Parameter | Value | Source of data |
Λ | 10 | [17] |
a | 0.01 | [17] |
β | 0.00034 | Assumed |
γ | 0.0001 | Assumed |
b | 0.26 | [6] |
k | 11 | [6] |
c | 0.1 | [6] |
μ1 | 0.3 | Assumed |
μ2 | 0.1 | Assumed |
Citation: Atefeh Afsar, Filipe Martins, Bruno M. P. M. Oliveira, Alberto A. Pinto. A fit of CD4+ T cell immune response to an infection by lymphocytic choriomeningitis virus[J]. Mathematical Biosciences and Engineering, 2019, 16(6): 7009-7021. doi: 10.3934/mbe.2019352
[1] | A. M. Elaiw, N. H. AlShamrani . Stability of HTLV/HIV dual infection model with mitosis and latency. Mathematical Biosciences and Engineering, 2021, 18(2): 1077-1120. doi: 10.3934/mbe.2021059 |
[2] | Baolin Kang, Xiang Hou, Bing Liu . Threshold control strategy for a Filippov model with group defense of pests and a constant-rate release of natural enemies. Mathematical Biosciences and Engineering, 2023, 20(7): 12076-12092. doi: 10.3934/mbe.2023537 |
[3] | Kangsen Huang, Zimin Wang . Research on robust fuzzy logic sliding mode control of Two-DOF intelligent underwater manipulators. Mathematical Biosciences and Engineering, 2023, 20(9): 16279-16303. doi: 10.3934/mbe.2023727 |
[4] | Yun Liu, Yuhong Huo . Predefined-time sliding mode control of chaotic systems based on disturbance observer. Mathematical Biosciences and Engineering, 2024, 21(4): 5032-5046. doi: 10.3934/mbe.2024222 |
[5] | A. M. Elaiw, N. H. AlShamrani, A. D. Hobiny . Stability of an adaptive immunity delayed HIV infection model with active and silent cell-to-cell spread. Mathematical Biosciences and Engineering, 2020, 17(6): 6401-6458. doi: 10.3934/mbe.2020337 |
[6] | Tinevimbo Shiri, Winston Garira, Senelani D. Musekwa . A two-strain HIV-1 mathematical model to assess the effects of chemotherapy on disease parameters. Mathematical Biosciences and Engineering, 2005, 2(4): 811-832. doi: 10.3934/mbe.2005.2.811 |
[7] | Shengqiang Liu, Lin Wang . Global stability of an HIV-1 model with distributed intracellular delays and a combination therapy. Mathematical Biosciences and Engineering, 2010, 7(3): 675-685. doi: 10.3934/mbe.2010.7.675 |
[8] | Andrew Omame, Sarafa A. Iyaniwura, Qing Han, Adeniyi Ebenezer, Nicola L. Bragazzi, Xiaoying Wang, Woldegebriel A. Woldegerima, Jude D. Kong . Dynamics of Mpox in an HIV endemic community: A mathematical modelling approach. Mathematical Biosciences and Engineering, 2025, 22(2): 225-259. doi: 10.3934/mbe.2025010 |
[9] | A. M. Elaiw, N. H. AlShamrani . Analysis of an HTLV/HIV dual infection model with diffusion. Mathematical Biosciences and Engineering, 2021, 18(6): 9430-9473. doi: 10.3934/mbe.2021464 |
[10] | Li Ma, Chang Cheng, Jianfeng Guo, Binhua Shi, Shihong Ding, Keqi Mei . Direct yaw-moment control of electric vehicles based on adaptive sliding mode. Mathematical Biosciences and Engineering, 2023, 20(7): 13334-13355. doi: 10.3934/mbe.2023594 |
Human immunodeficiency virus (HIV) takes the most important CD4+T lymphocytes in the human immune system as the main target, destroys people's CD4+T and makes the body lose its immune function [1]. HIV is a major problem facing the human race and poses a serious health threat to human society. While there has been remarkable advancement in the development of antiretroviral therapy (ART) and prevention strategies, currently there are still many people living with HIV [2]. From the National Health Commission of the People's Republic of China, to the end of 2020, a total of 1.053 million people in China were infected with HIV, 351,000 deaths were reported, and the number of infections is expected to increase to 1.6 million in 2022. Therefore, it is beneficial to study the effective control measures and find the optimal control by using mathematical models.
Early HIV models were primarily devoted to the virus-to-cell infection [3,4,5,6,7,8,9,10]. With the development of science and the improvement of medical standards, some studies show that virus can also spread by direct cell-to-cell transmission [11,12,13,14,15]. Thus, many researchers have begun working on an HIV model incorporating virus-to-cell and cell-to-cell transmission [16,17,18,19,20]. For example, by an HIV model, Wang et al. [16] showed the existence, positivity and boundedness of the model solution. Lai et al. [19] demonstrated global threshold dynamics by the basic reproduction number. For the control of acquired immune deficiency syndrome (AIDS), the main drug treatment is to prevent new HIV infections by blocking the transformation of viral RNA into DNA in T cells and reducing the number of viral particles [21]. With regard to AIDS control, there have already been some results [22,23,24]. Liu et al. [22], Akbari et al.[23] and Guo et al. [24] proposed an optimal control problem for an HIV infection model with cell-to-cell spread. However, these studies were focused on optimal control given over the entire infection period T, which will increase the cost of control measures. If threshold level Nt can be introduced, such that the control measures are not implemented when the total number of infected cells in an infected person is at a relatively low level (threshold level Nt), while otherwise effective measures are taken to suppress progression of viruses and infected cells, the time and cost will be reduced. The idea has been widely used in engineering [25,26], but little research has been done in epidemic diseases.
In addition, for the case that the number of infected cells in the body exceeds the threshold level (Nt), from the perspectives of epidemiology and economics, how to control the spread of virus is a valuable question. As indicated by discussion above, in this paper, we analyze the sliding mode dynamics and optimal control of the HIV model with virus-to-cell and cell-to-cell transmission. Here, because the timing of controlling infected cells and viruses is uncertain, dynamic programming is considered, which demonstrates that not only can infected cells and viruses be controlled in time, but also the goal of minimizing the concentration of infected cells and viruses with a low cost of application control can be achieved. The main novelties are summarized as follows:
● About the HIV model with cell-to-cell transmission, the majority of the existing results only discussed dynamics, this article proposes a piecewise control function concerning threshold policy and discusses sliding mode dynamics.
● Differently from previous works on optimal control over the time period [0,T] for the HIV model, in this paper, optimal control strategies for infected cells and viruses are achieved when the total number of infected cells in the body exceeds the certain tolerance threshold level (Nt). Moreover, our results generalize and improve some published results in the literature, such as [23,24].
The whole organization of this work is as follows. Section 2 describes the different components of the HIV model, and it then further extends a new three-dimensional Filippov model with two control measures. Section 3 investigates the sliding mode dynamics of the model and shows the existence of a unique positive pseudo-equilibrium. In Section 4, the optimal control problem is discussed. We first give the objective function and prove the uniqueness and existence of the viscosity solution of the HJB equation, and we then obtain the optimal control through the Hamiltonian function. The theoretical results are verified by numerical simulations in Section 5. Finally, conclusions and outlook for further work are given in Section 6.
Inspired by [17,19], the classical HIV model with virus-to-cell and cell-to-cell transmission is
{dx(t)dt=Λ−βx(t)v(t)−γx(t)y(t)−ax(t),dy(t)dt=βx(t)v(t)+γx(t)y(t)−by(t),dv(t)dt=ky(t)−cv(t), | (2.1) |
where x(t) denotes the concentration of uninfected target cells at time t, y(t) is the concentration of infected cells at time t, and v(t) denotes the concentration of virus particles at time t. Λ is the recruitment rate of healthy target cells, β is the rate at which an uninfected cell becomes infected by an infectious virus, γ represents the infection rate of productively infected cells, k shows the generation rate of virus particles, a expresses the loss rate of infected cells, b represents the natural death rate of uninfected cells, and c indicates the clearance rate of virions.
Based on the existing HIV models, we give the sliding control system to maintain the number of viruses and infected cells below the threshold level. Research [27] has found that current drug treatment consisting of five antiretroviral drugs can suppress viral replication to a low level or increase the CD4+T cell, the two main types of HIV resistance: reverse transcriptase inhibitors (RTIs) and protease inhibitors (PIs) [28,29]. RTIs prevent new HIV infections by blocking the transformation of viral RNA into DNA in T cells, and PIs reduce the number of viral particles produced by actively infected T cells [21]. We represent by μ1 the RTIs control variable and by μ2 the PIs control variable. The control system is given as follows:
{dx(t)dt=Λ−βx(t)v(t)−γx(t)y(t)−ax(t),dy(t)dt=βx(t)v(t)+γx(t)y(t)−by(t)−εμ1y(t),dv(t)dt=ky(t)−cv(t)−εμ2v(t), | (2.2) |
with
ε={0,y(t)−Nt<0,1,y(t)−Nt>0. | (2.3) |
μ1 is the culling rate of infected cells, μ2 is the clearance rate of virus particles, and they are constants. The critical level of the total number of infected cells is represented by Nt. For convenience, y(t)−Nt is defined as ω(M)=y(t)−Nt with vector M=(x(t),y(t),v(t))⊤∈R3+ and R3+={M=(x,y,v)|x≥0,y≥0,v≥0}.
Remark 2.1. When ε=0, model (2.2) becomes model (2.1) and the control measures are taken when ε=1.
In this section, according to system (2.2) with (2.3), we study the dynamics of the system. First, we define G1={M∈R3+|ω(M)<0}, G2={M∈R3+|ω(M)>0}. Furthermore, we describe the manifold Gs as Gs={M∈R3+|ω(M)=0} and the normal vector perpendicular to Gs is shown as n=(0,1,0)⊤. Then, we consider the following Filippov system:
˙M={f1(M),M∈G1,f2(M),M∈G2, | (3.1) |
where
f1(M)=(Λ−ax(t)−βx(t)v(t)−γx(t)y(t)βx(t)v(t)+γx(t)y(t)−by(t)ky(t)−cv(t)), |
f2(M)=(Λ−ax(t)−βx(t)v(t)−γx(t)y(t)βx(t)v(t)+γx(t)y(t)−by(t)−μ1y(t)ky(t)−cv(t)−μ2v(t)). |
Based on this, we present the following definitions of various equilibriums and sliding domain.
Definition 3.1. [30] If f1(M∗)=0 with ω(M∗)<0, or f2(M∗)=0 with ω(M∗)>0, the point M∗ is called a real equilibrium of system (3.1).
Definition 3.2. [30] If f1(M∗)=0 with ω(M∗)>0, or f2(M∗)=0 with ω(M∗)<0, the point M∗ is called a virtual equilibrium of system (3.1).
Definition 3.3. [30] If it is an equilibrium of the sliding mode of system (3.1), the point M∗ is called a pseudo-equilibrium.
Definition 3.4. [30] S is the sliding domain, if ⟨n,f1⟩>0 and ⟨n,f2⟩<0 on S⊂Gs.
In this subsection, we calculate the basic reproduction number and analyze the stability of equilibrium. The dynamics of system (3.1) in region G1 are indicated by
(x′(t)y′(t)v′(t))=(Λ−ax(t)−βx(t)v(t)−γx(t)y(t)βx(t)v(t)+γx(t)y(t)−by(t)ky(t)−cv(t)). | (3.2) |
In subsystem (3.2), the disease-free equilibrium E01=(Λa,0,0) and the endemic equilibrium E∗1=(x∗1,y∗1,v∗1) are given, where
x∗1=bckβ+cγ,y∗1=Λ(kβ+cγ)−acbb(kβ+cγ),v∗1=Λk(kβ+cγ)−ackbbc(kβ+cγ),R01=Λ(βk+cγ)acb. |
Next, the following theorem about the local asymptotic stability of equilibria E01 and E∗1 are given.
Theorem 3.5. The disease-free equilibrium E01 is locally asymptotically stable if R01<1; the endemic equilibrium E∗1 exists and is locally asymptotically stable if R01>1.
Proof. Using the next generation matrix, the basic reproduction number R01 of system (3.2) can be deduced, and system (3.2) can be represented as
M′=F(M)−V(M), |
where
F(M)=[βx(t)v(t)+γx(t)y(t)00],V(M)=[by(t)cv(t)−ky(t)ax(t)+βx(t)v(t)+γx(t)y(t)−Λ]. |
The Jacobian matrix of F(M) and V(M) at the equilibrium point E01 is
DF(E01)=[γΛaβΛa0000000],DV(E01)=[b00−kc0γΛaβΛaa], |
where
F=[γΛaβΛa00],V=[b0−kc]. |
Then, the generation matrix of system (3.2) is
FV−1=[Λ(βk+cγ)acbβΛac00]. |
So, the spectral radius of FV−1 is
ρ(FV−1)=Λ(βk+cγ)acb, |
where ρ is the spectral radius of a matrix. Therefore, the basic reproduction number of system (3.2) is
R01=Λ(βk+cγ)acb. |
For subsystem (3.2), The Jacobian matrix is
J1(x(t),y(t),v(t))=(−a−βv(t)−γy(t)−γx(t)−βx(t)βv(t)+γy(t)−b+γx(t)βx(t)0k−c). |
If R01<1, cb−cγΛ+kβΛa>0, c+b−γΛa>0 can be obtained, and (c+b−γΛa)2−4(cb−cγΛ+kβΛa)>0. The characteristic equation at E01 is
(λ+a)[(λ+b−γΛa)(λ+c)−kβΛa]=0, |
which indicates that
λ1=−a<0,λ2=−(c+b−γΛa)+√(c+b−γΛa)2−4(cb−cγΛ+kβΛa)2<0,λ3=−(c+b−γΛa)−√(c+b−γΛa)2−4(cb−cγΛ+kβΛa)2<0. |
Thus, the eigenvalues of J(E01) are negative when R01<1, so we can obtain that E01 is locally asymptotically stable.
If R01>1, then γcΛ+kβΛacb>1, and there exists the endemic equilibrium E∗1. Thus, the characteristic equation at E∗1 is
λ3+a1λ2+a2λ+a3=0, |
where
a1=a+c+b+βv∗1+γy∗1−γx∗1=a+c+b+βΛk(kβ+cγ)−ackbbc(kβ+cγ)+γΛ(kβ+cγ)−acbb(kβ+cγ)−γbckβ+cγ=c+a+bkβkβ+cγ>0,a2=bc+ac+ab−(γc+kβ+aγ)x∗1+(cβ+bβ)v∗1+(cγ+bγ)y∗1=bc+ac+ab−(γc+kβ+aγ)bckβ+cγ+(cβ+bβ)Λk(kβ+cγ)−ackbbc(kβ+cγ)+(cγ+bγ)Λ(kβ+cγ)−acbb(kβ+cγ)=ca+akbβkβ+cγ>0,a3=acb−(acγ+akβ)x∗1+bcβv∗1+bcγy∗1=acb−(acγ+akβ)bckβ+cγ+bcβΛk(kβ+cγ)−ackbbc(kβ+cγ)+bcγΛ(kβ+cγ)−acbb(kβ+cγ)=Λ(kβ+cγ)−acb>0. |
Moreover, we can get
a1a2−a3=[a+c+b+βv∗1+γy∗1−γx∗1][bc+ac+ab−(γc+kβ+aγ)x∗1+(cβ+bβ)v∗1+(cγ+bγ)y∗1]−[acb−(acγ+akβ)x∗1+bcβv∗1+bcγy∗1]=Λ(kβ+cγ)−acbbc(kβ+cγ)[bbkβ+akbβ+accγ+bckβ+cckβ+cccγ+cakβ+cacγ]+(cakβ+cacγ+abkβ)bkβ+ckβ+ccγ+akβ+acγ(kβ+cγ)2+2kβγ(b+c)Λ(kβ+cγ)−acbbc(kβ+cγ)⋅Λ(kβ+cγ)−acbb(kβ+cγ)+Λ(kβ+cγ)−acbbc(kβ+cγ)(ackβ+cacγ+akbβ)>0. |
According to the Routh-Hurwitz Criterion, all eigenvalues of J(E∗1) have negative real parts. Hence, this represents that E∗1 is locally asymptotic stable.
In region G2, we give the dynamics of system (3.1):
(x′(t)y′(t)v′(t))=(Λ−ax(t)−βx(t)v(t)−γx(t)y(t)βx(t)v(t)+γx(t)y(t)−by(t)−μ1y(t)ky(t)−cv(t)−μ2v(t)). | (3.3) |
In subsystem (3.3), the disease-free equilibrium E02=(Λa,0,0) and the endemic equilibrium E∗2=(x∗2,y∗2,v∗2) are given, where
x∗2=(c+μ2)(b+μ1)kβ+γ(c+μ2),y∗2=Λb+μ1−a(c+μ2)kβ+γ(c+μ2),v∗2=kΛ(c+μ2)(b+μ1)−akkβ+γ(c+μ2),R02=Λ(kβ+γ(c+μ2))a(c+μ2)(b+μ1). |
Next, the following theorem about the local asymptotic stability of equilibria E02 and E∗2 are given.
Theorem 3.6. The disease-free equilibrium E02 is locally asymptotically stable if R02<1. The endemic equilibrium E∗2 exists and is locally asymptotically stable if R02>1.
Proof. Using the next generation matrix, we deduce the basic reproduction number R02 of system (3.3):
R02=ρ([γΛaβΛa00][b+μ10−kc+μ2]−1)=Λ(kβ+γ(c+μ2))a(c+μ2)(b+μ1). |
For subsystem (3.3), the Jacobian matrix is
J2(x(t),y(t),v(t))=(−a−βv(t)−γy(t)−γx(t)−βx(t)βv(t)+γy(t)−b+γx(t)−μ1βx(t)0k−c−μ2). |
Let c+μ2=c1, b+μ1=b1. Similar to the proof of Theorem 3.5, if R02<1, then all eigenvalues of J(E02) are negative, and thus the local asymptotic stability of E02 can also be concluded. Furthermore, all eigenvalues of J(E∗2) have negative real parts, if R02>1, which shows that E∗2 is locally asymptotically stable.
Remark 3.7. The basic reproduction number R02 is related to control variables μ1 and μ2, and when μ1=μ2=0, that is, the control measures are not implemented, we have R01=R02.
In order to study the dynamics of sliding mode of system (3.1) in this subsection, we initially examine the existence of the sliding mode. The manifold Gs is defined as y(t)=Nt, and we have
⟨n,f1⟩=⟨(010),(Λ−ax(t)−βx(t)v(t)−γx(t)y(t)βx(t)v(t)+γx(t)y(t)−by(t)ky(t)−cv(t))⟩=βx(t)v(t)−bNt+γx(t)Nt=σ1(x(t),v(t)), | (3.4) |
and
⟨n,f2⟩=⟨(010),(Λ−ax(t)−βx(t)v(t)−γx(t)y(t)βx(t)v(t)+γx(t)y(t)−by(t)−μ1y(t)ky(t)−cv(t)−μ2v(t))⟩=βx(t)v(t)−bNt+γx(t)Nt−μ1Nt=σ2(x(t),v(t)), | (3.5) |
which shows that σ2(x(t),v(t))<σ1(x(t),v(t)). Because ⟨n,f1⟩>0 and ⟨n,f2⟩<0, we can obtain
⟨n,f1⟩>0,ifx(t)>bNtβv(t)+γNt=H1(v(t)),⟨n,f2⟩<0,ifx(t)<bNt+μ1Ntβv(t)+γNt=H2(v(t)), |
and
H2(v(t))>H1(v(t)). |
It is obvious that H1(v(t)) is positive on the interval (0,Nt), and so is H2(v(t)). Then, the sliding domain S⊂Gs is defined as follows:
S≜{(x(t),y(t),v(t))∈Gs:H1(v(t))<x(t)<H2(v(t)),y(t)=Nt}. | (3.6) |
Moreover, in order to obtain sliding mode equations in the region S, we use the Utkin equivalent control method, as used in [31]. First, according to ω(y(t))=y(t)−Nt=0 and dω/dt=0,
dωdt=∂ω∂y(t)dy(t)dt=βx(t)v(t)−by(t)+γx(t)y(t)−εμ1y(t)=0. | (3.7) |
By solving Eq (3.7), the function about ε is obtained:
ε(v(t))=βx(t)v(t)−by(t)+γx(t)y(t)μ1y(t)=βx(t)v(t)−bNt+γx(t)Ntμ1Nt. | (3.8) |
Substitute Eq (3.8) into system (2.2), and the dynamics of sliding mode in S can be given by differential equations as follows:
{dx(t)dt=Λ−ax(t)−βx(t)v(t)−γx(t)Nt,dy(t)dt=0,dv(t)dt=kμ1N2t−cμ1Ntv(t)+bμ2Ntv(t)−γμ2Ntx(t)v(t)−μ2βx(t)v2(t)μ1Nt. | (3.9) |
Next, we exhibit the following proposition about the existence of equilibrium of sliding mode (3.9).
Proposition 3.8. The existence of a unique positive pseudo-equilibrium E∗s=(Λa+βv∗+γNt,Nt,v∗) of sliding mode (3.9) is obtained if c−kb<μ2μ1<cb and E∗s∈S.
Proof. Let dx(t)dt=0 and dv(t)dt=0. It obtains that
x(t)=Λa+βv(t)+γNt, | (3.10) |
and
kμ1N2t−cμ1Ntv(t)+bμ2Ntv(t)−γμ2Ntx(t)v(t)−μ2βx(t)v2(t)=0. | (3.11) |
After substituting Eq (3.10) into (3.11), we get the following equation of v(t):
τ1v2(t)+τ2v(t)+τ3=0, | (3.12) |
where
τ1=βNt(bμ2−cμ1)−μ2βΛ,τ2=abμ2Nt−caμ1Nt−γΛμ2Nt+bγμ2N2t−cγμ1N2t+kβμ1N2t,τ3=akμ1N2t+kγμ1N3t,Δ=τ22−4τ1τ3=N2t[abμ2−caμ1−γΛμ2+bγμ2Nt−cγμ1Nt+kβμ1Nt]2−4(βNt(bμ2−cμ1)−μ2βΛ)[akμ1N2t+kγμ1N3t]. |
If c−kb<μ2μ1<cb, then τ1<0,τ3>0 and Δ>0 can be received. The Vieta Theorem shows
v∗1v∗2=τ3τ1<0. | (3.13) |
Due to Δ>0, the existence of the root is proven, and there are two roots, one positive root and one negative root from (3.13). We can conclude that Eq (3.12) has a unique positive root denoted by v∗. Thus, the equilibrium point E∗s can be obtained from (3.10):
E∗s=(x∗,y∗,v∗)=(Λa+βv∗+γNt,Nt,v∗). |
Furthermore, E∗s is a unique pseudo-equilibrium if E∗s∈S⊂Gs holds.
Similar to the previous theorem, it is worth considering the local asymptotical stability of E∗s, and then the following theorem is given.
Theorem 3.9. Under the same conditions of Proposition 3.8, E∗s is locally asymptotically stable on the sliding domain S.
Proof. In system (3.9), the Jacobian matrix of the first two equations is
Js(x∗,v∗)=(J11J12J21J22). | (3.14) |
From Proposition 3.8, c−kb<μ2μ1<cb, we set
P=dx(t)dt=Λ−ax(t)−βx(t)v(t)−γx(t)Nt,Q=dv(t)dt=kμ1N2t−cμ1Ntv(t)+bμ2Ntv(t)−γμ2Ntx(t)v(t)−μ2βx(t)v2(t)μ1Nt, | (3.15) |
and obtain
J11=∂P∂x|(x∗,v∗)=−a−βv∗−γNt<0,J12=∂P∂v|(x∗,v∗)=−βx∗<0,J21=∂Q∂x|(x∗,v∗)=−γμ2v∗Nt−μ2β(v∗)2μ1Nt<0,J22=∂Q∂v|(x∗,v∗)=−cμ1Nt+bμ2Nt−γμ2Ntx∗−2μ2βx∗v∗μ1Nt<0. | (3.16) |
Thus, we get
J11⋅J22−J12⋅J21>0, | (3.17) |
since J11+J22<0, and we can obtain that
J11⋅J22−J12⋅J21=[−a−βv∗−γNt]⋅−cμ1Nt+bμ2Nt−γμ2Ntx∗−2μ2βx∗v∗μ1Nt−(−βx∗)⋅−γμ2v∗Nt−μ2β(v∗)2μ1Nt=(a+βv∗+γNt)⋅(cμ1Nt−bμ2Nt)+(a+γNt)γμ2Ntx∗+(a+βv∗+γNt)μ2βx∗v∗μ1Nt>0. | (3.18) |
Therefore, all eigenvalues of (3.14) have negative real parts. That E∗s is locally asymptotically stable can be obtained easily.
Remark 3.10. The sliding domain S represents the region between the two lines associated with x(t),v(t) belonging to the plane y=Nt, and the local asymptotical stability of pseudo-equilibrium E∗s in sliding domain S is related to the control variables, which needs to satisfy the condition of the parameters.
In order to reduce the number of infected cells and viruses, while keeping the cost to apply the control at the minimum level at any time, we show an optimal control problem on the basis of system (2.2). In this section, any time refers to when the number of infected cells is greater than threshold level Nt, that is, when ε=1 in system (2.2). If the control variable in system (2.2) is a time dependent variable, how do we find the optimal control? Thus, we represent by u1(t) the RTIs control variable and by u2(t) the PIs control variable. First, we have u(t)=(u1(t),u2(t))⊤∈V[s,T]={u(⋅):[s,T]→U|u1(t)andu2(t)aremeasurable:0≤u1(t)≤1,0≤u2(t)≤1}, where U is a metric space and convex, T>0, and the control problem of the model is given by
{dx(t)dt=Λ−βx(t)v(t)−γx(t)y(t)−ax(t),dy(t)dt=βx(t)v(t)+γx(t)y(t)−by(t)−u1(t)y(t),t>0,dv(t)dt=ky(t)−cv(t)−u2(t)v(t). | (4.1) |
Let (s,M)∈[0,T)×R3, and we consider the following control system over [s,T]:
{dM(t)dt=b(t,M(t),u(t)),t∈[s,T],M(s)=M0, | (4.2) |
where M(t)=(x(t),y(t),v(t))⊤∈R3. Next, we construct the following objective functional:
J(u1(t),u2(t))=∫Ts(A1u21(t)+A2u22(t)+A3y(t)+A4v(t))dt+h(M(T)), | (4.3) |
where Ai(i=1,2,3,4) are weights to make the terms of the integrand keep balance. The term ∫Ts(A1u21(t)+A2u22(t))dt gives the total cost of using the control strategy, h(M(T)) is the penalty function corresponding to the terminal state, and ∫Ts(A3y(t)+A4v(t))dt represents the total number of infected cells and viruses over the time period T. It is important to find an optimal control pair (u∗1(t),u∗2(t)),t∈[s,T] such that
J(u∗1(t),u∗2(t))=minu1(t),u2(t)∈V[s,T]J(u1(t),u2(t)). | (4.4) |
Further, the value function is as follows:
{V(s,M0)=infu1(t),u2(t)∈V[s,T]J(s,M0;u1(t),u2(t)),∀(s,M0)∈[0,T)×R3,V(T,M0)=h(M0),∀M0∈R3. | (4.5) |
Before further study, we give the following assumption.
Assumption 1. (U,˜d) is a separable metric space.
Then, we would like to study V(⋅,⋅) in great detail and present the following results called Bellman's principle of optimality by [32].
Theorem 4.1. Let 1 hold, and U is convex. Then, for any (s,M0)∈[0,T)×R3, we have
V(s,M0)=infu1(t),u2(t)∈V[s,T]{∫ˆss(A1u21(t)+A2u22(t)+A3y(t)+A4v(t))dt+V(ˆs,M(ˆs;s,M0,u(⋅)))},∀0≤s≤ˆs≤T. | (4.6) |
Proof. Let us define
ˉV(s,M0)={∫ˆss(A1u21(t)+A2u22(t)+A3y(t)+A4v(t))dt+V(ˆs,M(ˆs;s,M0,u(⋅)))}. | (4.7) |
By (4.6), we obtain
ˉV(s,M0)≤J(s,M0;u(t))=∫ˆss(A1u21(t)+A2u22(t)+A3y(t)+A4v(t))dt+J(ˆs,M(ˆs);u(⋅))),∀u(t)∈V[s,T]. | (4.8) |
Therefore, we take the infimum over u(t)∈V[s,T] and get
V(s,M0)≤ˉV(s,M0), | (4.9) |
and there exists a uε(t)∈V[s,T] for ∀ε>0 such that
V(s,M0)+ε≥J(s,M0;uε(t))≥∫ˆss(A1u21(t)+A2u22(t)+A3y(t)+A4v(t))dt+J(ˆs,Mε(ˆs))≥ˉV(s,M0), | (4.10) |
where Mε(⋅)=M(⋅;s,M0,uε(⋅)). Combining (4.9) and (4.10), (4.6) is obvious. This completes the proof.
(4.6) is called the dynamic programming equation, and the equation is difficult to handle. Thus, we construct the Hamilton-Jacobi-Bellman (HJB) equation as follows.
Theorem 4.2. Let 1 hold, and U is convex. Suppose V∈C1([0,T]×R3). Then, V(s,M0) is a solution to the following terminal value problem of a first-order partial differential equation:
{0=−Vt+supu(t)∈UH(t,M,u(t),−VM),(t,M)∈[0,T]×R3,V|t=T=h(M),M∈R3, | (4.11) |
where
H(t,M,u(t),−VM)=−Vx(Λ−ax(t)−βx(t)v(t)−γx(t)y(t))−Vy(βx(t)v(t)+γx(t)y(t)−by(t)−u1(t)y(t))−Vv(ky(t)−cv(t)−u2(t)v(t))−A1u21(t)−A2u22(t)−A3y(t)−A4v(t). | (4.12) |
We call (4.11) the Hamilton-Jacobi-Bellman (HJB) equation associated with (4.5).
Proof. Fix a u∈U. Let M(t) be the state trajectory corresponding to the control u(t)≡u. According to (4.6) with ˆs↓s, we get
0≥−V(ˆs,M(ˆs))−V(s,M0)ˆs−s−1ˆs−s∫ˆss(A1u21(t)+A2u22(t)+A3y(t)+A4v(t))dt→−Vt(s,M0)−(Vx(Λ−ax(t)−βx(t)v(t)−γx(t)y(t))+Vy(βx(t)v(t)+γx(t)y(t)−by(t)−u1(t)y(t))+Vv(ky(t)−cv(t)−u2(t)v(t)))−(A1u21(t)+A2u22(t)+A3y(t)+A4v(t)),∀u∈U, | (4.13) |
which results in
0≥−Vt(s,M0)+supu∈U{−Vx(Λ−ax(t)−βx(t)v(t)−γx(t)y(t))−Vy(βx(t)v(t)+γx(t)y(t)−by(t)−u1(t)y(t))−Vv(ky(t)−cv(t)−u2(t)v(t))−A1u21(t)−A2u22(t)−A3y(t)−A4v(t)}. | (4.14) |
For any ε>0,0≤s≤ˆs≤T with ˆs−s>0 small enough, there exists a u=uε,ˆs(⋅)∈V[s,T] such that
V(s,M0)+ε(ˆs−s)≥∫ˆss(A1u21(t)+A2u22(t)+A3y(t)+A4v(t))dt+V(ˆs,M(ˆs)), | (4.15) |
so it follows that
−ε≤−V(ˆs,M(ˆs))−V(s,M0)ˆs−s−1ˆs−s∫ˆss(A1u21(t)+A2u22(t)+A3y(t)+A4v(t))dt=1ˆs−s∫ˆss{−Vt(t,M)−[Vx(Λ−ax(t)−βx(t)v(t)−γx(t)y(t))+Vy(βx(t)v(t)+γx(t)y(t)−by(t)−u1(t)y(t))+Vv(ky(t)−cv(t)−u2(t)v(t))]−[A1u21(t)+A2u22(t)+A3y(t)+A4v(t)]}dt≤1ˆs−s∫ˆss{−Vt(t,M)+supu∈U{−Vx(Λ−ax(t)−βx(t)v(t)−γx(t)y(t))−Vy(βx(t)v(t)+γx(t)y(t)−by(t)−u1(t)y(t))−Vv(ky(t)−cv(t)−u2(t)v(t))−A1u21(t)−A2u22(t)−A3y(t)−A4v(t)}}dt→−Vt(s,M0)+supu∈U{−Vx(Λ−ax(t)−βx(t)v(t)−γx(t)y(t))−Vy(βx(t)v(t)+γx(t)y(t)−by(t)−u1(t)y(t))−Vv(ky(t)−cv(t)−u2(t)v(t))−A1u21(t)−A2u22(t)−A3y(t)−A4v(t)}. | (4.16) |
Combining (4.14) and (4.16), we complete the proof.
For further study, the definition of viscosity solution of (4.11) by [32] is given.
Definition 4.3. A function V∈C[0,T]×R3) is called a viscosity subsolution (or supersolution) of (4.11) if
V(T,M)≤h(M),(V(T,M)≥h(M))∀M∈R3, | (4.17) |
and for any φ∈C1([0,T]×R), whenever V−φ attains a local maximum(or minimum) at (t,M)∈[0,T]×R3, we have
−φt(t,M)+supu(t)∈UH(t,M,u(t),−φM(T,M))≤0,(−φt(t,M)+supu(t)∈UH(t,M,u(t),−φM(T,M))≥0). | (4.18) |
In this case that V is both a viscosity subsolution and supersolution of (4.11), it is a viscosity solution of (4.11).
If we get the value function V by solving the HJB equation, then an optimal pair could be constructed. Thus, the characterization of the value function of the viscosity solution to the HJB equation is given below.
Theorem 4.4. Let 1 hold, and U is convex. Then, the value function V(⋅,⋅) satisfies
|V(s,M0)−V(ˉs,¯M0)|≤K|s−ˉs|,∀(s,M0),(ˉs,¯M0)∈[0,T]×R3, | (4.19) |
for some K>0. Moreover, V is the only viscosity solution of (4.11) in the class C([0,T]×R3).
Proof. We have
|V(s,M0)−V(ˉs,¯M0)|=|infu1,u2∈V[s,T]∫Ts[A1u21(t)+A2u22(t)+A3y(t)+A4v(t)]dt−infu1,u2∈V[s,T]∫Tˉs[A1u21(t)+A2u22(t)+A3y(t)+A4v(t)]dt|≤|∫ˉss[A1u21(t)+A2u22(t)+A3y(t)+A4v(t)]dt|≤K|s−ˉs|, | (4.20) |
where K=max{A1u21(t)+A2u22(t)+A3y(t)+A4v(t)}>0. Then, similar to [32, Theorem 2.5], the value function V clearly satisfies the condition and is the only viscosity solution of (4.11).
Next, in order to discuss the existence of the optimal control pair, we define H(t,M,u(t),p) as the Hamiltonian as follows:
H(t,M,u(t),p)=(Λ−ax(t)−βx(t)v(t)−γx(t)y(t))p1+(βx(t)v(t)+γx(t)y(t)−by(t)−u1(t)y(t))p2+(ky(t)−cv(t)−u2(t)v(t))p3+A1u21(t)+A2u22(t)+A3y(t)+A4v(t). | (4.21) |
Theorem 4.5. There exists an optimal control pair (u∗1(t),u∗2(t)) and a corresponding optimal state (x∗(t),y∗(t),v∗(t)) such that
J(u∗1(t),u∗2(t))=minu1(t),u2(t)∈UJ(u1(t),u2(t)). | (4.22) |
Proof. Note that both the control variables and state variables are nonnegative, and the objective functional (4.3) is convex with respect to the control variables. The convexity and closure of the control set U are obtained according to the definition of it. Furthermore, the optimal control is bounded. Therefore, we conclude that there exists an optimal control u∗1(t),u∗2(t) for J(u∗1(t),u∗2(t))=minu1(t),u2(t)∈UJ(u1(t),u2(t)).
The HJB equation and the value function have been discussed. Further, we figure out the optimal control as follows.
Theorem 4.6. Let u∗1(t),u∗2(t) be optimal control variables, and x∗(t),y∗(t),v∗(t) are corresponding optimal state variables. There exists adjoint process p(t)=(p1(t),p2(t),p3(t))⊤, satisfying the following adjoint equation:
{dp1(t)=−[(−a−βv(t)−γy(t))p1(t)+(βv(t)+γy(t))p2(t)]dtdp2(t)=−[(−γx(t))p1(t)+(γx(t)−b−u1(t))p2(t)+kp3(t)+A3]dtdp3(t)=−[(−βx)p1(t)+(βx)p2(t)+(−c−u2(t))p3(t)+A4]dtpi(T)=0,i=1,2,3. | (4.23) |
Furthermore, the optimal control is given as follows:
u∗i(t)=min{max{Bi,0},1},i=1,2, | (4.24) |
where B1=y∗(t)p2(t)2A1,B2=v∗(t)p3(t)2A2.
Proof. Because the Hamiltonian function H is given. Moreover, by using the optimal condition, and we obtain u∗1(t) and u∗2(t)
∂H∂u1=0,∂H∂u2=0. | (4.25) |
Hence,
u∗1(t)=y∗(t)p2(t)2A1,u∗2(t)=v∗(t)p3(t)2A2. | (4.26) |
So, B1=y∗(t)p2(t)2A1,B2=v∗(t)p3(t)2A2 are taken, and the optimal control u∗1(t),u∗2(t) follows:
u∗1(t)={0,ify∗(t)p2(t)2A1<0,y∗(t)p2(t)2A1,if0≤y∗(t)p2(t)2A1≤1,1,ify∗(t)p2(t)2A1>1,u∗2(t)={0,ifv∗(t)p3(t)2A2<0,v∗(t)p3(t)2A2,if0≤v∗(t)p3(t)2A2≤1,1,ifv∗(t)p3(t)2A2>1. | (4.27) |
Thus, the optimal value can be obtained.
Remark 4.7. In practical problems, when the parameters are known, the optimal control u∗1(t) and u∗2(t) can be calculated by using computer programming for Eqs (4.1), (4.23) and (4.27), that is, the dosage intensity of RTIs and PIs at each time can be calculated. If patients are treated in such a proportion, an optimal control strategy to minimize the costs and the number of viruses and infected cells can be obtained.
In this section, several numerical simulations are presented to prove and verify the above theoretical results by MATLAB software.
In this subsection, numerical simulations are performed to illustrate the theoretical results. We discuss the dynamics behavior of system (3.1). Since the choice of the threshold level Nt is different, system (3.1) will show various dynamics. There are three cases to consider. First, we discretize system (3.1) as follows:
{x(j+1)=x(j)+Δt(Λ−ax(j)−βx(j)v(j)−γx(j)y(j)),y(j+1)=y(j)+Δt(βx(j)v(j)+γx(j)y(j)−by(j)−εμ1y(j)),v(j+1)=v(j)+Δt(ky(j)−cv(j)−εμ2v(j)). | (5.1) |
In Table 1, all parameter values are exhibited. Some values are collected from different papers [6,17], and others are assumed.
Parameter | Value | Source of data |
Λ | 10 | [17] |
a | 0.01 | [17] |
β | 0.00034 | Assumed |
γ | 0.0001 | Assumed |
b | 0.26 | [6] |
k | 11 | [6] |
c | 0.1 | [6] |
μ1 | 0.3 | Assumed |
μ2 | 0.1 | Assumed |
Case 1: E∗1 is a real equilibrium, and E∗2 is a virtual equilibrium.
This situation is established. Suppose that the following conditions are satisfied:
y∗1<Ntandy∗2<Nt. | (5.2) |
In this case, both equilibriums are located on the same side of the plane Nt, namely, the region G1. Thus, when (5.2) is satisfied, we obtain that E∗1 can achieve stability. In Figure 1 E∗1∈G1 achieves stability with Nt=60, and the possible trajectories of this figure are shown: A trajectory that starts in region G1 will converge to E∗1 as t→+∞ with hitting and sliding down on the sliding domain S⊂Gs; a trajectory of the initial point in region G2 will cross the manifold Gs, then enter region G1 and finally converge to E∗1; a trajectory that begins in region G1 or G2 will hit and slide to the boundary of the sliding domain S⊂Gs before moving towards E∗1.
Case 2: E∗2 is a real equilibrium, and E∗1 is a virtual equilibrium.
This situation is established. Suppose that the following conditions are satisfied:
y∗1>Ntandy∗2>Nt. | (5.3) |
In this case, both equilibriums are located on the same side of the plane Nt, namely, the region G2. Thus, when (5.3) is satisfied, we obtain that E∗2 can achieve stability. In Figure 2 E∗2∈G2 achieves stability with Nt=15, and the possible trajectories of this figure are shown: A trajectory with initial condition in region G2 will converge to E∗2 as t→+∞ with hitting and sliding up on the sliding domain S⊂Gs; a trajectory with initial point in region G1 will pass through the manifold Gs from G1 to G2; a trajectory that starts in region G1 or G2 will hit and slide up on the boundary of the sliding domain S⊂Gs before moving towards E∗2.
Case 3: Both E∗1 and E∗2 are virtual equilibria.
This situation is established. Suppose that the following conditions are satisfied:
y∗1>Ntandy∗2<Nt. | (5.4) |
For this case, we can find that E∗1 and E∗2 belong to regions G2 and G1, respectively. If E∗S∈S⊂Gs, it is a pseudo-equilibrium. Thus, when (5.4) is satisfied, we obtain that E∗s is stable when it exists. Figure 3 E∗s∈S achieves stability with Nt=25, and the possible trajectories of this figure are shown: A trajectory with initial condition in region G1 will converge to E∗s as t→+∞; a trajectory that begins in region G2 will cross the manifold Gs, then enter region G1 and finally converge to E∗s; a trajectory with initial condition in region G1 or G2 will hit and slide down on S⊂Gs before converging to E∗s.
In this subsection, we also discretize the optimal control system (4.1) as
{x(j+1)=x(j)+Δt(Λ−ax(j)−βx(j)v(j)−γx(j)y(j)),y(j+1)=y(j)+Δt(βx(j)v(j)+γx(j)y(j)−by(j)−u1(j)y(j)),v(j+1)=v(j)+Δt(ky(j)−cv(j)−u2(j)v(j)). | (5.5) |
The parameter values Λ, a, b, β, γ, k and c are chosen as shown in Table 1, and A1=20,A2=50,A3=0.1,A4=0.1. Then, we compare the effects of different control intensities by the following figures. The cost of control strategies must be considered, and the cost of each measure is different, so we want to know the change under only one control strategy compared with two control strategies. Thus, we make simulations under only u1(t) (see Figure 4) or u2(t) (see Figure 5) and optimal control (see Figure 6).
When initial value (x0,y0,v0)=(50,150,800), Figure 4 illustrates that if the control strategy is only applied to infected cells (u2(t)=0), the concentration of infected cells and viruses will decrease. Meanwhile, the number of infected cells and viruses decreased with the increase of u1(t) intensity. Figure 5 shows that if the control strategy is only applied to viruses (u1(t)=0), the density of infected cells and viruses will also decrease, and the number of infected cells and viruses decreased with the increase of u2(t) intensity.
The expressions of optimal control u∗1(t) and u∗2(t) are obtained through calculation (in Eq (4.27)), and then we obtain the optimal control states of uninfected cells, infected cells and viruses in Figure 6, which shows that the concentration of infected cells and viruses decreases to some extent after the control is applied, and the combination of multi-drug works better than a single-drug approach. Figure 7 shows the values of control variables u1(t) and u2(t) in each time. Eventually, the concentration of infected cells and viruses and control intensity gradually decrease and stabilize. Thus, the control strategies in our model have significant influence on the spread of HIV.
In this paper, we have proposed an HIV model with cell-to-cell transmission and analyzed the sliding mode dynamics of the model and optimal control problem. First, we extended a novel Filippov model (3.1), which indicates that corresponding control measures (i.e., antiretroviral drugs RTIs and PIs) are triggered once the total number of infected cells reaches the threshold level Nt, where RTIs prevent new HIV infections by blocking the transformation of viral RNA into DNA in T cells, and PIs reduce the number of viral particles produced by actively infected T cells [21]. Further, the sliding domain and sliding mode dynamics of system (2.2) have been examined. In addition, the simulation results show that the model solution is either near a real equilibrium point or near a pseudo-equilibrium point according to the different threshold levels. It is worth mentioning that some parameters can be simulated through actual data, and the basic reproduction number can be calculated to judge the stability. In addition, the number of viruses and infected cells can reduce to a previously desired level when the threshold level is chosen properly. Because different patients have different initial viral loads, an individualized therapy is suggested, which shows that the choice of a treatment strategy for a given patient should depend on HIV viruses and infected cells and proposed threshold level.
Moreover, the cost of treatment is beyond the reach of many infected patients. Therefore, we have introduced an optimal therapy to minimize the cost of treatment and reduce the viral load and the number of infected cells. Then, the efficacies of RTIs and PIs and their combinations have been measured. In addition, we have discussed an efficient numerical method based on optimal control to determine the best treatment strategy for HIV infection. Our results indicate that with the increase of treatment intensity, the number of infected cells and viruses decreases, while the density of CD4+T increases. Due to the multiple transmission routes of HIV, the combined use of multiple drugs is better than the use of a single drug. From a biological point of view, it can be concluded that optimal control is adopted when the number of infected cells in the patient is higher than threshold level Nt, and at this time, the economic cost can be considered to select the optimal control measures. Control measures are not necessary when the concentration of infected cells is low. Optimal control path is shown in Figure 7.
The results of this paper have practical implications for controlling HIV transmission. However, our work is only a preliminary exploration of the impact of some control measures on HIV transmission, which can be improved in many aspects. The number of viruses in the infected person varies with age, so it is interesting and challenging to consider the sliding mode control of HIV model with age structure. We can also refer to the methods in the literature [33,34] to study the fractional-order HIV model. These issues will be the focus of our future research.
The research was supported by the Key Research and Development Program of Ningxia (2020BEB04007) and the Fundamental Research Funds for the Central Universities, North Minzu University (2020KYQD17).
The authors declare there is no conflict of interest.
[1] | J. Banchereau, F. Briere, C. Caux, et al., Immunobiology of dendritic cells, Annu. Rev. Immunol., 18 (2000), 767–811. |
[2] | J. Zhu and W. E. Paul, CD4 T cells: fates, functions, and faults, Blood, 112 (2008), 1557–1569. |
[3] | N. J. Burroughs, B. M. P. M. Oliveira and A. A. Pinto, Regulatory T cell adjustment of quorum growth thresholds and the control of local immune responses, J. Theor. Biol., 241 (2006), 134–141. |
[4] | A. L. Cava, Tregs are regulated by cytokines: Implications for autoimmunity, Autoimmun. Rev., 8 (2008), 83–87. |
[5] | B.-I. Moon, T. H. Kim and J.-Y. Seoh, Functional modulation of regulatory T cells by IL-2, PLoS One, 10 (2015), 1–13. |
[6] | E. M. Shevach, R. S. McHugh, C. A. Piccirillo, et al., Control of T-cell activation by CD4+ CD25+ suppressor T cells, Immunol. Rev., 182 (2001), 58–67. |
[7] | A. M. Thornton and E. M. Shevach, CD4+ CD25+ immunoregulatory T cells suppress polyclonal T cell activation in vitro by inhibiting interleukine 2 production, The Journal of Experimental Medicine, 188 (1998), 287–296. |
[8] | L. A. Turka and P. T. Walsh, IL-2 signaling and CD4+ CD25+ Foxp3+ regulatory T cells, Front. Biosci., 13 (2008), 1440–1446. |
[9] | S. Sakaguchi, Naturally arising CD4+ regulatory T cells for immunological self-tolerance and negative control of immune responses, Annu. Rev. Immunol., 22 (2004), 531–562. |
[10] | J. Zhu, H. Yamane and W. E. Paul, Differentiation of Effector CD4 T Cell Populations, Annu. Rev. Immunol., 28 (2010), 445–489. |
[11] | D. Homann, L. Teyton and M. B. Oldstone, Differential regulation of antiviral T-cell immunity results in stable CD8+ but declining CD4+ T-cell memory, Nat. Med., 7 (2001), 913–919. |
[12] | R. E. Callard and A. J. Yates, Immunology and mathematics: crossing the divide, Immunology, 115 (2005), 21–33. |
[13] | G. Lythe and C. Molina-París, Some deterministic and stochastic mathematical models of naïve T-cell homeostasis, Immunol. Rev., 285 (2018), 206–217. |
[14] | R. J. de Boer and P. Hogeweg, Immunological discrimination between self and non-self by precur-sor depletion and memory accumulation, J. Theor. Biol., 124 (1987), 343–369. |
[15] | A. Pinto, N. Burroughs, F. Ferreira, et al., Dynamics of immunological models, Acta Biotheor., 58 (2010), 391–404. |
[16] | K. Blyuss and L. Nicholson, The role of tunable activation thresholds in the dynamics of autoim-munity, J. Theor. Biol., 308 (2012), 45–55. |
[17] | S. Khailaie, F. Bahrami, M. Janahmadi, et al., A mathematical model of immune activation with a unified self-nonself concept, Front. Immunol., 4 (2013), 474. |
[18] | K. León, A. Lage, and J. Carneiro, Tolerance and immunity in a mathematical model of T-cell mediated suppression, J. Theor. Biol., 225 (2003), 107–126. |
[19] | C. Bianca and L. Brézin, Modeling the antigen recognition by B-cell and T-cell receptors through thermostatted kinetic theory methods, Int. J. Biomath., 10 (2017), 1750072. |
[20] | R. F. Alvarez, J. A. Barbuto and R. Venegeroles, A nonlinear mathematical model of cell-mediated immune response for tumor phenotypic heterogeneity, J. Theor. Biol., 471 (2019), 42–50. |
[21] | N. Burroughs, M. Ferreira, B. Oliveira, et al., Autoimmunity arising from bystander proliferation of T cells in an immune response model, Math. Comput. Modell., 53 (2011), 1389–1393. |
[22] | B. M. P. M. Oliveira, R. Trinchet, M. V. Otero-Espinar, et al., Modelling the suppression of au-toimmunity after pathogen infection, Math. Methods Appl. Sci., 41 (2018), 8565–8570. |
[23] | B. M. P. M. Oliveira, I. P. Figueiredo, N. J. Burroughs, et al., Approximate equilibria for a T cell and Treg model, Appl. Math. Inform. Sci., 9 (2015), 2221–2231. |
[24] | N. Burroughs, B. Oliveira, A. Pinto, et al., Immune response dynamics, Math. Comput. Modell., 53 (2011), 1410–1419. |
[25] | N. Burroughs, B. Oliveira, A. Pinto, et al., Sensibility of the quorum growth thresholds controlling local immune responses, Math. Comput. Modell., 47 (2008), 714–725. |
[26] | N. Burroughs, M. Ferreira, B. Oliveira, et al., A transcritical bifurcation in an immune response model, J. Differ. Eq. Appl., 17 (2011), 1101–1106. |
[27] | B. M. P. M. Oliveira, A. Yusuf, I. P. Figueiredo, et al., The effect of a linear tuning between the antigenic stimulations of T cells and Tregs. In Preparation. |
[28] | N. J. Burroughs, M. Ferreira, J. Martins, et al., Dynamics and biological thresholds, in Dynam-ics, Games and Science I, DYNA 2008, in Honor of Maur´ ıcio Peixoto and David Rand (editors A. A. Pinto, D. A. Rand, and M. M. Peixoto), volume 1 of Springer Proceedings in Mathematics, Springer-Verlag Berlin Heidelberg (2011), pp. 183–191. |
[29] | R. J. de Boer, D. Homann and A. S. Perelson, Different Dynamics of CD4+ and CD8+ T Cell Responses During and After Acute Lymphocytic Choriomeningitis Virus Infection, J. Immunol., 171 (2003), 3928–3935. |
[30] | R. J. de Boer and A. S. Perelson, Quantifying T lymphocyte turnover, J. Theor. Biol., 327 (2013), 45–87. |
[31] | R. E. Callard, J. Stark and A. J. Yates, Fratricide: a mechanism for T memory-cell homeostasis, Trends Immunol., 24 (2003), 370–375. |
[32] | S. Nagata, Fas ligand-induced apoptosis, Annu. Rev. Genet., 33 (1999), 29–55. |
[33] | P. M. Anderson and M. A. Sorenson, Effects of route and formulation on clinical pharmacokinetics of interleukine-2, Clin. Pharmacokinet., 27 (1994), 19–31. |
[34] | A. R. Mclean and C. A. Michie, In vivo estimates of division and death rates of human t lympho-cytes., Proceedings of the National Academy of Sciences, 92 (1995), 3707–3711. |
[35] | C. Michie, A. McLean, C. Alcock, et al., Life-span of human lymphocyte subsets defined by CD45 isoforms, Nature, 360 (1992), 264–265. |
[36] | D. Moskophidis, M. Battegay, M. Vandenbroek, et al., Role of virus and host variables in virus persistence or immunopathological disease caused by a non-cytolytic virus, J. Gen. Virol., 76 (1995), 381–391. |
[37] | H. Veiga-Fernandes, U. Walter, C. Bourgeois, et al., Response of naïve and memory CD8+ T cells to antigen stimulation in vivo, Nat. Immunol., 1 (2000), 47–53. 38. G. A. Seber and C. J. Wild, Nonlinear Regression, Wiley and Sons (2003). |
1. | Dan Shi, Mengqing Zhang, Qimin Zhang, Stationary distribution and near‐optimal control of a stochastic reaction–diffusion HIV model, 2024, 47, 0170-4214, 4381, 10.1002/mma.9819 |