Research article

An age-structured SIPC model of cervical cancer with immunotherapy

  • Received: 02 February 2024 Revised: 26 March 2024 Accepted: 01 April 2024 Published: 18 April 2024
  • MSC : 35F50, 35Q92, 92C37

  • Immunotherapy is a targeted therapy that can be applied to cervical cancer patients to prevent DNA damage caused by human papillomavirus (HPV). The HPV infects normal cervical cells withing a specific cell age interval, i.e., between the G1 to S phase of the cell cycle. In this study, we developed a new mathematical model of age-dependent immunotherapy for cervical cancer. The model is a four-dimensional first-order partial differential equation with time- and age-independent variables. The cell population is divided into four sub-populations, i.e., susceptible cells, cells infected by HPV, precancerous cells, and cancer cells. The immunotherapy term has been added to precancerous cells since these cells can experience regression if appointed by proper treatments. The immunotherapy process is closely related to the rate of T-cell division. The treatment works in the same cell cycle that stimulates and inhibits the immune system. In our model, immunotherapy is represented as a periodic function with a small amplitude. It is based on the fluctuating interaction between T-cells and precancerous cells. We have found that there are two types of steady-state conditions, i.e., infection-free and endemic. The local and global stability of an infection-free steady-state has been analyzed based on basic reproduction numbers. We have solved the Riccati differential equation to show the existence of an endemic steady-state. The stability analysis of the endemic steady-state has been determined by using the perturbation approach and solving integral equations. Some numerical simulations are also presented in this paper to illustrate the behavior of the solutions.

    Citation: Eminugroho Ratna Sari, Lina Aryati, Fajar Adi-Kusumo. An age-structured SIPC model of cervical cancer with immunotherapy[J]. AIMS Mathematics, 2024, 9(6): 14075-14105. doi: 10.3934/math.2024685

    Related Papers:

    [1] Anusmita Das, Kaushik Dehingia, Hemanta Kumar Sharmah, Choonkil Park, Jung Rye Lee, Khadijeh Sadri, Kamyar Hosseini, Soheil Salahshour . Optimal control of effector-tumor-normal cells dynamics in presence of adoptive immunotherapy. AIMS Mathematics, 2021, 6(9): 9813-9834. doi: 10.3934/math.2021570
    [2] Xingxiao Wu, Lidong Huang, Shan Zhang, Wenjie Qin . Dynamics analysis and optimal control of a fractional-order lung cancer model. AIMS Mathematics, 2024, 9(12): 35759-35799. doi: 10.3934/math.20241697
    [3] Kaushik Dehingia, Hemanta Kumar Sarmah, Kamyar Hosseini, Khadijeh Sadri, Soheil Salahshour, Choonkil Park . An optimal control problem of immuno-chemotherapy in presence of gene therapy. AIMS Mathematics, 2021, 6(10): 11530-11549. doi: 10.3934/math.2021669
    [4] Noufe H. Aljahdaly, Nouf A. Almushaity . A diffusive cancer model with virotherapy: Studying the immune response and its analytical simulation. AIMS Mathematics, 2023, 8(5): 10905-10928. doi: 10.3934/math.2023553
    [5] Kai Zhang, Yunpeng Ji, Qiuwei Pan, Yumei Wei, Yong Ye, Hua Liu . Sensitivity analysis and optimal treatment control for a mathematical model of Human Papillomavirus infection. AIMS Mathematics, 2020, 5(3): 2646-2670. doi: 10.3934/math.2020172
    [6] Salma M. Al-Tuwairqi, Asma A. Badrah . Modeling the dynamics of innate and adaptive immune response to Parkinson's disease with immunotherapy. AIMS Mathematics, 2023, 8(1): 1800-1832. doi: 10.3934/math.2023093
    [7] Ahmed J. Abougarair, Mohsen Bakouri, Abdulrahman Alduraywish, Omar G. Mrehel, Abdulrahman Alqahtani, Tariq Alqahtani, Yousef Alharbi, Md Samsuzzaman . Optimizing cancer treatment using optimal control theory. AIMS Mathematics, 2024, 9(11): 31740-31769. doi: 10.3934/math.20241526
    [8] Muhammad Farman, Aqeel Ahmad, Ali Akgül, Muhammad Umer Saleem, Kottakkaran Sooppy Nisar, Velusamy Vijayakumar . Dynamical behavior of tumor-immune system with fractal-fractional operator. AIMS Mathematics, 2022, 7(5): 8751-8773. doi: 10.3934/math.2022489
    [9] Yue Liu, Liming Chen, Liangting Lv, Pierre Failler . The impact of population aging on economic growth: a case study on China. AIMS Mathematics, 2023, 8(5): 10468-10485. doi: 10.3934/math.2023531
    [10] Zawar Hussain, Atif Akbar, Mohammed M. A. Almazah, A. Y. Al-Rezami, Fuad S. Al-Duais . Diagnostic power of some graphical methods in geometric regression model addressing cervical cancer data. AIMS Mathematics, 2024, 9(2): 4057-4075. doi: 10.3934/math.2024198
  • Immunotherapy is a targeted therapy that can be applied to cervical cancer patients to prevent DNA damage caused by human papillomavirus (HPV). The HPV infects normal cervical cells withing a specific cell age interval, i.e., between the G1 to S phase of the cell cycle. In this study, we developed a new mathematical model of age-dependent immunotherapy for cervical cancer. The model is a four-dimensional first-order partial differential equation with time- and age-independent variables. The cell population is divided into four sub-populations, i.e., susceptible cells, cells infected by HPV, precancerous cells, and cancer cells. The immunotherapy term has been added to precancerous cells since these cells can experience regression if appointed by proper treatments. The immunotherapy process is closely related to the rate of T-cell division. The treatment works in the same cell cycle that stimulates and inhibits the immune system. In our model, immunotherapy is represented as a periodic function with a small amplitude. It is based on the fluctuating interaction between T-cells and precancerous cells. We have found that there are two types of steady-state conditions, i.e., infection-free and endemic. The local and global stability of an infection-free steady-state has been analyzed based on basic reproduction numbers. We have solved the Riccati differential equation to show the existence of an endemic steady-state. The stability analysis of the endemic steady-state has been determined by using the perturbation approach and solving integral equations. Some numerical simulations are also presented in this paper to illustrate the behavior of the solutions.



    Cervical cancer is a type of cancer that can be prevented with screening and vaccines. However, it is still a significant type of cancer for women. Even though surgery, chemotherapy, and radiation provide the potential for survival, the side effects will burden the patient, such as the appearance of secondary infections, anemia, hair loss, skin rashes, and diarrhea [1].

    Like other cancer cells, cervical cancer cells are camouflaged to trick the T-cells so that they cannot recognize the presence of the cancer cells. Typically, the cells undergo a cyclic process that stimulates immune cells by strengthening and expanding the response of T-cells. Simultaneously, this process also considers the inhibiting factors that limit immunity and thus prevent autoimmunity. This cyclic process is known as the cancer-immunity cycle [2]. The cancer-immunity cycle can be divided into seven main steps. The first step is the antigen release from cancer cells. The second one involves the antigen being transported to T-cells. In the third step, the T-cells become active and initiate the immune system's fight against the cancer cells. The four-next steps involve active T-cells passing through the blood vessels to find cancer cells, infiltrate, recognize the cancer cells due to the generated antigens, and kill cancer cells. Identification of T-cell inhibitory signals by cancer cells drives the development of immunotherapy.

    Immunotherapy treats the cancer cells by stimulating the immune response, where the immunotherapy drugs will block the cancer cells from deactivating T-cells. There are several types of immunotherapy, i.e., immune checkpoint inhibitors, T-cell transfer therapy, monoclonal antibodies, treatment vaccines, and immune system modulators. The T-cell transfer therapy has three approaches, i.e., the use of tumor-infiltrating lymphocytes, chimeric antigen receptor T-cells, and engineered T-cell receptor T cells [3]. These procedures entail harvesting the immune cells, multiplying vast numbers of them ex vivo, and then returning the cells to the host via a vein needle.

    In terms of mathematical models, cervical cancer growth models have been developed with several approaches. The development of cervical cancer at the tissue level can be modeled by a system of ordinary differential equations; see [4,5,6]. In those cases, the cell population was divided into susceptible, infected, precancer, cancer cells, and the virus population. Since HPV infects cells within a specific age interval, the cervical cancer cell model can be described as age-structured; see [7,8,9,10]. The authors of [7] described the dynamics between age- and time-dependent cell population and the time-dependent free-floating virus population. Age-structured cervical cancer cell models have been developed and analyzed by numerically modeling the solution behavior [8,10].

    In fact, viruses that have entered the tissue will be reproduced by infected cells; after that, the infected cells release the virus particles into the extracellular environment, and the virus has the ability to infect the surrounding susceptible cells [11]. Therefore, the authors of [9] reconsidered the appearance of the free-floating virus compartment and proposed a new parameter to represent the infection rate of the viruses, called the force of infection. However, the authors of [7,9] considered the interaction between cells without treatment, which is important for the control of cancer cells in the tissue.

    Meanwhile, cancer models with immunotherapy and involve the use of systems of ordinary differential equations have been discussed in [12,13,14]. The authors of [12] incorporated immunotherapy by applying a constant number to represent how the cytokine interleukin-2 boosts the immune system to fight tumors. A mathematical analysis considering immune checkpoint inhibitors has been studied and described in [13]. Moreover, a group of scholars have modeled immunotherapy by using a perturbation term to take into account the protein fluctuations in the cells [14].

    Due to the fact that the role of the E2F1 gene is to preserve the code of proteins, E2F1 regulates cell cycle progression through the G1-to-S transition of the cell cycle, where HPV damages cells when cells are in this phase. E2F1 also plays an important role in DNA repair and apoptosis [15]. On the other hand, the administration of immunotherapy can activate the enhancer of zeste homolog 2 (EZH2) as a tumor suppressor gene. EZH2 and E2F1 can induce apoptosis in the abnormal cells [16]. Thus, the immunotherapy process works when the cells are in a specific phase of their life. Therefore, in our model, immunotherapy is considered as age-dependent.

    In this paper, we introduce a new mathematical model for cervical cancer that includes an age-structured immunotherapy program. We have extended the investigations in [9] by adding an immunotherapy term. According to the authors of [4], the precancerous cells can be controlled when the lesions become precancerous. Hence, we chose to apply the immunotherapy term to the precancerous cells. Motivated by the results in [14,17], since the levels of stimulation and inhibition of T-cells fluctuate throughout the cell cycle, we assume that the interaction between the T-cells and the precancerous cells can be represented by a periodic function with a small amplitude.

    Due to the addition of the immunotherapy term, the discussion on an endemic steady-state solution draws attention to the Riccati differential equation. The characteristics of the Riccati differential equation are as follows: the coefficient of the dependent variable is a function of the independent variable, and the highest power of the dependent variable is two. There are some methods for solving the Riccati differential equation. In [18], the Riccati equation was converted into a differential equation of the second order with variable coefficients. Commonly, the general solutions of the second order differential equation with variable coefficients can be calculated by applying a reduction of order, a change of independent variables, and an undetermined coefficient method. The restrictive conditions in [19,20] showed that the method can only be used in some instances. The authors of [21] provided another alternative for solving the second-order differential equation with variable coefficients by using an integral series. Apart from the existing procedures, we can solve the Riccati case in this study by manipulating the algebra such that the equation becomes linear-differential.

    Remarkable mathematical models can make contributions toward reducing abnormal cells in the tissue. It is promising to use age-dependent treatment to interact with age-dependent abnormal cells. Therefore, in this article, there are three main findings, as follows. The first one is a new mathematical model that contains the force of infection that represents the infection rate of the free-floating viruses. Second, the mathematical model considers the age-dependency of immunotherapy. Third, through theoretical and numerical analysis, we have drawn some conclusions, i.e., the infection rate plays an important role in reducing precancerous cells when no immunotherapy exists; the higher the infection rate, the more immunotherapy that is needed; and the best time to apply immunotherapy is the initial cell cycle phase.

    We present the content of this work as follows. Section 2 describes a new age-structured mathematical model of cervical cancer that incorporates a force of infection parameter and age-dependent immunotherapy. Section 3 shows that the solution exists and is unique. Detailed steps for obtaining the basic reproduction numbers are presented in Section 4. This value is important as a threshold to analyze the stability of infection-free and endemic steady-state solutions. In Section 5, we provide numerical simulations to verify the results in Section 4. Finally, we give a concluding remark in the last section.

    The progression of cervical cancer occurs as follows: HPV infecting susceptible cells, the transformation of infected cells to precancerous cells, and the formation of cancer. For background information regarding the interaction of cells at specific ages and with time-dependency, recall the model in [9]. The susceptible cells, infectious cells, precancerous cells, and cancer cells at age a and time t are respectively denoted by S(a,t), I(a,t),P(a,t), and C(a,t). The notation N(a,t) expresses the total population of cells at age a and time t in the cervical tissue. In this paper, the age of a cell that can be infected by the virus or become abnormal is denoted by a1, and the maximum cell's age is aσ.

    Infected cells will keep producing viruses, and then infect susceptible cells around them. Therefore, we use the force of infection parameter, instead of applying a virus compartment as in [4,7]. This parameter is denoted by β(a,t), which represents the infection rate of a susceptible cell due to the virus; it is as follows:

    β(a,t)=λ(a)aσa1h(b)[I(b,t)+C(b,t)]N(b,t)db, (2.1)

    where b[a1,aσ], λ(a) denotes the innate susceptibility of susceptible cells at age a, and h(b) represents the innate ability of infected and cancer cells to infect at age b. The infection rate of susceptible cells at age a due to interaction with infected cells and cancer cells at age b is denoted by the product of λ(a)h(b).

    We have extended the model in [9] by adding an immunotherapy parameter. The precancerous phase can regress with proper immunotherapy and return to the infected phase. In this model, the immunotherapy term was designed to depend on cell age to determine the appropriate cell age for drug administration. This study is focused on analyzing a cervical cancer model with immunotherapy. Immunotherapy works against abnormal cells by stimulating the immune response system. In this case, the immunotherapy drugs will activate T-cells. When T-cells are activated, they undergo proliferation to generate more T-cells. On the other side, the authors of [4] suggested reducing the risk of cervical cancer by making cells stay longer in the precancerous phase. In the precancerous phase, it is easier for cells to regress and prevent them from becoming cancer cells by increasing the body's immunity. Hence, we have added a new parameter to show the immunotherapy process that affects the reduction of precancerous cells. The dividing T-cells will be more controlled and become active against precancerous cells.

    Note that ε0 represents the rate of T-cell division after being stimulated by immunotherapy drugs. The interaction between T-cells and precancerous cells was set to fluctuate according to ε1. We write ε0(1+ε1cos(2πa)), where ε0>0 and 0<ε11. If ε0>0 and ε1=0, the treatment is constant since the rate of T-cell division is comparable to that for precancerous cells. Furthermore, δ(a) denotes the age-dependent rate of progression from infected to precancerous, γ(a) denotes the age-dependent rate of progression from precancerous to cancer, and μ(a) denotes the age-dependent death rate of each type of cell. Because of the cell division, we denote ΛN(a,t) as the density of new susceptible cells.

    Considering that the infection rate of HPV is based on cell age, the growth of the susceptible, the infected, the precancerous, and the cancer cell densities also depend on cell age. Moreover, the immunotherapy term and death rate of each cell type depend on the age of the cell as well. Thus, the densities of susceptible, infected, precancerous, and cancer cells are considered in terms of age and time. Figure 1 illustrates the interaction between the densities of susceptible, infected, precancerous, and cancer cells.

    Figure 1.  Transfer diagram of the HPV transmission.

    A mathematical model of cervical cancer progression that includes age-dependent immunotherapy can be formulated as a nonlinear system of partial differential equations as follows:

    S(a,t)a+S(a,t)t=ΛN(a,t)μ(a)S(a,t)β(a,t)S(a,t),I(a,t)a+I(a,t)t=β(a,t)S(a,t)(δ(a)+μ(a))I(a,t),P(a,t)a+P(a,t)t=δ(a)I(a,t)(γ(a)+μ(a))P(a,t)ε0(1+ε1cos(2πa))P(a,t),C(a,t)a+C(a,t)t=γ(a)P(a,t)μ(a)C(a,t), (2.2)

    where the boundary and initial conditions are as follows:

    S(a1,t)=μaσa1N(a,t)da,I(a1,t)=P(a1,t)=C(a1,t)=0,S(a,0)=S0(a),I(a,0)=I0(a),P(a,0)=P0(a),C(a,0)=C0(a). (2.3)

    The analysis of System (2.2) is complicated since there is a term that contains N(a,t). Dividing each of the variables S(a,t),I(a,t),P(a,t), and C(a,t) by the corresponding total population N(a,t) will transform System (2.2) into a non-dimensional system. It is important to simplify the analysis of the model, so we can remove the total population N(a,t) on the first line of System (2.2). Therefore, we use the following coordinate transformation to remove N(a,t) from the system:

    s(a,t)=S(a,t)N(a,t),i(a,t)=I(a,t)N(a,t),p(a,t)=P(a,t)N(a,t),c(a,t)=C(a,t)N(a,t), (2.4)

    where s(a,t)+i(a,t)+p(a,t)+c(a,t)=1. By applying the transformations of (2.4) to the force of infection parameter (2.1), we acquire

    β(a,t)=λ(a)aσa1h(b)[i(b,t)+c(b,t)]db. (2.5)

    If the transformations of (2.4) is substituted into System (2.2) and the boundary-initial conditions of (2.3), we obtain

    s(a,t)a+s(a,t)t=ΛΛs(a,t)β(a,t)s(a,t)+ε0(1+ε1cos(2πa))p(a,t)s(a,t),i(a,t)a+i(a,t)t=β(a,t)s(a,t)(δ(a)+Λ)i(a,t)+ε0(1+ε1cos(2πa))p(a,t)i(a,t),p(a,t)a+p(a,t)t=δ(a)i(a,t)(γ(a)+Λ)p(a,t)ε0(1+ε1cos(2πa))(1p(a,t))p(a,t),c(a,t)a+c(a,t)t=γ(a)p(a,t)Λc(a,t)+ε0(1+ε1cos(2πa))p(a,t)c(a,t), (2.6)

    and the boundary-initial values are given by

    s(a1,t)=1,i(a1,t)=0,p(a1,t)=0,c(a1,t)=0,s(a,0)=s0(a),i(a,0)=i0(a),p(a,0)=p0(a),c(a,0)=c0(a).

    As we know, N(a,t)=S(a,t)+I(a,t)+P(a,t)+C(a,t). Hence, System (2.2) analysis appears complex since the first equation of System (2.2) contains the dependent variables S(a,t),I(a,t),P(a,t), and C(a,t). Using the transformations of (2.4), we can obtain System (2.6), which no longer contains N(a,t); thus, the analysis of System (2.6) is obviously preferable. In the next section, we discuss the existence and uniqueness of solutions, as well as the steady-state solutions of System (2.6) and their stability. By applying System (2.6), we can obtain two types of steady-state conditions, i.e., infection-free and endemic. The local stability of the steady-state can be determined by using the perturbation approach and solving integral equations. Some numerical simulations are also presented in this paper to illustrate the behavior of the solutions.

    In this section, we will discuss whether the solution of System (2.6) exists and is unique. We use Banach spaces, which possess a complete norm. It implies that every Cauchy sequence in the space converges to a unique limit in the same space. This property allows for the development of many important theorems for partial differential equations. We explore the existence and uniqueness of the solutions of System (2.6) by using the techniques given in [22,23]. Let the Banach space Y=L1(a1,aσ)×L1(a1,aσ)×L1(a1,aσ)×L1(a1,aσ), which is the set of Lebesgue integral functions that are endowed with the norm

    X=aσa1|X1(a)|da+aσa1|X2(a)|da+aσa1|X3(a)|da+aσa1|X4(a)|da,

    where X(a)=(X1(a),X2(a),X3(a),X4(a))TY. In this study, L1 is applied since it can be interpreted as the total population. Transform System (2.6) into an abstract Cauchy problem by moving the derivative term for age a on the left-hand side of System (2.6) to the right-hand side such that the independent variable a is considered to be in the function space. Introducing ˇs(a,t)=s(a,t)1 into the transformed System (2.6) gives

    ˇs(a,t)t=ˇs(a,t)a+ΛΛ(ˇs(a,t)+1)β(a,t)(ˇs(a,t)+1)+ε0(1+ε1cos(2πa))p(a,t)(ˇs(a,t)+1),i(a,t)t=i(a,t)a+β(a,t)(ˇs(a,t)+1)(δ(a)+Λ)i(a,t)+ε0(1+ε1cos(2πa))p(a,t)i(a,t),p(a,t)t=p(a,t)a+δ(a)i(a,t)(γ(a)+Λ)p(a,t)ε0(1+ε1cos(2πa))(1p(a,t))p(a,t),c(a,t)t=c(a,t)a+γ(a)p(a,t)Λc(a,t)+ε0(1+ε1cos(2πa))p(a,t)c(a,t). (3.1)

    The associated boundary and initial conditions are as follows: ˇs(a1,t)=0,i(a1,t)=0,p(a1,t)=0,c(a1,t)=0,ˇs(a,0)=s0(a)1,i(a,0)=i0(a),p(a,0)=p0(a),c(a,0)=c0(a). For convenience, we denote ˉε(a)=ε0(1+ε1cos(2πa)). Suppose that the linear operator A is defined by

    AX(a)=(dX1(a)daΛX1(a)dX2(a)da(δ(a)+Λ)X2(a)dX3(a)da+δ(a)X2(a)(γ(a)+Λ)X3(a)ˉε(a)X3(a)dX4(a)da+γ(a)X4(a)ΛX4(a)) (3.2)

    and the domain of X(a) as follows:

    D(A)={XY:X1,X2,X3,X4 are absolutely continuous on [a1,aσ],X(0)=0}.

    Let G:YY be a nonlinear operator defined by

    G(X(a))=(λ(a)((ZX2)(a)+(ZX4)(a))(X1+1)+ˉε(a)X3(a)(X1(a)+1)λ(a)((ZX2)(a)+(ZX4)(a))(X1+1)+ˉε(a)X3(a)X2(a)ˉε(a)X23(a)ˉε(a)X3(a)X4(a)), (3.3)

    where Z is a bounded linear operator on L1(a1,aσ) that is defined by

    (ZX2)(a)=aσa1h(b)X2(b)dband(ZX4)(a)=aσa1h(b)X4(b)db.

    If x(t)=(ˇs(,t),i(,t),p(,t),c(,t))TY, then System (3.1) can be written as a Cauchy problem:

    dx(t)dt=Ax(t)+G(x(t)),x(0)=x0Y, (3.4)

    where x0(a)=(s0(a)1,i0(a),p0(a),c0(a)). According to the work in [23], we have the following results.

    Lemma 3.1. The operator A generates a C0 semigroup eAt.

    Definition 3.1. [24] Let Y be a Banach space and V be an open set in Y. An operator G:VY is Fréchet-differentiable at XV if there exists G1 such that

    limy0G(X+y)G(X)G1(y)y=0.

    Lemma 3.2. Let G be as shown in (3.3). Then, G is Fréchet-differentiable on Y.

    Lemma 3.3. For each x0Y, there exists a maximal interval of existence [0,t0) and a unique continuous (mild) solution tx(t,x0) from [0,t0) to Y for (2.6) such that

    x(t)=eAtx0+t0eA(tτ)G(x(τ))dτ.

    Proof of Lemmas 3.1–3.3 can be achieved by using the Hille-Yosida theorem and the Gâteaux derivative, which can be seen in [23,24,25,26]. Next, a unique global classical solution of (3.4) can be derived by applying the following theorem. Let Δ={(ˇs,i,p,c)Y:ˇs1,i,p,c0} and Δ0={(ˇs,i,p,c)Y:1ˇs0,0i,p,c1}.

    Theorem 3.1. The mild solution x(t)Δ of (3.4) where x0Δ, enters into Δ0 in finite time, and Δ0 is positively invariant.

    Generally, to prove Theorem 3.1, the first step is to find solutions s(a,t),p(a,t),c(a,t) for System (2.6) by taking into account the characteristic equation. The next step is to transform i(a,t) in System (3.1) into the Cauchy problem. Therefore, (3.4) or, equivalently, the initial-boundary value problem given by System (2.6) has a unique positive global solution on Y with respect to the positive initial value in D(A)Δ.

    This subsection will discuss an infection-free steady-state solution, i.e., no virus infection in the cell population. Let (ˆs(a),ˆi(a),ˆp(a),ˆc(a)) be the steady-state solution of System (2.6); then, we have

    dˆs(a)da=Λλ(a)Jˆs(a)Λˆs(a)+ε0(1+ε1cos(2πa))ˆp(a)ˆs(a),dˆi(a)da=λ(a)Jˆs(a)δ(a)ˆi(a)Λˆi(a)+ε0(1+ε1cos(2πa))ˆp(a)ˆi(a),dˆp(a)da=δ(a)ˆi(a)(γ(a)+Λ+ε0(1+ε1cos(2πa)))ˆp(a)+ε0(1+ε1cos(2πa))ˆp(a)ˆp(a),dˆc(a)da=γ(a)ˆp(a)Λˆc(a)+ε0(1+ε1cos(2πa))ˆp(a)ˆc(a), (4.1)

    where

    J=aσa1h(b)[ˆi(b)+ˆc(b)]db. (4.2)

    In the case of the absence of infection, the infection-free steady-state solution is E0=(1,0,0,0).

    The coefficient of the left hand side derivative with respect to a of System (2.6) is a diagonal matrix; then, System (2.6) is a hyperbolic system of partial differential equations [27]. Therefore, we apply a perturbation technique to analyze the stability of the infection-free steady-state solution, that is, (˜s(a)eξt,˜i(a)eξt,˜p(a)eξt,˜c(a)eξt) near the steady-state solution E0, i.e.,

    s(a,t)=1+˜s(a)eξt,i(a,t)=0+˜i(a)eξt=˜i(a)eξt,p(a,t)=0+˜p(a)eξt=˜p(a)eξt,c(a,t)=0+˜c(a)eξt=˜c(a)eξt,β(a,t)=λ(a)Ueξt, (4.3)

    where ξ is a real or complex number, and

    U=aσa1h(b)[˜i(b)+˜c(b)]db. (4.4)

    By substituting (4.3) and (4.4) into System (2.6), it follows that (˜s(a),˜i(a),˜p(a),˜c(a)) will satisfy the conditions of the following linearized system:

    d˜s(a)da=(Λ+ξ)˜s(a)λ(a)U+ε0(1+ε1cos2πa)˜p(a),d˜i(a)da=λ(a)U(δ(a)+Λ+ξ)˜i(a),d˜p(a)da=δ(a)˜i(a)(γ(a)+Λ+ξ)˜p(a)ε0(1+ε1cos2πa)˜p(a),d˜c(a)da=γ(a)˜p(a)(Λ+ξ)˜c(a). (4.5)

    The value of U in (4.4) is obtained by solving System (4.5). Considering System (4.5), we get

    ˜s(a)=aa1e(Λ+ξ)(aτ)[ε0(1+ε1cos2πτ)˜p(τ)λ(a)U]dτ, (4.6)
    ˜i(a)=aa1eaτ(δ(j)+Λ+ξ)djλ(τ)Udτ, (4.7)
    ˜p(a)=aa1eaw(ξ+γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ewτ(δ(j)+Λ+ξ)djλ(τ)Udτdw, (4.8)
    ˜c(a)=aa1eaν(Λ+ξ)djγ(ν)νa1eνw(ξ+γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ewτ(δ(j)+Λ+ξ)djλ(τ)Udτdwdν. (4.9)

    Substituting (4.7) and (4.9) into U, we have

    U=aσa1h(b)[ba1ebτ(δ(j)+Λ+ξ)djλ(τ)Udτ+ba1ebν(Λ+ξ)djγ(ν),νa1eνw(ξ+γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ewτ(δ(j)+Λ+ξ)djλ(τ)Udτdwdν]db. (4.10)

    We can rewrite (4.10) by applying U=UˆQ(ξ), where

    ˆQ(ξ)=aσa1h(b)[ba1ebτ(δ(j)+Λ+ξ)djλ(τ)dτ+ba1ebν(Λ+ξ)djγ(ν),νa1eνw(ξ+γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ewτ(δ(j)+Λ+ξ)djλ(τ)dτdwdν]db. (4.11)

    Note that U=UˆQ(ξ) is equivalent to U(1ˆQ(ξ))=0. The solution of this equation is U=0 or

    ˆQ(ξ)=1. (4.12)

    Solution U=0 shows the existence of the infection-free steady-state solution. Meanwhile, ˆQ(ξ)=1 corresponds to the characteristic equation. Considering (4.11), we have

    ˆQ(0)=aσa1h(b)[ba1ebτ(δ(j)+Λ)djλ(τ)dτ+ba1ebν(Λ)djγ(ν),νa1eνw(γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ewτ(δ(j)+Λ)djλ(τ)dτdwdν]db.

    Now, we will show the relationship between ξ and ˆQ(0) to determine the conditions of ξ that satisfy (4.12).

    Lemma 4.1. Let ξ satisfy ˆQ(ξ)=1.

    (1) If ˆQ(0)>1, then ξ>0,

    (2) If ˆQ(0)<1, then ξ<0.

    We use the reductio ad absurdum method to prove Lemma 4.1. By using Lemma 4.1, we show that the threshold criterion of ˆQ(ξ)=1 is valid, and we can define the basic reproduction number as follows:

    R0=aσa1h(b)[ba1ebτ(δ(j)+Λ)djλ(τ)dτ+ba1ebν(Λ)djγ(ν),νa1eνw(γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ewτ(δ(j)+Λ)djλ(τ)dτdwdν]db. (4.13)

    See [28,29] for the detailed method. Biologically, the value of R0 represents the average number of secondary infections by HPV that are produced by one single infected cell or cancer cell during the period of infection, assuming that all other cells in the population are uninfected.

    The first step in analyzing the stability of an infection-free steady-state solution is to show that (4.12) has only one real solution.

    Lemma 4.2. If 0<ε11, ε0,ξ,Λ>0, and the functions h(a),λ(a),δ(a),γ(a) are continuous, positive and defined in a1<τ<w<ν<b<aσ, then ˆQ(ξ)<0, limξˆQ(ξ)=0, and limξˆQ(ξ)=.

    Proof. Let ξ be a real number; then, we obtain

    dˆQdξ=ddξ(aσa1h(b)[ba1ebτ(δ(j)+Λ+ξ)djλ(τ)dτ+ba1ebν(Λ+ξ)djγ(ν)νa1eνw(ξ+γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ewτ(δ(j)+Λ+ξ)djλ(τ)dτdwdν]db)=aσa1h(b)[ba1((bτ)eξ(bτ))ebτ(δ(j)+μ(j))djλ(τ)dτ+ba1ebνμ(j)djγ(ν)νa1eνw(γ(j)+μ(j)+ε0(1+ε1cos2πj))djδ(w)wa1((bτ)eξ(bτ))ewτ(δ(j)+μ(j))djλ(τ)dτdwdν]db<0,

    where τ<b. Furthermore, limξˆQ(ξ)=0 and limξˆQ(ξ)=.

    According to Lemma 4.2, the function ˆQ(ξ) is decreasing monotonically. It implies that the characteristic Eq (4.12) has a real solution and it is unique, denoted by ξ. Next, we will show that (4.12) has a dominant root, where the real part of the other roots are smaller than ξ.

    Lemma 4.3. If ξ is the root of (4.12), then ξ is the dominant root.

    Proof. Suppose that ξ=x+iy is a complex root of (4.12); hence, ξ satisfies (4.12). Then

    ˆQ(ξ)=aσa1h(b)[ba1ebτ(δ(j)+Λ+ξ)djλ(τ)dτ+ba1ebν(Λ+ξ)djγ(ν)νa1eνw(ξ+γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ewτ(δ(j)+Λ+ξ)djλ(τ)dτdwdν]db=aσa1h(b)[ba1e(x+iy)(bτ)ebτ(δ(j)+Λ)djλ(τ)dτ+ba1ebνΛdjγ(ν)νa1eνw(γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1e(x+iy)(bτ)ewτ(δ(j)+Λ)djλ(τ)dτdwdν]db=aσa1h(b)[ba1ex(bτ)(cos((bτ)y)isin((bτ)y))ebτ(δ(j)+Λ)djλ(τ)dτ+ba1ebνΛdjγ(ν)νa1eνw(γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ex(bτ)(cos((bτ)y)isin((bτ)y))ewτ(δ(j)+Λ)djλ(τ)dτdwdν]db.

    By equating the real and imaginary parts, we observe that

    ˆQ(ξ)=aσa1h(b)[ba1ex(bτ)cos((bτ)y)ebτ(δ(j)+Λ)djλ(τ)dτ+ba1ebνΛdjγ(ν)νa1eνw(γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ex(bτ)cos((bτ)y)ewτ(δ(j)+Λ)djλ(τ)dτdwdν]dbaσa1h(b)[ba1ex(bτ)ebτ(δ(j)+Λ)djλ(τ)dτ+ba1ebνΛdjγ(ν)νa1eνw(γ(j)+Λ+ε0(1+ε1cos2πj))djδ(w)wa1ex(bτ)ewτ(δ(j)+Λ)djλ(τ)dτdwdν]db=ˆQ(x)=ˆQ(Re ξ).

    According to Lemma 4.2, for ˆQ(ξ)<0, we acquire that ˆQ(ξ)ˆQ(Reξ). Therefore, we obtain that Reξξ.

    Theorem 4.1. The infection-free steady-state solution E0=(1,0,0,0) is locally asymptotically stable if R0<1, and it is unstable if R0>1.

    Proof. If R0>1, according to Lemma 4.1, it follows that ˆQ(0)>1. Considering Lemma 4.2, let the unique real value root of (4.12) be denoted as ξ, which further implies that ξ>0. Since the characteristic equation has a positive root, the infection-free steady-state (E0) is unstable.

    On the other hand, if R0<1, it follows that ˆQ(0)<1 according to Lemma 4.1, and the root of (4.12) is real and negative. Let us assume that the root of the characteristic equation is ξ. Then, we have that ξ<0. Since, as per Lemma 4.3, ξ is the dominant root of (4.12), the infection-free steady-state (E0) is asymptotically stable.

    The term "clearance" is used when HPV can be eliminated from the tissue. This can be numerically modeled if the initial value of each subpopulation is in the basin of attraction at E0. It means that a global stability analysis needs to be carried out around E0 to determine the conditions that must be obtained for HPV clearance to occur.

    In order to demonstrate the global stability of E0, it suffices to show that, as time t approaches infinity, s(a,t) approaches 1, while i(a,t), p(a,t), and c(a,t) all approach 0. Equation (2.5) can be rewritten as β(a,t)=λ(a)B(t), where

    B(t)=aσa1h(b)[i(b,t)+c(b,t)]db. (4.14)

    Recall that we use ˉε(a)=ε0(1+ε1cos(2πa)) for convenience. Considering (4.14), System (2.6) can be solved for a<t as follows:

    s(a,t)=eaa1Λ+λ(j)B(j+ta)ˉε(j)p(j,ta+j)djaa1Λeτa1Λ+λ(j)B(j+ta)ˉε(j)p(j,ta+j)djdτ+eaa1Λ+λ(j)B(j+ta)ˉε(j)p(j,ta+j)dj, (4.15)
    i(a,t)=eaa1δ(j)+Λˉε(j)p(j,ta+j)djaa1eτa1δ(j)+Λˉε(j)p(j,ta+j)djλ(τ)B(τ+ta)s(τ,τ+ta)dτ, (4.16)
    p(a,t)=eaa1γ(j)+Λ+ˉε(j)(1p(j,ta+j))djaa1eηa1γ(j)+Λ+ˉε(j)(1p(j,ta+j))djδ(η)eηa1δ(j)+Λˉε(j)p(j,ta+j)djηa1eτa1δ(j)+Λˉε(j)p(j,ta+j)djλ(τ)B(τ+ta)s(τ,τ+ta)dτdη, (4.17)
    c(a,t)=eaa1Λˉε(j)p(j,ta+j)djaa1ewa1Λˉε(j)p(j,ta+j)djγ(w)ewa1γ(j)+Λ+ˉε(j)(1p(j,ta+j))djwa1eηa1γ(j)+Λ+ˉε(j)(1p(j,ta+j))djδ(η)eηa1δ(j)+Λˉε(j)p(j,ta+j)djηa1eτa1δ(j)+Λˉε(j)p(j,ta+j)djλ(τ)B(τ+ta)s(τ,τ+ta)dτdηdw. (4.18)

    Substituting (4.16) and (4.18) at E0 into (4.14), we obtain

    B(t)=aσa1h(b)[ba1ebτδ(j)+Λdjλ(τ)B(τ+tb)dτ+ba1ebwΛdjγ(w)wa1ewηγ(j)+Λ+ˉε(j)djδ(η)ηa1eητδ(j)+Λdjλ(τ)B(τ+tb)dτdηdw]db.

    Applying the supremum limit t, we get

    limsuptaσa1h(b)[ba1ebτδ(j)+Λdjλ(τ)B(τ+tb)dτ+ba1ebwΛdjγ(w)wa1ewηγ(j)+Λ+ˉε(j)djδ(η)ηa1eητδ(j)+Λdjλ(τ)B(τ+tb)dτdηdw]dbaσa1h(b)[ba1ebτδ(j)+Λdjλ(τ)limsuptB(τ+tb)dτ+ba1ebwΛdjγ(w)wa1ewηγ(j)+Λ+ˉε(j)djδ(η)ηa1eητδ(j)+Λdjλ(τ)limsuptB(τ+tb)dτdηdw]db.

    Let V(a)=λ(a)limsuptB(t); then,

    V(a)λ(a)C1, (4.19)

    where

    C1=aσa1h(b)[ba1ebτδ(j)+ΛdjV(τ)dτ+ba1ebwΛdjγ(w)wa1ewηγ(j)+Λ+ˉε(j)djδ(η)ηa1eητδ(j)+ΛdjV(τ)dτdηdw]db. (4.20)

    By applying (4.19) and (4.20), we have

    C1aσa1h(b)[ba1ebτδ(j)+Λdjλ(τ)C1dτ+ba1ebwΛdjγ(w)wa1ewηγ(j)+Λ+ˉε(j)djδ(η)ηa1eητδ(j)+Λdjλ(τ)C1dτdηdw]db. (4.21)

    Given (4.13), we note that (4.21) is equivalent to C1C1R0. Hence, it can be rewritten as C1(1R0)0. We maintain the last inequality to show global stability around the infection-free steady-state through Theorem 4.2.

    Theorem 4.2. If R0<1, then the infection-free steady-state E0=(1,0,0,0) is globally asymptotically stable.

    Proof. Since R0<1, for the inequality C1(1R0)0 to be satisfied, it must hold that C1=0. It follows that V(a)=0. As a consequence, we obtain that limsuptB(t)=0. Therefore, we acquire that at E0, the limti(a,t)=limtp(a,t)=limtc(a,t)=0 and limts(a,t)=1, as required.

    We have analyzed the stability when J=0, that is, in an infection-free steady-state (1,0,0,0). This means that there are only susceptible cells in the tissue. We analyze the condition for J>0 and its stability in the following subsections.

    The endemic steady-state solution is obtained by solving System (4.1). Let E1=(ˆs(a),ˆi(a),ˆp(a),ˆc(a)) be the endemic steady-state solution. System (4.1) can be considered as a system of first-order ordinary differential equations in a. Therefore, by multiplying each equation by a specific integration factor and performing simple algebraic calculations, we obtain the solution E1 given by (4.22)–(4.25):

    ˆs(a)=aa1eaτ(λ(j)J+Λε0(1+ε1cos(2πj))ˆp(j))djΛdτ+eaa1(λ(j)J+Λε0(1+ε1cos(2πj))ˆp(j))dj (4.22)
    ˆi(a)=aa1eaτ(δ(j)+Λε0(1+ε1cos(2πj))ˆp(j))djλ(τ)Jˆs(τ)dτ (4.23)
    ˆp(a)=aa1eaτ(γ(j)+Λ+ε0(1+ε1cos(2πj))ε0(1+ε1cos(2πj))ˆp(j))djδ(τ)ˆi(τ)dτ (4.24)
    ˆc(a)=aa1eaτ(Λε0(1+ε1cos(2πj))ˆp(j))djγ(τ)ˆp(τ)dτ. (4.25)

    The solution of E1 exists if ˆs(a),ˆi(a),ˆp(a), and ˆc(a) are positive. We can see that (4.22) is always positive. The solution ˆi(a) in (4.23) is positive for J that is positive. Moreover, ˆp(a)>0 and ˆc(a)>0 if ˆi(a)>0. Therefore, it is necessary to find biologically meaningful conditions for the positive value of J. Now, substituting ˆi(a) in (4.23) into ˆp(a) in (4.24), we can substitute the resulting ˆp(a) into ˆc(a) in (4.25); then, (4.2) becomes as follows:

    J=aσa1h(b)ba1ebw(δ(j)+Λε0(1+ε1cos(2πj))ˆp(j))djJλ(w)ˆs(w)dwdb+aσa1h(b)ba1ebw(Λε0(1+ε1cos(2πj))ˆp(j))djγ(w)wa1ewη(γ(j)+Λ+ε0(1+ε1cos(2πj))ε0(1+ε1cos(2πj))ˆp(j))djδ(η)ηa1eητ(δ(j)+Λε0(1+ε1cos(2πj))ˆp(j))djJλ(τ)ˆs(τ)dτdηdwdb. (4.26)

    We denote that

    ˆQ(J)=aσa1h(b)ba1ebw(δ(j)+Λε0(1+ε1cos(2πj))ˆp(j))djλ(w)ˆs(w)dwdb+aσa1h(b)ba1ebw(Λε0(1+ε1cos(2πj))ˆp(j))djγ(w)wa1ewη(γ(j)+Λ+ε0(1+ε1cos(2πj))ε0(1+ε1cos(2πj))ˆp(j))djδ(η)ηa1eητ(δ(j)+Λε0(1+ε1cos(2πj))ˆp(j))djλ(τ)ˆs(τ)dτdηdwdb. (4.27)

    Considering (4.26), (4.27) can be rewritten as J=JˆQ(J) or J(1ˆQ(J))=0. Thus, we have that J=0 or

    ˆQ(J)=1. (4.28)

    Note that, using (4.27), we obtain

    ˆQ(0)=aσa1h(b)ba1ebw(δ(j)+Λ)djλ(w)dwdb+aσa1h(b)ba1ebw(Λ)djγ(w)wa1ewη(γ(j)+Λ+ε0(1+ε1cos(2πj)))djδ(η)ηa1eητ(δ(j)+Λ)djλ(τ)dτdηdwdb.

    By applying (4.13), we can see that ˆQ(0)=R0. On the other hand, since ˆi(a)+ˆc(a)<1, then

    JˆQ(J)=aσa1h(b)(ˆi(b)+ˆc(b))db<aσa1h(b)db.

    If aσa1h(b)db=h+, then JˆQ(J)<h+. In particular, if J=h+, we have that h+ˆQ(h+)<h+, which means that

    ˆQ(h+)<1. (4.29)

    The existence of a positive endemic steady-state can be summarized as follows. Since we require the necessary condition of existence of the endemic steady-state, we will determine the condition for J to be positive. Considering (4.28), we have that ˆQ(J)=1. It is noteworthy that ˆQ(0)=R0 and the function ˆQ(J) is monotonically decreasing. By (4.29), we find that ˆQ(h+)<1. Therefore, the equation ˆQ(J)=1 will have a positive intersection if ˆQ(0)=R0 is greater than 1. The proof of this condition is in Lemma 4.4, and we note that (4.29) is the inequality used to prove Lemma 4.4.

    Lemma 4.4. If R0>1, then the endemic steady-state exists.

    Proof. Since R0=ˆQ(0)>1, (4.29) states that ˆQ(h+)<1, and ˆQ(J) is continuous, then the solution of (4.28) is J>0. Furthermore, for J>0, the solution that satisfies the steady-state system (4.1) is (ˆs(a),ˆi(a),ˆp(a),ˆc(a)), as given by (4.22)–(4.25).

    Stability analysis will be discussed by using the perturbation approach and using the exponential solutions (˜s(a)eξt,˜i(a)eξt,˜p(a)eξt,˜c(a)eξt) near the steady-state solution (ˆs(a),ˆi(a),ˆp(a),ˆc(a)), as follows:

    s(a,t)=ˆs(a)+˜s(a)eξt,i(a,t)=ˆi(a)+˜i(a)eξt,p(a,t)=ˆp(a)+˜p(a)eξt,c(a,t)=ˆc(a)+˜c(a)eξt, (4.30)

    where (ˆs(a),ˆi(a),ˆp(a),ˆc(a)), as shown in Lemma 4.4. We obtain

    β(a,t)=λ(a)aσa1h(b)(ˆi(b)+˜i(b)eξt+ˆc(b)+˜c(b)eξt)db=λ(a)J+λ(a)Ueξt, (4.31)

    where

    U=aσa1h(b)(˜i(b)+˜c(b))db. (4.32)

    Using (4.30)–(4.32), the functions (˜s(a),˜i(a),˜p(a),˜c(a)) will satisfy the conditions of the following linear system:

    d˜s(a)da=(Λ+λ(a)J+ξε0(1+ε1cos2πa)ˆp(a))˜s(a)λ(a)Uˆs(a)+ε0(1+ε1cos2πa)˜p(a)ˆs(a),d˜i(a)da=(λ(a)J˜s(a)+λ(a)Uˆs(a))(δ(a)+Λ+ξε0(1+ε1cos(2πa))ˆp(a))˜i(a)+ε0(1+ε1cos(2πa))˜p(a)ˆi(a),d˜p(a)da=δ(a)˜i(a)(ξ+γ(a)+Λ+ε0(1+ε1cos(2πa)))˜p(a)+2ε0(1+ε1cos(2πa))ˆp(a)˜p(a),d˜c(a)da=γ(a)˜p(a)(Λ+ξ)˜c(a)+ε0(1+ε1cos(2πa))ˆp(a)˜c(a)+ε0(1+ε1cos(2πa))˜p(a)ˆc(a).

    Since U0, to simplify the analysis, we can apply the following ratio:

    ˉs(a)=˜s(a)U,ˉi(a)=˜i(a)U,ˉp(a)=˜p(a)U,ˉc(a)=˜c(a)U.

    Furthermore, (4.32) becomes

    1=aσa1h(b)(ˉi(b)+ˉc(b))db, (4.33)

    and we obtain the following system:

    dˉs(a)da=(Λ+λ(a)J+ξε0(1+ε1cos2πa)ˆp(a))ˉs(a)λ(a)ˆs(a)+ε0(1+ε1cos2πa)ˉp(a)ˆs(a),dˉi(a)da=λ(a)Jˉs(a)+λ(a)ˆs(a)(δ(a)+Λ+ξε0(1+ε1cos(2πa))ˆp(a))ˉi(a)+ε0(1+ε1cos(2πa))ˉp(a)ˆi(a),dˉp(a)da=δ(a)ˉi(a)(ξ+γ(a)+Λ+ε0(1+ε1cos(2πa)))ˉp(a)+2ε0(1+ε1cos(2πa))ˆp(a)ˉp(a),dˉc(a)da=γ(a)ˉp(a)(Λ+ξ)ˉc(a)+ε0(1+ε1cos(2πa))ˆp(a)ˉc(a)+ε0(1+ε1cos(2πa))ˉp(a)ˆc(a). (4.34)

    The solution of System (4.34) is given by

    ˉs(a)=aa1e(Λ+ξ)(aτ)eaτ(λ(j)Jε0(1+ε1cos2πj)ˆp(j))dj(ε0(1+ε1cos2πτ)ˉp(τ)λ(τ))ˆs(τ)dτ, (4.35)
    ˉi(a)=aa1e(Λ+ξ)(aw)eaw(δ(j)ε0(1+ε1cos(2πj))ˆp(j))dj[λ(w)Jˉs(w)+λ(w)ˆs(w)+ε0(1+ε1cos(2πw))ˉp(w)ˆi(w)]dw, (4.36)
    ˉp(a)=aa1e(ξ+Λ)(aτ)eaτ(γ(j)+ε0(1+ε1cos(2πj)ε0(1+ε1cos(2πj))ˆp(j))dj(δ(τ)ˉi(τ)+ε0(1+ε1cos(2πj))ˆp(τ)ˉp(τ))dτ, (4.37)
    ˉc(a)=aa1e(Λ+ξ)(aw)eawε0(1+ε1cos(2πj))ˆp(j)djγ(w)ˉp(w)dw+aa1e(Λ+ξ)(aw)eawε0(1+ε1cos(2πj))ˆp(j)djε0(1+ε1cos(2πw))ˆc(w)ˉp(w)dw. (4.38)

    The value of ˉs(a) is negative if ε0(1+ε1cos2πa)ˉp(a)<λ(a). It can be interpreted that, if the immunotherapy treatment parameter is less than the virus infection parameter, then an endemic condition appears in the tissue. To simplify the function in subsequent calculations, (4.36) can be substituted into (4.37); we get

    ˉp(a)=aa1e(ξ+Λ)(aτ)eaτ(γ(j)+ε0(1+ε1cos(2πj)ε0(1+ε1cos(2πj))ˆp(j))djδ(τ)τa1eτwδ(j)dj[λ(w)Jˉs(w)+λ(w)ˆs(w)+ε0(1+ε1cos(2πw))ˉp(w)ˆi(w)]dwdτ+aa1e(ξ+Λ)(aτ)eaτ(γ(j)+ε0(1+ε1cos(2πj)ε0(1+ε1cos(2πj))ˆp(j))djε0(1+ε1cos(2πj))ˆp(τ)ˉp(τ)dτ. (4.39)

    If we substitute the solutions of ˉi(a) and ˉc(a) into (4.33), we obtain ˉQ(ξ) on the right-hand side of (4.33). Then, if we further substitute ˉs(a) from (4.35) into ˉQ(ξ), we obtain

    ˉQ(ξ)=aσa1h(b)ba1e(Λ+ξ)(bw)ebw(δ(j)ε0(1+ε1cos(2πj))ˆp(j))djε0(1+ε1cos(2πw))ˉp(w)ˆi(w)dwdb+aσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djλ(w)ˆs(w)(ebwδ(j)djJwτebτδ(j)djewτλ(j)Jdjλ(τ)dτ)dwdb+Jaσa1h(b)ba1e(Λ+ξ)(bw)ebw(δ(j)ε0(1+ε1cos(2πj))ˆp(j))djλ(w)wτewτλ(j)Jdjε0(1+ε1cos2πτ)ˉp(τ)ˆs(τ)dτdwdb+aσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djγ(w)wηewη(γ(j)+ε0(1+ε1cos(2πj))djδ(η)ηνeηwδ(j)djε0(1+ε1cos(2πν))ˉp(ν)ˆi(ν)dνdηdwdb+aσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djγ(w)wηewη(γ(j)+ε0(1+ε1cos(2πj))djδ(η)ηνλ(ν)ˆs(ν)(ebνδ(j)djJντebτδ(j)djeντλ(j)Jdjλ(τ)dτ)dνdηdwdb+Jaσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djγ(w)wηewη(γ(j)+ε0(1+ε1cos(2πj))djδ(η)ηνeηwδ(j)djλ(ν)ντeντλ(j)Jdjε0(1+ε1cos2πτ)ˉp(τ)ˆs(τ)dτdνdηdwdb+aσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djγ(w)wηewη(γ(j)+ε0(1+ε1cos(2πj))ε0(1+ε1cos(2πη))ˆp(η)ˉp(η)dηdwdb+aσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djε0(1+ε1cos(2πw))ˉp(w)ˆc(w)dwdb. (4.40)

    If Ψ(b,w)=ebw(δ(j))djJwa1λ(τ)ewτ(λ(j)Jε0(1+ε1cos2πj)ˆp(j))djebτδ(j)djdτ and we let

    L1(w)=ε0(1+ε1cos(2πw))ˉp(w)[ˆc(w)+ebwδ(j)djˆi(w)+wηewη(γ(j)+ε0(1+ε1cos(2πj))djγ(η)ˆp(η)dη+wηewη(γ(j)+ε0(1+ε1cos(2πj))djγ(η)δ(η)ηνeηwδ(j)djˆi(ν)dνdη],

    and

    L2(w)=ε0(1+ε1cos2πw)ˉp(w)[wτλ(τ)ebτδ(j)djewτλ(j)Jdjˆs(τ)dτ+wηewη(γ(j)+ε0(1+ε1cos(2πj))djγ(η)δ(η)ηνeηwδ(j)djντeντλ(j)Jdjλ(τ)ˆs(τ)dτdνdη],

    then

    ˉQ(ξ)=aσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djL1(w)dwdb+Jaσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djL2(w)dwdb+aσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djλ(w)ˆs(w)Ψ(b,w)dwdb+aσa1h(b)ba1e(Λ+ξ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djλ(w)ˆs(w)Ψ(b,w)wηewη(γ(j)+ε0(1+ε1cos(2πj))djγ(η)δ(η)ηνeηwδ(j)djdνdηdwdb.

    Thus, ˉQ(ξ) can be written as follows:

    ˉQ(ξ)=aσa1h(b)ba1e(Λ+ξ)(bw)F(b,w)dwdb, (4.41)

    where

    F(b,w)=ebwε0(1+ε1cos(2πj))ˆp(j)dj[L1(w)+JL2(w)+λ(w)ˆs(w)Ψ(b,w)(1+wηewη(γ(j)+ε0(1+ε1cos(2πj))djγ(η)δ(η)ηνeηwδ(j)djdνdη)].

    Lemma 4.5 proves that ˉQ(ξ) is a monotonically decreasing function. We prove that the first derivative of ˉQ(ξ) is negative, ˉQ(ξ) goes to zero as ξ tends to infinity, and ˉQ(ξ) goes to infinity as ξ tends to negative infinity.

    Lemma 4.5. If L1(w)<0 and F(b,w)0, then ˉQ(ξ)<0,limξˉQ(ξ)=0, limξˉQ(ξ)=, and ˉQ(0)<1.

    Proof. We will prove the first statement. If F(b,w)0, then ˉQ(ξ)0 and

    ddξˉQ(ξ)=ddξaσa1h(b)ba1e(Λ+ξ)(bw)F(b,w)dwdb=aσa1h(b)ba1(bw)e(Λ+ξ)(bw)F(b,w)dwdb<0,

    since b>w. For the second statement, it is easy to see that ˉQ(ξ)<0,limξˉQ(ξ)=0, and limξˉQ(ξ)=. Now, we will prove that ¯Q(0)<1. By applying (4.40), we have

    ˉQ(0)=aσa1h(b)ba1e(Λ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djλ(w)ˆs(w)ebwδ(j)dj+aσa1h(b)ba1e(Λ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djγ(w)wa1ewη(γ(j)+ε0(1+ε1cos(2πj))djδ(η)ηa1eηwδ(j)djλ(ν)ˆs(ν)dνdηdwdbJaσa1h(b)bwebτδ(j)djλ(w)wτe(Λ)(bw)ewτλ(j)Jdjebwε0(1+ε1cos(2πj))ˆp(j)djλ(τ)ˆs(τ)dτdwdb+Jaσa1h(b)bwebwδ(j)djλ(w)wτe(Λ)(bw)ebwλ(j)Jdjebwε0(1+ε1cos(2πj))ˆp(j)djε0(1+ε1cos2πτ)ˉp(τ)ˆs(τ)dτdwdbJaσa1h(b)bwebτδ(j)djγ(w)wηewη(γ(j)+ε0(1+ε1cos(2πj))djδ(η)ηνeηwδ(j)djλ(ν)ντeντλ(j)Jdje(Λ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djλ(τ)ˆs(τ)dτdνdηdwdb+Jaσa1h(b)ba1γ(w)wηewη(γ(j)+ε0(1+ε1cos(2πj))djδ(η)ηνeηwδ(j)djλ(ν)ντewτλ(j)Jdje(Λ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djε0(1+ε1cos2πτ)ˉp(τ)ˆs(τ)dτdνdηdwdb+aσa1h(b)ba1e(Λ)(bw)ebw(δ(j)ε0(1+ε1cos(2πj))ˆp(j))djε0(1+ε1cos(2πw))ˉp(w)ˆi(w)dwdb+aσa1h(b)ba1e(Λ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djγ(w)wa1ewη(γ(j)+ε0(1+ε1cos(2πj))djδ(η)ηa1eηwδ(j)djε0(1+ε1cos(2πν))ˉp(ν)ˆi(ν)dνdηdwdb+aσa1h(b)ba1e(Λ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djγ(w)wa1ewη(γ(j)+ε0(1+ε1cos(2πj))ε0(1+ε1cos(2πη))ˆp(η)ˉp(η)dηdwdb+aσa1h(b)ba1e(Λ)(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djε0(1+ε1cos(2πw))ˉp(w)ˆc(w)dwdb.

    Moreover, by applying (4.28), we notice that the two-first integrals are equal to one. Thus, ˉQ(0) can be written as

    ˉQ(0)=1+Jaσa1h(b)bwebτδ(j)djλ(w)ˉs(w)dwdb+Jaσa1h(b)bwebτδ(j)djγ(w)wηewη(γ(j)+ε0(1+ε1cos(2πj))djδ(η)ηνeηwδ(j)djλ(ν)ˉs(ν)dηdwdb+aσa1h(b)ba1eΛ(bw)ebwε0(1+ε1cos(2πj))ˆp(j)djL1(w)dwdb.

    Since ˉs(a)<0 and L1(w)<0, then ˉQ(0)<1.

    Theorem 4.3. If R0>1, then the endemic steady-state is locally asymptotically stable.

    Proof. By applying Lemma 4.4, we found that the endemic steady-state exists for R0>1. Lemma 4.5 guarantees that ˉQ(ξ) is a monotonically decreasing function such that ˉQ(ξ)=1 has a unique real root. On the other hand, considering Lemma 4.5, for ˉQ(0)<1, we obtain ξ<0 as the negative real root. The appearance of this negative real root means that the endemic steady-state of System (2.6) is locally asymptotically stable.

    In this simulation, we investigate the model in a constant scenario. This indicates that all parameters have fixed values. We used the parameter values shown in the Table 1. In this case, different values of h produce different values of R0. Recall that h denotes the infection rate associated with susceptible cells being transformed to infected cells; the larger the value of h, the larger the value of R0.

    Table 1.  Parameter value.
    Parameter Value Reference
    Λ 0.1 [30]
    γ 0.02 [7]
    δ 0.15 [7]
    h 0.07 Assumed
    λ 0.65 Assumed
    ε0 0.005-0.123 [31]
    ε1 0<ε11 [14]

     | Show Table
    DownLoad: CSV

    Figures 2a and 2b align with Theorem 4.1, which states that the solution is an infection-free steady-state solution if R0<1. According to Figures 2c and 2d, the solution tends to the endemic steady-state solution if R0>1, which is in line with Theorem 4.3. In this work, immunotherapy was applied as the treatment to reduce precancerous cells to control the endemic condition. In Figure 3, the graph for ε0=0 illustrates the density of precancerous cells when no immunotherapy treatment has been given. When the cell age is between 0 and 10, there is a significant increase in the density of precancerous cells. This is consistent with the fact that HPV will damage cells when cells are in the initial phase of their cycle, namely, between the G1 and S phases [32].

    Figure 2.  Given h=0.07, the value of R0=0.88<1. The cell population tends to infection-free steady-state (1,0,0,0). If h=0.65, then R0=8.17>1 and the cell population tends to the endemic steady-state (0.4588,0.2677,0.22,0.053).
    Figure 3.  Precancerous cell densitites with and without immunotherapy. In this case, ε0=0 is indicated by a dotted line, ε0=0.0801 is indicated by a solid line, ε0=0.09 is indicated by a dash-dotted line, and ε0=0.123 is indicated by dashed line. The larger the value of ε0, the faster the rate at which the precancerous cells decrease.

    For this reason, it is hoped that immunotherapy can suppress the growth of precancerous cells. According to the Lemma 4.4, an endemic steady-state condition is defined if R0>1. Furthermore, according to Theorem 4.3, the endemic steady-state conditions are locally asymptotically stable. It can be interpreted analytically that precancerous cells are still present in the cervical tissue. Figure 3 shows that precancerous cells are still present when ε0=0.0801. However, the density is less than the case without therapy. When ε0=0.09 (larger than before), the density of precancerous cells decrease even more. The precancerous cell population remains in cervical tissue, as indicated for ε0=0.123. Precancerous cells are not entirely eliminated since the R0 value is still larger than one.

    Moreover, as h increases, more susceptible cells become infected. Thus, it can be seen in Figure 4a that a larger ε0 is required through greater T-cell proliferation. If the infection rate for susceptible cells is h=0.5625 and the effects of immunotherapy on T-cell proliferation is represented by ε0=0.0179, then the infected, precancerous, and cancerous cells decrease rapidly, as shown in Figure 4b.

    Figure 4.  The rate of T-cell proliferation as a result of immunotherapy (ε0) is directly proportional to the infection rate (h). It is to be noted that, although R0>1 and there are some different initial values, the solutions of System (2.6) eventually converge to the same value at appropriate values of ε0 and h. It suggests that we can control the proliferation of T-cells to achieve better outcomes for patients.

    When R0=1, the dynamical behavior is related to the characteristic root, which has a value of zero. However, working directly with R0 can be complicated, so it is straightforward to observe the parameters that form R0. In this case, we chose to use the parameter h to form R0 and then determine the value of h that makes R0=1. On the other hand, the characteristic root is correlated with J in (4.2). If J=0, then the system is in an infection-free steady-state condition, while, if J>0, the system is in an endemic steady-state condition. To illustrate, we have provided Figure 5, which shows that h=0.079 results in R0=1. We also show in Figure 6 that a value of h less than or more than h causes the behavior of precancerous and cancer cells to change significantly.

    Figure 5.  The value h=0.079 leads to R0=1.
    Figure 6.  Precancerous and cancer cell densities for h<h and h>h. For h=0.5>h and h=0.6>h, the densities of precancerous and cancer cells indicate an endemic steady-state. Meanwhile, for h=0.03<h, the densities of precancerous and cancer cells indicate an infection-free steady-state.

    We propose a new mathematical model of age-structured development of cervical cancer cells. We consider an age-dependent immunotherapy treatment for the precancerous cell population. Regarding the basic reproduction number, we have examined the existence and stability of the infection-free and endemic steady-states. Determination of the basic reproduction number in this work began by proofing Lemma 4.1. Lemma 4.1 provides some insights into the threshold value of the basic reproduction number, which is important for the stability analysis of the model.

    The value of R0=1, as shown in Figure 5, can be observed as a bifurcation point since it marks the boundary between two distinct behaviors. It could happen due to the changes in stability. For this value of R0, we will have a degenerate case in which linearization near the infection-free steady-state condition yields a characteristic equation with a zero root. Intuitively, the endemic steady-state is forwardly bifurcating from the infection-free steady-state at R0=1; see [23,33].

    The results of the theoretical analysis and numerical simulation of the model suggest that immunotherapy is more effective in the initial precancerous phase. Significant changes in precancerous cell population behavior occur when the cell age is between 0 to 10. This means that the simulation results provide valuable insights into the behavior of precancerous cells and the effectiveness of immunotherapy in the early phase of its cycle, namely, the G1-to-S phase.

    In a previous model [14], the effects of representing immunotherapy treatment as a periodic function were examined. However, the model did not specifically provide the therapy based on cell age. On the other hand, the models in [7,9] were related to age structure, but the authors did not discuss the provision of therapy. In contrast, in this research, we indicate that administering therapy based on cell age is a promising approach that can lead to more optimal results. This study clearly demonstrates the potential benefits of administering therapy based on cell age, especially in the early phase. Therefore, if immunotherapy drugs are given when precancerous cells are in this phase, there is a high probability of success of immunotherapy. One way to see if a cell is in a particular phase of its cycle is to employ flow cytometry, which uses light scattering.

    However, our work does not take into account some considerable findings, i.e., optimal control, cervical cancer metastasis condition, and uniform persistence analysis. The concept of persistence in dynamical systems is of significant importance; see [34]. Persistence is referred to as the long-term survival of some or all system components when predicting the behavior of systems. Although we have not addressed this here, it could be addressed in future research.

    Moreover, in this research, immunotherapy term involved the assumption of T-cells as immune cells. Naturally, T-cells have the ability to remember past infection by pathogens upon subsequent exposures. On the other hand, by allowing the derivative order to be a non-integer value, fractional derivatives can capture the influence of past states and events on the current state of the system. Therefore, fractional derivatives can be used to model the dynamics of immune cells, including the memory response; see [35,36] for detailed explanations. Hence, for further research, the HPV infection model with immunotherapy can be developed with fractional derivatives.

    The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The first author would like to express gratitude to the Indonesian Endowment Fund for Education (LPDP) Indonesia for awarding the Doctoral scholarship. All authors extend their appreciation to Universitas Gadjah Mada for partially funding this work through "Rekognisi Tugas Akhir" 2023, with the assignment letter number 5075/UN1.P.II/Dit-Lit/PT.01.01/2023.

    The authors declare that they have no conflict of interest.



    [1] P. Zibako, N. Tsikai, S. Manyame, T. G. Ginindza, Cervical cancer management in Zimbabwe (2019–2020), Plos one, 17 (2022), e0274884. https://doi.org/10.1371/journal.pone.0274884 doi: 10.1371/journal.pone.0274884
    [2] D. S. Chen, I. Mellman, Oncology meets immunology: the cancer-immunity cycle, Immunity, 39 (2013), 1–10. https://doi.org/10.1016/j.immuni.2013.07.012 doi: 10.1016/j.immuni.2013.07.012
    [3] L. Ferrall, K. Y. Lin, R. B. S Roden, C. F. Hung, T. C. Wu, Cervical cancer immunotherapy: facts and hopes, Clin. Cancer Res., 27 (2021), 4953–4973. https://doi.org/10.1158/1078-0432.CCR-20-2833 doi: 10.1158/1078-0432.CCR-20-2833
    [4] T. S. N. Asih, S. Lenhart, S. Wise, L. Aryati, F. Adi-Kusumo, M. S. Hardianti, et al., The dynamics of HPV infection and cervical cancer cells, Bull. Math. Biol., 78 (2016), 4–20. https://doi.org/10.1007/s11538-015-0124-2 doi: 10.1007/s11538-015-0124-2
    [5] L. Aryati, T. S. Noor-Asih, F. Adi-Kusumo, M. S. Hardianti, Global stability of the disease free equilibrium in a cervical cancer model: a chance to recover, FJMS, 103 (2018), 1535–1546. https://doi.org/10.17654/MS103101535 doi: 10.17654/MS103101535
    [6] K. Allali, Stability analysis and optimal control of HPV infection model with early-stage cervical cancer, Biosystems, 199 (2021), 104321. https://doi.org/10.1016/j.biosystems.2020.104321 doi: 10.1016/j.biosystems.2020.104321
    [7] V. V. Akimenko, F. Adi-Kusumo, Stability analysis of an age-structured model of cervical cancer cells and HPV dynamics, Math. Biosci. Eng., 18 (2021), 6155–6177. https://doi.org/10.3934/mbe.2021308 doi: 10.3934/mbe.2021308
    [8] V. V. Akimenko, F. Adi-Kusumo, Age-structured delayed SIPCV epidemic model of HPV and cervical cancer cells dynamics I. Numerical method, Biomath, 10 (2021), 1–23. https://doi.org/10.11145/j.biomath.2021.10.027 doi: 10.11145/j.biomath.2021.10.027
    [9] E. R. Sari, F. Adi-Kusumo, L. Aryati, Mathematical analysis of a SIPC age-structured model of cervical cancer, Math. Biosci. Eng., 19 (2022), 6013–6039. https://doi.org/10.3934/mbe.2022281 doi: 10.3934/mbe.2022281
    [10] V. V. Akimenko, F. Adi-Kusumo, Age-structured delayed SIPCV epidemic model of HPV and cervical cancer cells dynamics II. Convergence of numerical solution, Biomath, 11 (2022), 1–20. https://doi.org/10.55630/j.biomath.2022.03.278 doi: 10.55630/j.biomath.2022.03.278
    [11] N. Cifuentes-Muñoz, R. E. Dutch, R. Cattaneo, Direct cell-to-cell transmission of respiratory viruses: The fast lanes, PLoS pathog., 14 (2018), e1007015. https://doi.org/10.1371/journal.ppat.1007015 doi: 10.1371/journal.ppat.1007015
    [12] D. Kirschner, J. C. Panetta, Modeling immunotherapy of the tumor-immune interaction, J. Math Biol., 37 (1998), 235–252. https://doi.org/10.1007/s002850050127 doi: 10.1007/s002850050127
    [13] E. Nikolopoulou, L. R. Johnson, D. Harris, J. D. Nagy, E. C. Stites, Y. Kuang, Tumour-immune dynamics with an immune checkpoint inhibitor, Lett. Biomath., 5 (2018), 137–159. https://doi.org/10.30707/LiB5.2Nikolopoulou doi: 10.30707/LiB5.2Nikolopoulou
    [14] F. Adi-Kusumo, R. S. Winanda, Bifurcation analysis of the cervical cancer cells, effector cells, and IL-2 compounds interaction model with immunotherapy, FJMS, 99 (2016), 869–883. http://dx.doi.org/10.17654/MS099060869 doi: 10.17654/MS099060869
    [15] A. Pavan, I. Attili, G. Pasello, V. Guarneri, P. F. Conte, L. Bonanno, Immunotherapy in small-cell lung cancer: from molecular promises to clinical challenges, J. Immunother. Cancer, 7 (2019), 205. https://doi.org/10.1186/s40425-019-0690-1 doi: 10.1186/s40425-019-0690-1
    [16] H. Tabbal, A. Septier, M. Mathieu, C. Drelon, S. Rodriguez, C. Djari, et al., EZH2 cooperates with E2F1 to stimulate expression of genes involved in adrenocortical carcinoma aggressiveness, Br. J. Cancer, 121 (2019), 384–394. https://doi.org/10.1038/s41416-019-0538-y doi: 10.1038/s41416-019-0538-y
    [17] M. Barberis, T. Helikar, P. Verbruggen, Simulation of stimulation: Cytokine dosage and cell cycle crosstalk driving timing-dependent T cell differentiation, Front. Physiol., 9 (2018), 879. https://doi.org/10.3389/fphys.2018.00879 doi: 10.3389/fphys.2018.00879
    [18] Y. Pala, M. O. Ertas, An analytical method for solving general Riccati equation, Int. J. Math. Comput. Sci., 11 (2017), 125–130. https://doi.org/10.5281/zenodo.1340124 doi: 10.5281/zenodo.1340124
    [19] Z. Ahmed, M. Kalim, A new transformation technique to find the analytical solution of general second order linear ordinary differential equation, Int. J. Adv. Appl. Sci., 5 (2018), 109–114. https://doi.org/10.21833/ijaas.2018.04.014 doi: 10.21833/ijaas.2018.04.014
    [20] D. P. Singh, A. Ujlayan, An alternative approach to write the general solution of a class of second-order linear differential equations, Resonance, 26 (2021), 705–714. https://doi.org/10.1007/s12045-021-1170-8 doi: 10.1007/s12045-021-1170-8
    [21] A. Wilmer Ⅲ, G. B. Costa, Solving second-order differential equations with variable coefficients, Int. J. Math. Educ. Sci., 39 (2008), 238–243. https://doi.org/10.1080/00207390701464709 doi: 10.1080/00207390701464709
    [22] A. Khan, G. Zaman, Global analysis of an age-structured SEIR endemic model, Chaos Soliton. Fract., 108 (2018), 154–165. https://doi.org/10.1016/j.chaos.2018.01.037 doi: 10.1016/j.chaos.2018.01.037
    [23] H. Inaba, Mathematical analysis of an age-structured SIR epidemic model with vertical transmission, Discrete Cont. Dyn. B, 6 (2006), 69–96. https://doi.org/10.3934/dcdsb.2006.6.69 doi: 10.3934/dcdsb.2006.6.69
    [24] A. H. Siddiqi, Functional analysis and applications, Singapore: Springer, 2018. https://doi.org/10.1007/978-981-10-3725-2
    [25] A. Pazy, Semigroups of linear operators and applications to partial differential equations, New York: Springer, 1983. https://doi.org/10.1007/978-1-4612-5561-1
    [26] G. F. Webb, Theory of nonlinear age-dependent population dynamics, Marcel Dekker, 1985.
    [27] L. C. Evans, Partial differential equations, American Mathematical Society, 2010.
    [28] O. Diekmann, J. A. P. Heesterbeek, J. A. Metz, On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations, J. Math. Biol., 28 (1990), 365–382. https://doi.org/10.1007/BF00178324 doi: 10.1007/BF00178324
    [29] C. Castillo-Chavez, H. W. Hethcote, V. Andreasen, S. A. Levin, W. M. Liu, Epidemiological models with age structure, proportionate mixing, and cross-immunity, J. Math. Biology, 27 (1989), 233–258. https://doi.org/10.1007/BF00275810 doi: 10.1007/BF00275810
    [30] A. K. Miller, K. Munger, F. R. Adler, A mathematical model of cell cycle dysregulation due to Human Papillomavirus infection, Bull. Math. Biol., 79 (2017), 1564–1585. https://doi.org/10.1007/s11538-017-0299-9 doi: 10.1007/s11538-017-0299-9
    [31] H. Cho, Z. Wang, D. Levy, Study of dose-dependent combination immunotherapy using engineered T cells and IL-2 in cervical cancer, J. Theor. Biol., 505 (2020), 110403. https://doi.org/10.1016/j.jtbi.2020.110403 doi: 10.1016/j.jtbi.2020.110403
    [32] S. Patil, R. S. Rao, N. Amrutha, D. S. Sanketh, Analysis of human papillomavirus in oral squamous cell carcinoma using p16: An immunohistochemical study, J. Int. Soc. Prev. Commu., 4 (2014), 61–66. https://doi.org/10.4103/2231-0762.131269 doi: 10.4103/2231-0762.131269
    [33] X. Z. Li, J. Yang, M. Martcheva, Age structured epidemic modeling, Switzerland: Springer Nature, 2020. https://doi.org/10.1007/978-3-030-42496-1
    [34] K. Hattaf, Y. Yang, Global dynamics of an age-structured viral infection model with general incidence function and absorption, Int. J. Biomath., 11 (2018), 1850065. https://doi.org/10.1142/S1793524518500651 doi: 10.1142/S1793524518500651
    [35] K. Hattaf, A new class of generalized fractal and fractal-fractional derivatives with non-singular kernels, Fractal Fract., 7 (2023), 395. https://doi.org/10.3390/fractalfract7050395 doi: 10.3390/fractalfract7050395
    [36] K. Hattaf, A new mixed fractional derivative with applications in computational biology, Computation, 12 (2024), 7. https://doi.org/10.3390/computation12010007 doi: 10.3390/computation12010007
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1035) PDF downloads(48) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog