continuous with respect to t, our assumption implies that t0>0 and i(t0,a)=0 for almost all a∈(0,ω). Let us fix a positive number 0<h<t0∧δ0, where δ0 is defined in Assumption 3.3. From (2.7), we have
i(t0,a)=0≥Γ(a)Γ(a−t)i(t0−h,a−h),a∈(h,ω), | (5.6) |
which shows that i(t0−h,a)=0 for almost all a∈(0,ω−h). Using (5.5), we obtain for almost all a∈(0,t0∧ω),
i(t0,a)=∫a0Γ(a)Γ(η)β1(η)ϕ(t0−a+η)s(t0−a+η,η)dη=0. | (5.7) |
From the Assumption 3.3 and the separable mixing assumption, we have β1(a)>0 for all a∈(0,ω) and s(t0−a+η,η)>0 for all η>0 because s(t,0)=1 for t≥0. Then we have ϕ(t0−a+η)=0 for almost all η∈(0,a) and a∈(0,t0∧ω). Hence we have ϕ(t)=0 for all t∈(0∨(t0−ω),t0). Observe that
ϕ(t)=∫ω0β2(σ)N(σ)i(t,σ)dσ≥β_∫ωω−δ0N(σ)i(t,σ)dσ. | (5.8) |
Then i(t,a)=0 for almost all a∈(ω−δ0,ω) and t∈(0∨(t0−ω),t0). Since ω−δ0<ω−h and t0−h∈(0∨(t0−ω),t0), we have i(t0−h,a)=0 for almost all a∈(0,ω), so ‖i(t0−h)‖E=0, which contradicts the definition of t0. Then we have t0=∞ and our conclusion.
Define the function ρ:X→R+ as
ρ(x1,x2,x3)=∫ω0N(σ)x2(σ)dσ. | (5.9) |
We decompose the state space X into two disjoint subsets:
X0:={x∈X∣ρ(x)>0},∂X0:={x∈X∣ρ(x)=0}. | (5.10) |
Then we obtain
Lemma 5.3. If ρ(s0,i0,r0)>0, then ρ(Ψ(t,(s0,i0,r0))>0 for all t>0.
Proof. If ρ(s0,i0,r0)>0, then ‖i0‖E>0. From of Lemma 5.2, ‖i(t,⋅)‖E>0 for all t>0. Thus we obtain ρ(Ψ(t,(s0,i0,r0))=∫ω0N(σ)i(t,σ)dσ>0 for all t>0.
Theorem 5.4. Suppose that the separable mixing assumption 3.8 holds. If R0>1, then the semiflow Φ is uniformly weak ρ-persistent. That is, there exists a number ϵ>0 such that
lim supt→∞∫ω0N(σ)i(t,σ)dσ>ϵ | (5.11) |
for all (s0,i0,r0)T∈X with ρ((s0,i0,r0))>0
Proof. Substituting (5.5) into the definition (5.4), we have
ϕ(t)=∫ω0β2(σ)N(σ)i(t,σ)dσ=∫ω0β2(σ)N(σ)∫σ0Γ(σ)Γ(η)β1(η)ϕ(t−σ+η)s(t−σ+η,η)dηdσ=∫ω0β2(σ)N(σ)∫σ0Γ(σ)Γ(σ−η)β1(σ−η)ϕ(t−η)s(t−η,σ−η)dηdσ=∫ω0ϕ(t−η)∫ωηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)s(t−η,σ−η)dσdη=∫ω0ϕ(t−η)∫tηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)s(t−η,σ−η)dσdη. | (5.12) |
For given b≥0, define ϕb(t):=ϕ(t+b). Here we adopt a convention that N(a)=0 for a>ω. For t>ω, we can observe that
ϕb(t)=∫ω0ϕb(t−η)∫t+bηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)s(t+b−η,σ−η)dσdη=∫ω0ϕb(t−η)∫tηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)s(t+b−η,σ−η)dσdη. | (5.13) |
Since
(∂∂t+∂∂a)s(t,a)≥−λ[a|i(t,⋅)]s(t,a), |
then s is estimated as
s(t,a)≥e−∫a0λ[z∣i(t−a+z,⋅)]dz for t>ω. | (5.14) |
Thus we obtain
ϕb(t)≥∫ω0ϕb(t−η)∫tηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)e−∫σ−η0β1(z)ϕ(t+b−σ+z)dzdσdη. | (5.15) |
Now we are going to prove (5.11) by contradiction. Assume that for all ϵ>0, there exist a time T0>0 and an initial data x0=(s0,i0,r0)T∈X such that ρ(Φ(t,x0))≤ϵ for all t≥T0. We set T0>ω without loss of generality. By (5.15), we obtain
ϕb(t)≥∫ω0ϕb(t−η)∫tηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)e−ϵβ∞2∫σ−η0β1(z)dzdσdη, | (5.16) |
where β∞2=supa∈[0,ω]β2(a). For sufficiently large T>ω and t>T0,
ϕb+T(t)≥∫t+T0ϕb+T(t−η)∫t+Tηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)e−ϵβ∞2∫σ−η0β1(z)dzdσdη≥∫t0ϕb+T(t−η)∫η+Tηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)e−ϵβ∞2M∫σ−η0β1(z)dzdσdη. | (5.17) |
Since ϕ is a bounded function, it is Laplace transformable on {λ>0}. By (5.17), we obtain
ˆϕb+T(λ)≥ˆϕb+T(λ)F(ϵ,λ,T), | (5.18) |
where
F(ϵ,λ,T)=∫∞0e−λη∫η+Tηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)e−ϵM∫σ−η0β1(z)dzdσdη. | (5.19) |
In particular, note that
F(0,0,T)=∫∞0∫η+Tηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)dσdη=∫ω0∫ωηβ2(σ)N(σ)Γ(σ)Γ(σ−η)β1(σ−η)dσdη=∫ω0∫ωηβ2(σ)N(σ)Γ(σ)Γ(η)β1(η)dσdη=R0>1. | (5.20) |
Because of the continuity of F(⋅,⋅,⋅), there are small positive number ϵ,λ, and large positive number T>ω such that F(ϵ,λ,T)>1 which implies ˆϕb+T(λ)=0 by (5.18). Then for sufficiently large T>ω, ϕb+T(t)=0 for all t∈(0,∞). If we take sufficiently large T1 such that T1>b+T,
0=ϕb+T(T1−b−T)=∫ω0β2(σ)N(σ)i(T1,σ)dσ, | (5.21) |
which implies i(T1,a)=0 for almost all a∈(ω−δ0,ω). Let us take positive numbers p and q such that 0<p<q<ω−δ0 with q−p<δ0, then (p+τ,q+τ)⊂(ω−δ0,ω) for ω−δ0−p<τ<ω−q. Hence i(T1+τ,a)=0 for almost all a∈(p+τ,q+τ), that is, i(T1,a)=0 for almost all a∈(p,q). Repeating to choose p and q finitely many times, we obtain i(T1,a)=0 for almost all a∈(0,ω−δ0). Consequently i(T1,a)=0 for almost all a∈(0,ω), so ∫ω0N(σ)i(T1,σ)dσ=0, which contradicts Lemma 5.3.
Let Φ:R×X→X be the semiflow induced by the mild solution of the system (2.7). Then it is easy to see that Φ is state-continuous, i.e., all maps Φ(t,⋅):X→X, t≥0 is continuous. According to [22], here we summarize some basic definitions for a semiflow Φ in a metric space X. A nonempty, compact, invariant set K⊂X is called a compact attractor of bounded sets if K attracts all bounded sets of X. The state-continuous semiflow Φ is called point-dissipative if there exists a bounded subset B of X which attracts all points in X. Φ is called asymptotically smooth if Φ is asymptotically compact on every forward invariant bounded closed set. Φ is called eventually bounded on a set M⊂X if Φ([r,∞)×M) is bounded for some r>0.
Theorem 5.5. Suppose the Assumption 3.8. There exists a compact attractor B of bounded sets in X.
Proof. From Theorem 2.33 of [22], the statement is proved if we can check that the semiflow Φ is point-dissipative, eventually bounded on every bounded sets in X and asymptotically smooth. The first and the second conditions hold trivially, because X itself is a bounded set. We are going to prove the eventual smoothness. As X itself is forward invariant bounded closed set, it is sufficient to show that Φ(t,⋅)X is compact. In order to check the Fréchet-Kolmogolov criterion for the compactness of sets in Lp(R) (Theorem B.1, [22]; [29]), it is sufficient to show the equi-continuity in L1 of Φ(t,⋅)X for a large t>0. By Assumption 3.3, for any ϵ>0 there exists some κ∈(0,ϵ) such that
∫ω0|β(a+h,η)−β(a,η)|da<ϵ uniformly for η∈R,∫ω0|δ(a+h)−δ(a)|da<ϵ,∫ω0|γ(a+h)−γ(a)|da<ϵ. | (5.22) |
whenever |h|<κ. For sufficiently large T0>ω, let B={Φ(T0;(s0,i0,r0))∣(s0,i0,r0)T∈X}. Integrating s, i and r in (2.7) along the characteristic line, for t>ω>a and |h|<κ,
|s(t,a+h)−s(t,a)|≤|e−∫a+h0λ[z∣i(t−a−h+z,⋅)]dz−e−∫a0λ[z∣i(t−a+z,⋅)]dz|+|∫a+h0e−∫a+hσλ[z∣i(t−a−h+z,⋅)]dzδ(σ)r(t−a−h+σ,σ)dσ−∫a0e−∫aσλ[z∣i(t−a+z,⋅)]dzδ(σ)r(t−a+σ,σ)dσ|≤|∫a+h0λ[z∣i(t−a−h+z,⋅)]dz−∫a0λ[z∣i(t−a+z,⋅)]dz|+|∫a0(e−∫a+hσ+hλ[z∣i(t−a−h+z,⋅)]dzδ(σ+h)r(t−a+σ,σ+h)dσ−e−∫aσλ[z∣i(t−a+z,⋅)]dzδ(σ)r(t−a+σ,σ))dσ|+|∫0−he−∫a+hσ+hλ[z∣i(t−a−h+z,⋅)]dzδ(σ+h)r(t−a+σ,σ+h)dσ|≤(1+δ∞+Bω(1+β∞+δ∞ω))ϵ+δ∞∫a0|r(t−a+σ,σ+h)−r(t−a+σ,σ)|dσ. | (5.23) |
By the same kind of calculation, we have
|r(t,a+h)−r(t,a)|≤γ∞∫a0|i(t−a+σ,σ+h)−i(t−a+σ,σ)|dσ+(1+γ∞+γ∞ω)ϵ,|i(t,a+h)−i(t,a)|≤β∞Bω∫a0|s(t−a+σ,σ+h)−s(t−a+σ,σ)|dσ+ωB(1+2β∞ω+β∞)ϵ. | (5.24) |
Thus for all t>3ω,
|s(t,a+h)−s(t,a)|≤c1ϵ+c2∫a0|s(t−a+ξ,ξ+h)−s(t−a+ξ,ξ)|dξ, | (5.25) |
where
c1:=1+δ∞+Bω(1+β1∞β2∞+δ∞ω)+δ∞ω(1+γ∞+γ∞ω)+12δ∞γ∞ω3B(1+2β∞ω+β∞),c2:=β∞γ∞δ∞ω. |
Define a function
h(b):=c2e−c2b∫b0|s(t−a+ξ,ξ+h)−s(t−a+ξ,ξ)|dξ, for b∈(0,ω). | (5.26) |
Then h′ satisfies
h′(b)=c2e−c2b[|s(t−a+b,b+h)−s(t−a+b,b)|−c2∫b0|s(t−a+ξ,ξ+h)−s(t−a+ξ,ξ)|dξ]≤c1c2e−c2bϵ. | (5.27) |
Therefore we know that h(b)≤c1(1−e−c2b)ϵ holds. Then we have
c2e−c2a∫a0|s(t−a+ξ,ξ+h)−s(t−a+ξ,ξ)|dξ≤c1(1−e−c2a)ϵ, | (5.28) |
which shows that
∫a0|s(t−a+ξ,ξ+h)−s(t−a+ξ,ξ)|dξ≤c1c2ec2ωϵ. | (5.29) |
Thus each PiB (i=1,2,3) is relatively compact subsets of E, where each Pi:X→E is the projection to the i-th component. This implies the set B is relatively compact set in X.
Theorem 5.6. Suppose the Assumption 3.8. If R0>1, then the semiflow Φ is uniformly strongly ρ-persistent.
Proof. To show this, it is needed to show that there is no total trajectory ϕ:R→B such that ρ(ϕ(0))=0 and ρ(ϕ(t))>0 for all t∈R∖{0} (see [22] Theorem 5.2.). Assume ρ(ϕ(0))=0. Then we obtain i0=0∈E. For t∈[0,ω], I(t):=∫ω0N(σ)i(t,σ)dσ is estimated as
I(t)=∫t0N(σ)i(t,σ)dσ+∫ωtN(σ)i(t,σ)dσ≤∫t0N(σ)∫σ0e−∫σηγ(z)dzλ[η∣i(t−σ+η,⋅)]s(t−σ+η,η)dηdσ+∫ωtN(σ)∫t0e−∫σηγ(z+σ−t)dzλ[σ−t+η∣i(η,⋅)]s(η,σ−t+η)dηdσ≤β∞B{∫t0∫σ0I(t−σ+η)dηdσ+∫ωt∫t0I(η)dηdσ}≤β∞Bω∫t0I(η)dη. | (5.30) |
From the Gronwall inequality, we conclude that I(t)=0 for all t∈[0,ω], which shows that there is no trajectory such as mentioned at the beginning.
In this section, we provide numerical examples that support our theoretical results. Let ω:=1 to normalize the age interval as [0,1]. Fix the age-specific death rate as μ(a):=[10(1−a)2]−1, a∈[0,1). We then see that μ∈L1loc,+(0,1), ∫10μ(σ)dσ=∞ and ℓ(a)=exp(−∫a0μ(σ)dσ)=exp(−a/[10(1−a)]), a∈[0,1) (see Figure 1 (a)). Let B:=1/∫10ℓ(a)da≈1.2527, and thus, the total population is normalized as ∫10N(a)da=∫10Bℓ(a)da=1. We fix the following parameter functions:
γ(a):=100,δ(a):=δ0[arctan20(a−12)+π2],a∈[0,1],β(a,σ)=β(a):=β0(√ae−2a+1100),(a,σ)∈[0,1]×[0,1], | (6.1) |
where β0>0 and δ0>0 are positive constants. The choice of these functions is based on the following biological assumptions: the average infectious period 1/γ(a)=1/100 is age-independent and 100 times smaller than the maximum attainable age; the loss-of-immunity rate δ(a) is high in the elderly people (see Figure 1 (b)); the transmission coefficient β(a,σ)=β(a) depends only on the age of susceptibles and the youth people are more likely to be infected (see Figure 1 (c)). We can easily check that all necessary assumptions on the parameter functions stated in the previous sections are satisfied. In particular, the basic reproduction number R0 is explicitly calculated as
R0=∫10∫1ηN(σ)Γ(σ)Γ(η)dσβ(η)dη≈1.2527 β0∫10∫1ηe−σ10(1−σ)e−100(σ−η)dσ(√ηe−2η+1100)dη. | (6.2) |
In what follows, we fix the following initial data x0=(s0,i0,r0):
s0(a):=1−i0(a),i0(a):=12e−100(a−12)2×10−3,r0(a):=0,a∈[0,1]. |
We first verify the threshold property of R0 under the fixed δ0=1. For β0=380, we obtain R0≈0.9812<1. Hence, by Theorem 4.6, we can expect that the disease-free steady state is globally asymptotically stable. In fact, Figure 2 (a) exhibits that the age distribution of infected population converges to zero as time evolves. On the other hand, for β0=400, we obtain R0≈1.0329>1. Hence, by Theorems 3.6, 4.5, 4.8 and 5.6, we can expect that the disease-free steady state is unstable, there exists a locally asymptotically stable endemic steady state and the semiflow Φ is uniformly strongly ρ-persistent. In fact, Figure 2 (b) exhibits that the age distribution of infected population converges to a positive distribution as time evolves. This example suggests that the endemic steady state is not only locally but also globally asymptotically stable in this case.
We next observe the effect of loss of immunity δ0 under the fixed β0=600. Note that R0 is independent of δ0 and fixed to 1.5493>1 in this case. In Figure 3, we see that the infected population increases in particular in the elderly age class as the effect of loss of immunity δ0 increases. Specifically, the total number of infected population ∫10N(σ)i(t,σ)dσ converges to 0.0095, 0.0338 and 0.0747 as time evolves for δ0=1, 20 and 70, respectively. This implies that R0 for the SIRS epidemic model does not reflect the intensity of the disease endemicity but it is the threshold for the eradication or persistence of the disease.
We here briefly consider a mass-vaccination effect on the basic system (2.1) to clear the implication of reinfection on R0 and the threshold results, because it has practically important implications that the reinfection phenomena would make disease control more difficult and complex. In fact, threshold results of the SIRS epidemic are similar to those of the SIR epidemic, but its controllability is very much different from the SIR epidemic. An important effect of vaccination policy is reduction of the effective size of the susceptible population, however in the reinfection model, there is a possibility that a disease can invade a fully vaccinated population, and we are naturally led to the idea of the reinfection threshold ([6,7]). In other words, for the SIRS reinfection model, mass-vaccination policy is not necessarily almighty.
Suppose that newborns in the virgin population are mass vaccinated with coverage ϵ∈[0,1] and the immunological status of newly vaccinated individuals is identical with that of the newly recovered individuals. Then it is easy to see that the boundary condition in (2.1) is replaced by s(t,0)=1−ϵ, i(t,0)=0 and r(t,0)=ϵ. Then it is easy to see that the disease-free steady state is a partially immunized state given by (s†,i†,r†)=(1−ϵΔ(a),0,ϵΔ(a)). Here let us introduce the effective next generation operator Kϵ as
(Kϵϕ)(a):=s†(a)∫ω0∫ωηβ(a,σ)N(σ)Γ(σ)Γ(η)dσϕ(η)dη. | (7.1) |
Then the effective reproduction number Rϵ is given by its spectral radius r(Kϵ). From the monotonicity of the spectral radius, we have Rϵ≤R0. Let σ:=R1/R0, where R1 is the effective reproduction number for the fully vaccinated population. Given that the qualitative change in the epidemiological implication occurs for the prevalence and controllability at R0=1/σ, Gomes et al. ([40,41]) referred to 1/σ as the reinfection threshold of R0. As seen above, the reinfection threshold of R0 corresponds to the fact that σR0=R1=1, i.e., R0=1/σ does not imply a bifurcation point of the basic system (2.1), but the threshold condition R1=1 of the fully vaccinated system. The disease is uncontrollable by the mass vaccination if R1=σR0>1, because the fully vaccinated population can be invaded by the disease.
As an example, we consider the same parameter functions as in Section 6 with β0=500 and δ0=30. In this case, we have R0≈1.2911, and thus, the disease will persist without vaccination. For ϵ=1, we obtain R1≈0.9696<1, and hence, the reinfection threshold is σ−1=R0/R1≈1.33>R0, the complete mass-vaccination policy is successful and the disease will be eradicated in this case (see Figure 4 (a)). On the other hand, for ϵ=0.8, we obtain Rϵ≈1.0339>1. This implies that the disease will persist even if 80 percent of newborns are successfully immunized by vaccination in this case (see Figure 4 (b)). In fact, as in (6.2), we can calculate Rϵ as
Rϵ=R0−ϵ∫10∫1ηN(σ)Γ(σ)Γ(η)dσΔ(η)β(η)dη, | (7.2) |
and hence, Rϵ<1 is equivalent to
ϵ>R0−1∫10∫1ηN(σ)Γ(σ)Γ(η)dσΔ(η)β(η)dη≈0.9054. | (7.3) |
That is, more than 90 percent of newborns should be immunized to control the disease in this case, and (7.3) is a much severe criterion than the usual critical coverage of immunization ϵ>1−1/R0 calculated for the SIR disease with permanent immunity.
In this paper, we have rigorously established the threshold property of R0 in the age-structured SIRS model that the disease will be naturally eradicated if R0<1, while it is strongly persistent and endemic steady states exists if R0>1. It is noted that different from the SIR model, we have not yet known whether the endemic steady state is unique or not even in the separable mixing case, because the characteristic equation satisfied by the force of infection at the endemic steady state is complex, and not monotone. The number of endemic steady states and their stability should be investigated in future. Although our main analysis depends on the separable mixing assumption, it is limited and could be relaxed to obtain the same kind of results.
Using numerical calculations, we have shown that the loss of immunity has a drastic effect on the critical coverage of immunization. In fact, if the basic reproduction number is grater than the reinfection threshold, we cannot control the disease by the mass vaccination to newborns. We have shown the critical coverage of immunization for the separable mixing case.
For future extension of our model and real-world applications, it is noted that if the vaccination effect is incomplete, the vaccinated individuals could be partially susceptible and their infection would lead partial infectivity. If the secondary infection will lead a longer infective period, the reproductivity enhancement would occur and we could expect that there exist subcritical endemic steady states [7]. In such a case, even the subcriticality R0<1 does not guarantee the eradication of the disease.
Another possible extension would be realized if we introduce the class age structure of recovered individuals, because the loss of immunity depends on the time since recovery. It is a future challenge to develop age-structured epidemic models that can describe more realistic, complex dynamics of susceptibility, infectivity and immunity within host individuals.
H. Inaba and K. Okuwa were supported by the Japan Society for the Promotion of Science (JSPS), Grant-in-Aid for Scientific Research (C) (No.16K05266) and T. Kuniya was supported by JSPS, Grant-in-Aid for Early-Career Scientists (No.19K14594).
The authors declare there is no conflict of interest.
[1] | W. O. Kermack and A. G. McKendrick, Contributions to the mathematical theory of epidemics II. The problem of endemicity, Proc. Roy. Soc., 138A(1932), 55–83; reprinted in Bull. Math. Biol. 53(1991), 57–87. |
[2] | W. O. Kermack and A. G. McKendrick, Contributions to the mathematical theory of epidemics III. Further studies of the problem of endemicity, Proc. Roy. Soc., 141A, 94–122; reprinted in Bull. Math. Biol., 53(1991), 89–118. |
[3] | H. Inaba, Kermack and McKendrick revisited: The variable susceptibility model for infectious diseases, Jap. J. Indust. Appl. Math., 18(2001), 273–292. |
[4] | H. Inaba, Endemic threshold and stability in an evolutionary epidemic model, Mathematical Approaches for Emerging and Reemerging Infectious Diseases, Castillo-Chaves, C. et al. (eds.), The IMA Volumes in Mathematics and its Applications 126, Springer, (2002), 337–359. |
[5] | H. R. Thieme and J. Yang, An endemic model with variable re–infection rate and applications to influenza, Math. Biosci., 180(2002), 207–235. |
[6] | H. Inaba, Endemic threshold analysis for the Kermack–McKendrick reinfection model, Josai Math. Monograph., 9(2016), 105–133. |
[7] | H. Inaba, Age-Structured Population Dynamics in Demography and Epidemiology, Springer, Singapore, 2017. |
[8] | H. W. Hethcote, Qualitative analysis of communicable disease models, Math. Biosci., 28(1976), 335–356. |
[9] | J. L. Aron, Dynamics of acquired immunity boosted by exposure to infection, Math. Biosci.,64(1983), 249–259. |
[10] | J. L. Aron, Acquired immunity dependent upon exposure in an SIRS epidemic model, Math. Biosci., 88(1988a), 37–47. |
[11] | J. L. Aron, Mathematical modelling of immunity of malaria, Math. Biosci., 90(1988b), 385–396. |
[12] | P. K. Tapaswi and J. Chattopadhyay, Global stability results of a "susceptible-infective-immune-susceptible" (SIRS) epidemic model, Ecol. Modell., 87(1996), 223–226. |
[13] | S. Busenberg and K. P. Hadeler, Demography and epidemics, Math. Biosci.. 101(1990), 63–74. |
[14] | S. Busenberg and P. van den Driessche, Analysis of a disease transmission model in a population with varying size, J. Math. Biol., 28, 257–270. |
[15] | Y. Nakata, Y. Enatsu, H. Inaba, et al., Stability of epidemic models with waning immunity, SUT J. Math., 50(2014), 205–245. |
[16] | D. Breda, O. Diekmann, W. F. de Graaf, et al., On the formulation of epidemic models (an appraisal of Kermack and McKendrick), J. Biol. Dyn., 6(2012), S103–S117. |
[17] | D. W. Tudor, An age-dependent epidemic model with application to measles, Math. Biosci., 73(1985), 131–147. |
[18] | H. Inaba, Threshold and stability results for an age-structured epidemic model, J. Math. Biol., 28(1990), 411–434. |
[19] | M. C. M. De Jong, O. Diekmann and H. Heesterbeek, How does transmission of infection depend on population size ?, Epidemic Models: Their Structure and Relation to Data, D. Mollison (ed.), Cambridge U. P., Cambridge, (1995), 84–94. |
[20] | M. van Boven, H. E. de Melker, J. F. P. Schellekens, et al., Waning immunity and sub-clinical infection in an epidemic model: implications for pertussis in the Netherlands, Math. Biosci., 164(2000), 161–182. |
[21] | S. Tsutsui, Mathematical analysis for an age-structured epidemic model with waning immunity and subclinical infection, Master Thesis, Graduate School of Mathematical Sciences, The Univer-sity of Tokyo, 2010. |
[22] | H. L. Smith and H. R. Thieme Dynamical Systems and Population Persistence, Graduate Studies in Mathematics 118, Amer. Math. Soc. Providence, Rhode Island, 2011. |
[23] | S. Busenberg, M. Iannelli, and H. Thieme, Global behaviour of an age-structured S-I-S epidemic model, SIAM J. Math. Anal., 22(1991), 1065–1080. |
[24] | O. Diekmann, J. A. P. Heesterbeek and J. A. Metz, On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations, J. Math. Biol., 28(1990), 365–382. |
[25] | O. Diekmann, J. A. P. Heesterbeek and T. Britton, Mathematical Tools for Understanding Infec-tious Disease Dynamics, Princeton University Press, Princeton and Oxford, 2013. |
[26] | H. Inaba, On a new perspective of the basic reproduction number in heterogeneous environments, J. Math. Biol., 65(2012), 309–348. |
[27] | C. Barril, À. Calsina and J. Ripoll, A practical approach to R0 in continuous-time ecological models, Math. Meth. Appl. Sci., 41(2017), 8432–8445. |
[28] | M. A. Krasnoselskii, Positive Solutions of Operator Equations, Noordhoff, Groningen, 1964. |
[29] | N. Dunford and J. T. Schwartz Linear Operators Part I: General Theory, New York: Interscience publishers, 1958. |
[30] | I. Sawashima, On spectral properties of some positive operators, Nat. Sci. Rep. Ochanomizu Univ., 15(1964), 53–64. |
[31] | I. Marek, Frobenius theory of positive operators: Comparison theorems and applications, SIAM J. Appl. Math., 19(1970), 607–628. |
[32] | H. Inaba, Mathematical analysis of an age-structured SIR epidemic model with vertical transmis-sion, Disc. Conti. Dyn. Sys., Series B, 6(2006), 69–96. |
[33] | H. J. Heijmans, The dynamical behaviour of the age-size-distribution of a cell population, The Dynamics of Physiologically Structured Populations, Springer, Berlin, Heidelberg, (1986), 185–202. |
[34] | K. J. Engel and R. Nagel, One-Parameter Semigroups for Linear Evolution Equations, Springer Science & Business Media, 1999. |
[35] | T. Kato, Perturbation Theory for Linear Operators, 2nd Edition, Springer, Berlin, 1984. |
[36] | M. Iannelli, Mathematical Theory of Age-Structured Population Dynamics, Giardini Editori e Stampatori in Pisa, 1995. |
[37] | M. Iannelli and F. Milner, The Basic Approach to Age-Structured Population Dynamics, Springer, The Netherlands, 2017. |
[38] | H. R. Thieme, Disease extinction and disease persistence in age structured epidemic models, Nonl. Anal., 47(2001), 6181–6194. |
[39] | H. R. Thieme, Mathematics in Population Biology, Princeton University Press, Princeton and Oxford, 2003. |
[40] | M. G. Gomes, L. J. White and G. F. Medley, Infection, reinfection, and vaccination under subop-timal immune protection: epidemiological perspectives, J. Theor. Biol., 228(2004), 539–549. |
[41] | M. G. Gomes, L. J. White and G. F. Medley, The reinfection threshold, J. Theor. Biol., 236(2005), 111-113. |
1. | Michael F. Good, Michael T. Hawkes, Alan Sher, The Interaction of Natural and Vaccine-Induced Immunity with Social Distancing Predicts the Evolution of the COVID-19 Pandemic, 2020, 11, 2150-7511, 10.1128/mBio.02617-20 | |
2. | Kaiyan Zhao, Shaojuan Ma, Qualitative analysis of a two-group SVIR epidemic model with random effect, 2021, 2021, 1687-1847, 10.1186/s13662-021-03332-w | |
3. | Huaixing Li, Jiaoyan Wang, Global Dynamics of an SEIR Model with the Age of Infection and Vaccination, 2021, 9, 2227-7390, 2195, 10.3390/math9182195 | |
4. | V. Skakauskas, The Kermack-McKendrick epidemic model with variable infectivity modified, 2022, 507, 0022247X, 125817, 10.1016/j.jmaa.2021.125817 | |
5. | Xiaoguang Li, Liming Cai, Mohammad Murshed, Jin Wang, Dynamical analysis of an age-structured dengue model with asymptomatic infection, 2023, 524, 0022247X, 127127, 10.1016/j.jmaa.2023.127127 | |
6. | Kento Okuwa, Hisashi Inaba, Toshikazu Kuniya, An age-structured epidemic model with boosting and waning of immune status, 2021, 18, 1551-0018, 5707, 10.3934/mbe.2021289 | |
7. | Sourav Kumar Sasmal, Yasuhiro Takeuchi, Editorial: Mathematical Modeling to Solve the Problems in Life Sciences, 2020, 17, 1551-0018, 2967, 10.3934/mbe.2020167 | |
8. | Toshikazu Kuniya, Structure of epidemic models: toward further applications in economics, 2021, 72, 1352-4739, 581, 10.1007/s42973-021-00094-8 | |
9. | Bukyoung Jhun, Hoyun Choi, Abrupt transition of the efficient vaccination strategy in a population with heterogeneous fatality rates, 2022, 32, 1054-1500, 093140, 10.1063/5.0087627 | |
10. | Candy Sonveaux, Joseph J. Winkin, Design of a vaccination law for an age-dependent epidemic model using state feedback, 2022, 55, 24058963, 65, 10.1016/j.ifacol.2022.10.378 | |
11. | Manoj Kumar, Syed Abbas, Age-Structured SIR Model for the Spread of Infectious Diseases Through Indirect Contacts, 2022, 19, 1660-5446, 10.1007/s00009-021-01925-z | |
12. | Michael T. Hawkes, Bonita E. Lee, Jamil N. Kanji, Nathan Zelyas, Kerry Wong, Michelle Barton, Shamir Mukhi, Joan L. Robinson, Seasonality of Respiratory Viruses at Northern Latitudes, 2021, 4, 2574-3805, e2124650, 10.1001/jamanetworkopen.2021.24650 | |
13. | Manoj Kumar, Syed Abbas, Analysis of steady state solutions to an age structured SEQIR model with optimal vaccination, 2022, 45, 0170-4214, 10718, 10.1002/mma.8414 | |
14. | Nurbek Azimaqin, Yingke Li, Xianning Liu, A heterogeneous continuous age-structured model of mumps with vaccine, 2025, 10, 24680427, 75, 10.1016/j.idm.2024.09.004 | |
15. | Candy Sonveaux, Joseph J. Winkin, State feedback control law design for an age-dependent SIR model, 2023, 158, 00051098, 111297, 10.1016/j.automatica.2023.111297 | |
16. | Riya Das, Dhiraj Kumar Das, T K Kar, Analysis of a chronological age-structured epidemic model with a pair of optimal treatment controls, 2024, 99, 0031-8949, 125240, 10.1088/1402-4896/ad8e0b | |
17. | Rukhsar Ikram, Amir Khan, Aeshah A. Raezah, Impact of supervise neural network on a stochastic epidemic model with Levy noise, 2024, 9, 2473-6988, 21273, 10.3934/math.20241033 | |
18. | Ralph Brinks, Annika Hoyer, Approximation of the infection-age-structured SIR model by the conventional SIR model of infectious disease epidemiology, 2024, 4, 2674-1199, 10.3389/fepid.2024.1429034 | |
19. | Xiaoxia Li, Yang Wang, Juping Zhang, Zhen Jin, Propagation Dynamic for a Physiological Age-Structured SIR Epidemic Model with Diffusion, 2025, 24, 1575-5460, 10.1007/s12346-024-01219-1 | |
20. | Peng Wu, Lan Zou, Shigui Ruan, An age-structured syphilis model, I: well-posedness and stability, 2025, 481, 1471-2946, 10.1098/rspa.2024.0218 | |
21. | Nurbek Azimaqin, Xianning Liu, Yangjiang Wei, Yingke Li, Explicit Formula of the Basic Reproduction Number for Heterogeneous Age‐Structured SIR Epidemic Model, 2025, 0170-4214, 10.1002/mma.10994 |