1,219
Views
7
CrossRef citations to date
0
Altmetric
Articles

A Lyapunov–Schmidt method for detecting backward bifurcation in age-structured population models

&
Pages 543-565 | Received 26 Dec 2019, Accepted 14 Jun 2020, Published online: 02 Jul 2020

Abstract

Backward bifurcation is an important property of infectious disease models. A centre manifold method has been developed by Castillo-Chavez and Song for detecting the presence of backward bifurcation and deriving a necessary and sufficient condition for its occurrence in Ordinary Differential Equations (ODE) models. In this paper, we extend this method to partial differential equation systems. First, we state a main theorem. Next we illustrate the application of the new method on a chronological age-structured Susceptible-Infected-Susceptible (SIS) model with density-dependent recovery rate, an age-since-infection structured HIV/AIDS model with standard incidence and an age-since-infection structured cholera model with vaccination.

AMS Subject Classification:

1. Introduction

Backward bifurcation plays an important role in infectious diseases and their control. Having methods for deriving necessary and sufficient conditions for backward bifurcation can delineate the reasons for its occurrence and help devise control strategies that combat its negative effects. A recent review article [Citation10] collects methods for analysis of backward bifurcation in ODE models and age-structured partial differential equation (PDE) models. The reader may find some mechanisms that create backward bifurcations for endemic steady states in [Citation7].

Most methods were first developed or applied in ODEs but have an analogue for age-structured PDEs. A very popular method for detecting backward bifurcation in ODE models was developed by Castillo-Chavez and Song [Citation2]. Their method is based on the centre manifold theorem and is one of the preferred methods for modern analysis of backward bifurcation in ODE population models. As developed in [Citation2], however, this method cannot be applied to PDEs and does not work for delay equations, age-structured population models, or diffusion models.

In this paper, our goal is to extend the method to infinite dimensional systems. In the next section, we state a theorem analogous to Castillo-Chavez and Song's theorem for infinite dimensional systems, which gives the necessary and sufficient condition for backward bifurcations to occur. In one of our remarks, we argue that the backward bifurcating solution is unstable, at least close to the bifurcating point, while the forward bifurcating solution is locally stable, also at least close to the bifurcating point. This is an expected universal property for transcritical bifurcation. In the rest of the paper, we focus on concrete epidemic examples. The first example we consider is a chronological age-structured SIS model with non-linear recovery rate. For this example, the theory applies directly because the boundary condition is linear, so the differential operator becomes a generator of a C0-semigroup. We also illustrate this example with a numerical simulation that shows the backward bifurcation, once the condition for it has been met. Example 2 is an age-since-infection HIV/AIDS model with standard incidence. It is known from the literature that this model does not exhibit backward bifurcation and the result of the analysis concludes exactly that. The importance of this example is the development of the tools to handle the non-linear, non-local boundary condition associated with age-since-infection structured models. Example 3 is an age-since-infection structured cholera model considered as a test example in [Citation10], where we derive a necessary and sufficient condition for backward bifurcation using several different methods. Here we use the newly proposed Lyapunov–Schmidt method and we reach a similar condition for backward bifurcation. Compared to the methods in [Citation10] the Lyapunov–Schmidt method developed here is somewhat easier to use but it is based on more abstract approach.

We anticipate that the methodology introduced here can be applied to other infinite dimensional models, such as diffusion epidemic models, or delay epidemic models. We expect that this methodology will be helpful to pave the road for deriving necessary and sufficient conditions for backward bifurcation in infinite dimensional systems.

2. Lyapunov–Schmidt method for PDEs

In this section, we introduce an analogue of the Castillo-Chavez and Song Theorem [Citation2] designed for PDEs. The proof is based on the Lyapunov–Schmidt theory.

Let X and Z be Banach spaces, and xX and αR is a parameter. We consider the following abstract differential equation (1) dxdt=F(x,α),F:X×RZ.(1) Without loss of generality, we assume that 0 is an equilibrium so F(0,α)=0forallαR.

Theorem 2.1

Assume:

  • A=DxF(0,α0) is the linearization around the equilibrium 0, evaluated at a critical value of parameter α0, such that A is a closed operator with a simple isolated eigenvalue zero and remaining eigenvalues having negative real part. Let v^0 be the unique (up to a constant) positive solution of Av=0.

  • F(x,α)C2(U0×I0,Z) for some neighbourhood U0 of 0 and interval I0 containing α0.

  • Assume Z is the dual of Z and , is the pairing between Z and Z. Assume v^0Z is the unique (up to a constant) positive vector satisfying Ax,v^0=0for all xX, that is dim(kerA)=1, where A is the adjoint of A, and kerA=span{v^0}.

  • Assume Dxα2F(0,α0)v^0,v^00 where DxαF(0,α0) is the second derivative of F with respect to x and α.

Then, the direction of the bifurcation is determined by the numbers (2) a=Dxx2F(0,α0)[v^0,v^0],v^0,b=Dxα2F(0,α0)v^0,v^0,(2) where Dxx2F(0,α0)[h1,h2] is the second derivative of F with respect to x applied to the functions h1 and h2. If b>0, the bifurcation is backward if and only if a>0 and forward if and only if a<0.

Proof Theorem 2.1.

(Theorem 1) To determine the direction of the bifurcation, we study the solution of F(x,α)=0 in a neighbourhood of (0,α0)X×R. Let A=DxF(0,α0). Since A is closed and has a simple isolated eigenvalue, it follows from [Citation4, p. 284] that A is a Fredholm operator of index zero, that is A satisfies:

  • dim(kerA)=1;

  • dim(Z/imA)=1;

  • imA is closed in Z

where kerA={xX:Ax=0}, and imA={yZ:y=Ax for some xX}. Because, dim(Z/imA)=1, we have Z/imA=span{[z^0]} where [z^0]=z^0+imA and z^0imA. Let Z0=span{z^0}. Then following [Citation3] we have: Z=Z0imA. Similarly, if kerA=span{v^0}, then X=kerAX0, where X0 is the topological complement of kerA in X.

It follows that there exist projections PL(X,kerA) and QL(Z,Z0) such that kerA=imP and imA=kerQ. Consequently we can write xX in the form x=v+y where v=PxkerA. That implies that there exists a constant c such that v=cv^0. In addition y=(IP)xX0. Using the projection Q we can decompose F(x,α)=0 into two equations (3) (IQ)F(v+y,α)=0,QF(v+y,α)=0.(3) Let G(v,y,α):=(IQ)F(v+y,α). Then the first equation in (Equation3) is written as G(v,y,α)=0. In particular, we have G(0,0,α0)=(IQ)F(0,α0)=0. Furthermore, DyG(0,0,α0)=(IQ)DxF(0,α0)=(IQ)A=A. The restriction A|X0 is a one-to-one mapping between X0 and imA and it is therefore invertible. Hence DyG(0,0,α0)|X0 is invertible. As a result we can apply the Implicit Function Theorem. From the Implicit Function Theorem, it follows that there exists a neighbourhood ΩkerA with 0Ω and an interval I with α0I and a C2 map ψ:Ω×IX0 with y=ψ(v,α) such that ψ(0,α0)=0 and (IQ)F(v+ψ(v,α),α)0forall(v,α)Ω×I. Furthermore, we define a map B:Ω×IZ0 by B(v,α):=QF(v+ψ(v,α),α). Then the second equation in (Equation3) can be written as (4) B(v,α)=0,(4) which is called the bifurcation equation.

The Lyapunov–Schmidt reduction says that there exists a neighbourhood U of 0Ω such that each solution of B(v,α)=0 corresponds one-to-one to some solution of F(x,α)=0. Furthermore, it is shown in [Citation8, Corollary I.2.4] that (5) B(0,α0)=0,Bv(0,α0)=0.(5) From F(0,α)=0 for all α, we have that ψ(0,α)=0forallαI. Furthermore, Dαψ(0,α)=0forallαI. Consequently, B(0,α)=0forallαI. This corresponds to the trivial solution. This last equality means that we can write B in the following way [Citation8]: B(v,α)=01ddtB(tv,α)dt. Alternatively, B(v,α)=01DvB(tv,α)vdtforall(v,α)Ω×I. For vΩ, we set v=sv^0, s(δ,δ) and we obtain the nontrivial solutions for s0 by solving B~(s,α)01DvB(tsv^0,α)v^0dt=0fors(δ,δ),s0. The assumptions on the smoothness of F give that B~C1((δ,δ)×I,Z0). Furthermore, by the properties (Equation5) we have B~(0,α0)=0. We would like to apply the Implicit Function Theorem and express α=ϕ(s). To do that, we need DαB~(0,α0)0. Computation similar to [Citation8, p. 19] give that (6) Dα(DvB(v,α)v^0)=Dα(QDxF(v+ψ(v,α),α)(v^0+Dvψ(v,α)v^0)=QDxx2F(v+ψ(v,α)α,α)[v^0+Dvψ(v,α)v^0,Dαψ(v,α)]+QDxF(v+ψ(v,α),α)Dvα2ψ(v,α)v^0+QDxα2F(v+ψ(v,α),α)(v^0+Dvψ(v,α)v^0).(6) Setting (v,α)=(0,α0) we find that the first term is equal to zero from the properties of ψ and the second term is also zero by the definition of the projection Q. Hence, we obtain (7) DαB~(0,α0)=QDxα2F(0,α0)v^0,(7) and QDxα2F(0,α0)v^00 because of assumptions A3 and A4. In particular, Dxα2F(0,α0)v^00 and Dxα2F(0,α0)v^0imA. The Implicit Function Theorem now gives the existence of ϕ(s) such that ϕ:(δ,δ)I with ϕ(0)=α0 and B~(s,ϕ(s))=0foralls(δ,δ). Then B(sv^0,ϕ(s))=sB~(s,ϕ(s))=0 for s(δ,δ). Consequently, F(x(s),α(s))=0fors(δ,δ) where (8) x(s)=sv^0+ψ(sv^0,ϕ(s)),α(s)=ϕ(s).(8) This is the essence of the Crandall-Rabinowitz Theorem [Citation8]. Next we claim that the tangent vector to the nontrivial solution curve (Equation8) at the bifurcation point (0,α0) is given by (v^0,α(0))X×R. Differentiating x(s) in (Equation8) with respect to s gives (9) ddsx(s)|s=0=v^0+Dvψ(0,α0)v^0+Dαψ(0,α0)α˙(0)=v^0.(9) The bifurcation at the critical value (0,α0) is backward if and only if α(0)<0. To obtain formulas to compute α(0), we recall that α(s)=ϕ(s) and therefore B~(s,α(s))=0 for all s(δ,δ). Thus, (10) ddsB~(s,α(s))|s=0=DsB~(0,α0)+DαB~(0,α0)α˙(0)=0,(10) where DαB~(0,α0)0 is given in (Equation7). Furthermore, (11) DsB~(0,α0)=01Dvv2B(0,α0)[v^0,tv^0]dt=QDxx2F(0,α0)[v^0,v^0]01tdt=12QDxx2F(0,α0)[v^0,v^0].(11) Using the dual pairing and from Equation (Equation10) we have 12QDxx2F(0,α0)[v^0,v^0],v^0+QDxα2F(0,α0)[v^0,v^0],v^0α˙(0)=0. Notice that (12) QDxx2F(0,α0)[v^0,v^0],v^0=Dxx2F(0,α0)[v^0,v^0],v^0,QDxα2F(0,α0)v^0,v^0=Dxα2F(0,α0)v^0,v^0.(12) These equalities hold because z,v^0=0 for all zimA. Hence, we have α(0)=12Dxx2F(0,α0)[v^0,v^0],v^0Dxα2F(0,α0)v^0,v^0=12ab. If we assume that b>0, the bifurcation curve bifurcates backward if α(0)<0, that is, if a>0. If α(0)>0, that is a<0, then the bifurcation is forwards.

Remark 2.1

The condition b>0 means that R0 increases with the parameter α. If b<0, that means that R0 decreases as α increases and here we will not consider this case as a backward bifurcation case.

Remark 2.2

As is usually observed in the ODE case, stability of the bifurcating solutions can be determined such that the backwardly bifurcated solution is unstable, while the forwardly bifurcated solution is locally stable as long as the parameter s is small enough. According to Britton [Citation1], let us show the Factorization Theorem. Let (x(s),α(s)) be the bifurcating solutions as (13) F(x(s),α(s))=0,s(δ,δ),(x(0),α(0))=(0,α0),(13) and ddsx(s)|s=0=v^0. Define A(s):=DxF(x(s),α(s)). Then A(0)=A. The dominant eigenvalue σ(s) and corresponding eigenvector are introduced as follows: (14) A(s)v^0(s)=σ^(s)v^0(s),(14) where σ^(0)=0,v^0(0)=v^0. Define the dual eigenvector v^0(s) as A(s)v^0(s)=σ^(s)v^0(s), where they are normalized as follows: v^0(s),v^0(s)=1. Especially, we write as v^0(0)=v^0.

Differentiating (Equation13) with respect to s, we have (15) DxF(x,α)x(s)+DαF(x,α)α(s)=0(15) From (Equation15), taking the duality pairing with v^0(s), it follows that (16) DxF(x,α)x(s),v^0(s)=A(s)x(s),v^0(s)=x(s),A(s)v^0(s)=σ^(s)x(s),v^0(s)=DαF(x,α)α(s),v^0(s)=α(s)DαF(x,α),v^0(s).(16) Then we arrive at an expression (17) σ^(s)=α(s)DαF(x,α),v^0(s)x(s),v^0(s)(17) Using the Taylor expansion around s = 0, we have DαF(x,α)=DxDαF(0,α0)x(s)s+O(s2),α(s)=α0+α(0)s+O(s2),x(s),v^0(s)=1+O(s), where we used DαF(0,α0)=0. Then it holds that (18) σ^(s)=(α(s)α0)DxDαF(0,α0)v^0,v^0+O(s2)=(α(s)α0)b+O(s2).(18) Therefore, if we assume that b>0, for small s>0, the backward bifurcation (α(0)<0) implies the bifurcated solution is unstable (σ^(s)>0), while the forward bifurcation (α(0)>0) implies the bifurcated solution is locally stable (σ^(s)<0) as long as we could assume that σ(s) is the dominant eigenvalue. In many concrete examples, we could show the dominance of σ(s) for small s using complex analysis arguments applied to the characteristic equation.

In the following sections, we illustrate how the above theory can be applied to concrete examples from epidemic modeling.

3. Example 1: age-structured SIS epidemic model

3.1. An SIS epidemic model with density-dependent recovery rate

First we consider the following SIS model with age-structure and non-linear recovery rate: (19) st+sa=κ(a)s(a,t)0ωβ(σ)i(σ,t)dσ+γ(a)i(a,t)A+0ωi(a,t)dait+ia=κ(a)s(a,t)0ωβ(σ)i(σ,t)dσγ(a)i(a,t)A+0ωi(a,t)da(19) where a is chronological age, t is time, and ω is the maximal age, assumed finite. Furthermore, s(a,t) and i(a,t) are the age-density functions of susceptibles and invectives, respectively, κ(a) is the susceptibility of susceptible individuals, β(a) is the age-specific infectivity of infective individuals, γ(a) is the standard recovery rate, and A is a non-zero constant. That is, it is assumed that the effective recovery rate is density-dependent, it is decreasing as the size of infecteds increases. Although this is a toy model to illustrate the bifurcation calculation, it would not necessarily be unrealistic, if the recovery needs medical supports and the medical resource is limited. Here we assume that the time scale of epidemic is so short that the natural death rate of host population can be neglectedFootnote1.

We assume that κ(a),β(a),γ(a)L(0,ω). Although model (Equation19) is with separable transmission rate, this approach can be applied in more general situations.

The model has the following initial and boundary conditions: (20) s(0,t)=Bi(0,t)=0s(a,0)=s0(a)i(a,0)=i0(a)(20) where B is the birth rate, assumed constant. Furthermore, we assume that s0(a)+i0(a)=B for all a[0,ω]. The equation of the total population size is nt+na=0. Hence, the total population is constant n(a,t)=B. Let, s(a,t)=Bi(a,t). Thus, we reduce the system to a unique equation: (21) it+ia=κ(a)(Bi(a,t))0ωβ(σ)i(σ,t)dσγ(a)i(a,t)A+0ωi(a,t)da,i(0,t)=0.(21)

3.2. Application of the Lyapunov–Schmidt method

Now, we consider the equation for the equilibria which is the base equation to which we will apply the Lyapunov–Schmidt backward bifurcation theorem. (22) dida+κ(a)(Bi(a))0ωβ(σ)i(σ)dσγ(a)i(a)A+0ωi(a)da=0,i(0)=0.(22) This equation clearly has the disease-free equilibrium i0(a)=0. Let us introduce a bifurcation parameter β¯ as β(a)=β¯β0(a), where a normalized function β0L+1(0,ω)L+(0,ω) will be determined below. Thus, we have a problem F(i,β¯)=0, where F(i,β¯):=(Ai)(a)+κ(a)(Bi(a))β¯0ωβ0(σ)i(σ)dσγ(a)i(a)A+0ωi(a)da, where A is the differential operator given by (Af)(a)=f(a) with domain D(A)={fAC(0,ω):f(0)=0}, where AC(0,ω) is the set of absolutely continuous functions on the interval (0,ω). The operator F is defined from D(A)×R to L1(0,ω).

To compute DiF(0,β¯) we consider a parameter s and a function h(a) with properties to be specified later and we compose φ(s)=F(i+sh,β¯). Taking φ(s) and setting s = 0 we obtain (23) DiF(i,β¯)h=(Ah)(a)κ(a)h(a)β¯0ωβ0(σ)i(σ)dσ+β¯κ(a)(Bi)0ωβ0(σ)h(σ)dσγ(a)h(a)A+0ωi(a)da+γ(a)i(a)(A+0ωi(a)da)20ωh(σ)dσ.(23) Setting i = 0, we have DiF(0,β¯)h=(Ah)(a)+β¯κ(a)B0ωβ0(σ)h(σ)dσγ(a)h(a)A. This is not exactly the operator A since the derivative is evaluated at an arbitrary value of the parameter β¯.

To see the stability of the disease-free equilibrium, we look for eigenvalues of the operator DiF(0,β¯)h=λh. This leads to the linearized equation of the disease-free equilibrium: (24) ha+λh+γ(a)Ah=κ(a)BH,h(0)=0,(24) where H=0ωβ(σ)h(σ)dσ. Solving the above differential equation we have h(a)=BH0aeλ(as)sa(γ(η)/A)dηκ(s)ds. Substituting h(a) in H and cancelling since we are looking for non-zero h, we obtain the following characteristic equation: 1=B0ωβ(a)0aeλ(as)sa(γ(η)/A)dηκ(s)dsda. From here we can define the basic reproduction number R0=β¯B0ωβ0(a)0aesa(γ(η)/A)dηκ(s)dsda. The normalized function β0 is assumed to satisfy (25) 1=B0ωβ0(a)0aesa(γ(η)/A)dηκ(s)dsda.(25) Then we have R0=β¯ and λ=0 is a simple isolated eigenvalue of the characteristic equation when β¯=1. It is not hard to see that all remaining eigenvalues have negative real part.

We define the operator A: A=DiF(0,1), which has the same domain as D(A). Looking for kerA, we need to solve Av=0. Following a procedure similar to obtaining h(a), we obtain v^0(a)=B0aesa(γ(η)/A)dηκ(s)ds. We notice that from (Equation25) we have 0ωβ0(a)v^0(a)da=1. To obtain v^0 we need to formally find the adjoint operator, and obtain its kernel. Since Z=L1(0,ω), the dual space is Z=L(0,ω). Let A be the adjoint with D(A)={ψ(a)AC(0,ω):ψ(ω)=0}. For ψD(A), we have after integration by parts (26) 0ω[ha+κ(a)B0ωβ0(s)h(s)dsγ(a)Ah(a)]ψ(a)da=0ω[ψa+B0ωκ(s)ψ(s)dsβ0(a)γ(a)Aψ(a)]h(a)da.(26) Hence, Aψ=ψa+B0ωκ(s)ψ(s)dsβ0(a)γ(a)Aψ(a) with ψ(ω)=0. To find v^0 we need to determine the kernel of the adjoint operator. For that purpose, we have to solve the problem: (27) ψa+Bβ0(a)Kγ(a)Aψ=0,ψ(ω)=0,(27) where K=0ωκ(s)ψ(s)ds. Solving this system as before, we obtain v^0(a)=Baωβ0(a)esa(γ(η)/A)dηds. It is not hard to see from integration by parts that 0ωκ(a)v^0(a)da=1. Next, we compute a and b. First we compute b. We have, Diβ¯2F(0,1)v^0=Bκ(a)0ωβ0(s)v^0(s)ds. Hence, we have (28) b=B0ωκ(a)v^0(a)da0ωβ0(s)v^0(s)ds=B>0.(28) Next, we compute a. First, we need to compute Dii2F(0,β¯)[h,h2]. To do that we consider (29) φ2(s)=DiF(i+sh2,β¯)h=haβ¯κ(a)h0ωβ0(a)(i+sh2)da+β¯κ(a)(Bish2)0ωβ0(a)h(a)daγ(a)h(a)A+0ω(i+sh2)da+γ(a)(i+sh2(a))(A+0ω(i+sh2)da)20ωh(η)dη.(29) Differentiating with respect to s and setting s = 0 and i = 0, we obtain (30) Dii2F(0,β¯)[h,h2]=β¯κ(a)h(a)0ωβ0(a)h2(a)daβ¯κ(a)h2(a)0ωβ0(a)h(a)da+γ(a)h(a)A20ωh2(a)da+γ(a)h2(a)A20ωh(a)da.(30) We notice that this form is symmetric with respect to h and h2. Thus, we have Dii2F(0,1)[v^0,v^0]=2κ(a)v^0(a)0ωβ0(a)v^0(a)da+2γ(a)v^0(a)A20ωv^0(a)da. If we recall 0ωβ0(a)v^0(a)da=1, multiplying with v^0 and integrating, we obtain (31) a=20ωκ(a)v^0(a)v^0(a)da+2A20ωγ(a)v^0(a)v^0(a)da0ωv^0(a)da.(31) Hence, the following is a necessary and sufficient condition for backward bifurcation (a>0) (32) 0ωκ(a)v^0(a)v^0(a)da<0ωγ(a)v^0(a)v^0A2da0ωv^0(a)da.(32)

4. Example 2: infection-age structured HIV/AIDS epidemic model

The example in this section is an HIV/AIDS epidemic model developed in [Citation11]. Although it is already known that this model does not possess backward bifurcation, it is instructive to apply the technique on this rather simple model under formal setting before we apply it to more complex models. In fact, if we extend this model to take into account chronological age structure, we can show that the backward bifurcation of endemic steady states can occur [Citation5,Citation6], and many age-structured epidemic model can be formulated as the same kind of the Cauchy problem with non-densely defined generator and non-local boundary condition.

Consider the infection-age (age-since-infection) structured HIV/AIDS model with the standard incidence: (33) S=Λβ¯SN0β0(τ)i(τ,t)dτμS,it+iτ=(μ+α(τ))i,i(0,t)=β¯S(t)N(t)0β0(τ)i(τ,t)dτ,(33) where S(t) is the number of susceptibles, i(τ,t) is the infection-age density of infectives and N(t)=S(t)+0i(τ,t)dτ. is the total size of sexually active population, which is not constant because of the disease-induced death rate α(τ). We assume that β0(τ) is normalized as (34) 0β0(τ)π(τ)dτ=1,(34) where π(τ):=exp(0τ(μ+α(σ))dσ), denotes the survival rate of infecteds, so we can use β¯ as a bifurcation parameter, and β¯ is no other than the basic reproduction number.

According to [Citation9], in order to take into account the boundary condition, we define the extended state space X:=R×X^ associated with X^:=L1(0,)×R. We also define the subspace by X0:=R×X^0, where X^0:=L1(0,)×{0}. Then X0 can be identified with the usual state space R×L1(0,). Let E0=(Λ/μ,0^)X be the disease-free steady state, where 0^=(0,0)TX^0. Define a linear operator A^:D(A^)X^0X^ by A^(u0)=(u(μ+α)uu(0)), for uW1,1(0,) with D(A^)=W1,1(0,)×{0}X^0.

Let i^(t)=(i(t,),0)TX^0. Then the basic problem (Equation36) can be formulated as a semilinear Cauchy problem in X: (35) dS(t)dt=μS(t)+G1(S(t),i^(t)),di^(t)dt=A^i^(t)+G2(S(t),i^(t)),(35) where for x=(x1,x^2)TX0, G1:X0R and G2:X0X^ are defined by G1(x):=Λβ¯x1N(x)0β0(τ)x2(τ)dτ,G2(x):=(0β¯x1N(x)0β0(τ)x2(τ)dτ) and N(x) is a positive functional of x=(x1,x^2)TX0 defined by N(x):=x1+0x2(τ)dτ. Moreover, define a linear operator A:D(A)X0X and a nonlinear operator G:X0X by A(x1x^2)=(μx1A^x^2),G(x):=(G1(x)G2(x)) with D(A)=R×D(A^). Then D(A)¯=X0 is not dense in X, but A satisfies the Hille-Yosida estimate. Then (Equation35) can be formulated as a well-posed abstract Cauchy problem in X as follows: (36) dxdt=Ax+G(x),x(0)X0,t0,(36) where x=(x1,x^2)TX0.

In order to examine the bifurcation at the disease-free steady state of (Equation39), let F(x,β¯):=Ax+G(x), for x=(x1,x^2)X0. Then F(x0,β¯)=0 if x0=(Λ/μ,0^) is the disease-free steady state. Then the linearized operator A:=DxF(x0,β¯) acting on X is calculated as follows: (37) Ax:=(μx1β¯0β0(τ)x2(τ)dτx2(τ)(μ+α(τ))x2(τ)x2(0)+β¯0β0(τ)x2(τ)dτ),(37) where its domain is given by D(A)=D(A)=R×W1,1(0,)×{0}. Let us consider the eigenvalues of A. Solving the differential equation from Ax=λx for xX0, we have x2(τ)=x2(0)π(τ)eλτ. Inserting the above expression into the third equation in Ax=λx, we have x2(0)+x2(0)β¯0β0(τ)π(τ)eλτdτ=0. Then we obtain the characteristic equation (38) 1=β¯0β0(τ)π(τ)eλτdτ,(38) which has a simple eigenvalue of zero when β¯=1. To find the eigenvector v^0 associated with eigenvalue zero, we need to solve Ax=0. Then we can choose x2(τ)=π(τ) and the boundary condition (the third equation of Ax=0) is trivially satisfied at β¯=1. From the first equation of Ax=0, we have x1=(1/μ) and v^0=(1μ,π(τ),1). To find the adjoint operator A, let x=(x1,x^2)TD(A)X0, we multiply Ax with the vector ψ=(ψ1,ψ2,ψ3)TX:=R×L(0,)×R. Then we have ψ,Ax=[μx1ψ1β¯0β0(τ)x2(τ)dτ]ψ1+0(x2(τ)(μ+α(τ))x2(τ))ψ2(τ)dτ+[x2(0)+β¯0β0(τ)x2(τ)dτ]ψ3. Observe that 0(x2(τ)(μ+α(τ))x2(τ))ψ2(τ)dτ=x2(τ)ψ2(τ)|0+0x2(τ)ψ2(τ)τdτ0(μ+α(τ))x2(τ)ψ2(τ)dτ=x2(0)ψ2(0)+0[ψ2(τ)(μ+α(τ))ψ2(τ)]x2(τ)dτ, where we assume that ψ2()=0. Hence, we have ψ,Ax=μx1ψ1+0(β¯β0(τ)(ψ1+ψ2(0))+ψ2(τ)(μ+α(τ))ψ2(τ))x2(τ)dτ+x2(0)(ψ3+ψ2(0))=Aψ,x, which should hold for ψD(A) and xD(A)X0.

Thus we choose the domain of A as D(A)={ψ=(ψ1,ψ2,ψ3)R×W1,×R:ψ2()=0,ψ3=ψ2(0)}. Then the adjoint operator A is defined on D(A)X=R×L×R as follows: Aψ=(μψ1β¯β0(ψ1+ψ2(0))+ψ2(τ)(μ+α(τ))ψ2ψ2(0)),ψD(A). To find v^0, we solve Aψ=0 with β¯=1. We have ψ1=0. Solving the differential equation in Aψ=0, we have (39) ψ2(τ)=ψ2(0)τβ0(s)π(s)π(τ)ds.(39) Hence, we have the eigenfunction v^0=(0,ψ2(τ),ψ2(0)), where ψ2(0)>0.

Computing the second derivative for F, we have DxDβ¯F(x0,1)v^0=(0β0(τ)π(τ)dτ00β0(τ)π(τ)dτ). Hence, we obtain (40) b=DxDβ¯F(x0,1)v^0,v^0=ψ2(0)0β0(τ)π(τ)dτ>0.(40) On the other hand, Dxx2F(x0,1)[v^0,v^0] can be computed as follows: Dxx2F(x0,1)[v^0,v^0]=2hkF(x0+(h+k)v^0,1)|(h,k)=(0,0)=(2μΛ0π(τ)dτ02μΛ0π(τ)dτ). Thus, we have (41) a=Dxx2F(x0,1)[v^0,v^0],v^0=ψ2(0)2μΛ0π(τ)dτ<0.(41) Hence, it follows from (Equation41) and (Equation40) that 12a/b>0, so the bifurcation at (x,β¯)=(x0,1) is a forward one, which is a conclusion that we expected.

Remark 4.1

In this example, just as in the ODE case [Citation2], we see that components of the eigenvector v^0, that correspond to positive entries in the disease-free equilibrium, may be negative. This is in general the case with the PDEs as well.

5. Example 3: age-since-infection structured cholera model

In this section, we consider a cholera model as an age-since-infection example with backward bifurcation. The model was first introduced in [Citation10]. Here we use the model and formal calculations along Theorem 1 to derive the necessary and sufficient condition for backward bifurcation associated with this model. In particular, if τ is the age-since-infection of infectious individuals, and θ is the age of the bacteria in the environment, we denote by i(τ,t) the density of the infectives and by u(θ,t) the density of the bacteria in the environment.

Let also S be the number of susceptibles, V the number of vaccinated, and R the number of recovered individuals. With these notations, the age-structured model becomes: (42) S=ΛSB+D0β(θ)u(θ,t)dθ(μ+ψ)S+wR,V=ψSσVB+D0β(θ)u(θ,t)dθμV,it+iτ=(μ+γ(τ))i,i(0,t)=S+σVB+D0β(θ)u(θ,t)dθ,R=0γ(τ)i(τ,t)dτ(μ+w)R,uθ+ut=δ(θ)u,u(0,t)=0η(τ)i(τ,t)dτ.(42) Again the basic system can be formulated as a Cauchy problem with non-densely defined generators, so here we skip repeating the abstract formulation by using the extended state space.

The parameters are described in [Citation10]; β and δ are functions of θ, while γ and η are structured by τ. B in this model denotes the concentration of bacteria in the environment B(t)=0u(θ,t)dθ. This system has a disease-free equilibrium E0=(S0,V0,0,0,0) where S0=Λμ+ψV0=ψμS0. We define the probability of survival of infected individual in the infectious class and the probability of survival of the bacteria in the environment: π(τ)=eμτ+0τγ(s)dsπd(θ)=e0θδ(s)ds. The control reproduction number is given by (derivation can be found in [Citation10]): (43) R(ψ)=(S0+σV0)D0β(θ)πd(θ)dθ0η(τ)π(τ)dτ.(43) To have a bifurcation parameter, we decompose β(τ) as β¯β0(τ) where maxβ0(τ)=1. We define X=R×R×L1(0,)×R×R×L1(0,)×R and X0=X=R×R×L1(0,)×{0}×R×L1(0,)×{0}. For an element ϕX0 with ϕ=(S,V,i,0,R,u,0) we define the non-linear operator F(ϕ,β¯)=(Λβ¯SB+D0β0(θ)u(θ)dθ(μ+ψ)S+wRψSσβ¯VB+D0β0(θ)u(θ)dθμViτ(μ+γ(τ))iβ¯S+σVB+D0β0(θ)u(θ)dθi(0)0γ(τ)i(τ)dτ(μ+w)Ruθδ(θ)u0η(τ)i(τ)dτu(0)). Next, we define the derivative of the operator F applied to the element h=(x,v,y,0,r,z,0). We have Fϕ(ϕ,β¯)h=(Q1(μ+ψ)x+wrψxQ2μvyτ(μ+γ(τ))yQ3y(0)0γ(τ)y(τ)dτ(μ+w)rzθδ(θ)z0η(τ)y(τ)dτz(0)). where Q1=β¯SB+D0β0(θ)z(θ)dθβ¯xB+D0β0(θ)u(θ)dθ+β¯Sb(B+D)20β0(θ)u(θ)dθ,Q2=σβ¯VB+D0β0(θ)z(θ)dθ+σβ¯vB+D0β0(θ)u(θ)dθσVb(B+D)20β(θ)u(θ)dθ,Q3=β¯S+σVB+D0β0(θ)z(θ)dθ+β¯x+σvB+D0β0(θ)u(θ)dθβ¯(S+σV)b(B+D)20β0(θ)u(θ)dθ, and b=0z(θ)dθ. Computing the derivative of F at the DFE, we have: Fϕ(ϕ,β¯)h=(β¯S0D0β0(τ)z(τ)dτ(μ+ψ)x+wrψSβ¯σV0D0β0(τ)z(τ)dτμViτ(μ+γ(τ))iβ¯S0+σV0D0β0(τ)z(τ)dτy(0)0γ(τ)y(τ)dτ(μ+w)rzθδ(θ)z0η(τ)y(τ)dτz(0)). We define the operator A=Fϕ(ϕ0,β¯) where β¯ is the value of β¯ that makes R(ψ)=1. We solve the system Ah=0. From the differential equations we obtain y(τ)=y(0)π(τ),z(τ)=z(0)πd(τ). We take y(0)=1. Then z(0)=y(0)0η(τ)π(τ)dτ=Σ, and y(0)=β¯S0+σV0DΣ0β0(θ)πd(θ)dθ=1. This is trivially satisfied. Thus, we obtain y(τ)=π(τ),z(τ)=Σπd(τ). From the equation for x we have x=S0DΣ0β(τ)πd(τ)dτ)+wμ+w0γ(τ)π(τ)dτμ+ψ. Further, from the equation for v we have v=ψxσV0DΣ0β(θ)πd(θ)dθμ. Finally, from the equation for r we have r=0γ(τ)π(τ)dτμ+w=Γμ+w. Thus, the eigenvector with x and v as above is v^0=(x,v,π(τ),0,Γμ+w,Σπd,0). To compute the adjoint operator, we take ξZ and consider Ah,ξ=S0D0β(θ)z(θ)dθξ1(μ+ψ)xξ1+wrξ1+ψxξ2σV0D0β(θ)z(θ)dθξ2μvξ20(yτ+(μ+γ(τ)y(τ))ξ3(τ)dτ+S0+σV0D0β(θ)z(θ)dθξ4y(0)ξ4+0γ(τ)y(τ)dτξ5(μ+w)rξ50zθ+δ(θz(θ)ξ6(θ)dθ+ξ70η(τ)y(τ)dτξ7z(0). Integrating the differential equations by parts, we have the following adjoint operator: Aξ=((μ+ψ)ξ1+ψξ2μξ2(ξ3)τ(μ+γ(τ))ξ3(τ)+γ(τ)ξ5+η(τ)ξ6(0)ξ3(0)wξ1(μ+w)ξ5(ξ6)θδ(θ)ξ6(θ)S0Dβ(θ)ξ1σV0Dβ(θ)ξ2+S0+σV0Dβ(θ)ξ3(0)ξ6(0)). To find eigenvector that generates the kernel, we set Aξ=0. From the second line, we have ξ2=0. From the first line we have ξ1=0. From the fifth line, we have ξ5=0. Solving the equation for ξ3 we have (ξ3)τ(μ+γ(τ))ξ3(τ)+η(τ)ξ6(0)=0. Then we have ξ3(τ)=ξ6(0)τη(s)eτs(μ+γ(p))dpds. Further, ξ3(0)=ξ6(0)0η(s)e0s(μ+γ(p))dpds=ξ6(0)Σ. Solving the equation for ξ6(θ) we have (ξ6)θδ(θ)ξ6(θ)+S0+σV0Dβ(θ)ξ6(0)Σ=0. Integrating, we obtain: ξ6(θ)=S0+σV0Dβ(θ)ξ6(0)Σθβ(θ)eθsδ(p)dpds. We may take ξ6(0)=1. We take ξ4=ξ3(0) and ξ7=ξ6(0). Then the eigenvector is v^0=(0,0,ξ3(τ),Σ,0,ξ5(θ),1). To compute a and b we compute the second derivatives: Next, we define the derivative of the operator F applied to the element h2=(x2,v2,y2,0,r2,z2,0). We have Fϕϕ(ϕ0,β¯)[h,h2]=(Q1[h,h2]Q2[h,h2]0Q3[h,h2]000). where Q1=β¯x2DSb2D20β0(θ)z(θ)dθβ¯xD0β0(θ)z2(θ)dθ+β¯Sb(D)20β0(θ)z2(θ)dθ,Q2=σβ¯v2DVb2D20β0(θ)z(θ)dθσβ¯vD0β0(θ)z2(θ)dθ+σVb(D)20β(θ)z2(θ)dθ,Q3=β¯(x2+σv2)D(S+σV)b2D20β0(θ)z(θ)dθ+β¯x+σvD0β0(θ)z2(θ)dθβ¯(S+σV)b(D)20β0(θ)z2(θ)dθ. and b2=0z2(θ)dθ.Fϕϕ(ϕ0,β¯)[h,h]=(2β¯xDS0bD20β0(θ)z(θ)dθ2σβ¯vDV0bD20β0(θ)z(θ)dθ02β¯(x+σv)D(S0+σV0)bD20β0(θ)z(θ)dθ000). Thus, (44) a=Fϕϕ(ϕ0,β¯)[v^0v^0],v^0=C1((1+σψμ)S0DΣ0β(τ)πd(τ)dτ)+wμ+w0γ(τ)π(τ)dτμ+ψ+σσV0DΣ0β(θ)πd(θ)dθμ(S0+σV0)ΣΠD).(44) where C1=2Σ2D0β(θ)πd(θ)dθ. Denote by B=0β(τ)πd(τ)dτ. and Γ=0γ(τ)π(τ)dτ. Then, if b:=Fϕβ¯(ϕ0,β¯)v^0,v^0 is positive, the condition for backward bifurcation becomes (45) Qw(μ+w)(μ+ψ)Γ>QS0D(μ+ψ)ΣB+σ2V0DμΣB+(S0+σV0)ΣΠD,(45) where Q=(1+σψμ). This is the same condition obtained in [Citation10] as a condition for backward bifurcation of this model. Finally we check the positivity of b. First we compute the derivative of Fϕ with respect to β¯. We have Fϕβ¯(ϕ0,β¯)h=(S0D0β0(θ)z(θ)dθσV0D0β0(θ)z(θ)dθ0,S0+σV0D0β0(θ)z(θ)dθ000). Then we have, (46) b=Fϕβ¯(ϕ0,β¯)v^0,v^0=S0+σV0DΣ20β0(θ)πd(θ)dθ>0.(46) This concludes the example.

Acknowledgements

The authors would like to thank the editors and two anonymous reviewers for their helpful comments, HI was supported by the japan society for promotion of Science Grant-in-Aid for Scientific Research (c) (No. 19K03614).

Disclosure statement

No potential conflict of interest was reported by the author(s).

Notes

1 Another possible interpretation for (Equation19) is that s and i are prevalencies, B = 1 and the survival rate is Type 1, that is, all deaths occur at a=ω.

References

  • N.F. Britton, Reaction-Diffusion Equations and Their Applications to Biology, Academic Press, London, 1986.
  • C. Castillo-Chavez and B Song, Dynamical models of tuberculosis and their applications, Math Biosciences Eng 1(2) (2004), pp. 361–404. doi: 10.3934/mbe.2004.1.361
  • S. Guo and J. Wu, Bifurcation Theory of Functional Differential Equations, Springer, New York, 2013.
  • M. Haragus and G. Iooss, Local Bifurcations, Center Manifolds and Normal Forms in Infinite-Dimensional Dynamical Systems, Springer, New York, 2011.
  • H. Inaba, Backward bifurcation in a HIV/AIDS epidemic model with age structure II: the Case of General Transmission Rate, RIMS Kokyuroku Vol. 1337, Mathematical Economics, Resarch Institute for Mathematical Sciences, Kyoto University, Kyoto, 2003, p. 103–111.
  • H. Inaba, Endemic threshold results for age-duration-structured population model for HIV infection, Math Biosci 201 (2006), pp. 15–47. doi: 10.1016/j.mbs.2005.12.017
  • H. Inaba, Age-Structured Population Dynamics in Demography and Epidemiology, Springer, Singapore, 2017.
  • H. Kielhöfer, Bifurcation Theory: An Introduction with Applications to Partial Differential Equations, 2nd ed., Springer, New York, 2012.
  • P. Magal, C.C. McCluskey, and G.F. Webb, Lyapunov functional and global asymptotic stability for an infection-age model, Applicable Anal 89(7) (July 2010), pp. 1109–1140. doi: 10.1080/00036810903208122
  • M. Martcheva, Methods for deriving necessary and sufficient conditions for backward bifurcation, JBD 13(1) (2019), pp. 538–566. doi: 10.1080/17513758.2019.1647359
  • H. Thieme and C. Castillo-Chavez, How may infection-age-dependent infectivity affect the dynamics of HIV/AIDS?, SIAM J Appl Math 53(5) (1993), pp. 1447–1479. doi: 10.1137/0153068