1,508
Views
3
CrossRef citations to date
0
Altmetric
Research Article

Threshold dynamics of a HCV model with virus to cell transmission in both liver with CTL immune response and the extrahepatic tissue

, &
Pages 19-34 | Received 09 Apr 2020, Accepted 16 Nov 2020, Published online: 24 Dec 2020

Abstract

In this paper, a deterministic model characterizing the within-host infection of Hepatitis C virus (HCV) in intrahepatic and extrahepatic tissues is presented. In addition, the model also includes the effect of the cytotoxic T lymphocyte (CTL) immunity described by a linear activation rate by infected cells. Firstly, the non-negativity and boundedness of solutions of the model are established. Secondly, the basic reproduction number R01 and immune reproduction number R02 are calculated, respectively. Three equilibria, namely, infection-free, CTL immune response-free and infected equilibrium with CTL immune response are discussed in terms of these two thresholds. Thirdly, the stability of these three equilibria is investigated theoretically as well as numerically. The results show that when R01<1, the virus will be cleared out eventually and the CTL immune response will also disappear; when R02<1<R01, the virus persists within the host, but the CTL immune response disappears eventually; when R02>1, both of the virus and the CTL immune response persist within the host. Finally, a brief discussion will be given.

2010 Mathematics Subject Classifications:

1. Introduction

Viral hepatitis affects approximately 500 million people around the world – more than 10 times the number affected by HIV/AIDS [Citation1]. Different viruses can cause various forms of viral hepatitis. It is estimated that about 71 million people or 1% of the global population are chronically infected with hepatitis C according to the World Health Organization(WHO) Global Hepatitis Report in 2017 [Citation21]. Therefore, it is important to understand the dynamics of HCV infection in order to manage control programmes efficiently.

Hepatitis C virus (HCV) infection can lead to two different outcomes [Citation8]: in a small fraction of patients (15% of cases), the infection can be controlled and cleared from the blood; the rest of the patients become chronic. Chronic HCV is the main cause of chronic liver diseases and cirrhosis leading to liver transplantation (LT) or death [Citation2]. Unfortunately, the early results of transplantation for patients with chronic HCV were discouraging. Mortality rate of liver transplant is very high, and reinfection of the liver graft often occurs [Citation20]. This inevitable post-transplant infection may be related to the existence of an auxiliary compartment [Citation11]. The presence of HCV replicative intermediates has been reported in serum [Citation7], oral mucosa [Citation3] and gastric mucosa [Citation5]. Dahari et al. [Citation4] studied viral loads of 30 patients undergoing liver transplantation and observed the existence of a second replication compartment.

Virus clearance after acute HCV infection is associated with strong and polyclonal CD4 T cell responses, as well as sustained CTL responses. Recently, molecular techniques have provided fundamental insights into the molecular mechanisms of the immune system for HCV infection [Citation16, Citation18]. In 1996, Nowak and Bangham [Citation15] proposed a simple mathematical model to explore the relation between antiviral immune response and virus load. In 2003, Wodarz [Citation22] extended the model in [Citation15] to investigate the role of CTL and antibody response in HCV infection dynamics and pathology. Zhou et al. [Citation24, Citation25] considered the CTL immune response against HCV infection. However, the mechanism of CTL action in HCV infection is still not fully understood [Citation6, Citation9].

Mathematical models have become important tools in analysing the spread and control of HCV epidemic. Dahari et al. [Citation4] constructed a few within-host HCV infection models to describe HCV viral dynamics from the beginning of the anhepatic phase until the first viral increase data point. These models included two compartments of infection, but did not describe the asymptotical viral dynamics after transplantation of the liver. Qesmi et al. [Citation17] proposed a mathematical model of ordinary differential equations to describe the dynamics of the HBV/HCV and its interaction with both liver and blood cells based on [Citation4, Citation13], and found that the system undergoes either a transcritical or a backward bifurcation. Wodarz and Jansen [Citation23] proposed a model containing infected cells, non-acitived antigen presenting cells (APCs), acitived APCs and CTL, and analysed its complex dynamics.

However, there are very few HCV infection models with two compartments. Based on the existence of a second replication compartment for HCV and the role of CTL immune response against HCV infection, in this paper, we propose a new mathematical model containing another compartment of HCV infection. Then by the analysis of golbal dynamics, these theoretical results will reveal the interaction between HCV and CLT immune response more completely.

This paper is organized as follows. In Section 2, we formulate a new HCV infection model with CTL immune response and give a positively invariant set. Section 3 deals with the existence of equilibria for the model and two important parameters thresholds will be defined. In Section 4, the global stability of equilibria is investigated by using the Routh-Hurwitz criterion and Lyapunov functions. Some numerical examples are shown in Section 5. Finally, the epidemiological meanings of the obtained results are discussed, and the basic reproduction numbers of HCV infection and CTL immune response are given in Section 6.

2. Model formulation

In this section, we formulate a dynamical model with two proliferative compartments of HCV, one of which is the liver, the other is the extrahepatic compartment including serum, peripheral blood mononuclear cells (PBMC), and perihepatic lymph nodes (PLN). No experiments have shown that CTL immune response has effect or no effect on the extrahepatic compartment, here, it is assumed that the CTL immune response takes part in clearing infected hepatocytes and plays no role for the second proliferative compartment. The flowchart of HCV infection is shown in Figure . Here, we denote the liver and the second proliferative compartment (extrahepatic compartment) as compartments C1 and C2, respectively. In compartment C1, there are uninfected hepatocytes (x1(t)), infected hepatocytes (y1(t)) and the CTL immune response (z(t)). In compartment C2, there are uninfected extrahepatic cells (x2(t)), infected extrahepatic cells (y2(t)) and free virus (v(t)). Following the transmission diagram in Figure , our model takes the form in (Equation1) (1) x1=λ1β1x1vd1x1,y1=β1x1va1y1py1z,x2=λ2β2x2vd2x2,y2=β2x2va2y2,v=k1y1+k2y2γv,z=qy1zrz,(1) Here, λi is the recruitment rate of healthy cells and 1di is the average lifespan of uninfected cells in compartment Ci(i=1,2). The healthy cells become infected by free virus at a rate βixiv; infected cells in compartment Ci(i=1,2) die at a rate aiyi, and infected cells in compartment C1 are cleared by the CTL immune response at a rate py1z; the CTL immune response is triggered at a rate qy1z, which in turn decays a rate rz. We assume that the parameters are positive and aidi(i=1,2) [Citation14] according to the biological meaning. Note that a saturated nonlinear function was used in Wodarz and Jansen [Citation23] to describe the activation of the CTL immune response by the virus. Since we are interested in the global dynamics of the model, for the sake of simplicity we use a linear function here.

Figure 1. Flowchart of the viral infection model with CTL immune response in intrahepatic and extrahepatic compartments.

Figure 1. Flowchart of the viral infection model with CTL immune response in intrahepatic and extrahepatic compartments.

We can see that solutions of model (Equation1) with the nonnegative initial conditions remain nonnegative. From the first equation of (Equation1), we have x1λ1d1x1,then lim suptx1(t)λ1/d1. From the first two equations of (Equation1), we obtain (x1+y1)=λ1d1x1a1y1py1zλ1d1(x1+y1),since a1d1, then lim supt[x1(t)+y1(t)]λ1/d1. Similarly, from the middle two equations of (Equation1), we have lim suptx2(t)λ2/d2, and lim supt[x2(t)+y2(t)]λ2/d2.

When yiλi/di(i=1,2), from the fifth equation of (Equation1) we have v(k1λ1d1+k2λ2d2)γv,then lim suptv(t)1γ(k1λ1d1+k2λ2d2).And z = 0 always satisfies the last equation in (Equation1). Therefore, the region Ω={(x1,y1,x2,y2,v,z)R+6:xi+yiλidi,v1γ(k1λ1d1+k2λ2d2)(i=1,2)}is positively invariant with respect to system (Equation1). Therefore, it is sufficient to study the dynamics of model (Equation1) with initial conditions in Ω.

3. Existence of equilibria

In this section, we discuss the existence of equilibria of model (Equation1) satisfying the following equations (2) λ1β1x1vd1x1=0,β1x1va1y1py1z=0,λ2β2x2vd2x2=0,β2x2va2y2=0,k1y1+k2y2γv=0,qy1zrz=0,(2) on the set Ω.

Model (Equation1) always has an infection-free equilibrium E0(x1(0),0,x2(0),0,0,0), where x1(0)=λ1d1,x2(0)=λ2d2. From the first and third equations of (Equation2), we obtain (3) x1=λ1β1v+d1,x2=λ2β2v+d2.(3) Substituting them into the second and fourth equations of (Equation2), they yields respectively (4) y1=β1va1+pzλ1β1v+d1,y2=β2va2λ2β2v+d2.(4) When v0 and z = 0, substituting y1 and y2 of (Equation4) into the fifth equation of (Equation2) yields (5) h(v):=1γ[k1β1λ1a1(β1v+d1)+k2β2λ2a2(β2v+d2)]=1.(5) We can see that function h(v) is decreasing with respect to v. Note that h(1γ(k1λ1d1+k2λ2d2))11+d1k2λ2d2k1λ1+γd12k1β1λ1+11+d2k1λ1d1k2λ2+γd22k2β2λ2:=h0,where aidi(i=1,2) is used. Since the inequality 11+m+11+n<1 holds for m, n>0 if and only if mn>1, we know that h0<1. Thus, by the monotonicity of function h(v), Equation (Equation5) has a unique positive root v(1) only when h(0)=1γ[k1β1λ1a1d1+k2β2λ2a2d2]>1. Furthermore, the corresponding xi(1) and yi(1) (i = 1, 2) can be obtained from (Equation3) and (Equation4). Thus, (Equation1) has a boundary equilibrium E1(x1(1),y1(1),x2(1), y2(1),v(1),0) when h(0)>1.

When z0, from the last equation of (Equation2) we have y1=rq:=y1(2). Substituting it and x1=λ1β1v+d1 in (Equation3) into the second equation of (Equation2) gives (6) z=qpr(β1v+d1)[β1(λ1a1rq)va1d1rq].(6) Then a necessary condition on the existence of the positive equilibrium is λ1>a1rq, and, for the positive equilibrium E2(x1(2),y1(2),x2(2),y2(2),v(2),z(2)), v(2)>a1d1rβ1(qλ1a1r):=v¯.

On the other hand, substituting y1=rq and y2=β2λ2va2(β2v+d2) in (Equation3) into the fifth equation of (Equation2) yields (7) g(v):=1γ[k1rqv+k2β2λ2a2(β2v+d2)]=1.(7) Under the case that λ1>a1rq (i.e. rq<λ1a1) and for aidi (i = 1, 2), we have g(1γ(k1λ1d1+k2λ2d2))=rqk1k1λ1d1+k2λ2d2+k2β2λ2a2[β2(k1λ1d1+k2λ2d2)+γd2]k1λ1a1(k1λ1d1+k2λ2d2)+k2β2λ2a2[β2(k1λ1d1+k2λ2d2)+γd2]11+d1k2λ2d2k1λ1+1d2k1λ1d1k2λ2+1+γd22k2β2λ2.Applying again the inequality that 11+m+11+n<1 holds for m, n>0 if and only if mn>1, we know that g(1γ(k1λ1d1+k2λ2d2))<1. Then, according to the monotonicity of function g(v), the fact that limv0+g(v)=+ implies that equation g(v)=1 has a unique root in the interval (v¯,1γ(k1λ1d1+k2λ2d2)) if and only if g(v¯)=1γ{k1β1(λ1qa1r)a1d1q+k2β2λ2a2[β2a1d1rβ1(λ1qa1r)+d2]}>1.Note that g(v¯)<1γ(k1β1λ1a1d1+k2β2λ2a2d2) as λ1>a1rq. Then there must exist a boundary equilibrium E1(x1(1),y1(1), x2(1),y2(1),v(1),0) if the positive equilibrium E2 exists.

We claim that for λ1>a1rq, g(v¯)>1 is equivalent to the inequality qβ1λ1v(1)ra1(β1v(1)+d1)>1, i.e. v(1)>v¯. In fact, since v=v(1) is the root of equation h(v)=1, then the monotonicity of h(v) implies that h(v¯)>1 as v(1)>v¯. On the other hand, direct calculation shows that g(v¯)=h(v¯). Hence, according the existence of the equilibrium E1, model (Equation1) has a unique positive equilibrium E2 as 1γ(k1β1λ1a1d1+k2β2λ2a2d2)>1, λ1>a1rq and qβ1λ1v(1)ra1(β1v(1)+d1)>1, i.e. qy1(1)r>1.

Notice that, when 1γ(k1β1λ1a1d1+k2β2λ2a2d2)>1, qβ1λ1v(1)ra1(β1v(1)+d1)>1 implies that λ1>a1rq. Then, summarizing the above discussion, we have the following theorem.

Theorem 3.1

Denote R01=1γ(k1β1λ1a1d1+k2β2λ2a2d2),R02=qy1(1)r.The existence of equilibria in system (Equation1) can be summarized below.

(a)

The infection-free equilibrium E0(x1(0),0,x2(0),0,0,0) on the set Ω, where x1(0)=λ1d1 and x2(0)=λ2d2, always exists.

(b)

When R01>1, in addition to the infection-free equilibrium E0, system (Equation1) also has the immune response-free equilibrium E1(x1(1),y1(1),x2(1), y2(1),v(1),0), where x1(1)=λ1β1v(1)+d1,y1(1)=β1λ1v(1)a1(β1v(1)+d1),x2(1)=λ2β2v(1)+d2,y2(1)=β2λ2v(1)a2(β2v(1)+d2),and v(1) is the positive root of (Equation5).

(c)

When R01>1 and R02>1, system (Equation1) has a unique positive equilibrium E2(x1(2),y1(2),x2(2),y2(2), v(2),z(2)), where x1(2)=λ1β1v(2)+d1,y1(2)=rq,x2(2)=λ2β2v(2)+d2,y2(2)=β2v(2)a2λ2β2v(2)+d2,z(2)=a1p(R021)and v(2) is the positive root of (Equation7).

4. Stability of equilibria

In this section, we discuss the global stability of equilibria of (Equation1). We first present two propositions for the infection-free equilibrium E0.

Proposition 4.1

When R01<1, the infection-free equilibrium E0 is locally asymptotically stable; when R01>1, it is unstable.

Proof.

The Jacobian matrix of system (Equation1) at E0 is J(E0)=(d1000β1λ1d100a100β1λ1d1000d20β2λ2d20000a2β2λ2d200k10k2γ000000r).The eigenvalues of J(E0) are d1, d2, r, and the roots of the equation (8) λ3+b1λ2+b2λ+b3=0,(8) where b1=a1+a2+γ>0,b2=a1a2+(a1+a2)γβ1λ1k1d1β2λ2k2d2=a1a2+a1γ(1β1λ1k1a1d1γ)+a2γ(1β2λ2k2a1d2γ),b3=a1a2γ(1R01).Since R01<1 implies that β1λ1k1a1d1γ<1andβ2λ2k2a1d2γ<1,we have b2>0 and b3>0 when R01<1.

Furthermore, we have b1b2b3=(a1+a2)a1a2+a1a2γR01+(a1+a2+γ)[a1γ(1β1λ1k1γd1a1)+a2γ(1β2λ2k2γd2a2)]>0as R01<1. It follows from the Routh-Hurwitz criterion that all roots of (Equation8) have negative real parts if R01<1. Thus, the infection-free equilibrium E0 is locally asymptotically stable when R01<1. Since R01>1 is equivalent to b3<0, we know that E0 is unstable as R01>1.

Proposition 4.2

When R01<1, limty1(t)=limty2(t)=limtv(t)=limtz(t)=0.

Proof.

Since R01<1, we have k1β1λ1a1d1γ<1 and k2β2λ2a2d2γ<1, that is, β1λ1d1<a1γk1 and β2λ2d2<a2γk2. Moreover, direct calculation shows that R01<1 is equivalent to the following inequality 0<qβ1λ1pd1a2γk2β2λ2d2<qp(a1γk1β1λ1d1)β2λ2d2.So we can choose a positive number m1 satisfying the inequality (9) qβ1λ1pd1a2γk2β2λ2d2<m1<qp(a1γk1β1λ1d1)β2λ2d2,(9) that is, 1γ(qβ1λ1pd1+m1β2λ2d2)<m1a2k2and 1γ(qβ1λ1pd1+m1β2λ2d2)<qa1pk1.Further, for the given m1, we choose a positive number m2 satisfying the inequality (10) 1γ(qβ1λ1pd1+m1β2λ2d2)<m2<min{m1a2k2,qa1pk1}.(10) When m1 and m2 are given, we define a function V1=qpy1+m1y2+m2v+z.Then x1λ1d1 and x2λ2d2 imply that the derivative of V1 along solutions of model (Equation1) is given by V1=(m2k1qpa1)y1+(m2k2m1a2)y2+(qpβ1x1+m1β2x2m2γ)vrz(qpa1m2k1)y1(m1a2m2k2)y2(m2γqpβ1λ1d1m1β2λ2d2)vrz.It follows from (Equation9) and (Equation10) that qpa1m2k1>0,m1a2m2k2>0,m2γqpβ1λ1d1m1β2λ2d2>0.Then ρ=min{pq(qpa1m2k1),1m1(m1a2m2k2),1m2(m2γqpβ1λ1d1m1β2λ2d2),r}>0.Thus, we have V1ρV1. It implies that limtV1(t)=0, that is, limty1(t)=limty2(t)=limtv(t)=limtz(t)=0.

For the global stability of equilibria of (Equation1), we have the following results.

Theorem 4.1

When R01<1, the infection-free equilibrium E0 of model (Equation1) is globally stable in Ω; when R02<1<R01, the immune response-free equilibrium E1 of model (Equation1) is globally stable in Ω{E0}; when R01>1 and R02>1, the infection equilibrium E2 of model (Equation1) is globally stable in the set Ω.

Proof.

When R01<1, by Proposition 4.2 and the theory of asymptotic autonomous systems [Citation19, Theorem 1.2], it then follows from the first and third equations of (Equation1) that x1(t)λ1/d1 and x2(t)λ2/d2 as t+. Furthermore, Proposition 4.1 implies that the infection-free equilibrium E0 is globally stable in the set Ω when R01<1.

Next, we consider the global stability of the equilibrium E1(x1(1),y1(1),x2(1), y2(1),v(1),0). Define a Lyapunov function V2=k1a1(x1x1(1)x1(1)lnx1x1(1)+y1y1(1)y1(1)lny1y1(1))+k2a2(x2x2(1)x2(1)lnx2x2(1)+y2y2(1)y2(1)lny2y2(1))+(vv(1)v(1)lnvv(1))+pk1qa1z,then the derivative of V2 along solutions of system (Equation1) is given by dV2dt=k1a1[(1x1(1)x1)(λ1β1x1vd1x1)+(1y1(1)y1)(β1x1va1y1py1z)]+k2a2[(1x2(1)x2)(λ2β2x2vd2x2)+(1y2(1)y2)(β2x2va2y2)]+(1v(1)v)(k1y1+k2y2γv)+pk1qa1(qy1r)z.Since x1(1),y1(1),x2(1),y2(1), and v(1) satisfy the following equations λ1=β1x1v+d1x1,β1x1v=a1y1,λ2=β2x2v+d2x2,β2x2v=a2y2,k1y1+k2y2=γv,then dV2/dt can be rewritten as follows dV2dt=k1d1x1(1)a1(2x1(1)x1x1x1(1))+k1y1(1)(3x1(1)x1x1vy1(1)x1(1)v(1)y1v(1)y1vy1(1))+k2d2x2(1)a2(2x2(1)x2x2x2(1))+k2y2(1)(3x2(1)x2x2vy2(1)x2(1)v(1)y2v(1)y2vy2(1))+k1pra1q(R021)z.Since the arithmetical mean is greater than or equal to the geometrical mean, for x1,y1,x2,y2,v>0, we have x1(1)x1+x1x1(1)20, and the equality holds if and only if x1=x1(1); x2(1)x2+x2x2(1)20, and the equality holds if and only if x2=x2(1); x1(1)x1+x1vy1(1)x1(1)v(1)y1+v(1)y1vy1(1)30, and the equality holds if and only if x1=x1(1) and v/v(1)=y1/y1(1); x2(1)x2+x2vy2(1)x2(1)v(1)y2+v(1)y2vy2(1)30, and the equality holds if and only if x2=x2(1) and v/v(1)=y2/y2(1).

Therefore, when R02<1<R01, we have that dV2/dt0, and that the equality holds if and only if x1=x1(1), x2=x2(1), z = 0, and y1/y1(1)=y2/y2(1)=v/v(1). The largest invariant set of system (Equation1) on the region {(x1,y1,x2,y2,v,z)Ω:dV2/dt=0} is the singleton {E1} when R02<1<R01. Thus, it follows from LaSalle Invariance Principal [Citation10] that the boundary equilibrium E1 is globally asymptotically stable in the region Ω.

Lastly, we discuss the global stability of the positive equilibrium E2(x1(2),y1(2),x2(2),y2(2),v(2),z(2)). Define a Lyapunov function V3=k1a1+pz(2)(x1x1(2)x1(2)lnx1x1(2)+y1y1(2)y1(2)lny1y1(2))+k2a2(x2x2(2)x2(2)lnx2x2(2)+y2y2(2)y2(2)lny2y2(2))+(vv(2)v(2)lnvv(2))+pk1q(a1+pz(2))(zz(2)z(2)lnzz(2)),then the derivative of V3 along solutions of system (Equation1) is given by dV3dt=k1a1+pz(2)×[(1x1(2)x1)(λ1β1x1vd1x1)+(1y1(2)y1)(β1x1va1y1py1z)]+k2a2[(1x2(1)x2)(λ2β2x2vd2x2)+(1y2(2)y2)(β2x2va2y2)]+(1v(2)v)(k1y1+k2y2γv)+pk1q(a1+pz(2))(1z(2)z)(qy1r)z.Using the equalities λ1=β1x1(2)v(2)+d1x1(2), β1x1(2)v(2)=a1y1(2)+py1(2)z(2), λ2=β2x2(2)v(2)+d2x2(2), β2x2(2)v(2)=a2y2(2), k1y1(2)+k2y2(2)=γv(2), and qy1(2)=r, we can rewrite dV3/dt as follows: dV3dt=k1d1x1(2)a1+pz(2)(2x1(2)x1x1x1(2))+k1y1(2)(3x1(2)x1x1vy1(2)x1(2)v(2)y1v(2)y1vy1(2))+k2d2x2(2)a2(2x2(2)x2x2x2(2))+k2y2(2)(3x2(2)x2x2vy2(2)x2(2)v(2)y2v(2)y2vy2(2)).Similar to the proof of the global stability of the equilibrium E1, we have that dV3/dt0, and that the equality holds if and only if x1=x1(2), x2=x2(2), and y1/y1(2)=y2/y2(2)=v/v(2). In addition, the largest invariant set of system (Equation1) on the set {(x1,y1,x2,y2,v,z)Ω: dV3/dt=0} is the singleton {E2} when R02>1. By LaSalle Invariance Principal [Citation10], the positive equilibrium E2 is globally asymptotically stable in the region Ω.

5. Numerical simulations

In the previous sections, we have investigated the existence and global stability of the equilibria through some theoretical analysis. In this section, we will carry out some numerical simulations of (Equation1) with parameter values λ1=1,β1=0.08,d1=0.2,a1=0.2,p=0.3,λ2=0.8,β2=0.1,d2=0.2,a2=0.2,k1=1,k2=1.2,r=0.2, except for γ and q. Initial values are fixed in Figure  at (5.5,4.7,2.2,2.5,2.2,2.8), (5,4,1.6,1.3,1.0,0.5), (3.8,3,1.2,1.2,0.8,1) and (3.5,2.0,1.3,1.8,1.4,1.2). We choose the different values of γ and q to represent different dynamic behaviours. These parameter values chosen here are consistent with those in the models [Citation12, Citation22, Citation23].

Figure 2. The infection-free equilibrium E0 is globally asymptotically stable, here γ=5 and q = 0.1. The other parameter values are fixed as λ1=1,β1=0.08,d1=0.2,a1=0.2,p=0.3,λ2=0.8,β2=0.1,d2=0.2,a2=0.2,k1=1,k2=1.2,r=0.2, and then R01=0.88<1.

Figure 2. The infection-free equilibrium E0 is globally asymptotically stable, here γ=5 and q = 0.1. The other parameter values are fixed as λ1=1,β1=0.08,d1=0.2,a1=0.2,p=0.3,λ2=0.8,β2=0.1,d2=0.2,a2=0.2,k1=1,k2=1.2,r=0.2, and then R01=0.88<1.

Figure 3. The immune response-free equilibrium E1 is globally asymptotically stable, here γ=2 and q = 0.06. The other parameter values are identical with those in Figure , and then R01=2.2>1 and R02=0.7729<1.

Figure 3. The immune response-free equilibrium E1 is globally asymptotically stable, here γ=2 and q = 0.06. The other parameter values are identical with those in Figure 2, and then R01=2.2>1 and R02=0.7729<1.

Figure 4. The infection equilibrium E2 is globally asymptotically stable, here γ=2 and q = 0.1. The other parameter values are identical with those in Figure , and then R01=2.2>1 and R02=1.2882>1.

Figure 4. The infection equilibrium E2 is globally asymptotically stable, here γ=2 and q = 0.1. The other parameter values are identical with those in Figure 2, and then R01=2.2>1 and R02=1.2882>1.

When γ=5 and q = 0.1, we calculate R01=0.88<1. From Theorem 4.1, it follows that the equilibrium E0 of model (Equation1) is globally asymptotically stable (see Figure ). When γ=2 and q = 0.06, we obtain R01=2.2>1 and R02=0.7729<1. From Theorem 4.1, we know that the equilibrium E1 of model (Equation1) is globally asymptotically stable (see Figure ). In Figure , with parameters γ=2 and q = 0.1, the thresholds R01=2.2>1 and R02=1.2882>1. The infection equilibrium E2 of model (Equation1) is globally asymptotically stable, which is consistent with Theorem 4.1.

6. Conclusion and discussion

The novelty of our study is that we introduced a new compartment (extrahepatic compartment)) into the classical within-host hepatitis C virus infection models (see Wozard and Jansen [Citation23]) and provided results on the gobal dynamics of the model. According to Theorems 3.1 and 4.1, R01 and R02 are two thresholds determining the dynamical behaviours of system (Equation1) (see Figure ). When R01<1, system (Equation1) has a unique equilibrium E0, which is globally stable in the set Ω; when R02<1<R01, besides the boundary equilibrium E0, system (Equation1) also has another boundary equilibrium E1, which is globally stable in the set Ω; when R02>1, in addition to the boundary equilibria E0 and E1, system (Equation1) has a unique infection equilibrium E2 which is globally stable in the set Ω.

Figure 5. Existence and stability region of equilibria in terms of the thresholds R01 and R02.

Figure 5. Existence and stability region of equilibria in terms of the thresholds R01 and R02.

Notice that R01=1γ(k1β1λ1a1d1+k2β2λ2a2d2)is the basic reproduction number of Hepatitis C virus infection within the host. In fact, λ1d1 is the number of healthy cells at the steady state in compartment C1 in the absence of infection, β1λ1d1 is the number of new infected cells in unit time by per capital virion in compartment C1, k1a1 is the number of new released virion by an infected cell, 1γ is the average infectious period of virus, then k1β1λ1a1d1γ is the basic reproduction number of Hepatitis C virus within compartment C1. Similarly, k2β2λ2a2d2γ is the one within compartment C2.

Notice that although R02 is the threshold determining the existence and stability of the positive equilibrium E2, it is not the basic reproduction number. From Theorem 3.1, qy1(1)r, denoted by R02, can also be thought as a threshold when R01>1, which plays the same role as the threshold R02 in determining the dynamical behaviours of system (Equation1). Since y1(1) is the number of infected cells in compartment C1 at the steady state when R01>1, qy1(1) is the CTL immune response existed in a unit time by per CTL response, and 1r is the average decaying period of CTL response, then qy1(1)r represents the basic reproduction number of the CTL immune response.

For system (Equation1), we have discussed the existence and stability of three equilibria E0, E1 and E2. In the sense of viral dynamics, the boundary equilibrium E0 represents the infection-free steady state within the host; the other boundary equilibrium E1 represents the steady state at which the host is infected, and the CTL immune response plays no role in protecting the host from infection; the positive equilibrium E2 represents the steady state at which the host is infected and the CTL immune response plays a certain role in protecting the host from infection, but it cannot clear completely the infected cells in compartment C1. Therefore, by Theorems 3.1 and 4.1, when R01<1, the virus will be cleared eventually and the CTL immune response also disappears; when R02<1<R01, the virus persists within the host, but the CTL immune response disappears eventually, this implies that the immune system in the body of patient has no effect on neutralizing the infection of the virus; when R02>1, both the virus and the CTL immune response persist within the host, but the CTL immune response does not suffice to clear the virus completely.

According to the expression of the basic reproduction number of the CTL response R02, increasing the coefficient of producing the CTL response q implies the increase of R02. Since increasing the value of q results in the decrease of the values of y1(2), v(2) and y2(2) for the infetion equilibrium E2, increasing the ability of producing the CTL response can relieve the infetion from viruses.

Disclosure statement

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

Additional information

Funding

This work was partially supported by the National Natural Science Foundation of China (No.12061079 and No.11971281),China Postdoctoral Science Foundation [2019M653529], Scientific Research Program funded by Shaanxi Provincial Education Department (No.18JK0336), Cultive Scientific Research Excellence Programs of Higher Education Institutions in Shanxi Provincial Education Department (No.2020KJ020), and Scientific and Technological Innovation Programs of Higher Education Institutions in Shanxi Provincial Education Department (STIP) [2019L0861].

References

  • http://www.who.int/bulletin/volumes/88/11/10-011110/en/index.html.
  • M.J. Alter, H.S. Margolis, K. Krawczynski, F.N. Judson, A. Mares, W.J. Alexander, P.Y. Hu, J.K. Miller, M.A. Gerber, R.E. Sampliner, E.L. Meeks, and M.J. Beach, The natural history of community-acquired hepatitis C in the United States, N. Engl. J. Med. 327 (1992), pp. 1899–1905.
  • M. Carrozzo, R. Quadri, P. Latorre, M. Pentenero, S. Paganin, G. Bertolusso, S. Gandolfo, and F.Negro, Molecular evidence that the hepatitis C virus replicates in the oral mucosa, J. Hepatol. 37 (2002), pp. 364–369.
  • H. Dahari, A. Feliu, M. Garcia-Retortillo, X. Forns, and A.U. Neumann, Second hepatitis C replication compartment indicated by viral dynamics during liver transplantation, J. Hepatol. 42 (2005), pp. 491–498.
  • S. De Vita, V. De Re, D. Sansonno, D. Sorrentino, R.L. Corte, B. Pivetta, D. Gasparotto, V. Racanelli, A. Marzotto, A. Labombarda, A. Gloghini, G. Ferraccioli, A. Monteverde, A. Carbone, F. Dammacco, and M. Boiocchi, Gastric mucosa as an additional extrahepatic localization of hepatitis C virus: viral detection in gastric low-grade lymphoma associated with autoimmune disease and in chronic gastritis, Hepatology 31 (2000), pp. 182–189.
  • P. Farci, A. Shimoda, A. Coiana, G. Diaz, G. Peddis, J.C. Melpolder, A. Strazzera, D.Y. Chien, S.J.Munoz, A. Balestrieri, R.H. Purcell, and H.J. Alter, The outcome of acute hepatitis C predicted by the evolution of the viral quasispecies, Science. 288 (2000), pp. 339–344.
  • T.L. Fong, M. Shindo, S.M. Feinstone, J.H. Hoofnagle, and A.M. Di Bisceglie, Detection of replicative intermediates of hepatitis C viral RNA in liver and serum of patients with chronic hepatitis C, J. Clin. Invest. 88 (1991), pp. 1058–1060.
  • J.H. Hoofnagle, Management of hepatitis C: current and future perspectives, J. Hepatol 31 (1999), pp. 264–268.
  • P. Klenerman, F. Lechner, M. Kantzanou, A. Ciurea, H. Hengartner, and R. Zinkernagel, Viral escape and the failure of cellular immune responses, Science. 289 (2000), p. 2003.
  • J.P. LaSalle, The Stability of Dynamical Systems, SIAM, Philadelphia, 1976.
  • T. Laskus, M. Radkowski, J. Wilkinson, H. Vargas, and J. Rakela, The origin of hepatitis C virus reinfecting transplanted livers: serum-derived versus peripheral blood mononuclear cell-derived virus, J. Infect. Dis. 185 (2002), pp. 417–421.
  • J. Li, K. Men, Y. Yang, and D. Li, Dynamical analysis on a chronic hepattis C virus infection model with immune response, J. Theoret. Biol. 365 (2015), pp. 337–346.
  • Y. Liu and X. Liu, Global properties and bifurcation analysis of an HIV-1 infection model with two target cells, Comp. Appl. Math. 37 (2018), pp. 3455–3472.
  • A.U. Neumann, N.P. Lam, H. Dahari, D.R. Gretch, T.E. Wiley, T.J. Layden, and A.S. Perelson, Hepatitis C viral dynamics in vivo and the antiviral efficacy of interferon-α therapy, Science. 282 (1998), pp. 103–107.
  • M.A. Nowak and C.R.M. Bangham, Population dynamics of immune responses to persistent viruses, Science. 272 (1996), pp. 74–79.
  • A. Plauzolles, M. Lucas, and S. Gaudieri, Hepatitis C virus adaptation to T-cell immune pressure, Sci. World J. 2013 (2013), Article ID 673240.
  • R. Qesmi, J. Wu, J. Wu, and J.M. Heffernan, Influence of backward bifurcation in a model of hepatitis B and C viruses, Math. Biosci. 224 (2010), pp. 118–125.
  • S.D. Sharma, Hepatitis C virus: molecular biology and current therapeutic options, Indian J. Med. Res. 131 (2010), pp. 17–34.
  • H.R. Thieme, Convergence results and a Poincaré-Bendison trichotomy for asymptotical autonomous differential equations, J. Math. Biol. 30 (1992), pp. 755–763.
  • M. Tsiang, J. Rooney, J. Toole, and C. Gibbs, Biphasic clearance kinetics of hepatitis B virus from patients during adefovir dipivoxil therapy, Hepatology 29 (1999), pp. 1863–1869.
  • WHO, Global Hepatitis Report, 2017. France: WHO; 2017.
  • D. Wodarz, Hepatitis C virus dynamics and pathology: the role of CTL and antibody responses, J. Gen. Virol. 84 (2003), pp. 1743–1750.
  • D. Wodarz and V.A.A. Jansen, A dynamical perspective of CTL cross-priming and regulation: implications for cancer immunology, Immunol. Lett. 86 (2003), pp. 213–227.
  • X. Zhou, X. Shi, Z. Zhang, and X. Song, Dynamical behavior of a virus dynamics model with CTL immune response, Appl. Math. Comput. 213 (2009), pp. 329–347.
  • Y. Zhou, Y. Zhang, Z. Yao, J.P. Moorman, and Z. Jia, Dendritic cell-based immunity and vaccination agains the patitis C virus infection, Immunology 136 (2012), pp. 385–396.