1,353
Views
1
CrossRef citations to date
0
Altmetric
Research Article

Stability and bifurcation analysis of an HIV-1 infection model with a general incidence and CTL immune response

, &
Pages 367-394 | Received 20 Nov 2020, Accepted 12 Jun 2021, Published online: 12 Jul 2021

Abstract

In this paper, with eclipse stage in consideration, we propose an HIV-1 infection model with a general incidence rate and CTL immune response. We first study the existence and local stability of equilibria, which is characterized by the basic infection reproduction number R0 and the basic immunity reproduction number R1. The local stability analysis indicates the occurrence of transcritical bifurcations of equilibria. We confirm the bifurcations at the disease-free equilibrium and the infected immune-free equilibrium with transmission rate and the decay rate of CTLs as bifurcation parameters, respectively. Then we apply the approach of Lyapunov functions to establish the global stability of the equilibria, which is determined by the two basic reproduction numbers. These theoretical results are supported with numerical simulations. Moreover, we also identify the high sensitivity parameters by carrying out the sensitivity analysis of the two basic reproduction numbers to the model parameters.

1. Introduction

HIV (Human Immunodeficiency Virus) is the pathogen causing AIDS (Acquired ImmunoDeficiency Syndrome). Now, HIV has been a major global public health issue. There were approximately 37.9 million people living with HIV at the end of 2018 with 1.7 million people becoming newly infected in 2018 globally [Citation34]. Though an estimated entry into the human population in the early twentieth century [Citation9], HIV was identified as responsible for AIDS only until 1984 [Citation10]. Since then, HIV infection has been intensively studied and massive drug development efforts have been made. “Yet despite these advances, we still do not have a clear explanation for the pathogenesis of infection nor therapies that can permanently cure the infection” [Citation12]. Mathematical models have played an important role in enhancing understanding of HIV infection dynamics and helping generating ideas on how to treat the infection. Here we only focus on within-host HIV models. We refer to [Citation25] for a recent review for such models.

In the pioneering work [Citation22], Nowak and Bangham proposed some simple models to explore the relation between antiviral immune responses, virus load, and virus diversity. Since then, these models have been modified to reflect the effects of many other factors such as delay, eclipse phase, treatment, drug resistance, nonlinear incidence, transmission modes, multi-strain of viruses, co-infection, super-infection, vaccination, age/stage structure, spatial heterogeneity, and so on. We refer to [Citation1–3, Citation8, Citation14, Citation15, Citation17, Citation19–21, Citation23, Citation24, Citation28, Citation29, Citation32] as samples of works in the last five years to indicate how active the study on HIV-1 infection still is.

In particular, the HIV eclipse phase is the time period from a virus entering a target cell to the infected cell becoming infectious and producing new viruses. Of the four phases of HIV infection (acute, eclipse, chronic, and AIDS), the first three are currently the window of opportunity for viral clearance. There are many ways to describe the eclipse phase, which include delay models and age-structured models (see, for example, the references in [Citation26]). In [Citation26], Rong and his coworkers introduced a stage-structured HIV-1 infection model. Their model is described by the following system of four ordinary differential equations, (1) {x˙(t)=λμx(t)βx(t)v(t)+δw(t),w˙(t)=βx(t)v(t)(δ+η+q)w(t),y˙(t)=qw(t)αy(t),v˙(t)=σy(t)γv(t),(1) where x(t), w(t), y(t), and v(t) represent the concentrations of uninfected CD4+ T cells, infected cells in the eclipse stage, productively infected cells, and free viruses at time t, respectively. One feature of model (Equation1) is that a portion of infected cells in the eclipse phase may revert to the uninfected cells. The biological meanings of the parameters are summarized in Table .

Table 1. The biological meanings of the parameters in model (Equation1).

Rong et al. analysed the local stability of the infected equilibrium of (Equation1). Later on, Buonomo and Vargas-De-Léon [Citation4] provided some sufficient conditions on its global stability by using the Lyapunov direct method and a geometric method based on compounded matrices.

On the one hand, the incidence in (Equation1) is the bilinear one. However, experiments [Citation7] have strongly suggested that the incidence is an increasing function of the parasite dose, and is usually sigmoidal in shape (see, for example, [Citation30]). As a result, Hesaaraki and Sabzevari [Citation11] modified (Equation1) by replacing βxv with βxv1+av. Using the same approaches of Buonomo and Vargas-De-Léon [Citation4], they obtained sufficient conditions on the global stability of the equilibria of the resulting system. On the other hand, in most virus infections, cytotoxic T lymphocytes (CTLs) play a critical role in antiviral defence by attacking the productively infected cells. In [Citation18], Lv et al. modified (Equation1) to study the following system with Beddington-DeAngelis incidence and CTL immune response, (2) {x˙=λμxβxv1+mx+nv+δw,w˙=βxv1+mx+nv(δ+η+q)w,y˙=qwαypyz,v˙=σyγv,z˙=kyzcz.(2) Here z is the concentration of CTLs; p is the rate of productively infected cells removed by CTLs; k is the number of virions produced by a productively infected cell; c is the decay rate of CTLs; and the meanings of the other parameters are the same as those in Table . With the Lyapunov direct method, established are conditions on the global stability of equilibria in terms of the basic reproduction number and the immune response reproduction number. Later on, Maziane et al. [Citation20] considered a variation of (Equation2) by replacing βxv1+mx+nv with βxv1+α1x+α2v+α3xv and similar results were derived.

Motivated by the above discussion, in this paper, we investigate the following HIV-1 infection model with eclipse phase and CTL immune response, (3) {x˙(t)=λμx(t)x(t)f(v(t))+δw(t),w˙(t)=x(t)f(v(t))(δ+η+q)w(t),y˙(t)=qw(t)αy(t)py(t)z(t),v˙(t)=σy(t)γv(t),z˙(t)=ky(t)z(t)cz(t).(3) Here the meanings of the parameters are the same as before. The function f in the nonlinear incidence xf(v) is continuously differentiable on R+ and satisfies

  1. f(0)=0;

  2. f(v)>0 for v0;

  3. f is concave down on R+.

Biologically, assumptions (H1)–(H3) mean that the disease transmission rate is monotonically increasing, but subject to saturation effects. Note that (H1) and (H2) combined imply that f(v)>0 for v>0. It is clear that incidence functions such as the bilinear incidence with f(v)=βv and the Holling-type II incidence with f(v)=βv1+αv satisfy (H1)–(H3).

The main results of this paper are about the global stability of equilibria of (Equation3). The paper is organized as follows. In Section 2, we give some preliminary results of (Equation3), which include the boundedness of solutions and the existence of equilibria. The existence of equilibria is characterized by two basic reproduction numbers. Then we discuss the local stability of equilibria. It turns out that transcritical bifurcations of equilibria occur. We confirm them with the transmission rate and decay rate of CTLs as bifurcation parameters. After that, we employ the Lyapunov direction method to establish the global stability of the equilibria. Finally, we illustrate these theoretical results with numerical simulations and carry out the sensitivity analysis of the two basic reproduction numbers to the parameters. The paper ends with a brief summary and discussion.

2. Preliminaries

The initial conditions for system (Equation3) are (4) x(00,w(0)0,y(0)0,v(0)0,z(0)0.(4) With standard arguments, one can easily see that (Equation3) with the initial condition (Equation4) has a unique solution (x(t),w(t),y(t),v(t),z(t)) on R+. Moreover, the solution stays nonnegative on R+.

For any solution of (Equation3), we have x˙+w˙+y˙+pkz˙=λμxηwαypckzλb(x+w+y+pkz),where b=min{c,α,η,μ}. It follows that (5) lim supt(x+w+y+pkz)λb(5) and if x(t0)+w(t0)+y(t0)+pkz(t0)λb for some t0R+ then x(t)+w(t)+y(t)+pkz(t)λb for all tt0. The fourth equation of system (Equation3) together with (Equation5) implies that lim suptvλσbγ.

Proposition 2.1

The set Γ defined by Γ={(x,w,y,v,z)R+5:x+w+y+pkzλb,vλσbγ}is an attracting and positively invariant set of (Equation3).

Proof.

The discussion before the statement of Proposition 2.1 tells us that Γ attracts every solution of (Equation3). Now, let (x(0),w(0),y(0),v(0),z(0))Γ. Then x(t)+w(t)+y(t)+pkz(t)λb for every tR+ as mentioned before. We claim that v(t)λσbγ for tR+. If this is not true, then there exists t0R+ such that v(t0)>λσbγ. Let t1=inf{st0:v(t)>λσbγon[s,t0]}. It follows that v(t1)=λσbγ. Obviously, there exists t2(t1,t0) such that v˙(t2)>0. However, v˙(t2)=σy(t2)γv(t2)<σλbγλσbγ=0,a contradiction. This proves the claim and hence Γ is positively invariant.

Now, we consider the existence of equilibria of (Equation3).

System (Equation3) always has an infection-free equilibrium E0=(x0,0,0,0,0), where x0=λμ. In fact, E0 is the only infection-free equilibrium. Employing the next-generation matrix method in van den Driessche and Watmough [Citation33], one can easily get the expression of the basic infection reproduction number R0 of system (Equation3), R0=λσqμαγ(δ+η+q)f(0).An equilibrium of (Equation3) satisfies kyzcz = 0 and hence either z = 0 or y=ck.

We first consider the case where z = 0, which gives immune-free equilibria. At an immune-free equilibrium, we have (6a) λμxxf(v)+δw=0,(6a) (6b) xf(v)(δ+η+q)w=0,(6b) (6c) qwαy=0,(6c) (6d) σyγv=0.(6d) Adding (Equation6a) and (Equation6b) yields x=λ(η+q)wμ,which also implies that w<λη+q. Furthermore, it follows from (Equation6c) and (Equation6d), respectively that y=qwαandv=σqwαγ.Substituting the expressions of x and v above into (Equation6b), we see that (7) g(w)λ(η+q)wμf(σqαγw)(δ+η+q)w=0.(7) Clearly, g(0)=0, which gives the infection-free equilibrium E0. In the following, we consider infected immune-free equilibria. Note that g(0)=λσqμαγf(0)(δ+η+q)=(δ+η+q)(R01).If R0>1, then g(0)>0. This, combined with g(0)=0, implies that g(w)>0 for w>0 sufficiently small. Noting g(λη+q)=λ(δ+η+q)η+q<0, we see that g has at least one positive zero in (0,λη+q), which gives an infected immune-free equilibrium. Note that if E1=(x1,w1,y1,v1,0) is an infected immune-free equilibrium, then g(w1)=η+qμf(v1)+σqαγx1f(v1)(δ+η+q)=η+qμf(v1)+σqαγx1f(v1)σqx1f(v1)αγv1=η+qμf(v1)+σqx1αγv1(v1f(v1)f(v1))<0.To obtain the last inequality, we have used assumption (H3) that f is concave down. This fact immediately tells us that there can be at most one infected immune-free equilibrium and hence there is a unique infected immune-free equilibrium, denoted by E1=(x1,w1,y1,v1,0), when R0>1. Now assume that R01. Then g(0)0. This, together with g(0)=0 and g(0)=η+qμf(0)+λ(σq)2μ(αγ)2f(0)<0,tells us that g(w)<0 for w>0 sufficiently small. We claim that g has no zero in (0,λη+q). Otherwise, at the first zero of g, the derivative of g is larger than or equal to zero, a contradiction to the fact that the derivative is less than zero. Therefore, if R01, then there is no infected immune-free equilibrium.

Next, we consider the case where y=ck and z0, that is, infected equilibria with immunity. In this case, an equilibrium satisfies (8a) λμxxf(v)+δw=0,(8a) (8b) xf(v)(δ+η+q)w=0,(8b) (8c) qwαckpckz=0,(8c) (8d) σckγv=0.(8d) Clearly, (Equation8d) gives v=σckγ, and as before x=λ(η+q)wμ from (Equation8a) and (Equation8b). Substituting both into (Equation8b) yields (9) w=λf(σcγk)(η+q)f(σcγk)+μ(δ+η+q)(9) and hence x=λ(δ+η+q)(η+q)f(σcγk)+μ(δ+η+q). From (Equation9) and (Equation8c), we see that there is an infected equilibrium with immunity (and if exists then there is a unique one) if and only if R1>1, where R1=kqλf(σcγk)αc[(η+q)f(σcγk)+μ(δ+η+q)].In this case, z=αp(R11). R1 is called the basic immunity reproduction number.

Note that R1<R0 as shown below. From the concavity of f, we know that f(v)vf(0) for all v0. Thus R1=λkqf(v)μαc(δ+η+q)+αc(η+q)f(v)λkqf(0)vμαc(δ+η+q)+αc(η+q)f(v)<λσqf(0)μαγ(δ+η+q)=R0.R1<R0 implies that, for an infected equilibrium with immunity to exist, it is necessary that an infected immunity-free equilibrium exists. This agrees with the biological process. Actually, one can obtain R1 in the same manner as for R0 by the next generation matrix method at E1.

In summary, we have shown the following result on the existence of equilibria.

Theorem 2.1

  1. If R01, then (Equation3) only has the infection-free equilibrium E0.

  2. If R11<R0, then besides E0, (Equation3) also has an infected immune-free equilibrium E1=(x1,w1,y1,v1,0), where x1=λ(η+q)w1μ, y1=qw1α, v1=σqw1αγ, and w1 is the unique positive zero of g defined by (Equation7) in (0,λη+q).

  3. If 1<R1(<R0), then in addition to E0 and E1, there is a unique infected equilibrium with immunity E2=(x2,w2,y2,v2,z2) for (Equation3), where x2=λ(δ+η+q)(η+q)f(σcγk)+μ(δ+η+q), w2=λf(σcγk)(η+q)f(σcγk)+μ(δ+η+q), y2=ck, v2=σcγk, and z2=αp(R11).

Note that (10) g(cαqk)=cα[(η+q)f(cσγk)+μ(δ+η+q)](R11)μqk.(10) According to the proof of Theorem 2.1, we see that R1<1 (=1, >1) if w1<cαqk (w1=cαqk, w1>cαqk, respectively). As y1=qw1α, we easily obtain the following result.

Proposition 2.2

The basic immunity reproduction number R1<1 (=1, >1) if c>ky1 (c=ky1, c<ky1, respectively).

3. Stability and bifurcation analysis

3.1. Local stability of equilibria

We start with the local stability of equilibria through linearization.

Theorem 3.1

The following statements on the stability of the equilibria E0, E1, and E2 of (Equation3) obtained in Theorem 2.1 are true.

  1. The infection-free equilibrium E0 is locally asymptotically stable if R0<1 while unstable if R0>1.

  2. The infected immune-free equilibrium E1 is locally asymptotically stable if R1<1<R0 while unstable if R1>1.

  3. The infected equilibrium with immunity E2 is locally asymptotically stable when it exists, i.e. when R1>1.

Proof.

(i) The characteristic equation of the Jacobian matrix of (Equation3) at E0, J(E0)=(μδ0x0f(0)00(δ+η+q)0x0f(0)00qα0000σγ00000c),is (11) (ξ+μ)(ξ+c)(ξ3+a1ξ2+a2ξ+a3)=0,(11) where a1=α+δ+η+q+γ>0,a2=(α+δ+η+q)γ+α(δ+η+q)>0,a3=αγ(δ+η+q)[1λσqf(0)μαγ(δ+η+q)]=αγ(δ+η+q)(1R0).As ξ=μ and ξ=c are two negative eigenvalues of (Equation11), the local stability of E0 is determined by the roots of (12) ξ3+a1ξ2+a2ξ+a3=0.(12) First, assume R0<1. Then a3>0 and a1a2a3=(α+δ+η+q+γ)[(α+δ+η+q)γ+α(δ+η+q)]αγ(δ+η+q)(1R0)=(α+δ+η+q+γ)(α+δ+η+q)γ+α(δ+η+q)(α+δ+η+q)+αγ(δ+η+q)R0>0.According to the Routh-Hurwitz criterion, all roots of (Equation12) have negative real parts and hence E0 is locally asymptotically stable. Now assume that R0>1. Then a3<0. By the Intermediate Value Theorem, (Equation12) has a positive root, which implies that E0 is unstable if R0>1.

(ii) This time, the Jacobian matrix at E1 is J(E1)=(μf(v1)δ0x1f(v1)0f(v1)(δ+η+q)0x1f(v1)00qα0py100σγ00000ky1c)with the characteristic equation being (13) (ξ+cky1)(ξ4+b1ξ3+b2ξ2+b3ξ+b4)=0,(13) where b1=μ+δ+η+q+α+γ+f(v1)(>0),b2=αγ+(α+γ)(μ+η+q+f(v1))+(μ+f(v1))(η+q)+δ(α+γ+μ)(>0),b3=αγ(μ+η+q+f(v1))+(α+γ)(μ+f(v1))(η+q)+δ(αγ+μγ+αμ)qσx1f(v1),b4=αγ(μ+f(v1))(η+q)+αγδμμqσx1f(v1).By Proposition 2.2, ξ=ky1c>0 is an eigenvalue of J(E1) when R1>1 and hence E1 is unstable if R1>1. Now assume that R1<1. Then the eigenvalue ξ=ky1c<0 and the stability of E1 is determined by the following equation, (14) ξ4+b1ξ3+b2ξ2+b3ξ+b4=0.(14) Using x1f(v1)=(δ+η+q)ω1, ω1=αγqσv1, and the assumption on f, we can obtain qσx1f(v1)=αγ(δ+η+q)v1f(v1)f(v1)αγ(δ+η+q),which implies that b3>0 and b4>0. Furthermore, a simple calculation gives b1b2b3>0. To apply the Routh-Hurwitz criterion, we need to check that b3(b1b2b3)b4b12>0. For simplicity of notation, let L=μ(δ+η+q)+(η+q)f(v1),M=μ+(δ+η+q)+f(v1),N=qσx1f(v1).Then Nαγ(δ+η+q)<αγM and b3(b1b2b3)b4b12=[(α+γ)L+αγMN]×[ML+(α+γ)M2+(α+γ)2M+αγ(α+γ)+N](αγLμN)(α+γ+M)2=(α+γ)L2M+(α+γ)LN+μM2N+2μ(α+γ)MN+μ(α+γ)2N+LM[(α2+γ2)M+2αγMN]+(α+γ)LM[(α+γ)22αγ]+[(α+γ)M2+(α+γ)2M+(α+γ)αγ+N](αγMN)>0.By the Routh-Hurwitz criterion, all roots of (Equation14) have negative real parts and hence E1 is locally asymptotically stable if R1<1.

(iii) The Jacobian matrix at E2 is given by J(E2)=(μf(v2)δ0x2f(v2)0f(v2)(δ+η+q)0x2f(v2)00qαpz20py200σγ000kz20ky2c),whose characteristic equation is (15) ξ5+d1ξ4+d2ξ3+d3ξ2+d4ξ+d5=0,(15) where d1=μ+δ+η+q+αR1+γ+f(v2)(>0),d2=αc(R11)+αγR1+(αR1+γ)[μ+δ+η+q+f(v2)]+μ(δ+η+q)+f(v2)(η+q)(>0),d3=αcγ(R11)+[μ(δ+η+q)+f(v2)(η+q)](αR1+γ)+[μ+δ+η+q+f(v2)][αc(R11)+αγR1]qσx2f(v2),d4=αcγ(R11)[μ+δ+η+q+f(v2)]+[μ(δ+η+q)+f(v2)(η+q)][αc(R11)+αγR1]μqσx2f(v2),d5=αcγ(R11)[μ(δ+η+q)+f(v2)(η+q)](>0).For the convenience of notation, denote A=μ(δ+η+q)+(η+q)f(v2),B=μ+(δ+η+q)+f(v2),D=qσx2f(v2).According to the concavity assumption on f, we have Dαγ(δ+η+q)R1<αγR1B. Then it is easy to see that d3>0 and d4>0. To apply the Routh-Hurwitz criterion, we need to further check that d1d2d3>0, d3(d1d2d3)d4d12>0, and (d1d2d3)(d3d4d2d5)(d1d4d5)2>0.

It follows from R1>1 that d1d2d3=[αc(R11)+αγR1]B+(αR1+γ)B2+AB+αc(R11)(αR1+γ)+αγR1(αR1+γ)+(αR1+γ)2B+(αR1+γ)Aαcγ(R11)(αR1+γ)A[αc(R11)+αγR1]B+D=αc(R11)αR1+αγR1(αR1+γ)+AB+(αR1+γ)2B+(αR1+γ)B2+D>0.Also note that B22A>0. Then d3(d1d2d3)d4d12={αcγ(R11)+(αR1+γ)A+[αc(R11)+αγR1]BD}×[αc(R11)αR1+αγR1(αR1+γ)+AB+(αR1+γ)2B+(αR1+γ)B2+D][αcγ(R11)B+(αc(R11)+αγR1)AμD][(αR1+γ)+B]2=α3c2γ(R11)2R1+α2γ2cR1(R11)(αR1+γ)+αcγ(R11)D+(αR1+γ)αc(R11)γA+(αR1+γ)A2B+(αR1+γ)AD+α2c2(R11)2αR1B+αc(R11)BD+(αR1+γ)2μD+2(αR1+γ)μBD+μB2D+αc(R11)αR1B(B22A)+αc(R11)αR1B2+AB(αR1+γ)[(αR1+γ)22αγR1]+AB[(αR1+γ)2BD]+(αγR1BD)[αc(R11)αR1+αγR1(αR1+γ)+(αR1+γ)2B+(αR1+γ)B2+D]>0.Similarly, we can obtain (d1d2d3)(d3d4d2d5)(d1d4d5)2>0. As the writing is long and complex, we omit the detail here. Therefore, by the Routh-Hurwitz criterion, all roots of (Equation15) have negative real parts and hence the infected equilibrium with immunity E2 is locally asymptotically stable when exists. This completes the proof.

From the proof of Theorem 3.1, we can observe that

  1. for E0, when R01, besides the simple eigenvalue 0 when R0=1 and a positive eigenvalue when R0>1, the other four eigenvalues of (Equation11) all have negative real parts;

  2. for E1, when R11, in addition to the simple eigenvalue 0 when R1=1 and a positive eigenvalue ky1c when R1>1, the other four eigenvalues of (Equation15) all have negative real parts.

This observation means that there is only transcritical bifurcation of equilibria and no Hopf bifurcation can occur. In what follows, we investigate the bifurcation analysis around the equilibria when R0=1 and R1=1, which is based on the centre manifold theory [Citation5, Citation6] and Sotomayor's theorem [Citation31].

3.2. Transcritical bifurcations

For readers' convenience, we first cite a theoretical result on local dynamics near equilibria.

Consider the following system of autonomous differential equations, (16) dxdt=f(x,θ),f:Rn×RandfC2(Rn×R),(16) where θ is a bifurcation parameter and x=(x1,x2,,xn)Rn. Without loss of generality, it is assumed that 0 is an equilibrium of (Equation16) for all values of the parameter θ, that is f(0,θ)0 for all θ.

Lemma 3.1

Castillo-Chavez and Song [Citation5]

Suppose the Jacobian matrix A=Dxf(0,0) of (Equation16) at 0 with θ=0 satisfies the following two hypotheses.

  1. Zero is a simple eigenvalue of A and all other eigenvalues of A have negative real parts;

  2. A has a nonnegative right eigenvector w and a left eigenvector v corresponding to the zero eigenvalue.

Let fk be the kth component of f and a=k,i,j=1nvkwiwj2fkxixj(0,0),b=k,i=1nvkwi2fkxiθ(0,0).Then the local dynamics of (Equation16) around 0 is totally determined by a and b as follows.

  1. Suppose a>0 and b>0. When θ<0 with |θ|1, 0 is locally asymptotically stable and there exists a positive unstable equilibrium; when 0<θ1, 0 is unstable and there exists a negative and locally asymptotically stable equilibrium;

  2. Suppose a<0 and b<0. When θ<0 with |θ|1, 0 is unstable; when 0<θ1, 0 is locally asymptotically stable and there exists a positive unstable equilibrium;

  3. Suppose a>0 and b<0. When θ<0 with |θ|1, 0 is unstable and there exists a locally asymptotically stable negative equilibrium; when 0<θ1, 0 is stable and a positive unstable equilibrium appears;

  4. Suppose a<0 and b>0. When θ changes from negative to positive, 0 changes its stability from stable to unstable. Correspondingly a negative unstable equilibrium becomes positive and locally asymptotically stable.

Recall that R0=λσqμαγ(δ+η+q)f(0).As incidence is a very important factor in infection dynamics, we choose the related transmission rate as the bifurcation parameter to analyse the bifurcation at the infection-free equilibrium E0. For this purpose, we rewrite f(v)=βh(v), where h satisfies the same assumptions on f with h(0)=1. With R0=1, we get β=μαγ(δ+η+α)λσqβ.

Theorem 3.2

System (Equation3) has a forward transcritical bifurcation at the infection-free equilibrium E0 when β=β.

Proof.

To apply the centre manifold theory, let U=(u1,u2,u3,u4,u5)T, where x=u1, w=u2, y=u3, v=u4, z=u5, Then (Equation3) can be rewritten as (17) dUdt=F(U)=(F1,F2,F3,F4,F5)T,(17) where F(U)=λμu1(t)βu1(t)h(u4(t))+δu2(t),F2(U)=βu1(t)h(u4(t))(δ+η+q)u2(t),F3(U)=qu2αu3(t)pu3(t)u5(t),F4(U)=σu3(t)γu4(t),F5(U)=ku3(t)u5(t)cu5(t).Now the Jacobian matrix of system (Equation17) evaluated at E0 and the critical value β is J(E0,β)=(μδ0βλμ00(δ+η+q)0βλμ00qα0000σγ00000c).It is easy to see that J(E0,β) has zero as a simple eigenvalue and the other eigenvalues are real and negative.

A right eigenvector w=(w1,w2,w3,w4,w5)T of the Jacobian matrix J(E0,β) associated with the zero eigenvalue satisfies {μw1+δw2βλμw4=0,(δ+η+q)w2+βλμw4=0,qw2αw3=0,σw3γw4=0,cw5=0.Letting w4=1, we can get a right eigenvector given by w=(αγ(η+q)μσq,αγσq,γσ,1,0)T. Furthermore, let v be the left eigenvector of J(E0,β) associated with the zero eigenvalue satisfying vw=1. Then {μv1=0,δv1(δ+η+q)v2+qv3=0,αv3+σv4=0,βλh(0)μv1+βλh(0)μv2γv4=0,cv5=0.It follows that v=(0,σq(α+γ)(δ+η+q)+αγ,σ(δ+η+q)(α+γ)(δ+η+q)+αγ,α(δ+η+q)(α+γ)(δ+η+q)+αγ,0).

Evaluating the non-zero partial derivatives of Fi at E0, we obtain 2F2u1u4=2F2u4u1=β,2F2u42=βλμh(0),2F2u4β=λμ.By Lemma 3.1, the coefficients a and b for system (Equation17) are given by a=k,i,j=15vkwiwj2Fkuiuj=2v2w1w42F2u1u4+v2w422F2u42=2αγ(η+q)β[(α+γ)(δ+η+q)+αγ]μ+qλσβh(0)[(α+γ)(δ+η+q)+αγ]μand b=k,i=15vkwi2Fkuiβ=v2w42F2u4β=qλσ[(α+γ)(δ+η+q)+αγ]μ.By the assumptions on h, we have a<0 and b>0. Therefore, the system (Equation3) undergoes a transcritical forward bifurcation at the infection-free equilibrium E0.

By Proposition 2.2, R1=1 when c=ky1c. As a result, with c as a bifurcation parameter, we can get the following result.

Theorem 3.3

System (Equation3) undergoes a transcritical bifurcation at the infected immune-free equilibrium E1 when c=c.

Proof.

Recall that the Jacobian matrix at E1 when c=c is J(E1,c)=(μf(v1)δ0x1f(v1)0f(v1)(δ+η+q)0x1f(v1)00qα0py100σγ000000),which has a simple eigenvalue zero and the other four eigenvalues all have negative real parts. As in the proof of Theorem 3.2, we can show the right eigenvector W=((η+q)(αγ+py1)qμ,αγ+py1q,γ,σ,1)T and the left eigenvector V=(0,0,0,0,1) satisfy VW=1. With (Equation17), we have Fc(E1,c)=(0,0,0,0,0)T. Then VFc(E1,c)=0, which implies that system (Equation3) has no saddle-node bifurcation at E1. Now, DFc(E1,c)W=(00001)and hence V[DFc(E1,c)W]=1<0. Furthermore, we have D2F(E1,c)(W,W)=(xf(v)v422f(v)v1v4xf(v)v42+2f(v)v1v42pv3v502kv3v5).Then V[D2F(E1,c)(W,W)]=2kv3v5>0. With the help of Sotomayor's theorem[Citation31], we know that system (Equation3) has a transcritical bifurcation near E1 when c=c.

3.3. Global stability of equilibria

In this subsection, we study the global stability of equilibria by the Lyapunov direct method. To construct an appropriate Lyapunov function, we need the Volterra-type function h:(0,)xx1lnx. Note that h(x)0 and h(x)=0 if and only if x = 1.

We first study the global stability of E0.

Theorem 3.4

Suppose R01. Then the infection-free equilibrium E0 is globally attractive in R+5.

Proof.

First note that for any solution of (Equation3), one can easily see that x(t)>0 for t>0. Thus the Lyapunov function V1=x0h(xx0)+w+δ2(μ+η+q)x0((xx0)+w)2+δ+η+qqy+α(δ+η+q)σqv+p(δ+η+q)kqzis well defined on solutions of (Equation3) (without loss of generality). Substituting λ=μx0, we can compute the derivative of V1 along solutions of (Equation3) to get V1˙=xx0xx˙+w˙+δ(μ+η+q)x0((xx0)+w)(x˙+w˙)+δ+η+qqy˙+α(δ+η+q)σqv˙+p(δ+η+q)kqz˙=xx0x(λμxxf(v)+δw)+(xf(v)(δ+η+q)w)+δ(μ+η+q)x0((xx0)+w)(λμx(η+q)w)+δ+η+qq(qwαypyz)+α(δ+η+q)σq(σyγv)+p(δ+η+q)kq(kyzcz)=xx0x[μ(xx0)+δw]xf(v)(1x0x)+xf(v)+δ(μ+η+q)x0[μ(xx0)2(μ+η+q)w(xx0)(η+q)w2]αγ(δ+η+q)σqvpc(δ+η+q)kqz.Noting xx0x=(xx0)2xx0+xx0x0 and f(v)vf(0) for v0, we have V1˙=(μx0+δw+μδxμ+η+q)(xx0)2xx0δ(η+q)w2(μ+η+q)x0pc(δ+η+q)kqz+αγ(δ+η+q)σq(λσqf(v)μαγ(δ+η+q)v1)v(μx0+δw+μδxμ+η+q)(xx0)2xx0δ(η+q)w2(μ+η+q)x0pc(δ+η+q)kqz+αγ(δ+η+q)σq(R01)v.Since R01, it follows that V1˙0. Moreover, V1˙=0 if and only if x=x0 and w = z = 0. Then it is easy to see that the largest invariant set in {V1˙=0} is the singleton {E0}. By LaSalle's invariance principle [Citation16], the infection-free equilibrium E0 is globally attractive.

Next, denote X0={(x,w,y,v,z)R+5:w+y+z>0}.It is easy to see that every solution of (Equation3) with the initial condition in R+5X0 tends to E0 as t. Moreover, all the components except z of every solution of (Equation3) with the initial condition in X0 are positive for all t>0.

Theorem 3.5

Suppose R1<1<R0. If μx1δw1, then the infected immune-free equilibrium E1 is globally attractive in X0.

Proof.

Define the Lyapunov function V2=x1h(xx1)+w1h(ww1)+δ2(μ+η+q)x1[(xx1)+(ww1)]2+x1f(v1)qw1y1h(yy1)+αx1f(v1)σqw1v1h(vv1)+px1f(v1)kqw1z.As argued before the statement of the theorem, without loss of generality, we can assume that V2 is well-defined on every solution of (Equation3) with initial condition in X0. Note that λ=μx1+x1f(v1)δw1=μx1+(η+q)w1,(δ+η+q)=x1f(v1)w1.Then similarly as for the proof of Theorem 3.4, we can calculate the derivative of V2 along solutions of (Equation3), V2˙=xx1xx˙+ww1ww˙+δ(μ+η+q)x1[(xx1)+(ww1)](x˙+w˙)+x1f(v1)qw1yy1yy˙+αx1f(v1)σqw1vv1vv˙+px1f(v1)kqw1z˙=xx1x[μ(xx1)xf(v)+x1f(v1)+δ(ww1)]+ww1w[xf(v)x1f(v1)ww1]+x1f(v1)qw1yy1y(qwαypyz)+δ(μ+η+q)x1[(xx1)+(ww1)][μ(xx1)(η+q)(ww1)]+αx1f(v1)σqw1vv1v(σyγv)+px1f(v1)kqw1(kyzcz)=μ(xx1)2x+x1f(v)+x1f(v1)(xx1)x+δ(ww1)xx1xw1wxf(v)+x1f(v1)μδ(xx1)2(μ+η+q)x1δ(ww1)(xx1)x1δ(η+q)(ww1)2(μ+η+q)x1x1y1f(v1)qw1y(qwαypyz)αγx1f(v1)σqw1vαv1x1f(v1)σqvw1(σyγv)px1f(v1)ckqw1z.Using qw1=αy1, σy1=γv1, and xx1x=(xx1)2xx1+xx1x1, we can obtain V2˙=(μx1+δwδw1+μδxμ+η+q)(xx1)2xx1δ(η+q)(μ+η+q)x1(ww1)2+x1f(v1)[4x1xy1ww1yvv1v1yvy1+f(v)f(v1)w1xf(v)wx1f(v1)]+px1f(v1)qw1(y1ck)z.Note that x1x+y1ww1y+v1yvy1+w1xf(v)wx1f(v1)4v1f(v)vf(v1)4=4+lnv1f(v)vf(v1)+4h(v1f(v)vf(v1)4)4lnvf(v1)v1f(v)=5vf(v1)v1f(v)+h(vf(v1)v1f(v))5vf(v1)v1f(v)and the concavity of f implies that f(v)v is decreasing. It follows that 4x1xy1ww1yvv1v1yvy1+f(v)f(v1)w1xf(v)wx1f(v1)f(v)f(v1)+vf(v1)v1f(v)vv11=(1f(v1)f(v))(f(v)f(v1)vv1)0.By Proposition 2.2 and μx1σw1, we have V2˙0. It is easy to see that V2˙=0 if x=x1, w=w1, y=y1 and v=v1. Then the largest invariant set in {V˙2=0} is the singleton {E1}. By the LaSalle's invariance principle, the infected immune-free equilibrium E1 is globally attractive in X0.

Theorem 3.1 combined with Theorem 3.4 implies that the infection-free equilibrium E0 is globally asymptotically stable if R0<1 while Theorem 3.1 combined with Theorem 3.5 implies that the infected immune-free equilibrium E1 is globally asymptotically stable if R1<1<R0 and μx1δw1.

Lastly, we consider the stability of the infected equilibrium with immunity E2. For this purpose, we denote X00={(x,w,y,v,z)X0:z>0}.

Theorem 3.6

Suppose R1>1 and μx2δw2. Then the infection equilibrium with immunity E2 is global asymptotically stable in X00.

Proof.

We first note that any solution of (Equation3) with initial condition in R+5X00 tends to E0 or E1 as t and every component of the solution with the initial condition in X00 is positive for t>0. Thus we define the Lyapunov function V3=x2h(xx2)+w2h(ww2)+δ2(μ+η+q)x2[(xx2)+(ww2)]2+x2f(v2)qw2y2h(yy2)+x2f(v2)(α+qz2)σqw2v2h(vv2)+px2f(v2)kqw2z2h(zz2),which, without loss of generality, is assumed to be well-defined on solutions with initial conditions in X00. Using λ=μx2+x2f(v2)δw2,x2f(v2)=(δ+η+q)w2,qw2=αy2+py2z2,σy2=γv2,y2=ck,and similar calculations as those in the proof of Theorem 3.5, we can get the derivative of V2 along solutions of (Equation3), V3˙=(μx2+δwδw2+μδxμ+η+q)(xx2)2xx2δ(η+q)(μ+η+q)x2(ww2)2+x2f(v2)[4x2xy2ww2yvv2v2yvy2+f(v)f(v2)w2xf(v)wx2f(v2)](μx2+δwδw2+μδxμ+η+q)(xx2)2xx2δ(η+q)(μ+η+q)x2(ww2)2+x2f(v2)(1f(v2)f(v))(f(v)f(v2)vv2)0.Moreover, the equality holds if and only if x=x2, w=w2, y=y2, and v=v2. This will immediately imply that the largest invariant set in {V˙3=0} is the singleton {E2}. Again, with the help of the LaSalle's invariance principle and Theorem 3.1, we know that E2 is globally asymptotically stable in X00.

We mention that though there is no Hopf bifurcation, we can only establish the global stability of the infected immune-free equilibrium E1 and the infected equilibrium with immunity E2 under technical conditions μx1δw1 and μx2δw2, respectively. This is very common in the literature for such similar models. How to relax these constraints to obtain the result that local stability implies global stability is challenging. We are still working on it.

4. Numerical simulations

In this section, we carry out some numerical simulations to demonstrate the stability results obtained in Section 3. For this purpose, we take f(v)=βv1+ρv with ρ0, namely, we consider (18) {x˙=λμxβxv1+ρv+δw,w˙=βxv1+ρv(δ+η+q)w,y˙=qwαypyz,v˙=σyγv,z˙=kyzcz.(18) System (Equation18) always has the infection-free equilibrium E0=(λμ,0,0,0,0). We can easily get the basic infection reproduction number R0=λσβqμαγ(δ+η+q).If R0>1, then, in addition to E0, system (Equation18) also has a unique infected immune-free equilibrium E1=(x1,w1,y1,v1,0), where x1=λ(η+q)w1μ,w1=μαγ(δ+η+q)(R01)σq[ρμ(δ+η+q)+β(η+q)],y1=qw1α,v1=qσw1αγ.Also one can deduce that μx1δw1 is equivalent to R01+qμρλσ+αγμ(η+q)μδαγ. Moreover, the basic immune reproduction number is R1=kqλβσα[cβσ(η+q)+μ(kγ+ρcσ)(δ+η+q)].When R1>1, the unique infected equilibrium with immunity is E2=(x2,w2,y2,v2,z2), where x2=λ(kγ+cρσ)(δ+η+q)cβσ(η+q)+μ(kγ+cρσ)(δ+η+q),w2=cλβσcβσ(η+q)+μ(kγ+cρσ)(δ+η+q),y2=ck,v2=cσkγ,z2=αp(R11).One can also deduce that μx2δw2 is equivalent to μ(kγ+ρcσ)(δ+η+q)βδσc.

Firstly, we choose parameter values listed in Table . Then R0=0.9118<1. By Theorem 3.4, the infection-free equilibrium E0=(7×105,0,0,0,0) of (Equation18) is globally asymptotically stable (see Figure ). In the meanwhile, we also observe that x(t) decreases first, then tends to v0=7×105. w(t), y(t), and v(t) start with fluctuations, but in the following trend, w(t) decreases slowly and approaches zero, while y(t) and v(t) are both close to be stable first, then decrease, and finally tend to zero. z(t) decreases rapidly and coincides with the time axis.

Figure 1. The infection-free equilibrium E0 of (Equation18) is globally asymptotically stable when R0<1. See the text for the values of the parameters.

Figure 1. The infection-free equilibrium E0 of (Equation18(18) {x˙=λ−μx−βxv1+ρv+δw,w˙=βxv1+ρv−(δ+η+q)w,y˙=qw−αy−pyz,v˙=σy−γv,z˙=kyz−cz.(18) ) is globally asymptotically stable when R0<1. See the text for the values of the parameters.

Table 2. Parameters used for simulation purpose.

Secondly, we keep the values of the parameters as Table  except that δ is changed to 0.5 and k is changed to 5×104. Then R0=1.6659<1+qμρλσ+αγμ(η+q)μδαγ=2.7075, R1=0.8644<1, and c=3>ky1=2.1542. Therefore, all the assumptions of Theorem 3.5 are satisfied and hence the infected immune-free equilibrium E1=(4.2492×105,6.6286×103,4.3086×103,7.4929×105,0) is globally asymptotically stable. Figure  strongly supports this. In addition, we note that x(t), w(t), y(t), and v(t) start with severe fluctuations, and finally all tend to the stable values. While z(t) decreases rapidly and is close to zero. This means that when the virus level in the body is low at the beginning, the body continues to carry out normal immune response, and healthy cells will continue to increase. With the increase, more healthy cells will be infected and the virus level will increase, which will lead to the decrease of healthy cells.

Figure 2. The infected immune-free equilibrium E1 of (Equation3) is globally asymptotically stable when R1<1<R0. We refer to the text for details on parameter values and verification of other conditions.

Figure 2. The infected immune-free equilibrium E1 of (Equation3(3) {x˙(t)=λ−μx(t)−x(t)f(v(t))+δw(t),w˙(t)=x(t)f(v(t))−(δ+η+q)w(t),y˙(t)=qw(t)−αy(t)−py(t)z(t),v˙(t)=σy(t)−γv(t),z˙(t)=ky(t)z(t)−cz(t).(3) ) is globally asymptotically stable when R1<1<R0. We refer to the text for details on parameter values and verification of other conditions.

Thirdly, we take the parameter values listed in Table . In this case, we have R1=3.8929 and μ(kγ+ρcσ)(δ+η+q)=6.045×104βδσc=6×106. Therefore, Theorem 3.6 is applicable and we see that the infected equilibrium with immunity E2=(196.0992,1.9464,0.05,2.4997,0.1446) is globally asymptotically stable as demonstrated in Figure . Clearly, the values of x(t), w(t), y(t), v(t), and z(t) are all greater than zero. We notice that x(t) increases first and is close to be stable. w(t) fluctuates obviously at the beginning, and then is stabilized. y(t) and v(t) decrease first, then fluctuate slightly, and tend to respective positive stable values. z(t) increases first, then decreases and is close to the time axis, and finally tends towards the stable value after slight fluctuation. Unlike the previous two scenarios, z(t) no longer coincides with the time axis. It shows that the patient's body produces a lot of CTLs.

Figure 3. With the parameter values in the text, we have R1>1 and μx2δw2. Therefore, the infected equilibrium with immunity E2 of (Equation18) is globally asymptotically stable.

Figure 3. With the parameter values in the text, we have R1>1 and μx2≥δw2. Therefore, the infected equilibrium with immunity E2 of (Equation18(18) {x˙=λ−μx−βxv1+ρv+δw,w˙=βxv1+ρv−(δ+η+q)w,y˙=qw−αy−pyz,v˙=σy−γv,z˙=kyz−cz.(18) ) is globally asymptotically stable.

Table 3. Parameters used for simulation purpose.

Finally, in Figure , we see that the infected immune-free equilibrium E1=(5.7285,97.1357,9.7136,485.6787,0) is globally asymptotically stable. However, the condition μx1δw1 is not satisfied with R0=40>1, R1=0.3294<1 and 1+qμρλσ+αγμ(η+q)μδαγ=3.3. Similarly, in Figure , we observe that the infected equilibrium with immunity E2 is globally asymptotically stable. However, the condition μx2δw2 is not satisfied with R1=9.7525>1 and μ(kγ+ρcσ)(δ+η+q)βδσc=5.6955×104<0. This may suggest that the technical conditions μx1δw1 and μx2δw2 may not be necessary for the global stability of E1 and E2, respectively (Tables  and ).

Figure 4. With the parameter values in the Table , it is easy to verify that R0=40>1+qμρλσ+αγμ(η+q)μδαγ=3.3 and R1=0.3294<1. Therefore, the infected immune-free equilibrium E1 of (Equation3) is globally asymptotically stable.

Figure 4. With the parameter values in the Table 4, it is easy to verify that R0=40>1+qμρλσ+αγμ(η+q)μδαγ=3.3 and R1=0.3294<1. Therefore, the infected immune-free equilibrium E1 of (Equation3(3) {x˙(t)=λ−μx(t)−x(t)f(v(t))+δw(t),w˙(t)=x(t)f(v(t))−(δ+η+q)w(t),y˙(t)=qw(t)−αy(t)−py(t)z(t),v˙(t)=σy(t)−γv(t),z˙(t)=ky(t)z(t)−cz(t).(3) ) is globally asymptotically stable.

Figure 5. With the parameter values in the Table , it is easy to check that R1=9.7525>1 and the condition μx2δw2 fails to hold.

Figure 5. With the parameter values in the Table 5, it is easy to check that R1=9.7525>1 and the condition μx2≥δw2 fails to hold.

Table 4. Parameters used for simulation purpose.

Table 5. Parameters used for simulation purpose.

To conclude this section, we investigate the sensitivity of the basic reproduction number R0 and the basic immune reproduction R1 to all the parameters. The normalized forward sensitivity index of Ri with respect to a parameter p is defined as ζp=RippRi,where i = 0, 1. Table  summarizes the sensitivity indices of each parameter value. According to these indices, we obtain the relative changes of the basic reproduction numbers when the parameters of the model change, some of which are shown in Figure . For R0, the most sensitive parameters are λ, μ, β, α, σ and γ, followed by q, δ, and η, while k, ρ, and c have no effect on R0. But for R1, things are a little different. This time, all parameters affect R1, but only λ and α are the most sensitive parameters. The parameters μ, β, σ, γ, k, and c have almost the same effect. Furthermore, R1 is least sensitive to ρ.

Figure 6. Variations of the basic reproduction numbers with respect to the crucial parameters. All the parameter values are taken from Table . (a) λ. (b) α. (c) β. (d) σ. (e) η. (f) c

Figure 6. Variations of the basic reproduction numbers with respect to the crucial parameters. All the parameter values are taken from Table 2. (a) λ. (b) α. (c) β. (d) σ. (e) η. (f) c

Table 6. The sensitivity indices of the basic reproduction number to the parameters.

5. Conclusion

With the eclipse stage during HIV infection in consideration, we proposed a within-host HIV-1 infection model with a general incidence rate and CTL immune response, which can also be used to study the efficacy of treatment strategies by changing the corresponding parameters. The model is described by a system of five ordinary differential equations, which includes some in the literature as special cases, for example, those in [Citation13, Citation18]. We found that the local dynamics is completely determined by the values of the two basic reproduction numbers, the basic infection reproduction number R0 and the basic immunity reproduction number R1. If R01, there is only the infection-free equilibrium E0, which is locally asymptotically stable when R0<1; If R11<R0, besides E0, there is also a unique infected immunity-free equilibrium E1=(x1,w1,y1,v1,0), which is locally asymptotically stable when R1<1<R0; if R1>1 (in this case, it is necessarily that R0>1), in addition to E0 and E1, there is a unique infection equilibrium with immunity E2=(x2,w2,y2,v2,z2), which is locally asymptotically stable. Though we believe that the local dynamics also determines the global dynamics, we cannot confirm this so far. With the approach of Lyapunov functions, we have shown the global attractivity of the equilibria. If R01 then E0 is globally attractive and hence the infection will be cleared; if R0>1 then the infection is persistent; if R1>1 then both infection and immunity will persist. Moreover, under some technical assumptions, we established the global stability of E1 and E2. More precisely, if R1<1<R0 and μx1δw1, then E1 is globally asymptotically stable. In this case, immunity cannot be established. But if R1>1 and μx2δw2, then E2 is globally asymptotically stable and hence infection and immunity will coexist. We should emphasize that, at that moment, the global stability of E1 and E2 are established under the technical conditions μx1δw1 and μx2δw2, respectively. When δ is very small, these technical conditions are satisfied. We also conducted some experiments without these technical conditions (see Figures  and ). The numerical simulations suggest that both equilibria E1 and E2 may be globally asymptotically stable without the technical conditions. Unfortunately, we cannot confirm this yet and we leave it for future work.

We mention that some factors are ignored in our model. For example, some CTL populations can also attack infected cells during their eclipse stage [Citation27]. With these factors incorporated, the analysis will be very complicated and so far we have not obtained any practicable theoretical results. This is one of our future works. On the other hand, minimizing the loss due to infections and the cost of treatments is also an important topic in infection dynamics. We shall construct reasonable objective functionals to investigate optimal problems.

Acknowledgments

This work was supported partially by the Funding for Outstanding Doctoral Dissertation in NUAA (No. BCXJ18-09), Postgraduate Research & Practice Innovation Program of Jiangsu Province (No. KYCX18_0234), and NSERC of Canada.

Disclosure statement

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

Additional information

Funding

This work was supported by Postgraduate Research & Practice Innovation Program of Jiangsu Province [grant number KYCX18_0234] and NSERC of Canada and Funding for Outstanding Doctoral Dissertation in NUAA [grant number BCXJ18-09].

References

  • P. Aavani and L.J.S. Allen, The role of CD4 T cells in immune system activation and viral reproduction in a simple model for HIV infection, Appl. Math. Model. 75 (2019), pp. 210–222.
  • S. Ali, A.A. Raina, J. Iqbal, and R. Mathur, Mathematical modeling and stability analysis of HIV/AIDS-TB co-infection, Palest. J. Math. 8 (2019), pp. 380–391.
  • D. Biswas and S. Palb, Stability analysis of a non-linear HIV/AIDS epidemic model with vaccination and antiretroviral therapy, Int. J. Adv. Appl. Math. Mech. 5 (2017), pp. 41–50.
  • B. Buonomo and C. Vargas-De-León, Global stability for an HIV-1 infection model including an eclipse stage of infected cells, J. Math. Anal. Appl. 385 (2012), pp. 709–720.
  • C. Castillo-Chavez and B. Song, Dynamical models of tuberculosis and their applications, Math. Biosci. Eng. 1 (2004), pp. 361–404.
  • Q. Chen, Z. Teng, L. Wang, and H. Jiang, The existence of codimension-two bifurcation in a discrete SIS epidemic model with standard incidence, Nonlinear Dyn. 71 (2013), pp. 55–73.
  • D. Ebert, C.D. Zschokke-Rohringer, and H.J. Carius, Dose effects and density-dependent regulation of two microparasites of Daphania magna, Oecologia 122 (2000), pp. 200–209.
  • A.M. Elaiw and A.D. Al Agha, Stability of a general HIV-1 reaction-diffusion model with multiple delays and immune response, Phys. A 536 (2019), pp. 122593.
  • N.R. Faria, A. Rambaut, M.A. Suchard, G. Baele, T. Bedford, M.J. Ward, A.J. Tatem, J.D. Sousa, N. Arinaminpathy, J. Pépin, D. Posada, M. Peeters, O.G. Pybus, and P. Lemey, The early spread and epidemic ignition of HIV-1 in human populations, Science 346 (2014), pp. 56–61.
  • R.C. Gallo and L. Montagnier, The discovery of HIV as the cause of AIDS, N. Engl. J. Med. 349 (2003), pp. 2283–2285.
  • M. Hesaaraki and M. Sabzevari, Global properties for an HIV-1 infection model including an eclipse stage of infected cells and saturation infection, Differ. Equ. Control Process. 4 (2013), pp. 67–83.
  • A.L. Hill, D.I.S. Rosenbloom, M.A. Nowak, and R.F. Siliciano, Insight into treatment of HIV infection from viral dynamics models, Immunol. Rev. 285 (2018), pp. 9–25.
  • Z. Hu, W. Pang, F. Liao, and W. Ma, Analysis of a CD4 + T cell viral infection model with a class of saturated infection rate, Discrete Contin. Dyn. Syst. Ser. B. 19 (2014), pp. 735–745.
  • G. Kapitanov, C. Alvey, K. Vogt-Geisse, and Z. Feng, An age-structured model for the coupled dynamics of HIV and HSV-2, Math. Biosci. Eng. 12 (2015), pp. 803–840.
  • X. Lai and X. Zou, Modeling cell-to-cell spread of HIV-1 with logistic target cell growth, J. Math. Anal. Appl. 426 (2015), pp. 563–584.
  • J.P. LaSalle and S. Lefschetz, Stability by Lyapunov's Direct Method with Application, Academic Press, New York, 1961.
  • B. Li, Y. Chen, X. Lu, and S. Liu, A delayed HIV-1 model with virus waning term, Math. Biosci. Eng. 13 (2016), pp. 135–157.
  • C. Lv, L. Huang, and Z. Yuan, Global stability for an HIV-1 infection model with Beddington-DeAngelis incidence rate and CTL immune response, Commun. Nonlinear Sci. Numer. Simul. 19 (2014), pp. 121–127.
  • N.J. Malunguza, S.D. Hove-Musekwa, S. Dube, and Z. Mukandavire, Dynamical properties and thresholds of an HIV model with super-infection, Math. Med. Biol. 34 (2017), pp. 493–522.
  • M. Maziane, K. Hattaf, and N. Yousfi, Global stability of a class of HIV infection models with cure of infected cells in eclipse stage and CTL immune response, Int. J. Dynam. Control 5 (2017), pp. 1035–1045.
  • P. Ngina, R.W. Mbogo, and L.S. Luboobi, HIV drug resistance: insights from mathematical modelling, Appl. Math. Model. 75 (2019), pp. 141–161.
  • M. Nowak and C. Bangham, Population dynamics of immune responses to persistent viruses, Science272 (1996), pp. 74–79.
  • E. Numfor, Optimal treatment in a multi-strain within-host model of HIV with age structure, J. Math. Anal. Appl. 480(2) (2019), pp. 123410.
  • A. Nwankwo and D. Okuonghae, Mathematical analysis of the transmission dynamics of HIV syphilis co-infection in the presence of treatment for syphilis, Bull. Math. Biol. 80 (2018), pp. 437–492.
  • A.S. Perelson and R.M. Ribeiro, Modeling the within-host dynamics of HIV infection, BMC Biol. 11 (2013), pp. 123. DOI:https://doi.org/10.1186/1741-7007-11-96.
  • L. Rong, M.A. Gilchrist, Z. Feng, and A.S. Perelson, Modeling within-host HIV-1 dynamics and the evolution of drug resistance: trade-offs between viral enzyme function and drug susceptibility, J. Theoret. Biol. 247 (2007), pp. 804–818.
  • J.B. Sacha, C. Chung, E.G. Rakasz, S.P. Spencer, A.K. Jonas, A.T. Bean, W. Lee, B.J. Burwitz, J.J.Stephany, J.T. Loffredo, D.B. Allison, S. Adnan, A. Hoji, N.A. Wilson, T.C. Friedrich, J.D. Lifson, O.O.Yang, and D.I. Watkins, Gag-specific CD8 + T lymphocytes recognize infected cells before AIDS-virus integration and viral protein expression, J. Immunol. 178(5) (2007), pp. 2746–2754.
  • S.K. Sahani and Yashi, Effects of eclipse phase and delay on the dynamics of HIV infection, J. Biol. Syst. 26 (2018), pp. 421–454.
  • M. Shen, Y. Xiao, and L. Rong, Global stability of an infection-age structured HIV-1 model linking within-host and between-host dynamics, Math. Biosci. 263 (2015), pp. 37–50.
  • X. Song and A.U. Neumann, Global stability and periodic solution of the viral dynamics, J. Math. Anal. Appl. 329 (2007), pp. 281–297.
  • J. Sotomayor, Generic Bifurcations of Dynamical Systems, Dynamical systems, Academic Press, 1973.
  • N. Tarfulea, Drug therapy model with time delays for HIV infection with virus-to-cell and cell-to-cell transmissions, J. Appl. Math. Comput. 59 (2019), pp. 677–691.
  • P. van den Driessche and J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci. 180 (2002), pp. 29–48.
  • World Health Organization, HIV/AIDS, software. Available at https://www.who.int/news-room/fact-sheets/detail/hiv-aids.