1,511
Views
16
CrossRef citations to date
0
Altmetric
Articles

Stability and Hopf bifurcation for a five-dimensional virus infection model with Beddington–DeAngelis incidence and three delays

, &
Pages 146-170 | Received 06 Nov 2016, Accepted 17 Nov 2017, Published online: 03 Dec 2017

ABSTRACT

In this paper, the dynamical behaviours for a five-dimensional virus infection model with three delays which describes the interactions of antibody, cytotoxic T-lymphocyte (CTL) immune responses and Beddington–DeAngelis incidence are investigated. The reproduction numbers for virus infection, antibody immune response, CTL immune response, CTL immune competition and antibody immune competition, respectively, are calculated. By using the Lyapunov functionals and linearization method, the threshold conditions on the local and global stability of the equilibria for infection-free, immune-free, antibody response, CTL response and interior, respectively, are established. The existence of Hopf bifurcation with immune delay as a bifurcation parameter is investigated by using the bifurcation theory. Numerical simulations are presented to justify the analytical results.

1. Introduction

In recent years, the virus infection models provide comprehensive views for our understanding of diseases, such as HIV, influenza, HBV, Ebola, HTLV and HCV (see [Citation1–21,Citation23,Citation24]). Theoretical analysis for these mathematical models are important to obtain complete insights for the viral dynamics in vivo. In particular, the stability and the bifurcation will provide specific information for our understanding about disease control.

The adaptive immune system reacts against virus and infected cells during virus infections. The antibody and cytotoxic T-lymphocyte (CTL) responses play the significant role in preventing infections. Hence, effective strategies to prevent virus infection need both antibody and CTL responses (see [Citation1,Citation14,Citation17,Citation19]). In [Citation17], Wodarz proposed a basic model to describe the interactions of antibody and CTL immune responses with bilinear incidence which includes uninfected target cells x(t), productively infected cells y(t), free virus v(t), CTL response cells z(t) and antibody response cells w(t). Then, Yousfi et al. [Citation20] gave the global analysis for this model. Yan and Wang [Citation19] incorporated an intracellular delay into the infected cell equation in the model and studied the effect of the delay on the global dynamics. However, Wang and Xu [Citation15] suggested that the incidence rate is probably not linear over the large number of virus and susceptible cells. Balasubramaniam et al. [Citation1] developed a HIV model with Beddington–DeAngelis incidence and investigated the global stability and the existence of Hopf bifurcation.

However, only single immune response delay is considered in [Citation1]. We know that there are usually three delays in a virus infection disease with the interactions of antibody and CTL immune responses: the intracellular delay, virus replication delay and immune response delay in the transmission process of virus infection. An important and interesting problem is how dynamical properties in virus infection disease will befallen when three delays exist simultaneously. Particularly, how stability properties will occur at possible equilibrium stations.

Therefore, in this paper we consider a five-dimensional virus infection model with three time delays which describes the interactions of antibody, CTL immune responses and Beddington–DeAngelis incidence rate (1) dx(t)dt=Λdx(t)βx(t)v(t)1+ax(t)+bv(t),dy(t)dt=βemτ1x(tτ1)v(tτ1)1+ax(tτ1)+bv(tτ1)ry(t)py(t)z(t),dv(t)dt=kenτ2y(tτ2)uv(t)qv(t)w(t),dz(t)dt=cy(tτ3)z(tτ3)hz(t),dw(t)dt=gv(t)w(t)αw(t),(1) where Λ, k, c and g are the birth rate of the uninfected cells, the virus, the CTL responses and the antibody responses, respectively; β is the infection rate; d,r,u,h and α represent the death rate of uninfected target cells, productively infected cells, virus, CTL responses and antibody responses, respectively; p represents the killing rate of infected cells by CTL response cells; q is the B cells neutralize rate; τ1 denotes the intracellular delay, and emτ1 denotes the surviving rate of infected cells during delay period [tτ1,t] (see [Citation4,Citation7–10,Citation12,Citation19,Citation24]); τ2 is virus replication delay, and enτ2 denotes the surviving rate of virus during delay period [tτ2,t] (see [Citation5,Citation18]) and τ3 denotes immune response delay which is suggested in [Citation1,Citation8,Citation11,Citation23].

In this paper, our purpose is to investigate the dynamical properties of model (Equation1), expressly the stability of equilibria and the existence of Hopf bifurcation. The organization of our paper is as follows. In Section 2, the basic properties of model (Equation1) for the non-negativity and boundedness of solutions, the threshold values and the existence of five equilibria are discussed. In Section 3, the threshold conditions on the global stability and instability for the infection-free equilibrium, infection equilibrium without immune response and infection equilibrium with only antibody response are stated. When τ3=0, the threshold conditions on the global stability and instability for the infection equilibrium with only CTL response and infection equilibrium with both CTL and antibody responses are proved. In Section 4, when τ3>0, the sufficient conditions on the existence of Hopf bifurcation for the infection equilibrium with only CTL response and infection equilibrium with both CTL and antibody responses are established. In Section 5, the numerical simulations are presented to further illustrate the dynamical behaviour of the model. Finally, we will give a conclusion.

2. Boundedness and equilibrium

Let τ=max{τ1,τ2,τ3} and R+5={(x1,x2,x3,x4,x5):xi0,i=1,2,3,4,5}. C([τ,0],R+5) denotes the space of continuous functions mapping interval [τ,0] into R+5 with norm φ=supτt0{|φ(t)|} for any φC([τ,0],R+5).

The initial conditions for model (Equation1) are given as follows (2) (x(θ),y(θ),v(θ),z(θ),w(θ))=(φ1(θ),φ2(θ),φ3(θ),φ4(θ),φ5(θ)),φi(θ)0, θ[τ,0), φi(0)>0 (i=1,2,3,4,5),(2) where (φ1(θ),φ2(θ),φ3(θ),φ4(θ),φ5(θ))C([τ,0],R+5). By the fundamental theory of functional differential equation [Citation6], It is easy to see that model (Equation1) admits a unique solution (x(t),y(t),v(t),z(t),w(t)) satisfying initial conditions (Equation2). We have the following basic result of model (Equation1).

Theorem 2.1

Let (x(t),y(t),v(t),z(t),w(t)) be the solution of model (Equation1) satisfying initial conditions (Equation2), then x(t),y(t),v(t),z(t) and w(t) are positive and ultimately bounded.

Proof.

It is now easy to show that all solutions of model (Equation1) with initial conditions (Equation2) are defined on R+5 and remain positive for all t0. Denote N(t)=emτ1x(tτ1)+y(t)+renτ22kv(t+τ2)+pcz(t+τ3)+rqenτ22kgw(t+τ2). Calculating the derivative of N(t) along solutions of model (Equation1), we have N˙(t)=Λemτ1demτ1x(tτ1)r2y(t)ruenτ22kv(t+τ2)phcz(t+τ3)αrqenτ22kgw(t+τ2)Λemτ1sN(t), where s=min{d,r/2,u,h,α}. This implies that N(t) is ultimately bounded for large t. So, x(t),y(t),v(t),z(t) and w(t) are also ultimately bounded.

Next, we discuss the existence of equilibria of model (Equation1). Firstly, we directly obtain that model (Equation1) always has a unique infection-free equilibrium E0=(x0,0,0,0,0) with x0=Λ/d.

The basic reproductive number of viral infection for model (Equation1) is R0=k1uenτ2βΛd(1+aΛd)emτ11r. Here, k is the rate of the newly virus produced by infected cells, 1/u is the surviving period of virus, enτ2 is the surviving rate of newly virus in [tτ2,t], β(Λ/d)/(1+a(Λ/d)) denotes the newly infected cells which are infected by the first virus, emτ1 is the the surviving rate of newly infected cells in [tτ1,t], and 1/r is the surviving period of infected cells. Therefore, we easily see that R0 denotes the average number of the free virus released by the infected cells which are infected by the first virus.

Obviously, R0>1 implies that kβΛemτ1nτ2ur(d+aΛ)>0 and k(β+bd)auremτ1+nτ2>0.

When R0>1, model (Equation1) has a unique immune-free equilibrium E1=(x1,y1,v1, 0,0), where x1=Λkb+uremτ1+nτ2k(β+bd)auremτ1+nτ2,y1=kβΛemτ1urenτ2(d+aΛ)r[k(β+bd)auremτ1+nτ2],v1=Λk2βemτ1nτ2kur(aΛ+d)ur[k(β+bd)auremτ1+nτ2].

The antibody immune reproductive number for model (Equation1) is R1=gΛk2βemτ1nτ2kur(aΛ+d)ur[k(β+bd)auremτ1+nτ2]1α. Note that when R0>1 model (Equation1) has a unique immune-free equilibrium E1=(x1,y1,v1, 0,0). This shows that virus infection is successful and the number of free virus at equilibrium E1 is Λk2βemτ1nτ2kur(aΛ+d)/ur[k(β+bd)auremτ1+nτ2]. Furthermore, we have that 1/α is the average life-span of antibody cells, g is birth rate of the antibody response. Hence, R1 denotes the average number of the antibody immune cells activated by virus when virus infection is successful and CTL responses have not been established.

The CTL immune reproductive number for model (Equation1) is R2=ckβΛemτ1urenτ2(d+aΛ)r[k(β+bd)auremτ1+nτ2]1h. Here, R2 denotes the average number of the CTL immune cells activated by infected cells when virus infection is successful and antibody immune responses have not been established. Note that the number of infected cells at equilibrium E1 is kβΛemτ1urenτ2(d+aΛ)/r[k(β+bd)auremτ1+nτ2], 1/h is the average life-span of CTL cells and c is the rate at which the CTL responses are produced.

We see that R1>1 is equivalent to αgv1<0, and R2>1 is equivalent to hcy1<0.

When R1>1, model (Equation1) has a unique infection equilibrium with only antibody response E2=(x2,y2,v2,0,w2), where y2=(Λdx2)emτ1r,v2=αg,w2=kgemτ1nτ2(Λdx2)αurqαr and x2 is the unique positive zero point of the following function L(x)=adx2+d1+αbg+αβgaΛxΛ1+αbg.

In fact, from kΛemτ1nτ2/ur>v1, by R1>1 we obtain kΛgαuremτ1+nτ2>0. From the expression of w2 it follows that the existence of equilibrium E2 is equivalent to x2(0,kgΛαuremτ1+nτ2/kgd). Noticing that L(x) is a quadratic function and L(0)<0, we know that the existence and uniqueness of equilibrium E2 is equivalent to LkgΛαuremτ1+nτ2kgd>0. Since LkgΛαuremτ1+nτ2kgd=aruαemτ1+nτ2(Λkgαuremτ1+nτ2)k2g2d+kαβ(Λkgαuremτ1+nτ2)(g+αb)emτ1+nτ2αurkdk2g2d, from R1>1, we have k2βΛgemτ1nτ2>urkg(d+aΛ)+urαk(β+bd)aurαuremτ1+nτ2. Therefore, when R1>1, we get LkgΛαuremτ1+nτ2kgd>0.

When R2>1, model (Equation1) has a unique infection equilibrium with only CTL response E3=(x3,y3,v3,z3,0), where y3=hc,v3=khenτ2uc,z3=cemτ1(Λdx3)rhph and x3 is the unique positive zero point of the following function L(x)=adx2+d1+kbhenτ2uc+kβhenτ2ucaΛxΛ1+kbhenτ2uc.

In fact, since Λemτ1/r>y1, by R2>1 we obtain cΛrhemτ1>0. From the expression of z3 it follows that the existence of CTL-present infection equilibrium E3 is equivalent to x3(0,cΛrhemτ1/cd). Noticing that L(x) is a quadratic function and L(0)<0, we know that the existence and uniqueness of CTL-present equilibrium E3 is equivalent to LcΛrhemτ1cd>0. Since LcΛrhemτ1cd=arhuemτ1(Λcrhemτ1)+khβ(Λchremτ1)enτ2c2du(uc+kbhenτ2)rhdemτ1c2du,

from R2>1, we have Λkβcemτ1nτ2>ruc(d+aΛ)+hrk(β+bd)enτ2ahur2emτ1. Therefore, when R2>1, we get LcΛrhemτ1cd>0.

The CTL immune competitive reproductive number for model (Equation1) is R3=c(Λdx2)emτ1r1h. In fact, when R1>1, model (Equation1) has a unique infection equilibrium with only antibody response E2=(x2,y2,v2,0,w2). This predicates that CTL immune responses have been established, and the number of infected cells at equilibrium E2 is (Λdx2)emτ1/r. Hence, R3 denotes the average number of the CTL immune cells activated by infected cells under the condition that antibody immune responses have been established.

The antibody immune competitive reproductive number for model (Equation1) is R4=gkhenτ2uc1α. In fact, when R2>1, model (Equation1) has a unique infection equilibrium with only CTL response E3=(x3,y3,v3,z3,0). This predicates that antibody immune responses have been established, and the number of the virus at equilibrium E3 is khenτ2/uc. Hence, R4 denotes the average number of the antibody immune cells activated by viruses under the condition that CTL immune responses have been established.

When R3>1 and R4>1, model (Equation1) has a unique infection equilibrium with CTL and antibody responses E4=(x4,y4,v4,z4,w4), where y4=hc,v4=αg,z4=cemτ1(Λdx4)rhph,w4=khgenτ2αucαqc.

In fact, from the above discussion on the existence of equilibrium E2, we directly have v4=v2 and x4=x2. From R3>1, we have y2>h/c. From the expression of y2, we further have x4=x2<(cΛrhemτ1)/cd. Hence, crmτ1(Λdx4)rh>0, which implies z4>0. Furthermore, from R4>1 we also have w4>0. This shows that equilibrium E4 uniquely exists.

3. Stability analysis

Theorem 3.1

  1. If R01, then the infection-free equilibrium E0 is globally asymptotically stable.

  2. If R0>1, then the equilibrium E0 is unstable.

Proof.

Consider conclusion (a). Define a Lyapunov functional V1(t)=x01+ax0x(t)x01lnx(t)x0+emτ1y(t)+remτ1+nτ2kv(t)+pemτ1cz(t)+rqemτ1+nτ2kgw(t)+tτ1tβv(θ)x(θ)1+ax(θ)+bv(θ)dθ+pemτ1tτ3ty(θ)z(θ)dθ+remτ1tτ2ty(θ)dθ. Calculating the derivative of V1(t) along any positive solution of model (Equation1) and noting that x0=Λ/d, we can obtain dV1(t)dt=d(x(t)x0)2x(t)(1+ax0)+uremτ1+nτ2v(t)(1+ax(t))k(1+ax(t)+bv(t))(R01)buremτ1+nτ2v2(t)k(1+ax(t)+bv(t))phemτ1cz(t)rqαemτ1+nτ2kgw(t). If R01, then dV1(t)/dt0 for any (x(t),y(t),v(t),z(t),w(t)). We have dV1(t)/dt=0 if and only if x=x0, v=0, z=0 and w=0. From the LaSalle's invariance principle [Citation6], we have that E0 is globally asymptotically stable when R01.

Next, we consider conclusion (b). The characteristic equation of the linearized system of model (Equation1) at the equilibrium E0 is (s+h)(s+α)(s+d)f(s)=0, where f(s)=s2+(r+u)s+rukβΛd+aΛe(m+s)τ1(n+s)τ2. If R0>1, we have f(0)=ru(kβΛ/(d+aΛ))emτ1nτ2<0 and lims+f(s)=+. Hence, there is at least a positive s such that f(s)=0. Therefore, when R0>1, E0 is unstable. This completes the proof.

Remark 3.2

Theorem 3.1 shows that if only infection-free equilibrium E0 exists, then it is globally asymptotically stable, and delays τ1, τ2 and τ3 do not impact the stability of E0. Biologically, we see that in this case the virus is cleared up.

Theorem 3.3

Let R0>1.

  1. If R11 and R21, then the immune-free equilibrium E1 is globally asymptotically stable.

  2. If R1>1 or R2>1, then the equilibrium E1 is unstable.

Proof.

Let H(ξ)=ξ1lnξ, we have H(ξ)0 for all ξ>0 and H(ξ)=0 if and only if ξ=1. Consider conclusion (a). Define a Lyapunov functional V2(t)=emτ1(x(t)x1x1x(t)1+aθ+bv11+ax1+bv1x1θdθ)+y1Hy(t)y1+pcz(t)+renτ2v1kHv(t)v1+ry10τ1Hemτ1βx(tθ1)v(tθ1)ry1(1+ax(tθ1)+bv(tθ1))dθ1+tτ2tHy(θ)y1dθ+ptτ3ty(s)z(s)ds+rqenτ2kgw(t). Calculating the derivative of V2(t) along the solution of model (Equation1) gives dV2(t)dt=demτ1(x(t)x1)2(1+bv1)x(t)(1+ax1+bv1)+qrenτ2kv1αgw(t)+ry1lny(tτ2)y(t)+ry1lnx(tτ1)v(tτ1)1+ax(tτ1)+bv(tτ1)1+ax(t)+bv(t)x(t)v(t)+ry13x1(1+ax(t)+bv1)x(t)(1+ax1+bv1)y(tτ2)v1y1v(t)1+ax1+bv1x1v1y1y(t)x(tτ1)v(tτ1)1+ax(tτ1)+bv(tτ1)+ry1v(t)v11+ax(t)+bv11+ax(t)+bv(t)v(t)v1+pz(t)y1hc=demτ1(x(t)x1)2(1+bv1)x(t)(1+ax1+bv1)ry1Hx1(1+ax(t)+bv1)x(t)(1+ax1+bv1)+Hy(tτ2)v1y1v(t)+H1+ax1+bv1x1v1y1y(t)x(tτ1)v(tτ1)1+ax(tτ1)+bv(tτ1)+H1+ax(t)+bv(t)1+ax(t)+bv1+ry11+1+ax(t)+bv(t)1+ax(t)+bv1v(t)v1+v(t)v11+ax(t)+bv11+ax(t)+bv(t)+pcz(t)(cy1h)+qrenτ2kv1αgw(t). Notice that ry11+1+ax(t)+bv(t)1+ax(t)+bv1v(t)v1+v(t)v11+ax(t)+bv11+ax(t)+bv(t)=bry1(1+ax(t))(v(t)v1)2v1(1+ax(t)+bv1)(1+ax(t)+bv(t)). Hence, dV2(t)/dt0 and dV2(t)/dt=0 if and only if x(t)=x1,y(t)=y1,v(t)=v1, z(t)=0 and w(t)=0. From the LaSalle's invariance principle [Citation6], we have that E1 is globally asymptotically stable when R0>1, R11 and R21.

Next, we consider conclusion (b). The characteristic equation of the linearized system of model (Equation1) at the equilibrium E1 is (s+αgv1)f1(s)f2(s)=0, where f1(s)=s+hcy1esτ3,f2(s)=s+d+βv1(1+bv1)(1+ax1+bv1)20βx1(1+ax1)(1+ax1+bv1)2e(m+s)τ1βv1(1+bv1)(1+ax1+bv1)2s+re(m+s)τ1βx1(1+ax1)(1+ax1+bv1)20ke(n+s)τ2s+u. When R1>1, we have αgv1<0. Hence, there is a positive root s=gv1α. When R2>1, we have f1(0)=hcy1<0 and lims+f1(s)=+. Hence, there is at least a positive root s such that f1(s)=0. Therefore, when R1>1 or R2>1, E1 is unstable. This completes the proof.

Remark 3.4

Theorem 3.3 shows that delays τ1, τ2 and τ3 do not impact the stability of E1. Biologically, we see that when R0>1, R11 and R21 then the establishments of both CTLs and antibody immune responses are unsuccessful.

Theorem 3.5

Let R0>1 and R1>1.

  1. If R31, then the infection equilibrium E2 with only antibody response is globally asymptotically stable;

  2. If R3>1, then the equilibrium E2 is unstable.

Proof.

Consider conclusion (a). Define a Lyapunov functional V3(t)=emτ1x(t)x2x2x(t)1+aθ+bv21+ax2+bv2x2θdθ+rqenτ2w2kgHw(t)w2+y2Hy(t)y2+rv2enτ2kHv(t)v2+pcz(t)+ptτ3ty(s)z(s)ds+ry20τ1H(emτ1βx(tθ1)v(tθ1)ry1(1+ax(tθ1)+bv(tθ1)))dθ1+tτ2tHy(θ)y2dθ. Calculating the derivative of V3(t) along the solution of model (Equation1), it follows that dV3(t)dt=demτ1(x(t)x2)2(1+bv2)x(t)(1+ax2+bv2)ry2Hx2(1+ax(t)+bv2)x(t)(1+ax2+bv2)+Hy(tτ2)v2y2v(t)+H1+ax2+bv2x2v2y2y(t)x(tτ1)v(tτ1)1+ax(tτ1)+bv(tτ1)+H1+ax(t)+bv(t)1+ax(t)+bv2+ry21+1+ax(t)+bv(t)1+ax(t)+bv2v(t)v2+v(t)v21+ax(t)+bv21+ax(t)+bv(t)+pcz(t)(cy2h). Notice that ry21+1+ax(t)+bv(t)1+ax(t)+bv2v(t)v2+v(t)v21+ax(t)+bv21+ax(t)+bv(t)=bry2(1+ax(t))(v(t)v2)2v2(1+ax(t)+bv2)(1+ax(t)+bv(t)). Hence, dV3(t)/dt0 and dV3(t)/dt=0 if and only if x(t)=x2,y(t)=y2,v(t)=v2 and z(t)=0. From the LaSalle's invariance principle [Citation6], we have that E2 is globally asymptotically stable when R0>1, R1>1 and R31.

Next, we consider conclusion (b). The characteristic equation of the linearized system of model (Equation1) at the equilibrium E2 is f1(s)f2(s)=0, where f1(s)=s+hcy2esτ3,f2(s)=a110a130a21s+ra2300ke(n+s)τ2s+u+qw2qv200gw2s+αgv2, where a11=s+d+βv2(1+bv2)(1+ax2+bv2)2,a13=βx2(1+ax2)(1+ax2+bv2)2,a21=e(m+s)τ1βv2(1+bv2)(1+ax2+bv2)2,a23=e(m+s)τ1βx2(1+ax2)(1+ax2+bv2)2. If R3>1, then we have f1(0)=hcy2<0 and lims+f1(s)=+. Hence, there is at least a positive root s such that f1(s)=0. Therefore, when R3>1, E2 is unstable. This completes the proof.

Remark 3.6

From Theorem 3.5 we see that delays τ1, τ2 and τ3 do not impact the stability of E2. Biologically, Theorem 3.5 implies that when R0>1, R1>1 and R31, the antibody immune response can be established, but the infected cells are too weak so that it can not stimulate CTL immune response.

Theorem 3.7

Let R0>1 and R2>1.

  1. If R41 and τ3=0, then the infection equilibrium E3 with only CTL response is globally asymptotically stable;

  2. If R4>1, then the equilibrium E3 is unstable.

Proof.

Consider conclusion (a). Define a Lyapunov functional V4(t)=emτ1x(t)x3x3x(t)1+aθ+bv31+ax3+bv3x3θdθ+y3Hy(t)y3+(r+pz3)enτ2v3kHv(t)v3+pz3cHz(t)z3+q(r+pz3)enτ2kgw(t)+(r+pz3)y30τ1Hemτ1βx(tθ1)v(tθ1)(r+pz3)y3(1+ax(tθ1)+bv(tθ1))dθ1+tτ2tHy(θ)y3dθ+ptτ3ty(s)z(s)ds. Calculating the derivative of V4(t) along the solution of model (Equation1), we have dV4(t)dt=demτ1(x(t)x3)2(1+bv3)x(t)(1+ax3+bv3)(r+pz3)y3Hx3(1+ax(t)+bv3)x(t)(1+ax3+bv3)+Hy(tτ2)v3y3v(t)+H1+ax3+bv3x3v3y3y(t)x(tτ1)v(tτ1)1+ax(tτ1)+bv(tτ1)+H1+ax(t)+bv(t)1+ax(t)+bv3+(r+pz3)y31+1+ax(t)+bv(t)1+ax(t)+bv3v(t)v3+v(t)v31+ax(t)+bv31+ax(t)+bv(t)+q(r+pz3)enτ2kv3αgw(t). Notice that (r+pz3)y31+1+ax(t)+bv(t)1+ax(t)+bv3v(t)v3+v(t)v31+ax(t)+bv31+ax(t)+bv(t)=b(r+pz3)y3(1+ax(t))(v(t)v3)2v3(1+ax(t)+bv3)(1+ax(t)+bv(t)). Hence, dV4(t)/dt0 and dV4(t)/dt=0 if and only if x(t)=x3,y(t)=y3,v(t)=v3 and w(t)=0. From the LaSalle's invariance principle [Citation6], we have that the equilibrium E3 is globally asymptotically stable when τ3=0, R0>1, R2>1 and R41.

Next, we consider conclusion (b). The characteristic equation of the linearization system of model (Equation1) at the equilibrium E3 is (s+αgv3)f(s)=0, where f(s)=a110a130a21s+r+pz3a23py30ke(n+s)τ2s+u00cesτ3z30s+hcy3esτ3, where a11=s+d+βv3(1+bv3)(1+ax3+bv3)2,a13=βx3(1+ax3)(1+ax3+bv3)2,a21=e(m+s)τ1βv3(1+bv3)(1+ax3+bv3)2,a23=e(m+s)τ1βx3(1+ax3)(1+ax3+bv3)2. If R4>1, then we have a positive root s=gv3α. Therefore, when R4>1, E3 is unstable for any τ10, τ20 and τ3=0. This completes the proof.

Remark 3.8

Theorem 3.7 shows that delays τ1 and τ2 do not impact the stability of E3. Biologically, we see that, when τ3=0, for any τ10 and τ20 as long as R0>1, R2>1, R41 then the CTL immune response can be determined, but the virus loads are so small that it can not activate the antibody immune responses.

Theorem 3.9

If τ3=0, R0>1, R1>1, R3>1 and R4>1, then the infection equilibrium E4 with CTL and antibody responses is globally asymptotically stable.

Proof.

Define a Lyapunov functional V5(t)=emτ1(x(t)x4x4x(t)1+aθ+bv41+ax4+bv4x4θdθ)+y4Hy(t)y4+(r+pz4)enτ2v4kHv(t)v4+q(r+pz4)enτ2w4kgHw(t)w4+pz4cHz(t)z4+(r+pz4)y40τ1Hemτ1βx(tθ1)v(tθ1)(r+pz2)y2(1+ax(tθ1)+bv(tθ1))dθ1+tτ2tHy(θ)y4dθ+ptτ3ty(s)z(s)ds. Calculating the derivative of V5(t) along the solution of model (Equation1), it follows that dV4(t)dt=demτ1(x(t)x4)2(1+bv4)x(t)(1+ax4+bv4)(r+pz4)y4Hx4(1+ax(t)+bv4)x(t)(1+ax4+bv4)+Hy(tτ2)v4y4v(t)+H1+ax4+bv4x4v4y4y(t)x(tτ1)v(tτ1)1+ax(tτ1)+bv(tτ1)+H1+ax(t)+bv(t)1+ax(t)+bv4+(r+pz4)y41+1+ax(t)+bv(t)1+ax(t)+bv4v(t)v4+v(t)v41+ax(t)+bv41+ax(t)+bv(t). Notice that (r+pz4)y41+1+ax(t)+bv(t)1+ax(t)+bv4v(t)v4+v(t)v41+ax(t)+bv41+ax(t)+bv(t)=b(r+pz4)y4(1+ax(t))(v(t)v4)2v4(1+ax(t)+bv4)(1+ax(t)+bv(t)). Hence, dV5(t)/dt0 and dV5(t)/dt=0 if and only if x(t)=x4, y(t)=y4, v(t)=v4. From the LaSalle's invariance principle [Citation6], we finally have that the equilibrium E4 is globally asymptotically stable when τ3=0, R0>1, R1>1, R3>1 and R4>1. This completes the proof.

Remark 3.10

From Theorem 3.9 we see that delays τ1 and τ2 do not impact the stability of E4 when delay τ3=0. Biologically, Theorem 3.9 implies that, when τ3=0, for any τ10 and τ20 as long as R0>1, R1>1, R3>1 and R4>1 then the susceptible cells, infected cells, free virus, CTL immune responses and antibody immune responses can coexist in vivo.

4. Hopf bifurcation analysis

We first discuss Hopf bifurcation at the equilibrium E3. By Theorem 3.7, we obtain the globally asymptotically stability of the equilibrium E3 when τ10, τ20 and τ3=0. However, what kind of complicated dynamic behaviour will appear at the equilibrium E3 when τ3>0? By computing the characteristic equation for the corresponding linearized system of model (Equation1) at the equilibrium E3 is given by (3) s5+m1~s4+m2~s3+m3~s2+m4~s+m5~+(n1~s4+n2~s3+n3~s2+n4~s+n5~)esτ3+(p1~s3+p2~s2+p3~s+p4~)es(τ1+τ2)+(q1~s2+q2~s+q3~)es(τ1+τ2+τ3)=0,(3) where m1~=αgv3+d+h+u+r+pz3+βv3(1+bv3)(1+ax3+bv3)2,m2~=h(u+r+pz3)+(αgv3)d+h+u+r+pz3+βv3(1+bv3)(1+ax3+bv3)2+(r+pz3)u+(h+u+r+pz3)d+βv3(1+bv3)(1+ax3+bv3)2,m3~=(αgv3)[h(u+r+pz3)+(h+u+r+pz3)d+βv3(1+bv3)(1+ax3+bv3)2+(r+pz3)u]+hu(r+pz3)+d+βv3(1+bv3)(1+ax3+bv3)2[(r+pz3)(u+h)+hu], m4~=(αgv3)hu(r+pz3)+d+βv3(1+bv3)(1+ax3+bv3)2((r+pz3)(u+h)+hu)+hu(r+pz3)d+βv3(1+bv3)(1+ax3+bv3)2,m5~=αgv3)hu(r+pz3)d+βv3(1+bv3)(1+ax3+bv3)2,n1~=h,n2~=hαgv3+u+r+d+βv3(1+bv3)(1+ax3+bv3)2, n3~=h(αgv3)u+r+d+βv3(1+bv3)(1+ax3+bv3)2+ur+d+βv3(1+bv3)(1+ax3+bv3)2(u+r),n4~=h(αgv3)ur+d+βv3(1+bv3)(1+ax3+bv3)2(u+r)+urd+βv3(1+bv3)(1+ax3+bv3)2,n5~=hur(αgv3)d+βv3(1+bv3)(1+ax3+bv3)2,p1~=kβx3(1+ax3)emτ1nτ2(1+ax3+bv3)2, p2~=kβx3(1+ax3)emτ1nτ2(1+ax3+bv3)2(αgv3+h+d),p3~=kβx3(1+ax3)emτ1nτ2(1+ax3+bv3)2[hd+(h+d)(αgv3)],p4~=kβx3(1+ax3)hd(αgv3)emτ1nτ2(1+ax3+bv3)2,q1~=kβx3h(1+ax3)emτ1nτ2(1+ax3+bv3)2,q2~=kβx3h(1+ax3)emτ1nτ2(1+ax3+bv3)2(αgv3+d),q3~=kβx3hd(αgv3)(1+ax3)emτ1nτ2(1+ax3+bv3)2. However, when τ1>0 or τ2>0 Equation (Equation3) is too complicated so that it can not make a good conclusion. Therefore, in the following discussions, we assume τ1=τ2=0. For Equation (Equation3), simple manipulation leads to (4) s5+m1s4+m2s3+m3s2+m4s+m5+(n1s4+n2s3+n3s2+n4s+n5)esτ3=0,(4) where m1=m1~,m2=m2~+p1~,m3=m3~+p2~,m4=m4~+p3~,m5=m5~+p4~,n1=n1~,n2=n2~,n3=n3~+q1~,n4=n4~+q2~,n5=n5~+q3~. Let s=iω(ω>0) be a purely imaginary root of (Equation4). Separating real and imaginary parts, it follows that (5) m1ω4m3ω2+m5=(n1ω4n3ω2+n5)cosωτ3(n2ω3+n4ω)sinωτ3,ω5m2ω3+m4ω=(n1ω4n3ω2+n5)sinωτ3(n2ω3+qn4ω)cosωτ3.(5) Squaring and adding the two equations of (Equation5), it follows that (6) ω10+a0ω8+b0ω6+c0ω4+d0ω2+e0=0,(6) where a0=m122m2n12,b0=m22+2m42m1m3n22+2n1n3,c0=m32+2m1m52m4m22n1n5+2n4n2n32,d0=2m3m5+m42+2n3n5n42,e0=m52n52. Let λ=ω2, then Equation (Equation6) becomes (7) H(λ)=λ5+a0λ4+b0λ3+c0λ2+d0λ+e0=0.(7) Denote a1=625a02+35b0,b1=625a0b0+25c0+8125a03,c1=3625a04+3125a02b0225a0c0+15d0,Δ0=a124c1,a2=13a124c1,b2=227a13+83a1c1b12,Δ1=127a23+14b22,d0=b22+Δ13+b22Δ13+a13,Δ2=d0a1+2b1d0a1,Δ3=d0a12b1d0a1. Applying the results given in [Citation22] on the distribution of roots for five degree polynomial equation, we have the following results.

Lemma 4.1

For the polynomial equation (Equation7), the following results are true:

  1. If e0<0, then Equation (Equation7) has at least one positive root;

  2. Assume that e00 and b1=0,

    1. If Δ0<0, then Equation (Equation7) has no positive real root;

    2. If Δ00, a10 and c1>0, then Equation (Equation7) has no positive real root;

    3. If (a) and (b) are not satisfied, then Equation (Equation7) has positive real root if and only if there exists at least one λ{λ1,λ2,λ3,λ4} such that λ>0 and H(λ)0, where λi=δia0/5, i=1,2,3,4 and δ1=a1+Δ02,δ2=a1+Δ02,δ3=a1Δ02,δ4=a1Δ02.

    4. Assume that e00, b10 and d0>b1,

      1. If Δ2<0 and Δ3<0, then Equation (Equation7) has no positive real root;

      2. If (a) is not satisfied, then Equation (Equation7) has positive real root if and only if there exists at least one λ{λ1,λ2,λ3,λ4} such that λ>0 and H(λ)0, where λi=δia0/5, i=1,2,3,4 and δ1=d0a1+Δ22,δ2=d0a1Δ22,δ3=d0a1+Δ32,δ4=d0a1Δ32.

    5. Assume that e00 and b10 and d0<b1, then Equation (Equation7) has positive real root if and only if b12/4(a1d0)2+12d0=0 and λ¯>0 and H(λ¯)0, where λ¯=b1/2(a1d0)15a0.

Without loss of generality, we assume that Equation (Equation7) has m positive roots with m{1,2,3,4,5}, denoted by λk, k=1,2,m. Then Equation (Equation6) has m positive roots, say ωk=λk,k=1,2,m.

From Equation (Equation5) we have (8) sinωτ3=(ω5m2ω3+m4ω)(n1ω4n3ω2+n5)(n1ω4n3ω2+n5)2+(n2ω3n4ω)2(m1ω4m3ω2+m5)(n2ω3+n4ω)(n1ω4n3ω2+n5)2+(n2ω3n4ω)2γ(ω).(8) Let ω=ωk (k=1,2,,m), we solve τ3 from Equation (Equation8) to obtain that τk(j)=1ωkarcsinγ(ωk)+2πjωk, where k=1,2,,m, j=0,1,. Therefore, when τ=τk(j), k=1,2,,m. j=0,1,, ±iωk is a pair of purely imaginary roots of Equation (Equation4). Clearly, for every k=1,2,,m, {τk(j)} is monotonically increasing for j=0,1,2, and limj+τk(j)=. Therefore, there is a k0{1,2,,m} and j0{0,1,2,} such that τk0(j0)=min{τk(j):k=1,2,,m, j=0,1,2,}. Define (9) τ0=τk0(j0),ω0=ωk0,λ0=λk0.(9)

Let s(τ3)=ξ(τ3)+iω(τ3) be a root of Equation (Equation4) satisfying ξ(τ0)=0,ω(τ0)=ω0,λ0=ω02. Differentiating Equation (Equation4) with respect to τ3, we get (5s4+4m1s3+3m2s2+2m3s+m4)dsdτ3+(4n1s3+3n2s2+2n3s+n4)esτ3dsdτ3+(n1s4+n2s3+n3s2+n4s+n5)sτ3dsdτ3esτ3=0. This gives dsdτ31=5s4+4m1s3+3m2s2+2m3s+m4s(s5+m1s4+m2s3+m3s2+m4s+m5)+4n1s3+3n2s2+2n3s+n4s(n1s4+n2s3+n3s2+n4s+n5)τ3s. We have signdRe(s(τ3))dτ3τ3=τk(j)=signRedsdτ31s=iω0=sign5ω08+ω06(8m2+4m12)+ω04(6m4+3m226m1m3)ω02(ω04+m2ω02m4)2+(m1ω04m3ω02+m5)2+ω02(4m2m4+2m1m5+2m32)+(m422m3m5)ω02(ω04+m2ω02m4)2+(m1ω04m3ω02+m5)2+4n12ω06+(3n22+6n1n3)ω04+ω02(4n2n44n1n52n32)+(n42+2n3n5)ω02(n2ω02n4)2+(n1ω04n3ω02+n5)2. From Equation (Equation6), we get ω02(ω04+m2ω02m4)2+(m1ω04m3ω02+m5)2=ω02(n2ω02n4)2+(n1ω04n3ω02+n5)2. It follows that signdRe(s(τ3))dτ3τ3=τk(j)=signH(λk)ω2(n2ω02n4)2+(n1ω04n3ω02+n5)2. We conclude that dRe(s(τ3))/dτ3|τ3=τk(j) and H(λk) have the same sign. Sum up the above discussion, we get the following conclusion.

Theorem 4.2

Let λ0 and ω0, τ0 be defined by Equation (Equation9), respectively.

  1. If the following conditions are all not satisfied: (a) e0<0; (b) e00, b1=0, Δ00 and a1<0 or c10, and there exists λ{λ1,λ2,λ3,λ4} such that λ>0 and H(λ)0; (c) e00, b10, d0>b1 and Δ20 or Δ30 and there exists λ{λ1,λ2,λ3,λ4} such that λ>0 and H(λ)0; (d) e00, b10, d0<b1, b12/4(a1d0)2+12d0=0, λ¯>0 and H(λ¯)0, where λ¯=b1/2(a1d0)15a0, then equilibrium E3 is locally asymptotically stable for any τ30.

  2. If one of the conditions given in (i) is satisfied, then equilibrium E3 is locally asymptotically stable for τ3[0,τ0).

  3. If one of the conditions given in (i) holds and H(λ0)0, then model (Equation1) undergoes Hopf bifurcation from equilibrium E3 as τ3 passes through the critical value τ0.

Next, we discuss Hopf bifurcation of the equilibrium E4. By Theorem 3.9, we only obtain the global asymptotic stability of the equilibrium E4 when τ10, τ20 and τ3=0. When τ3>0, by the following theoretical analysis we will see that the Hopf bifurcation occurs at the equilibrium E4. Since when τ1>0 or τ2>0 the discussions are very complicated, we here only discuss the case τ1=τ2=0 and τ3>0.

At the equilibrium E4, the characteristic equation for the linearized system of model (Equation1) is given by (10) s5+p1s4+p2s3+p3s2+p4s+p5+(q1s4+q2s3+q3s2+q4s+q5)esτ3=0,(10) where p1=d+h+r+pz4+u+qw4+βv4(1+bv4)(1+ax4+bv4)2,p2=d+βv4(1+bv4)(1+ax4+bv4)2(h+r+pz4+u+qw4)+h(r+pz4+u+qw4)+αqw4+(r+pz4)(u+qw4)kβx4(1+ax4)(1+ax4+bv4)2,p3=d+βv4(1+bv4)(1+ax4+bv4)2[h(r+pz4+u+qw4)+αqw4+(r+pz4)(u+qw4)]+h[αqw4+(r+pz4)(u+qw4)](d+h)kβx4(1+ax4)(1+ax4+bv4)2+αqw4(r+pz4),p4=d+βv4(1+bv4)(1+ax4+bv4)2{αqw4(r+pz4)+h[αqw4+(r+pz4)(u+qw4)]}hdkβx4(1+ax4)(1+ax4+bv4)2+αqhw4(r+pz4),p5=αqhw4(r+pz4)(d+βv4(1+bv4)(1+ax4+bv4)2),q1=h,q2=h(d+r+u+qw4+βv4(1+bv4)(1+ax4+bv4)2),q3=h(r+u+qw4)d+βv4(1+bv4)(1+ax4+bv4)2+αqw4+r(u+qw4)kβx4(1+ax4)(1+ax4+bv4)2,q4=hd+βv4(1+bv4)(1+ax4+bv4)2[αqw4+r(u+qw4)]αhrqw4+hdkβx4(1+ax4)(1+ax4+bv4)2,q5=αhrqw4d+βv4(1+bv4)(1+ax4+bv4)2. Let s=iω(ω>0) be a purely imaginary root of (Equation10). Separating real and imaginary parts, it follows that (11) p1ω4p3ω2+p5=(q1ω4q3ω2+q5)cosωτ3(q2ω3+q4ω)sinωτ3,ω5p2ω3+p4ω=(q1ω4q3ω2+q5)sinωτ3(q2ω3+q4ω)cosωτ3.(11) Squaring and adding the two equations of (Equation11), it follows that (12) ω10+p0ω8+q0ω6+r0ω4+u0ω2+η0=0,(12) where p0=p122p2q12,q0=2p1p3+p22+2p4+2q1q3q22,r0=p32+2p1p52p2p4q322q1q5+2q2q4,u0=p422p3p5+2q3q5q42,η0=p52q52. Let λ=ω2, then Equation (Equation12) becomes (13) F(λ)=λ5+p0λ4+q0λ3+r0λ2+u0λ+η0=0.(13) Denote p1=625p02+35q0,q1=625p0q0+25r0+8125p03,r1=3625p04+3125q0p02225p0r0+15u0,Δ0=p124r1,p2=13p124r1,q2=227p13+83p1r1q12,Δ1=127p2314q22,u0=q22+Δ13+q22Δ13+p13,Δ2=u0p1+2q1u0p1,Δ3=u0p12q1u0p1.

A similar argument as in Lemma 4.1 we can define λi (i=1,2,3,4) for the polynomial equation (Equation13), and using the same method as in Equation (Equation9) we can further define τ0, ω0 and λ0. Therefore, we have the following results.

Theorem 4.3

  1. If the following conditions are all not satisfied: (a) η0<0; (b) η00, q1=0, Δ00, p1<0 or r10, and there exists λ{λ1,λ2,λ3,λ4} such that λ>0 and F(λ)0; (c) η00, q10, u0>q1, Δ20 or Δ30 and there exists λ{λ1,λ2,λ3,λ4} such that λ>0 and F(λ)0; (d) η00, q10, u0<q1, q12/4(p1u0)2+12u0=0, λ¯>0 and F(λ¯)0, where λ¯=q1/2(p1u0)15p0, then E4 is locally asymptotically stable for any τ30.

  2. If one of the conditions given in (i) is satisfied, then E4 is locally asymptotically stable for τ3[0,τ0).

  3. If one of the conditions given in (i) holds and F(λ0)0, then model (Equation1) undergoes Hopf bifurcation from E4 as τ3 passes through critical value τ0.

Remark 4.4

We here only establish the criteria on the existence of Hopf bifurcations at equilibria E3 and E4 for model (Equation1) in the case of delays τ1=τ2=0 and τ3>0. However, when τ1>0 or τ2>0 whether we also can obtain similar results still is a very interesting and estimable problem. In the following section, we will give a discussion by means of the numerical simulations.

5. Numerical simulations

In the above sections, we establish the global asymptotic stability of equilibria E3 and E4 when τ10, τ20 and τ3=0, and by using the theory of bifurcation, we obtain the existence of the Hopf bifurcation and stability switches at equilibria E3 and E4 when τ1=0, τ2=0 and τ30. However, aim at the case: τ10, τ20 and τ30, the theoretical analysis is very complicated. In this section, by using the numerical simulation, it is shown that the Hopf bifurcation and stability switches occur at these equilibria as τ3 increases. In model (Equation1), we choose a, α, β, τ1, τ2 and τ3 as free parameters and fix all other parameters as displayed in Table .

Table 1. List of parameters.

Example 5.1

Take a=0, we have that Beddington–DeAngelis incidence βx(t)v(t)/(1+ax(t)+bv(t)) is simplified to saturation incidence βx(t)v(t)/(1+bv(t)). Let β=0.5lday1, α=1day1, τ1=8.5 and τ2=0.1. All other parameter values are the same as in Table . From Figures  we see that as τ3 increases the dynamical behaviours of equilibrium E3 will occur: locally asymptotically stable → unstable and Hopf bifurcation appears → locally asymptotically stable → unstable and Hopf bifurcation appears.

Figure 1. Taking τ3=0.02, we have R2=12.1443>1 and R4=0.2997<1, infection equilibrium with only CTL response E3 is asymptotically stable.

Figure 1. Taking τ3=0.02, we have R2=12.1443>1 and R4=0.2997<1, infection equilibrium with only CTL response E3 is asymptotically stable.

Figure 2. Taking τ3=1.2, we have R2=12.1443>1 and R4=0.2997<1, infection equilibrium with only CTL response E3 occurs Hopf bifurcation.

Figure 2. Taking τ3=1.2, we have R2=12.1443>1 and R4=0.2997<1, infection equilibrium with only CTL response E3 occurs Hopf bifurcation.

Figure 3. Taking τ3=5.8, we have R2=12.1443>1 and R4=0.2997<1, infection equilibrium with only CTL response E3 is asymptotically stable.

Figure 3. Taking τ3=5.8, we have R2=12.1443>1 and R4=0.2997<1, infection equilibrium with only CTL response E3 is asymptotically stable.

Figure 4. Taking τ3=9.5, we have R2=12.1443>1 and R4=0.2997<1, infection equilibrium with only CTL response E3 occurs Hopf bifurcation.

Figure 4. Taking τ3=9.5, we have R2=12.1443>1 and R4=0.2997<1, infection equilibrium with only CTL response E3 occurs Hopf bifurcation.

In Figures , (a), (b) and (c) are denoted time-series figures of x(t), y(t), v(t), z(t) and w(t).

Example 5.2

Consider the Beddington–DeAngelis incidence βx(t)v(t)/(1+ax(t)+bv(t)). Let a=0.01, β=0.25μlday1, α=0.1day1, τ1=2 and τ2=5. All other parameter values are the same as in Table . From Figures  we see that as τ3 increases from zero the dynamical behaviours of equilibrium E4 will occur: locally asymptotically stable → unstable and Hopf bifurcation appears → locally asymptotically stable → unstable and Hopf bifurcation appears.

Figure 5. Taking τ3=1.285, we have R3=12.8126>1 and R4=2.8537>1, infection equilibrium with both antibody and CTL responses E4 is asymptotically stable.

Figure 5. Taking τ3=1.285, we have R3=12.8126>1 and R4=2.8537>1, infection equilibrium with both antibody and CTL responses E4 is asymptotically stable.

Figure 6. Taking τ3=6.365, we have R3=12.8126>1 and R4=2.8537>1, infection equilibrium with both antibody and CTL responses E4 occurs Hopf bifurcation.

Figure 6. Taking τ3=6.365, we have R3=12.8126>1 and R4=2.8537>1, infection equilibrium with both antibody and CTL responses E4 occurs Hopf bifurcation.

Figure 7. Taking τ3=51.578, we have R3=12.8126>1 and R4=2.8537>1, infection equilibrium with both antibody and CTL responses E4 is asymptotically stable.

Figure 7. Taking τ3=51.578, we have R3=12.8126>1 and R4=2.8537>1, infection equilibrium with both antibody and CTL responses E4 is asymptotically stable.

Figure 8. Taking τ3=65.163, we have R3=12.8126>1 and R4=2.8537>1, infection equilibrium with both antibody and CTL responses E4 occurs Hopf bifurcation.

Figure 8. Taking τ3=65.163, we have R3=12.8126>1 and R4=2.8537>1, infection equilibrium with both antibody and CTL responses E4 occurs Hopf bifurcation.

6. Discussion

In this paper, we have investigated a virus infection model (Equation1) with intracellular delay τ1, virus replication delay τ2 and immune response delay τ3. We assume that the production of CTL immune response depends on the infected cells and CTL cells based above important biological meaning. We see that similar assumption also is given in [Citation1,Citation9,Citation11,Citation12,Citation16,Citation18,Citation24]. Similarly, the production of antibody response depends on the virus and antibody (see [Citation1,Citation13,Citation14,Citation16]). Dynamical analysis shows that τ1, τ2 and τ3 play different roles in the stability of the model.

By the analysis, model (Equation1) has five possible equilibria, an infection-free equilibrium E0, immune-free equilibrium E1, infection equilibrium E2 with only antibody response, infection equilibrium E3 with only CTL response and infection equilibrium E4 with both CTL and antibody responses. A combination of basic reproductive ratio of viral infection R0, for antibody response R1, for CTL immune response R2, for CTL immune response competitive R3 and for antibody response competitive R4 determines the existence of these equilibria. Furthermore, they also determine the global properties of the model. We have shown that when R01, E0 is globally asymptotically stable, which means that the viruses are cleared and immune is not active. When R0>1, R11 and R21, E1 is globally asymptotically stable, which means that the infection becomes chronic but with no persistent CTL immune responses and antibody responses. When R1>1 and R31, E2 is globally asymptotically stable, which means that the infection becomes chronic with persistent antibody responses, but the infected cells can not stimulate and activate CTL immune responses. As respect to the analysis of E3, we consider special case τ10, τ20 and τ3=0, when R2>1 and R41, E3 is globally asymptotically stable, which means that the infection becomes chronic with persistent CTL immune responses, but the virus loads can not activate the antibody responses. About the stability of E4, we have obtained that for special case τ10, τ20 and τ3=0, when R3>1 and R4>1, E4 is globally asymptotically stable, that is, susceptible cells, infected cells, free virus, antibody responses and CTL responses coexist in vivo. We see that τ1 and τ2 do not affect the stability of the equilibria.

When τ3>0, by using the bifurcation theory, we obtain the sufficient conditions on the existence of Hopf bifurcation at E3 and E4. Meanwhile, by means of numerical simulations, it is shown that the Hopf bifurcation and stability switches occur at E3 and E4 as τ3 increases. Figures  indicate that E3 remains stable as τ3>0 is small, and along with the increase of τ3, E3 becomes unstable and periodic oscillations appear. It shows that stability switches occur as τ3 increases. Similarly, from Figures , we see that along with the increase of τ3 the dynamical behaviours of model (Equation1) at E4 appear very large diversification. Particularly, when τ3 is small enough, E4 is asymptotically stable and when τ3 is increasing, the stability switches occur at E4, and when E4 is unstable, Hopf bifurcation occurs. Finally, when τ3 is large enough, E4 is always unstable. Summarizing these discussions, we point out that τ3 affects markedly the stability of E3 and E4. This illustrates the fact that τ3 plays a negative part in the disease prevalence and control. Motivated by the above discussion, one can realize that the emergence in the broadly neutralizing antibodies reacted to change in infected cells that could kill viruses. On the whole, this paper somehow provides better understanding about neutralizing antibodies and might help to design a powerful vaccine, which prevents at least uninfected peoples from ever becoming infected with virus. However, by considering some other factors such as the antibody delay, diffusion and a time-varying drug concentration, whether we also can obtain that the global asymptotic stability of equilibria will also be a very estimable and significative subject. This is left to the future research.

Disclosure statement

The authors declare that there is no conflict of interests regarding the publication of this paper.

Additional information

Funding

This work was supported by the Natural Science Foundation of Shanxi University of Finance and Economics Starting Fund for the Shanxi University of Finance and Economics doctoral graduates research [grant No. Z18116], the National Natural Science Foundation of China [grant nos. 11771373 and 11661076], the Natural Science Foundation of Xinjiang [grant no. 2016D03022].

References

  • P. Balasubramaniam, P. Tamilalagan, and M. Prakash, Bifurcation analysis of HIV infection model with antibody and cytotoxic T-lymphocyte immune responses and Beddington–DeAngelis functional response, Math. Methods Appl. Sci. 38 (2015), pp. 1330–1341. doi: 10.1002/mma.3148
  • K. Hattaf, N. Yousfi, and A. Tridane, Mathematical analysis of a virus dynamics model with general incidence rate and cure rate, Nonlinear Anal. Real World Appl. 13 (2012), pp. 1866–1872. doi: 10.1016/j.nonrwa.2011.12.015
  • K. Hattaf, N. Yousfi, and A. Tridane, Stability analysis of a virus dynamics model with general incidence rate and two delays, Appl. Math. Comput. 221 (2013), pp. 514–521.
  • G. Huang, W. Ma, and Y. Takeuchi, Global analysis for delay virus dynamics model with Beddington–DeAngelis function response, Appl. Math. Lett. 24 (2011), pp. 1199–1203. doi: 10.1016/j.aml.2011.02.007
  • Y. Ji, Global stability of a multiple delayed viral infection model with general incidence rate and an application to HIV infection, Math. Biosci. Eng. 12 (2015), pp. 525–536. doi: 10.3934/mbe.2015.12.525
  • Y. Kuang, Delay Differential Equations with Applications in Population Dynamics, Academic Press, San Diego, 1993.
  • M.Y. Li and H. Shu, Global dynamics of an in-host viral model with intracellular delay, Bull. Math. Biol. 72 (2010), pp. 1492–1505. doi: 10.1007/s11538-010-9503-x
  • K.A. Pawelek, S. Liu, F. Pahlevani, and L. Rong, A model of HIV-1 infection with two time delays: Mathematical analysis and comparison with patient data, Math. Biosci. 235 (2012), pp. 98–109. doi: 10.1016/j.mbs.2011.11.002
  • H. Shu, L. Wang, and J. Watmoughs, Global stability of a nonlinear viral infection model with infinitely distributed intracellular delays and CTL immune responses, SIAM J. Appl. Math. 73 (2013), pp. 1280–1302. doi: 10.1137/120896463
  • X. Song, X. Zhou, and X. Zhao, Properties of stability and Hopf bifurcation for a HIV infection model with time delay, Appl. Math. Model. 34 (2010), pp. 1511–1523. doi: 10.1016/j.apm.2009.09.006
  • X. Tian and R. Xu, Global stability and Hopf bifurcation of an HIV-1 infection model with saturation incidence and delayed CTL immune response, Appl. Math. Comput. 237 (2014), pp. 146–154.
  • X. Wang, A. Elaiw, and X. Song, Global properties of a delayed HIV infection model with CTL immune response, Appl. Math. Comput. 218 (2012), pp. 9405–9414.
  • T. Wang, Z. Hu, and F. Liao, Stability and Hopf bifurcation for a virus infection model with delayed humoral immunity response, J. Math. Anal. Appl. 411 (2014), pp. 63–74. doi: 10.1016/j.jmaa.2013.09.035
  • J. Wang, J. Pang, T. Kuniya, and Y. Enatsu, Global threshold dynamics in a five-dimensional virus model with cell-mediated, humoral immune responses and distributed delays, Appl. Math. Comput. 241 (2014), pp. 298–316.
  • Z. Wang and R. Xu, Stability and Hopf bifurcation in a viral infection model with nonlinear incidence rate and delayed immune response, Commun. Nonlinear Sci. Numer. Simul. 17 (2012), pp. 964–978. doi: 10.1016/j.cnsns.2011.06.024
  • Y. Wang, Y. Zhou, F. Brauer, and J.M. Heffernan, Viral dynamics model with CTL immune response incorporating antiretroviral therapy, J. Math. Biol. 67 (2013), pp. 901–934. doi: 10.1007/s00285-012-0580-3
  • D. Wodarz, Hepatitis C virus dynamics and pathology: The role of CTL and antibody responses, J. Gen. Virol. 84 (2003), pp. 1743–1750. doi: 10.1099/vir.0.19118-0
  • H. Xiang, L. Feng, and H. Huo, Stability of the virus dynamics model with Beddington–DeAngelis functional response and delays, Appl. Math. Model. 37 (2013), pp. 5414–5423. doi: 10.1016/j.apm.2012.10.033
  • Y. Yan and W. Wang, Global stability of a five-dimensional model with immune responses and delay, Discrete Contin. Dyn. Syst. B 17 (2012), pp. 401–416. doi: 10.3934/dcdsb.2012.17.401
  • N. Yousfi, K. Hattaf, and M. Rachik, Analysis of a HCV model with CTL and antibody responses, Appl. Math. Sci. 57 (2009), pp. 2835–2845.
  • Z. Yuan and X. Zou, Global threshold dynamics in an HIV virus model with nonlinear infection rate and distributed invasion and production delays, Math. Biosci. Eng. 10 (2013), pp. 483–498. doi: 10.3934/mbe.2013.10.483
  • T. Zhang, H. Jiang, and Z. Teng, On the distribution of the roots of a fifth degree exponential polynomial with application to a delayed neural network model, Neurocomputing 72 (2009), pp. 1098–1104. doi: 10.1016/j.neucom.2008.03.003
  • H. Zhu, Y. Luo, and M. Chen, Stability and Hopf bifurcation of a HIV infection model with CTL-response delay, Comput. Math. Appl. 62 (2011), pp. 3091–3102. doi: 10.1016/j.camwa.2011.08.022
  • H. Zhu and X. Zou, Dynamics of a HIV-1 infection model with cell-mediated immune response and intracellular delay, Discrete Contin. Dyn. Syst. Ser. B 12 (2009), pp. 511–524. doi: 10.3934/dcdsb.2009.12.511