962
Views
10
CrossRef citations to date
0
Altmetric
Articles

Bifurcation analysis of an age-structured alcoholism model

, &
Pages 987-1011 | Received 10 Mar 2018, Accepted 07 Oct 2018, Published online: 01 Nov 2018

ABSTRACT

In this paper, we investigate a new alcoholism model in which alcoholics have age structure. We rewrite the model as an abstract non-densely defined Cauchy problem and obtain the condition which guarantees the existence of the unique positive steady state. By linearizing the model at steady state and analyzing the associated characteristic transcendental equations, we study the local asymptotic stability of the steady state. Furthermore, by using Hopf bifurcation theorem in Liu et al. (Z. Angew. Math. Phys. 62 (2011) 191–222), we show that Hopf bifurcation occurs at the positive steady state when bifurcating parameter crosses some critical values. Finally, we perform some numerical simulations to illustrate our theoretical results and give a brief conclusion.

1. Introduction

Globally, alcohol consumption results in approximately 3.3 million deaths each year(or 5.9% of all deaths), this is greater than, for example, the proportion of deaths from HIV/AIDS (2.8%), violence (0.9%) or tuberculosis (1.7%) [Citation28]. Alcohol consumption has been identified as a component cause for more than 200 diseases, injuries and other health conditions as described in the International Statistical Classification of Diseases and Related Health Problems (ICD) 10th Revision (WHO, 1992), more than 30 include alcohol in their name or definition. This indicates that these disease conditions do not exist at all in the absence of alcohol consumption. A strong association exists between alcohol consumption and HIV infection, sexually transmitted diseases [Citation27]. Alcohol consumption can result in harm to other individuals, such as assault, homicide (intentional) or traffic crash, workplace accident (unintentional). Moreover, alcohol consumption results in a significant economic burden on society at large. 5.1% of the global burden of disease and injury is attributable to alcohol [Citation28]. As discussed above, alcohol consumption has a severe effect on the health and wellbeing of individuals and populations.

Recently, it has been realized that mathematical models are important to understand the process of drinking. Mulone et al. [Citation24] studied a two-stage model for youths with serious drinking problems and their treatment. The youths with alcohol problems were divided into two component, namely those who admitted to having the problem and those who did not admit. The stability of two steady states was analysed. Xiang et al. [Citation32] developed a drinking model with public health educational campaigns. Mathematical analyses established that the global asymptotically stability of the equilibria were determined by the basic reproduction number R0. If R01, the alcohol-free equilibrium was globally asymptotically stable, and if R0>1, the alcohol present equilibrium was globally asymptotically stable, and they concluded that the public health educational campaigns of drinking individuals could slow down the drinking dynamics. Huo et al. [Citation11] introduced a novel alcoholism model which involved the impact of Twitter. Stability of all the equilibria were obtained in terms of the basic reproductive number R0. Backward and forward bifurcation, Hopf bifurcation were also analysed. For other alcoholism or epidemoc models, we referred to [Citation2, Citation12–15, Citation30, Citation33, Citation34].

The age-structured models have been studied by many authors, such as Webb [Citation31], Iannelli [Citation16] and Cushing [Citation7]. Age-structured epidemic models were known to be effective tools for studying epidemic dynamics [Citation3–5, Citation17, Citation19, Citation35] and many authors regarded alcoholism as a social epidemic disease [Citation11, Citation30]. According to the WHO report, alcohol-related deaths were related to age (see Figure [1]). Therefore, establishment of alcoholism model should consider age factor and the model becomes more reasonable. However, so far, works on alcoholism models with age structure are very scarce.

Figure 1. Proportion of alcohol-attributable deaths (%) of total deaths by age group, 2012.

Figure 1. Proportion of alcohol-attributable deaths (%) of total deaths by age group, 2012.

In 1990, Thieme [Citation26] observed that age-structured models could be regarded as non-densely defined Cauchy problems, subsequently, Magal and Ruan [Citation22], Liu et al. [Citation18] developed center manifold theory and Hopf bifurcation theorem for non-densely defined Cauchy problems, respectively. By using the above theory and theorem, Liu et al. [Citation20] showed that age-structured model of consumer-resource mutualism exhibited Hopf bifurcation at the positive equilibrium under some conditions. Wang and Liu [Citation29] considered an age-structured compartmental pest-pathogen model. Their results showed that Hopf bifurcation occurred at a positive steady state as bifurcating parameter passed values. Tang and Liu [Citation25] studied a new predator-prey model with age structure and exhibited Hopf bifurcation at a positive steady state.

Motivated by above works, the aim of this paper is to investigate the existence of Hopf bifurcation for an alcoholism model with age-structure. Our model consists of three variables : susceptible drinkers at time t are denoted by S(t), who do not drink or drink only moderately; alcoholics at time t with alcoholism age a are denoted by A(t,a), who strongly desire to consume alcohol, having difficulties in controlling of its use, persisting in its use despite harmful consequences; and the people who recover from alcoholism after treatment are denoted by R(t). Alcoholism as a long-standing social epidemic disease, it is difficult to eliminate for a short time. So we put the alcoholism problem into the population growth model to study. So we suppose new recruits enter the population at a rate r(S(t)+0+A(t,a)da+R(t)). The population flow among those classes is shown in the following diagram (Figure ).

Figure 2. The flow diagram of our model, where [A],[βA] represent 0+A(t,a)da, 0+β(a)A(t,a)da, respectively.

Figure 2. The flow diagram of our model, where [A],[βA] represent ∫0+∞A(t,a)da, ∫0+∞β(a)A(t,a)da, respectively.

The flow diagram leads to the following alcoholism model: (1) dS(t)dt=r(S(t)+0+A(t,a)da+R(t))S(t)0+β(a)A(t,a)da(μ+α)S(t),A(t,a)t+A(t,a)a=(μ+a1+δ)A(t,a),dR(t)dt=δ0+A(t,a)da(μ+a2+ρ)R(t),A(t,0)=S(t)0+β(a)A(t,a)da+ρR(t)+αS(t),A(0,a)=A0(a)L+1((0,+),R),S(0)=s0>0,R(0)=r0>0.(1) where alcoholics are assumed to be age-structured, whereas susceptible drinkers and recuperator are not age-structured. r is the birth rate, μ is the natural death rate, a1,a2 are the death rates of excessive drinking, respectively. δ is the transfer rate from alcoholics to recovered individuals, ρ is the relapse rate from recovered individuals to alcoholics. The coefficient α is the fraction of susceptible drinkers S(t) develop into alcoholics because of some of their own reasons, such as losses of earnings, unemployment or family problems, etc. The incidence rate at time t and alcoholism age a is S(t)β(a)A(t,a), where β(a) is the transmission rate due to pressure from alcoholics with alcoholism age a, β(a) is defined by (2) β(a)=β,aτ,0,otherwise,(2) and K0:=0+β(a)e(μ+a1+δ)ada, i.e. β=(μ+a1+δ)K0e(μ+a1+δ)τ, where τ represents the time span that a person from the initial alcoholism to a potential inviter, who invites susceptible drinkers to increase alcohol consumption. The K0 represents the total number of secondary alcoholics produced by a primary alcoholic. For the convenience of computation, we assume that K0 is a constant. The β is the rate that a alcoholic with alcoholism age a (aτ) will successfully infect a susceptible drinker.

The remainder of this paper is organized as follows. Section 2, we summarize the main results on Hopf bifurcation theorem, which obtained in [Citation18]. In Section 3, the stability of the steady state and the existence of Hopf bifurcation are investigated. In Section 4, we perform numerical simulations to verify our analytical results. Finally, a brief conclusion is given.

2. Preliminaries

We recall the Hopf bifurcation theorem in Liu et al. [Citation18] for the following non-densely defined abstract Cauchy problem: (3) du(t)dt=Au(t)+F(μ,u(t)),t0,u(0)D(A)¯,(3) where μR is the bifurcation parameter, A:D(A)XX is a linear operator on a Banach space X with D(A) not dense in X and A not necessary to be a Hille-Yosida operator, F:R×D(A)¯X is a Ck map with k4. Set X0:=D(A)¯, and A0 is the part of A in X0, which is defined by A0x=Ax,xD(A0)={xD(A):AxX0}. We denote by {TA(t)}t0 the C0semigroup of bounded linear operators on X (respectively {SA(t)}t0 the integrated semigroup) generated by A.

Definition 2.1

Let L:D(L)XX be the infinitesimal generator of a linear C0semigroup {TL(t)}t0 on a Banach space X. Define the growth bound ω0(L)(,+) of L by ω0(L):=ln(TL(t)L(X))t. The essential growth bound ω0,ess(L)(,+) of L by ω0,ess(L):=ln(TL(t)ess)t, where L(X) is the space of bounded linear operators from X into X, TL(t)ess is the essential norm of TL(t) defined by TL(t)ess=κ(TL(t)BX(0,1)), where BX(0,1)={xX:xX1}, and each bounded set BX, κ(B)=inf{ε>0:B can be covered by a finite number of balls of radius ε} is the Kuratovsky measure of non-compactness.

We make the following assumptions on the linear operator A and the nonlinear map F.

Assumption 2.1

Assume that A:D(A)XX is a linear operator on a Banach space (X,), there exist two constants ωAR and MA1, such that (ωA,+)ρ(A) and the following properties are satisfied

  1. limλ+(λIA)1x=0, xX;

  2. (λIA)kL(X0)MA/(λωA)k,  λ>ωA, k1.

Assumption 2.1 implies that A0 is the infinitesimal generator of a C0-semigroup {TA0(t)}t0 on X0 and TA0(t)L(X0)MAeωAt,t0. By Proposition 2.6 in Magal and Ruan [Citation22], we know if Assumption 2.1 is satisfied, then A generates a unique integrated semigroup {SA(t)}t0. If we assume in addition that A is a Hille-Yosida operator, then we have SA(t)SA(s)L(X)MAstewAldl, t,s[0,+)with ts0. Next, we consider the non-homogeneous Cauchy problem (4) du(t)dt=Au(t)+f(t),t0,u(0)D(A)¯,(4) where fL1((0,τ),X).

Assumption 2.2

There exists a map δ(t):[0,+)[0,+) with limt0+δ(t)=0, such that for each τ>0 and fC([0,τ],X), t0tSA(ts)f(s)ds is continuously differentiable and ddt0tSA(ts)f(s)dsδ(t)sups[0,t]f(s), t[0,τ].

By Corollary 2.11 in Magal and Ruan [Citation22], we know if Assumption 2.1 and 2.2 hold, then for each τ>0, fC([0,τ],X0), and xX0, (Equation4) has a unique integrated solution u(t)C([0,τ],X0). Meanwhile, if A is a Hille-Yosida operator, then Assumption 2.8 in [Citation23] holds. By Theorem 2.9 in [Citation23], we know Assumption 2.2 holds.

Assumption 2.3

Let ε>0 and FCk((ε,ε)×BX0(0,ε);X) for some k4. Assume that the following conditions are satisfied:

  1. F(μ,0)=0,  μ(ε,ε) and xF(0,0)=0;

  2. (Transversality condition) For each μ(ε,ε), there exists a pair of conjugated simple eigenvalues of (A+xF(μ,0))0, denoted by λ(μ) and λ(μ)¯, such that λ(μ)=α(μ)+iω(μ), the map μλ(μ) is continuously differentiable, ω(0)>0,α(0)=0,dα(0)dμ0, and σ(A0)iR={λ(0),λ(0)¯}.

  3. The essential growth rate of {TA0(t)}t0 is strictly negative, that is, ω0,ess(A0)<0.

Base on the above discussion and assumptions, now we can state the following Hopf bifurcation theorem [Citation18].

Theorem 2.1

Let Assumptions 2.1–2.3 be satisfied. Then there exist ε>0, three Ck1 maps, εμ(ε) from (0,ε) into R, εxε from (0,ε) into D(A)¯, and εγ(ε) from (0,ε) into R, such that for each ε(0,ε) there exists a γ(ε)-periodic function uεCk(R,X0), which is an integrated solution of (2.3) with the parameter value equals μ(ε) and the initial value equals xε. So for each t0, uε satisfies uε(t)=xε+A0tuε(l)dl+0tF(μ(ε),uε(l))dl. Moreover, we have the following properties

(a) There exist a neighborhood N of 0 in X0 and an open interval I in R containing 0, such that for μˆI and any periodic solution μˆ in N with minimal period γˆ close to 2π/ω(0) of (2.3) for the parameter value μˆ, there exists ε(0,ε) such that uˆ(t)=uε(t+θ) (for some θ[0,γ(ε)), μ(ε)=μˆ andγ(ε)=γˆ;

(b) The map εμ(ε) is a Ck1 function and we have the Taylor expansion μ(ε)=n=1[(k2)/2]μ2nε2n+O(εk1), ε(0,ε), where [(k2)/2] is the integer part of (k2)/2.

(c) The period γ(ε) of tuε(t) is a Ck1 function and γ(ε)=2πω(0)1+n=1[(k2)/2]γ2nε2n+O(εk1), ε(0,ε), where ω(0) is the imaginary part of λ(0) defined in Assumption 2.3.

3. Stability of equilibria and existence of Hopf bifurcation

In this section, we will study stability of equilibria and existence of Hopf bifurcation for (Equation1).

3.1. The transformation of the Cauchy problem

In system (Equation1), let S(t)=0+s(t,a)da,R(t)=0+v(t,a)da, we can rewrite system (Equation1) as the following age-structured model (5) x(t,a)t+x(t,a)a=Dx(t,a),x(t,0)=B(x(t,)),x(0,a)=x0L+1((0,+),R3),(5) where x(t,a)=(s(t,a),A(t,a),v(t,a))T, D=diag(m1,m2,m3), m1=μ+α, m2=μ+a1+δ, m3=μ+a2+ρ. B(x(t,))=s(t,0)A(t,0)v(t,0)=r0+s(t,a)da+0+A(t,a)da+0+v(t,a)da0+s(t,a)da0+β(a)A(t,a)da0+s(t,a)da0+β(a)A(t,a)da+ρ0+v(t,a)da+α0+s(t,a)daδ0+s(t,a)da. Consider the Banach space (X,) X=R3×L1((0,+),R3) with αϕ=αR3+ϕL1((0,+),R3). Define the linear operator L:D(L)X by L0ϕ=ϕ(0)ϕDϕ with D(L)={0R3}×W1,1((0,+),R3), we notice L is non-densely defined since X0:=D(L)¯={0R3}×L1((0,+),R3)X. Define the nonlinear operator F:D(L)¯X by F0ϕ=B(ϕ)0. Set w(t)=0x(t,). Now we can rewrite PDEs system (Equation5) as the following non-densely defined abstract Cauchy problem (6) dw(t)dt=Lw(t)+F(w(t)),ω(0)=0x0D(L)¯.(6) The global existence and uniqueness of solutions of system (Equation6) follow from the results of Magal [Citation21] and Magal and Ruan [Citation23].

3.2. Existence of equilibria

If w¯(a)=0x¯(a)X0 is an equilibrium of (Equation6), we have 0x¯(a)D(L)andL0x¯(a)+F0x¯(a)=0, which is equivalent to (7) x¯(a)Dx¯(a)=0,x¯(0)+B(x¯())=0.(7) Hence we obtain (8) x¯(a)=s¯(a)A¯(a)v¯(a)=eDax¯(0)andx¯(0)=r(S¯+A¯+V¯)S¯0+β(a)A¯(a)daS¯0+β(a)A¯(a)da+ρV¯+αS¯δA¯,(8) where S¯=0+s¯(a)da, A¯=0+A¯(a)da, V¯=0+v¯(a)da. From (Equation8) we can get m1S¯=r(S¯+A¯+V¯)K0m2S¯A¯,m2A¯=K0m2S¯A¯+ρV¯+αS¯,m3V¯=δA¯. It is easy to see the system (Equation7) has always trivial equilibrium x¯1(a)=eDa(0,0,0)T, Furthermore, if (H1)r>m1andm2ρδm3>r1+δm3 holds, system (Equation7) has a unique positive equilibrium x¯2(a)=s¯(a)A¯(a)v¯(a)=eDam1S¯m2A¯m3V¯, where S¯=(rm1)(m2ρδm3)+αr(1+δm3)(rμ)K0m2,A¯=(rμ)S¯(m2ρδm3)r(1+δm3),V¯=δm3A¯.

Lemma 3.1

The system (Equation1) has always trivial equilibrium E1(a)=(0,0L1((0,+),R),0)T. If r>m1andm2ρδm3>r1+δm3 hold, there exists a unique positive equilibrium of system (Equation1) E(a)=S¯A¯(a)V¯.

3.3. Linearized equation

Let y(t):=w(t)w¯(a), then system (Equation6) is equivalent to the following system (9) dy(t)dt=Ly(t)+F(y(t)+w¯(a))F(w¯(a)),y(0)=0x0x¯(a)y0X0,(9) and the equilibrium w¯(a) of system (Equation7) is transformed into the zero equilibrium of system (Equation9).

The linearized system of system (Equation9) at the equilibrium 0 is as follows dy(t)dt=Ay(t),y0X0t0, where A=L+DF(w¯). Then system (Equation9) can be written as dy(t)dt=Ay(t)+H(y(t)),y0X0t0, where H(y(t))=F(y(t)+w¯(a))F(w¯(a))DF(w¯)y(t) satisfying H(0)=0, and DH(0)=0.

Denote ξ:=min{m1,m2,m3},andΩ:={λC:Re(λ)>ξ}. By applying the results of Liu, Magal and Ruan [Citation18], we obtain the following result.

Lemma 3.2

For λΩ,λρ(L), we have the following formula (λIL)1αψ=0ϕ where αψX, 0ϕD(L) and ϕ(a)=e(λI+D)aα+0ae(λI+D)(as)ψ(s)ds.

It is readily checked that (λIL)11(Re(λ)+ξ), Re(λ)>ξ, so L is a Hille-Yosida operator and (10) (λIL)n1(Re(λ)+ξ)n, Re(λ)>ξ,n1.(10) Define the part of L in D(L)¯ by L0, L0:D(L0)XX with L0x=Lxfor xD(L0)={xD(L):LxD(L)¯}, and we know D(L0)=0ϕ{0R3}×W1,1((0,+)×,R3):ϕ(0)=0. Then, we can claim that L0 is the infinitesimal generator of a C0-semigroup {TL0(t)}t0 on D(L)¯. and for each t0 the liear operator TL0(t) is defined by TL0(t)0ϕ=0TˆL0(t)ϕ, where (11) TˆL0(t)(ϕ)(a)=eDtϕ(at),if at,0,otherwise.(11) Now we estimate the essential growth bound of the C0semigroup generated by A0 which is the part of A in D(A)¯. We observe that for any 0ϕD(L), DF(w¯)0ϕ=DB(x¯)(ϕ)0 where DB(x¯)(ϕ)=rK0m2A¯rrK0m2A¯+α0ρ0δ00+ϕ(a)da+0S¯00S¯00000+β(a)ϕ(a)da. Then DF(w¯):D(L)¯XX is a compact bounded linear operator. From (Equation10) we obtain TL0(t)eξt. Then we have ω0,ess(L0)ω0(L0)ξ<0. By applying the perturbation results in Ducrot, Liu and Magal [Citation6], we obtain ω0,ess(A0)ξ<0. Thus, by the above discussion and Theorem 3.5.5 in [Citation1], we obtain the following proposition.

Proposition 3.1

The linear operator A is a Hille-Yoside operator, and the essential growth rate of the strongly continuous semigroup generated by A0 is strictly negative, that is, ω0,ess(A0)<0.

In order to apply Theorem 2.1, we remain to precise the spectral properties of A0. Setting C:=DF(w¯), and let λΩ. Since (λIL) is invertible, it follows that λI(L+C) is invertible if and only if IC(λIL)1 is invertible. If IC(λIL)1 is invertible, we obtain (λI(L+C))1=(λIL)1(IC(λIL)1)1. Consider the equation (IC(λIL)1)αϕ=αˆϕˆ, that is αϕC0e(λI+D)aα+0ae(λI+D)(as)ϕ(s)ds=αˆϕˆ. Then, we obtain the system αDB(x¯)e(λI+D)aα+0ae(λI+D)(as)ϕ(s)ds=αˆ,ϕ(a)=ϕˆ(a), this system can be written as αDB(x¯)(e(λI+D)aα)=αˆ+DB(x¯)0ae(λI+D)(as)ϕ(s)ds,ϕ=ϕˆ. From the formula of DB(x¯), we know αDB(x¯)(e(λI+D)aα)=M(λ)α, where M(λ)=IrK0m2A¯rrK0m2A¯+α0ρ0δ00+e(λI+D)ada0S¯00S¯00000+β(a)e(λI+D)ada. Denote S(λ,ϕˆ)=DB(x¯)0ae(λI+D)(as)ϕˆ(s)ds. Then M(λ)α=αˆ+S(λ,ϕˆ). When M(λ) is invertible, we have α=M(λ)1(αˆ+S(λ,ϕˆ)). From the above discussion and by using the proof of Lemma 3.5 in [Citation29], we obtain the following lemma.

Lemma 3.3

The following results hold:

  1. σ(L+C)Ω=σP(L+C)Ω={λΩ:det(M(λ))=0};

  2. If λρ(L+C)Ω, we have the following formula for the resolvent (λI(L+C))1αϕ=0ϕˆ, where ϕˆ(a)=e(λI+D)aM(λ)1(α+S(λ,ϕ))+0ae(λI+D)(as)ϕ(s)ds, and M(λ),S(λ,ϕ) defined as above.

From the above discussion, we know that the linear operator A satisfies Assumptions 2.1–2.3(c) holds.

3.4. Stability of the trivial equilibrium

Now, we consider the stability of the trivial equilibrium E1(a)=(0,0L1((0,+),R),0)T, we obtain M(λ)=Irrrα0ρ0δ00+e(λI+D)ada0000000000+β(a)e(λI+D)ada. Thus, we obtain the characteristic equation det(M(λ))=f0(λ)(λ+m1)(λ+m2)(λ+m3)=0 where f0(λ)=λ3+a0λ2+b0λ+c0, and a0=(m1+m2+m3r),b0=(m1r)(m2+m3)+(m2m3ρδ)rα, c0=(m1r)(m2m3ρδ)(m3+δ)rα.

It is easy to see that {λΩ:detM(λ)=0}={λΩ:f0(λ)=0}. If c0<0 holds, then f0(λ)=0 has at least one root with a real part. Hence, the equilibrium E1(a) is unstable.

If c0>0 holds, then a0=m1+m2+m3r>0,b0>(m1r)(m2+m3)rα>rα(m3+δ)(m2+m3)m2m3ρδ1>0,a0b0c0=(m1+m2+m3r)((m1r)(m2+m3)+(m2m3ρδ)rα)(m1r)(m2m3ρδ)+(m3+δ)rα>(m1+m2+m3r)(m2m3ρδ)(m1r)(m2m3ρδ)+(m3+δ)rα=(m2+m3)(m2m3ρδ)+(m3+δ)rα>0, by the Routh-Hurwitz criterion, all the roots of f0(λ)=0 have negative real parts. Hence, the equilibrium E1(a) is stable.

Theorem 3.4

If c0>0, then E1(a) is locally asymptotically stable for all τ0. If c0<0, then E1(a) is unstable for all τ0.

Remark 3.1

c0>0 means r<μ+α(m3+δ)rα/(m2m3ρδ). α is small, then the birth rate r may be less than the mortality rate μ, the whole population is easily extinct. So in reality, the E1(a) is more likely to be unstable.

3.5. Stability of the positive equilibrium and Hopf bifurcation

When (H1) holds, the characteristic equation of system (Equation1) about the positive equilibrium E(a) can be rewritten as det(M(λ))=λ3+aλ2+bλ+c+(e+fλ+gλ2)eλτ(λ+m1)(λ+m2)(λ+m3)=:f(λ)s(λ)=0, where a=m1+m2+m3r+K0m2A¯,b=ρδ+m1m2+m1m3+m2m3K0m2A¯rr(m2+m3)+K0m2A¯(m2+m3)rα,c=ρδm1+(rK0m2A¯)ρδK0m2A¯rδ+m1m2m3K0m2A¯rm3rm2m3+K0m2A¯m2m3rδαrαm3=(rm1)(m2m3ρδ)+K0m2A¯((m2m3ρδ)r(δ+m3))rα(δ+m3)=(rm1)(m2m3ρδ)+K0m2S¯m3(rμ)rα(δ+m3)=0,e=K0m2S¯(m1rα)m3=K0m2S¯(rμ)m3>0,f=K0m2S¯(m1+m3rα),g=K0m2S¯. Since K0m2S¯=(rm1)(m2ρδm3)+αr(1+δm3)(rμ),K0m2A¯=(rμ)K0m2S¯(m2ρδm3)r(1+δm3), we know the coefficients a,b,c,e,f,g have nothing to do with τ.

It is easy to see that {λΩ:detM(λ)=0}={λΩ:f(λ)=0}. If τ=0, then f(λ)=λ3+(a+g)λ2+(b+f)λ+e=0. Since a+g=m1+m2+m3r+K0m2A¯K0m2S¯>(m1+m2+m3)r+r+r(A¯+V¯)S¯m1m2ρδm3=m3+r(A¯+V¯)S¯+ρδm3>0,b+f=ρδ+m1m2+m1m3+m2m3K0m2A¯rr(m2+m3)+K0m2A¯(m2+m3)rαK0m2S¯(m1+m3rα)=(m2m3ρδ)+m1(m2+m3)K0m2A¯rr(m2+m3)+r+r(A¯+V¯)S¯m1(m2+m3)rα+K0m2S¯(rμm3)>K0m2A¯r+(m2m3ρδ)+r(A¯+V¯)S¯(m2+m3)rα+(rm1)m2ρδm3+rα1+δm3m3m2ρδm3=K0m2A¯r+r(A¯+V¯)S¯(m2+m3)+(rm1)(m2ρδm3)+rαδm3=1S¯(r(ρV¯+αS¯m2A¯)+r(A¯+V¯)(m2+m3))+(rm1)m2ρδm3+rαδm3>0,e>0. By the Routh-Hurwitz criterion, when τ=0 all the roots of f(λ)=0 have negative real parts if and only if (H2)(a+g)(b+f)>e holds.

If τ0, let λ=iω (ω>0) be a purely imaginary roots of f(λ)=0. Then, we have iω3aω2+ibω+eeiωτ+ifωeiωτgω2eiωτ=0. Separating the real part and the imaginary part in the above equation, we can obtain (12) ω3+bω=(egω2)sin(ωτ)fωcos(ωτ),aω2=(gω2e)cos(ωτ)fωsin(ωτ).(12) Thus, we have (13) (ω3+bω)2+(aω2)2=(egω2)2+(fω)2,(13) i.e. (14) ω6+(a22bg2)ω4+(b2f2+2ge)ω2e2=0.(14) Denote z=ω2, (Equation14) becomes (15) z3+(a22bg2)z2+(b2f2+2ge)ze2=0.(15) Let z1,z2 and z3 be three roots of Equation (Equation15). Since a22bg2=(m1+m2+m3r+K0m2A¯)22(ρδ+m1m2+m1m3+m2m3K0m2A¯rr(m2+m3)+K0m2A¯(m2+m3)rα)(K0m2S¯)2=(rm1)2+m22+m32+2ρδ+(K0m2A¯)2+2K0m2A¯m1+2rα(K0m2S¯)2>(rm1)2+m22+m32+2ρδ+(K0m2A¯)2+2K0m2A¯m1+2rαm2ρδm32>(rm1)2+m22+m32+2ρδ+(K0m2A¯)2+2K0m2A¯m1+2rαm22>0, we know a22bg2>0,e2<0, then z1+z2+z3=(a22bg2)<0andz1z2z3=e2>0, it is easy to know that (Equation15) has only one positive real root. We denote this positive real root by z. Then (Equation14) has only one positive real root ω0=z.

Let g(z)=z3+(a22bg2)z2+(b2f2+2ge)ze2, it is easy to know that g(z)|z=z>0, then we have 3z2+(2a24b2g2)z+(b2f2+2eg)|z=z>0, i.e. (16) 3ω4+(2a24b2g2)ω2+(b2f2+2eg)|ω=ω0>0.(16) From (Equation12), we know that f(λ)=0 with τ=τk, has a pair of purely imaginary roots ±iω0, where (17) τk=1ω0(arccos(fag)ω04+(aebf)ω02f2ω02+(gω02e)2+2kπ)=0,η01ω0(arccos(fag)ω04+(aebf)ω02f2ω02+(gω02e)2+2(k+1)π),otherwish,(17) for k=0,1,2, and with η=(gω02e)(ω03bω0)+afω03.

Lemma 3.5

Assume that (H1) be satisfied, then df(λ)dλλ=iω00. Therefore λ=iω0 is a simple root of f(λ)=0.

Proof.

Differentiating the equation f(λ)=0 with respect to λ, we have df(λ)dλλ=iω0=(3λ2+2aλ+b+(2gλ+f)eτλτ(gλ2+fλ+e)eτλ)|λ=iω0 Thus if df(λ)dλλ=iω0=0. Separating the real part and the imaginary part in the above equation, we can obtain (18) 3ω02+b=2gω0sin(ω0τk)fcos(ω0τk)+τk(egω02)cos(ω0τk)+τkfω0sin(ω0τk),2aω0=2gω0cos(ω0τk)+fsin(ω0τk)τk(egω02)sin(ω0τk)+τkfω0cos(ω0τk).(18) Now defining G(ω)=g(ω2)=(ω3bω)2(egω2)2+a2ω4f2ω2, then G(ω0)=0. G(ω)=2(ω3bω)(3ω2b)+2(egω2)(2gω)+4a2ω32f2ω. According to Equations (Equation12), we know (19) ω03+bω0=(egω02)sin(ω0τk)fω0cos(ω0τk),aω02=(gω02e)cos(ω0τk)fω0sin(ω0τk).(19) With the help of Equations (Equation18) and (Equation19) we deduce that G(ω0)=0.

However, G(ω0)=2ω0g(ω02)=2ω0g(z)>0, thus df(λ)dλλ=iω00. This completes the proof.

Lemma 3.6

Assume that (H1) be satisfied, denote the root λ(τ)=α(τ)+iω(τ) of f(λ)=0, satisfying α(τk)=0, ω(τk)=ω0, where τk is defined in (Equation17). Then α(τk)=Redλdττ=τk>0.

Proof.

Taking the derivative of λ with respect to τ in f(λ)=0, it is easy to have (3λ2+2aλ+b)dλdτ+(2gλ+f)eλτdλdτ(gλ2+fλ+e)eλττdλdτ+λ=0,dλdτ1τ=τk=(3λ2+2aλ+b)λ(λ3+aλ2+bλ)+(2gλ+f)(gλ2+fλ+e)λτλτ=τk. By using (Equation13), we obtain Redλdτ1τ=τk=3ω04+(2a24b2g2)ω02+(b2f2+2eg)(fω)02+(egω02)2. By using (Equation16), it is easy to know that Redλdλ1τ=τk>0. Noting that signRedλdττ=τk=signRedλdτ1τ=τk, thus Redλdττ=τk>0. This completes the proof.

Theorem 3.7

Suppose that (H1), (H2) hold, then the following results hold:

  1. If τ[0,τ0) then the positive equilibrium E(a) of system (Equation1) is asymptotically stable.

  2. If τ>τ0, the positive equilibrium E(a) of system (Equation1) is unstable.

By Theorem 2.1, the above results can be summarized as the following Hopf bifurcation theorem for system (Equation1)

Theorem 3.8

If (H1) hold. Then there exist τk>0, k=0,1,2,, where τk is defined in (Equation17), such that age-structured acoholism model (Equation1) undergos a Hopf bifurcation at the positive equilibrium E(a). In particular, when τ=τk, a non-trivial periodic solution bifurcates from the equilibrium E(a).

4. Numerical simulations

In this section, we present some numerical simulation to verify the main results by using MATLAB programming. Some of the parameters are taken from [Citation24] μ=0.25, δ=0.05, ρ=0.8. Other parameters are estimated. First, we estimate other parameters as follows r=0.4, α=0.05, a1=0.5, a2=0.35, K0=0.01, τ=2, we can calculate c0=0.137<0, by Theorems 3.4 we can expect that the trivial equilibrium E1(a)=(0,0L1((0,+),R),0)T is unstable. see Figure . On the other hand, we estimate other parameters r=0.25, α=0.05, a1=0.5, a2=0.35, K0=0.01, τ=2, we can calculate c0=0.0359>0, by Theorems 3.4 we can expect that the trivial equilibrium E1(a)=(0,0L1((0,+),R),0)T is local asymptotically stable. see Figure .

Figure 3. The trajectory of susceptible drinkers S(t), alcoholics 0+A(t,a)da, recuperator R(t) versus time with the initial condition (50,20e0.5a,30). When c0<0 the equilibrium point E1(a) is unstable.

Figure 3. The trajectory of susceptible drinkers S(t), alcoholics ∫0+∞A(t,a)da, recuperator R(t) versus time with the initial condition (50,20e−0.5a,30). When c0<0 the equilibrium point E1(a) is unstable.

Figure 4. The trajectory of susceptible drinkers S(t), alcoholics 0+A(t,a)da, recuperator R(t) versus time with the initial condition (50,20e0.5a,30). When c0>0 the equilibrium point E1(a) is local asymptotically stable.

Figure 4. The trajectory of susceptible drinkers S(t), alcoholics ∫0+∞A(t,a)da, recuperator R(t) versus time with the initial condition (50,20e−0.5a,30). When c0>0 the equilibrium point E1(a) is local asymptotically stable.

Moreover, we estimate other parameters r=0.4, α=0.05, a1=0.5, a2=0.35, K0=0.01, we can calculate (H1) holds, then (Equation1) has only one positive equilibrium E(a)=(81.5476, 27.4e0.8a, 1.2232)T. By algorithms in the previous sections, we obtain ω0=0.1382,τ0=4.9203, and readily checked that the hypothesis of (H2) holds. Thus, E(a) is stable when τ[0,τ0), as depicted in Figure . From the numerical simulations we can see the solution of system (Equation1) with the initial condition (50,20e0.5a,4) and parameter τ=4 asymptotically tends to the positive equilibrium E(a). When τ pass through the critical value τ0, E(a) loss its stability and Hopf bifurcation occurs at τ=τ0=4.9203, as depicted in Figure .

Figure 5. The trajectory of susceptible drinkers and recuperator versus time with the initial condition (50,20e0.5a,4), when τ=4<τ0.

Figure 5. The trajectory of susceptible drinkers and recuperator versus time with the initial condition (50,20e−0.5a,4), when τ=4<τ0.

Figure 6. The surface of alcoholics A(t,a) versus time t and age a with the initial condition (50,20e0.5a,4), when τ=4<τ0.

Figure 6. The surface of alcoholics A(t,a) versus time t and age a with the initial condition (50,20e−0.5a,4), when τ=4<τ0.

Figure 7. The phase portrait of susceptible drinkers S(t), alcoholics 0+A(t,a)da, recuperator R(t) with the initial condition (50,20e0.5a,4), when τ=4<τ0.

Figure 7. The phase portrait of susceptible drinkers S(t), alcoholics ∫0+∞A(t,a)da, recuperator R(t) with the initial condition (50,20e−0.5a,4), when τ=4<τ0.

Figure 8. The trajectory of susceptible drinkers and recuperator versus time with the initial condition (50,20e0.5a,4), when τ=5.2>τ0.

Figure 8. The trajectory of susceptible drinkers and recuperator versus time with the initial condition (50,20e−0.5a,4), when τ=5.2>τ0.

Figure 9. The surface of alcoholics A(t,a) versus time t and age a with the initial condition (50,20e0.5a,5), when τ=5.2>τ0.

Figure 9. The surface of alcoholics A(t,a) versus time t and age a with the initial condition (50,20e−0.5a,5), when τ=5.2>τ0.

Figure 10. The phase portrait of susceptible drinkers S(t), alcoholics 0+A(t,a)da, recuperator R(t) with the initial condition (50,20e0.5a,4), when τ=5.2>τ0.

Figure 10. The phase portrait of susceptible drinkers S(t), alcoholics ∫0+∞A(t,a)da, recuperator R(t) with the initial condition (50,20e−0.5a,4), when τ=5.2>τ0.

5. Conclusion

In this paper, Hopf bifurcation of an age-structured alcoholism model is investigated. Real epidemic data indicate regular periodic fluctuations in disease incidence [Citation8–10] but most models for epidemic diseases predict convergence to a unique stable endemic equilibrium, so it is important to examine under what conditions periodic fluctuations in disease incidence can occur. Because there are people who alcoholism for their own reasons, alcoholism, as a social epidemic disease, is somewhat different from common epidemic diseases. However, in our analysis, we found that regular periodic fluctuations can still arise.

By choosing τ as the bifurcation parameter and analyzing the corresponding characteristic equation, we can conclude that the local asymptotically stability of the trivial equilibrium E1(a) is determined by c0. If c0<0, then the trivial equilibrium E1(a) is unstable for all τ0. If c0>0 the trivial equilibrium E1(a) is local asymptotically stable for all τ0. For the positive equilibrium, if (H1) holds, we establish conditions to ensure the local stability. The parameter τ does not affect the stability of the trivial equilibrium, but can change the stability of the positive equilibrium. By using the center manifold theory [Citation22] and Hopf bifurcation theorem [Citation18], which developed for non-densely defined Cauchy problems, the existence of Hopf bifurcation at the positive equilibrium is obtained. In particular, a non-trivial periodic solution bifurcates from the positive equilibrium when bifurcation parameter τ passes through the critical values τk (k=0,1,2,). Our analytical results indicate that introduction of parameter τ can affect the dynamic behavior of the system (Equation1).

Moreover, due to alcoholism is often associated with some of the alcoholics' own reasons, such as losses of earnings, unemployment or family problems, etc. In this sense, as long as the population is not extinct (according to the analysis of Remark 3.1, the E1(a) is more likely to be unstable), alcoholics will always exist. Therefore, we hope that alcohol-attributable socioeconomic burden is minimal. This optimal control problem, we will study in future work.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This work is supported by the NNSF of China (11861044 and 11661050), the NSF of Gansu Province (148RJZA024), and the HongLiu first-class disciplines Development Program of Lanzhou University of Technology.

References

  • W. Arendt, C. Batty, M. Hieber, and F. Neubrander, Vector-Valued Laplace Transforms and Cauchy Problems, Birkhäser Verlag, Basel, 2001.
  • Y. Cai, J. Jiao, Z. Gui, Y. Liu, and W. Wang, Environmental variability in a stochastic epidemic model, Appl. Math. Comput. 329 (2018), pp. 210–226.
  • Y. Chen, S. Zou, and J. Yang, Global analysis of an SIR epidemic model with infection age and saturated incidence, Nonlinear Anal. 30 (2016), pp. 16–31. doi: 10.1016/j.nonrwa.2015.11.001
  • X. Duan, S. Yuan, and X. Li, Global stability of an SVIR model with age of vaccination, Appl. Math. Comput. 226 (2014), pp. 528–540.
  • X. Duan, S. Yuan, Z. Qiu, and J. Ma, Global stability of an SVEIR epidemic model with ages of vaccination and latency, Comput. Math. Appl. 68(3) (2014), pp. 288–308. doi: 10.1016/j.camwa.2014.06.002
  • A. Ducrot, Z. Liu, and P. Magal, Essential growth rate for bounded linear perturbation of non densely defined cauchy problems, J. Math. Anal. Appl. 341(1) (2008), pp. 501–518. doi: 10.1016/j.jmaa.2007.09.074
  • J. Cushing, An Introduction to Structured Population Dynamics, Society for Industrial and Applied Mathematics, Philadelphia, PA, 1998.
  • P.E.M. Fine and J.A. Clarkson, Measles in england and wales I: An analysis of factors underlying seasonal patterns, Int. J. Epidemiol. 11(1) (1982), pp. 5–14. doi: 10.1093/ije/11.1.5
  • P.E.M. Fine and J.A. Clarkson, Measles in england and wales II. The impact of the measles vaccination programme on the distribution of immunity in the population, Int. J. Epidemiol. 11(1) (1982), pp. 15–25. doi: 10.1093/ije/11.1.15
  • P.E.M. Fine and J.A. Clarkson, Measles in england and wales III. Assessing published predictions of the impact of vaccination of incidence, Int. J. Epidemiol. 12(3) (1983), pp. 332–339. doi: 10.1093/ije/12.3.332
  • H.F. Huo and X.M. Zhang, Complex dynamics in an alcoholism model with the impact of twitter, Math. Biosci. 281 (2016), pp. 24–35. doi: 10.1016/j.mbs.2016.08.009
  • H.F. Huo, R. Chen, and X.Y. Wang, Modelling and stability of hiv/aids epidemic model with treatment, Appl. Math. Model. 40 (2016), pp. 6550–6559. doi: 10.1016/j.apm.2016.01.054
  • H.F. Huo, S.R. Huang, X.Y. Wang, and H. Xiang, Optimal control of a social epidemic model with media coverage, J. Biol. Dyn. 11 (2017), pp. 226–243. doi: 10.1080/17513758.2017.1321792
  • H.F. Huo, Y.L. Chen, and H. Xiang, Stability of a binge drinking model with delay, J. Biol. Dyn. 11 (2017), pp. 210–225. doi: 10.1080/17513758.2017.1301579
  • H.F. Huo, F.F. Cui, and H. Xiang, Dynamics of an saits alcoholism model on unweighted and weighted networks, Phys. A 496 (2018), pp. 249–262. doi: 10.1016/j.physa.2018.01.003
  • M. Iannelli, Mathematical Theory of Age-Structured Population Dynamics, Giadini Editori e Stampatori, Pisa, 1994.
  • X.Z. Li, J. Wang, and M. Ghosh, Stability and bifurcation of an SIVS epidemic model with treatment and age of vaccination, Appl. Math. Model. 34(2) (2010), pp. 437–450. doi: 10.1016/j.apm.2009.06.002
  • Z. Liu, P. Magal, and S. Ruan, Hopf bifurcation for non-densely defined cauchy problems, Z. Angew. Math. Phys. 62(2) (2011), pp. 191–222. doi: 10.1007/s00033-010-0088-x
  • L. Liu, J. Wang, and X. Liu, Global stability of an SEIR epidemic model with age-dependent latency and relapse, Nonlinear Anal. 24 (2015), pp. 18–35. doi: 10.1016/j.nonrwa.2015.01.001
  • Z. Liu, P. Magal, and S. Ruan, Oscillations in age-structured models of consumer-resource mutualisms, Discrete Contin. Dyn. Syst. Ser. B 21(2) (2016), pp. 537–555. doi: 10.3934/dcdsb.2016.21.537
  • P. Magal, Compact attractors for time-periodic age structured population models, Electron. J. Differential Equations 2001(65) (2001), pp. 1–35.
  • P. Magal, S. Ruan. Center manifolds for semilinear equations with non-dense domain and applications on hopf bifurcation in age structured models, Mem. Amer. Math. Soc. 202(2009), pp. 1–67. No. 951.
  • P. Magal and S. Ruan, On semilinear cauchy problems with non-dense domain, Adv. Differential Equations 14(11–12) (2009), pp. 1041–1084.
  • G. Mulone and B. Straughan, Modeling binge drinking, Int. J. Biomath. 5(1) (2012), p. 1250005. doi: 10.1142/S1793524511001453
  • H. Tang and Z. Liu, Hopf bifurcation for a predator-prey model with age structure, Appl. Math. Model. 40(2) (2016), pp. 726–737. doi: 10.1016/j.apm.2015.09.015
  • H.R. Thieme, Semiflows generated by lipschitz perturbations of non-densely defined operators, Differential Integral Equations 3(6) (1990), pp. 1035–1066.
  • G. Thomas and E.M. Lungu, A two-sex model for the influence of heavy alcohol consumption on the spread of HIV/AIDS, Math. Biosci. Eng. 7(4) (2010), pp. 871–904. doi: 10.3934/mbe.2010.7.871
  • W. H. Organization, Global status report on alcohol and health, 2014.
  • Z. Wang and Z. Liu, Hopf bifurcation of an age-structured compartmental pest-pathogen model, J. Math. Anal. Appl. 385(2) (2012), pp. 1134–1150. doi: 10.1016/j.jmaa.2011.07.038
  • X.-Y. Wang, K. Hattaf, H.-F. Huo, and H. Xiang, Stability analysis of a delayed social epidemics model with general contact rate and its optimal control, J. Ind. Manag. Optim. 12(4) (2016), pp. 1267–1285. doi: 10.3934/jimo.2016.12.1267
  • G. Webb, Theory of Nonlinear Age-Dependent Population Dynamics, Marcel Dekker, New York, NY, 1985.
  • H. Xiang, N.N. Song, and H.F. Feng, Modelling effects of public health educational campaigns on drinking dynamics, J. Biol. Dyn. 10(1) (2016), pp. 164–178. doi: 10.1080/17513758.2015.1115562
  • H. Xiang, Y.L. Tang, and H.F. Huo, A viral model with intracellular delay and humoral immunity, Bull. Malays. Math. Sci. Soc. 40 (2017), pp. 1011–1023. doi: 10.1007/s40840-016-0326-2
  • H. Xiang, Y.-P. Liu, and H.-F. Huo, Stability of an SAIRS alcoholism model on scale-free networks, Phys. A 473 (2017), pp. 276–292. doi: 10.1016/j.physa.2017.01.012
  • J. Yang, Y. Chen, and F. Xu, Effect of infection age on an SIS epidemic model on complex networks, J. Math. Biol. 73(5) (2016), pp. 1227–1249. doi: 10.1007/s00285-016-0991-7