Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Analysis of yellow fever prevention strategy from the perspective of mathematical model and cost-effectiveness analysis


  • We developed a new mathematical model for yellow fever under three types of intervention strategies: vaccination, hospitalization, and fumigation. Additionally, the side effects of the yellow fever vaccine were also considered in our model. To analyze the best intervention strategies, we constructed our model as an optimal control model. The stability of the equilibrium points and basic reproduction number of the model are presented. Our model indicates that when yellow fever becomes endemic or disappears from the population, it depends on the value of the basic reproduction number, whether it larger or smaller than one. Using the Pontryagin maximum principle, we characterized our optimal control problem. From numerical experiments, we show that the optimal levels of each control must be justified, depending on the strategies chosen to optimally control the spread of yellow fever.

    Citation: Bevina D. Handari, Dipo Aldila, Bunga O. Dewi, Hanna Rosuliyana, Sarbaz H. A. Khosnaw. Analysis of yellow fever prevention strategy from the perspective of mathematical model and cost-effectiveness analysis[J]. Mathematical Biosciences and Engineering, 2022, 19(2): 1786-1824. doi: 10.3934/mbe.2022084

    Related Papers:

    [1] Djamila Moulay, M. A. Aziz-Alaoui, Hee-Dae Kwon . Optimal control of chikungunya disease: Larvae reduction, treatment and prevention. Mathematical Biosciences and Engineering, 2012, 9(2): 369-392. doi: 10.3934/mbe.2012.9.369
    [2] Lili Liu, Xi Wang, Yazhi Li . Mathematical analysis and optimal control of an epidemic model with vaccination and different infectivity. Mathematical Biosciences and Engineering, 2023, 20(12): 20914-20938. doi: 10.3934/mbe.2023925
    [3] Maeve L. McCarthy, Dorothy I. Wallace . Optimal control of a tick population with a view to control of Rocky Mountain Spotted Fever. Mathematical Biosciences and Engineering, 2023, 20(10): 18916-18938. doi: 10.3934/mbe.2023837
    [4] Holly Gaff, Elsa Schaefer . Optimal control applied to vaccination and treatment strategies for various epidemiological models. Mathematical Biosciences and Engineering, 2009, 6(3): 469-492. doi: 10.3934/mbe.2009.6.469
    [5] Bruno Buonomo . A simple analysis of vaccination strategies for rubella. Mathematical Biosciences and Engineering, 2011, 8(3): 677-687. doi: 10.3934/mbe.2011.8.677
    [6] Andras Balogh, Tamer Oraby . Stochastic games of parental vaccination decision making and bounded rationality. Mathematical Biosciences and Engineering, 2025, 22(2): 355-388. doi: 10.3934/mbe.2025014
    [7] Haitao Song, Dan Tian, Chunhua Shan . Modeling the effect of temperature on dengue virus transmission with periodic delay differential equations. Mathematical Biosciences and Engineering, 2020, 17(4): 4147-4164. doi: 10.3934/mbe.2020230
    [8] Hamed Karami, Pejman Sanaei, Alexandra Smirnova . Balancing mitigation strategies for viral outbreaks. Mathematical Biosciences and Engineering, 2024, 21(12): 7650-7687. doi: 10.3934/mbe.2024337
    [9] Majid Jaberi-Douraki, Seyed M. Moghadas . Optimal control of vaccination dynamics during an influenza epidemic. Mathematical Biosciences and Engineering, 2014, 11(5): 1045-1063. doi: 10.3934/mbe.2014.11.1045
    [10] Rinaldo M. Colombo, Mauro Garavello . Optimizing vaccination strategies in an age structured SIR model. Mathematical Biosciences and Engineering, 2020, 17(2): 1074-1089. doi: 10.3934/mbe.2020057
  • We developed a new mathematical model for yellow fever under three types of intervention strategies: vaccination, hospitalization, and fumigation. Additionally, the side effects of the yellow fever vaccine were also considered in our model. To analyze the best intervention strategies, we constructed our model as an optimal control model. The stability of the equilibrium points and basic reproduction number of the model are presented. Our model indicates that when yellow fever becomes endemic or disappears from the population, it depends on the value of the basic reproduction number, whether it larger or smaller than one. Using the Pontryagin maximum principle, we characterized our optimal control problem. From numerical experiments, we show that the optimal levels of each control must be justified, depending on the strategies chosen to optimally control the spread of yellow fever.



    Yellow fever is a hemorrhagic fever disease that is transmitted by the Aedes or Haemagogus aegypti infected mosquito, which contains the yellow fever virus. The virus that causes this disease belongs to the genus Flavivirus, a large group of RNA viruses. It is one of the most dangerous infectious diseases. The mortality rate of this disease ranges from 20 to 50%, but in severe cases, it can exceed 50%. There is no specific treatment for this disease, but it can be reduced through vaccination [1]. Although the vaccine for yellow fever is lifelong, a booster is needed to guarantee the efficacy of this vaccine to protect humans from infection.

    Various interventions can be performed to suppress the spread of yellow fever. Based on [2], fumigation is the most recommended form of intervention to control mosquito populations. Even so, this intervention needs to be carefully considered in its application in the field because it can trigger mosquito resistance to fumigation materials [3]. In addition to fumigation, vaccination is highly recommended, especially for tourists entering areas where yellow fever is endemic. Although it is claimed that yellow fever vaccination lasts forever, a vaccine booster must maintain the vaccine's ability to protect the individual from infection at maximum levels [4]. Although the yellow fever vaccine is claimed to reach the maximum protection at 4th week after the injection, the implementation of the yellow fever vaccine has some side effects for high-risk groups. These groups include infants aged 6–8 months, individuals more than 60 years old, pregnant, and breastfeeding women [3]. Despite the long history of safety, there are some cases of worsening effect of the use of yellow fever vaccine [1]. The author of [1] expresses more concern about yellow fever vaccine-associated neurotropic disease (YEL-AND) because it is fatal, despite the low number of reported cases.

    Figure 1.  Type of yellow fever based on the transmission process.

    Mathematical model have been used by many authors to model the dynamic of disease, whether it in human [5,6,7,8,9], animals [10], or plants [11,12].Compared with other vector-borne diseases such as malaria or dengue, as far as we know, there are a few of mathematical models discussing yellow fever. The author of [13] introduced a yellow fever transmission model involving two types of hosts, namely, humans and primates. They found that their basic reproduction number consists of the addition of reproduction from human-mosquito interaction and primate (monkey)-mosquito interaction. Additionally, [14] introduced a yellow fever model that considers the vaccination of the human population and aquatic phase of the mosquito. Recently, [15] modified the model by [14] by adding a vertical transmission in the mosquito population to their model. Furthermore, they also involved treated bed nets, larvicides, and adulticides as alternative means of intervention. They showed that their model exhibits backward bifurcation phenomena that make the eradication program for yellow fever more difficult. A mathematical model was used to analyze the impact of vaccination for yellow fever in Angola, as analyzed by [16].

    Based on the above description, we introduce a new mathematical model for yellow fever transmission, considering vaccination, fumigation, and treatment. We also include the possible side effects of the yellow fever vaccine, which may lead to YEL-AND phenomena. Our model includes the aforementioned strategies as time-dependent variables. We aim to minimize the number of infected individuals keeping the cost of interaction as low as possible. Hence, the most cost-effective possible strategies are then analyzed using the cost-effectiveness method [17]. To our knowledge, this work is the first yellow fever model to consider the side effects of the yellow fever vaccine with a time-dependent intervention.

    The remainder of this paper is organized as follows. In Section 2, we develop our model using time-dependent interventions. Treating the interventions as a constant parameter, we analyze our special case model regarding its equilibrium points and local stability criteria in Section 3. The characterization of the optimal control problem and its numerical experiments are presented in Section 4. The conclusions of our study are presented in Section 5.

    In this study, we introduced a new mathematical model to describe the spread of yellow fever. Our model includes vaccination (u1(t)), hospitalization (u2(t)), and fumigation (u3(t)) as time-dependent control variables. We aimed to understand the impact of the worsening effect of vaccination intervention on the dynamics of yellow fever and determine the best possible combination of cost-effective strategies in the yellow fever eradication policy effort.

    First, we divide the total human population (denoted by N(t)) based on their health and vaccination status into eight compartments, namely, susceptible humans (S(t)), vaccinated humans (V(t)), unvaccinated exposed humans (E1(T)), vaccinated exposed humans (E2(t)), infected humans without vaccination worsening effect (I(t)), infected humans with worsening effect of vaccine (T1(t)), infected humans who get worsening effect from the vaccine and were hospitalized (T2(t)), and recovered humans (R(t)). Additionally, we divide the mosquito population into susceptible and infected mosquitoes, denoted by U(t) and W(t), respectively. Owing to the short life expectancy of mosquitoes, we exclude the recovered compartments of mosquitoes. This is because once the mosquitoes are infected by the yellow fever virus, they will stay infected forever. Hence, the total human and mosquito population at time t is expressed as follows:

    N(t)=S(t)+V(t)+E1(t)+E2(t)+I(t)+T1(t)+T2(t)+R(t),M(t)=U(t)+W(t).

    We use the transmission diagram in Figure 2 to construct our yellow fever model, and the description of the parameters is given in Table 1. For written simplification, we use S,V,E1,E2,I,T1,T2,R,U, and W instead of S(t),V(t),E1(t),E2(t),I(t),T1(t),T2(t),R(t),U(t), and W(t).

    Figure 2.  Transmission diagram of yellow fever for model in system (2.1).
    Table 1.  Description of parameters in system (2.1).
    Par Description Range Ref.
    Λh Human recruitment rate 1000065×365humanday Assumed
    Λv Mosquitoes recruitment rate 2000030mosquitoesday Assumed
    u1 Vaccination rate [0,1]1day [19,20,21]
    u2 Hospitalization rate [0,1]1day Assumed
    u3 Fumigation rate [0,1]1day [22,23]
    ωh Waning rate of vaccine (0,1)1day [19]
    βh Mosquito-human infection rate [0.03,0.75]1041mosquito×day [16,24]
    1ξ Vaccine efficacy [0.8-0.99] [15,20,25]
    βv Human-mosquito infection rate [0.15,1]1051human×day [16,24]
    μh Natural death rate of human 165×3651day [26]
    μv Natural death rate of mosquito 1301day [26]
    γ1 Progression rate of exposed humans who are not vaccinated 0.3311day [27]
    γ2 Progression rate of exposed human who are already vaccinated 0.311day [27]
    p Proportion of exposed individuals who progressed into infectious but experienced side effects of the vaccine [0, 1] Assumed
    δ1/δ2/δ3 Recovery rate of I/T1/T2 [0,0.143]1day [27]
    α1/α2/α3 Death rate because of yellow fever in I/T1/T2 [0.2,0.5]1031day Assumed

     | Show Table
    DownLoad: CSV

    In our model, we assume that all recruitment rates enter the susceptible population through a constant rate Λh. Owing to a vaccination rate of u1, the susceptible population will be transferred into compartment V. We assume that after a period of ω1h, the effect of the vaccine disappears. Hence, the vaccinated individual was sent back to S. In our model, we assume that vaccination cannot provide perfect protection against yellow fever infection. Hence, there is an efficacy level of the vaccine, which is denoted by 1ξ. Thus, when a susceptible individual is infected with a success rate of βh, vaccinated individuals will have a success probability of ξβh.

    Susceptible individuals who are already infected will be classified into the exposed compartment E1, which has no chance of experiencing any worsening vaccine effect. However, vaccinated individuals who are already infected and experienced any worsening vaccine effect will be classified into the exposed compartment E2. In our model, we assume that a yellow fever vaccine can have side effects on several types of individuals, including severely immunodeficient individuals, elderly individuals, pregnant women, newborns under six months of age, and individuals who already have thymic diseases, including thymoma and thymectomy [18]. These side effects can result in severe symptoms of infection. Because of this worsening effect, after the incubation period of γ12, an exposed individual E2 will progress into infected stage I with a proportion of 1p, where he/she has no worsening effect of the vaccine, and a proportion of p will progress to T1 when he/she has the worsening effect of the vaccine. In contrast, the exposed individual E1 progresses to I after the incubation period of γ11. Note that γ11γ12.

    Infected individuals in T1 are assumed to have more severe symptoms than I individuals. Hence, individuals in T1 must receive more intensive care in hospital. Therefore, we assume that there is a hospitalization rate of u2 that will bring an individual in T1 to a hospitalized individual in T2. The infected individuals in I are assumed to recover from yellow fever after δ11, whereas T1 and T2 are assumed to recover after δ11 and δ12, respectively. We assume that yellow fever may cause death in infected individuals I,T1, and T2 with a constant rate α1,α2, and α3, respectively.

    However, the mosquito population will increase because of a constant natural birth rate of Λv and decrease because of the natural death rate μv and fumigation u3. All newborns in the mosquito population were assumed to be susceptible. Additionally, we assume that susceptible mosquitoes may get infected by biting infected individuals who do not go to the hospital, I and T1, with a constant infection rate βv.

    Our aim is to understand how the spread of yellow fever may be reduced under three types of intervention: vaccination (u1), hospitalization (u2), and fumigation (u3). In so many realities of disease eradication in the field, the limited costs of implementing interventions are always an unavoidable limitation. Therefore, instead of treating the above mentioned interventions as constant parameters, it would be better to make the intervention parameter both time-dependent and adaptive to the number of infected people and environmental conditions at time t. One approach is to find the optimal value of u1 for each time t. This can be achieved by constructing a model to be discussed using an optimal control approach. Therefore, we discuss the time-dependent value of ui, namely, u1(t),u2(t), and u3(t).

    Based on the above descriptions and formulations, the mathematical model of yellow fever under vaccination, hospitalization, and fumigation, as well as the worsening effect of vaccination, is given by the following system of ordinary differential equations:

    dSdt=Λh+ωhVβhSWu1(t)SμhS,dVdt=u1(t)SξβhVWμhVωhV,dE1dt=βhSWγ1E1μhE1,dE2dt=ξβhVW(1p)γ2E2pγ2E2μhE2,dIdt=γ1E1+(1p)γ2E2(μh+α1)Iδ1I,dT1dt=pγ2E2δ2T1u2(t)T1(μh+α2)T1,dT2dt=u2(t)T1δ3T2(μh+α3)T2,dRdt=δ1I+δ2T1+δ3T2μhR,dUdt=ΛvβvU(I+T1)(μv+u3(t))U,dWdt=βvU(I+T1)(μv+u3(t))W. (2.1)

    Note that the yellow fever model in system (2.1) is completed with non-negative initial conditions. Additionally, the parameters in system (2.1) are assumed to be non-negative.

    This study aimed to minimize the number of infected individuals as much as possible while keeping the costs of vaccination, hospitalization, and fumigation as low as possible. Because we only want to minimize the number of infected individuals, we set the weight cost for S,V,R, and U, denoted by ω1,ω2,ω8, and ω9, respectively, to zero. Hence, we describe our aim in the following cost function:

    J(ui(t))=T0(ω3E1+ω4E2+ω5I+ω6T1+ω7T2+ω10W+φ1u21+φ2u22+φ3u23)dt, (2.2)

    where φi for i=1,2,3 denotes the weight cost for the control variables u1,u2,, and u3, respectively. The interval [0,T] represents the duration of the simulation time for the intervention that must be implemented. The total cost for interventions, such as vaccination, hospitalization, and fumigation, is presented as follows:

    T0(φ1u21+φ2u22+φ3u23)dt.

    The quadratic cost function was chosen because of a convex function that ensures the existence of an optimal solution. On the other hand, the quadratic cost function is biologically interpreted as a nonlinear cost for intervention. For example, the cost of fumigation in a small area may only require a cost for buying the fumigation material and hiring an operator. However, when the area that needs to be fumigated becomes much wider, the cost includes not only the purchase of fumigation material and the hiring of operator, but also coordination costs, counseling before implementation, etc. Another quadratic cost function for another type of epidemiological optimal control model can be found in [28,29]. On the other hand, all costs associated with the high number of infected humans and mosquitoes, which are not related to vaccination, hospitalization, or fumigation intervention were assumed to be proportional to the number of infected individuals, which is presented as follows:

    T0(ω3E1+ω4E2+ω5I+ω6T1+ω7T2+ω10W)dt.

    Our model's aim is to minimize the number of infected individuals in human and mosquito populations while keeping the costs of vaccination, hospitalization, and fumigation as low as possible. Therefore, we seek to optimize the value of ^ui for i=1,2,3 such that

    J(ˆu1,ˆu2,ˆu3)=minu1,u2,u3{J(u1,u2,u3)|uiU}, (2.3)

    where

    U={(u1,u2,u3)|ui:[0,T][umini,umaxi] is Lebesgue measurable,i=1,2,3,4}

    is the admissible control set. Note that umini and umaxi for i=1,2,3 represent the lower and upper bounds of each control variable, respectively.

    In this section, we analyze the autonomous version of system (2.1) regarding its existence and local stability criteria of equilibrium points, the form of the basic reproduction number, and bifurcation analysis. To accomplish this, let us assume that all control variables remain constant over time. Hence, using u1(t)=u1,u2(t)=u2, and u3(t)=u3, the autonomous version of system (2.1) is given by:

    dSdt=Λh+ωhVβhSWu1SμhS,dVdt=u1SξβhVWμhVωhV,dE1dt=βhSWγ1E1μhE1,dE2dt=ξβhVW(1p)γ2E2pγ2E2μhE2,dIdt=γ1E1+(1p)γ2E2(μh+α1)Iδ1I,dT1dt=pγ2E2δ2T1u2T1(μh+α2)T1,dT2dt=u2T1δ3T2(μh+α3)T2,dRdt=δ1I+δ2T1+δ3T2μhR,dUdt=ΛvβvU(I+T1)(μv+u3)U,dWdt=βvU(I+T1)(μv+u3)W. (3.1)

    Hence, the dynamics of the total human population is given by

    dNdt=ΛhIα1T1α2T2α3μhN,

    and the dynamics of the total mosquito population is given by

    dMdt=Λv(μv+u3)M.

    Before we begin analyzing the model in system (3.1) regarding the existence and stability of its equilibrium points, it is essential to ensure that the model in system (3.1) is well-defined biologically and mathematically. That is, the solution in system (3.1) must always be non-negative for all t>0. Properties related to these are written in the theorems below, where Theorem 1 deals with the positivity of the solution of each variable in system (3.1), and Theorem 2 discusses the upper bound of the total human and mosquito populations.

    Theorem 1. Given that the initial values of the system (2.1) are S(0), V(0), E1(0), E2(0), I(0), T1(0), T2(0), R(0), U(0), and W(0)0, the solution of S(t), V(t), E1(t), E2(t), I(t), T1(t), T2(t), R(t), U(t), and W(t) will always be non-negative for all time t0.

    Proof. Please see Appendix A for the proof.

    Theorem 2. The total human (N) and mosquito (M) population in system (2.1) is eventually bounded as t.

    Proof. The proof can be found in Appendix B.

    Based on Theorems 1 and 2, we can conclude that each variable in system (3.1) is also bounded.

    The purpose of this section is to determine the possible final state of the dynamics of system (3.1), which at times tends to infinity and is known as the equilibrium point. The equilibrium of system (3.1) is obtained by setting the right-hand side of system (3.1) to zero and solving this non-linear set of equations with respect to each variable on human and mosquito populations. The first equilibrium point of system (3.1) is the trivial point, known as the yellow fever-free equilibrium point (E1). This equilibrium point represents a situation in which all infected populations in humans and mosquitoes, as well as a recovered compartment in humans, do not exist. In this study, E1 is given as follows:

    E1=(S,V,E1,E2,I,T1,T2,R,U,W),=(Λh(μh+ωh)μh(μh+ωh+u1),Λhu1μh(μh+ωh+u1),0,0,0,0,0,0,Λvμv+u3,0). (3.2)

    From the expressions of S and V, S+V=Λhμh, which means that the existence of vaccination does not change the size of the total population at E1. Instead, it depends only on the number of newborns and the natural death rate. On the other hand, note that SV=μh+ωhu1. Our aim was to avoid the rapid spread of yellow fever by expecting many vaccinated individuals. In this study, this aim can be read as minimizing this ratio. To reduce the ratio of S to V, we can either increase the number of vaccinated individuals (increasing the value of u1) or improve vaccine quality by extending vaccine duration (reducing the value of ωh). Additionally, note that the total number of mosquitoes at E1 is inversely proportional to the fumigation rate u3. Hence, it is important to provide a large (but controlled) rate of fumigation to reduce the number of aedes mosquitoes. However, in many reports, uncontrolled fumigation may trigger a mosquito resistance to some fumigants [30].

    Having the yellow fever-free equilibrium point in hand, as shown in Eq (3.2), we can calculate the basic reproduction number (R0) of system (3.1). In the context of our problem, the basic reproduction number presents an expected number of secondary yellow fever cases caused by one primary case during one infection period in a completely susceptible population. Several studies have demonstrated that basic reproduction numbers play an important role in determining the existence or extinction of diseases in many epidemiological models [28,29,31]. These studies demonstrated that the disease will likely disappear if R0<1, and it must appear if R0>1. However, it is not always the case that R0<1 will always end in a situation where the disease disappears in the field. One of the causes is the occurrence of a backward bifurcation phenomenon in the model. When backward bifurcation occurs, R0 is no longer becomes a sufficient condition for the disappearance of a disease because the condition R0<1 causes a bistability phenomenon to occur. More examples of backward bifurcation phenomena arising from epidemiological models are available in [32].

    Using the next-generation matrix approach [33], we have the basic reproduction number of the system (3.1) as follows:

    R0=μh(α2+δ2+μh+u2)(γ1+μh)M1μh(μh+ωh+u1)(γ2+μh)(μv+u3)(γ1+μh)(α2+δ2+μh+u2)(α1+δ1+μh), (3.3)

    with

    M1=(γ1μ3h+((ξu1+γ1)γ2+γ1(δ2+ωh+u2+α2))μ2h+(((ξu1+α2+δ2+ωh+u2)γ1+u1(pα1+pδ1+(α2+δ2+u2)(1p))ξ)γ2+γ1ωh(α2+δ2+u2))μh+γ1γ2(u1(pα1+pδ1+(α2+δ2+u2)(1p))ξ+ωh(α2+δ2+u2)))Λvβv(μh+ωh+u1)(γ2+μh)βh(α1+δ1+μh)Λh.

    The derivation of R0 is available in Appendix C. Using Theorem 2 in [34], the local stability criterion of E1 is given by the following theorem:

    Theorem 3. The yellow fev-free equilibrium point of system (3.1) is locally asymptotically stable if R0<1, and unstable if R0>1.

    Proof. Please see Appendix D for the proof.

    In this section, we analyze the existence of the non-trivial equilibrium point of system (3.1). The other equilibrium point of system (3.1) besides the yellow fever-free equilibrium point is the yellow fever endemic equilibrium point, which is given by the following expression:

    E2=(S,V,E1,E2,I,T1,T2,R,U,W),

    where

    S=Λh(Wξβh+μh+ωh)W2ξβh2+((ξ+1)μh+ξu1+ωh)Wβh+μh(μh+ωh+u1),V=u1ΛhW2ξβh2+((ξ+1)μh+ξu1+ωh)Wβh+μh(μh+ωh+u1),E1=WΛhβh(Wξβh+μh+ωh)(γ1+μh)(μh2+(W(ξ+1)βh+ωh+u1))μh+Wβh(Wξβh+ξu1+ωh),E2=u1βhΛhξW(γ2+μh)(W2ξβh2+((ξ+1)μh+ξu1+ωh))Wβh+μh(μh+ωh+u1),I=(γ1μh2+((u1(1p)ξ+γ1)γ2+γ1(Wξβh+ωh)))μh+γ1(Wξβh+u1(1p)ξ+ωh)γ2)βhWΛh(γ1+μh)(α1+δ1+μh)(γ2+μh)(μh2+(W(ξ+1)βh+ωh+u1))μh+Wβh(Wξβh+ξu1+ωh),T1=pγ2u1βhΛhξW(μh2+(W(ξ+1)βh+ωh+u1)μh+Wβh)((Wξβh+ξu1+ωh))(α2+δ2+μh+u2)(γ2+μh),T2=WpξΛhβhγ2u1u2(α3+δ3+μh)(μh2+(W(ξ+1)βh+ωh+u1))μh+Wβh(Wξβh+ξu1+ωh),R=δ1I+δ2T1+δ3T2μh,U=ΛvIβv+T1βv+μv+u3, (3.4)

    where W is taken from the positive roots of the following quadratic polynomial of W:

    F(W)=a2W2+a1W+a0=0, (3.5)

    with

    a2=ξβh2(μv+u3)(γ2+μh)(α2+δ2+μh+u2)(Λhβvγ1+α1γ1μv+α1γ1u3+α1μhμv+α1μhu3+δ1γ1μv+δ1γ1u3+δ1μhμv+δ1μhu3+γ1μhμv+γ1μhu3+μh2μv+μh2u3),a1=βhβvΛh((((μv+u3)μh2+(((μv+u3)u1Λvβh)ξ+(μv+u3)(α2+δ2+ωh+u2))μh+((μv+u3)((1p)(α2+δ2+u2)+p(α1+δ1))u1Λvβh(α2+δ2+u2))ξ+ωh(μv+u3)(α2+δ2+u2))γ2+μh(α2+δ2+μh+u2)((μv+u3)μhξβhΛv+ωh(μv+u3)))γ1+μh(μv+u3)u1γ2ξ(μh+(1p)(α2+δ2+u2)+p(α1+δ1))),a0=μh(μv+u3)2(γ2+μh)(γ1+μh)(α2+δ2+μh+u2)(u1+μh+ωh)(α1+δ1+μh)(1R20).

    Based on the expression of F(W), we have the following theorem regarding the existence criteria of E2 in R6+:

    Theorem 4. The yellow fever endemic equilibrium (E2) of system (3.1) always exists if R0>1, and no yellow fever endemic equilibrium otherwise.

    Proof. Please see Appendix E for the proof.

    In this section, we analyze the type of bifurcation of the yellow fever model in (3.1). In many classical models, forward bifurcation in epidemiological models indicates that R0 is a necessary condition to ensure disease extinction ([5,31,35,36,37]). However, in several epidemiological models [28,38,39,40], another type of bifurcation, namely, backward bifurcation, may appear. Under these circumstances, the condition of the basic reproduction number being less than unity is no longer became sufficient to guarantee the extinction of the disease. This is because when backward bifurcation phenomena arise, bistability phenomena may appear where stable endemic equilibrium exists together with the stable disease-free equilibrium.

    To analyze the type of bifurcation of our model in (3.1), we use the result of Theorem 11 in [41], which is commonly known as the Castillo-Song bifurcation theorem. To apply this theorem, let us re-symbolize our variable from S,V,E1,E2,I,T1,T2, and R into xi for i=1,210, respectively. Hence, we now have the system (3.1) that reads as follows:

    g1:=Λhβhx1x10μhx1+ωhx2u1x1,g2:=u1x1ξβhx2x10μhx2ωhx2,g3:=βhx1x10γ1x3μhx3,g4:=ξβhx2x10(1p)γ2x4pγ2x4μhx4,g5:=γ1x3+(1p)γ2x4(μh+α1)x5δ1x5,g6:=pγ2x4δ2x6u2x6(μh+α2)x6,g7:=u2x6δ3x7(μh+α3)x7,g8:=δ1x5+δ2x6+δ3x7μhx8,g9:=Λvβvx9(x5+x6)(μv+u3)x9,g10:=βvx9(x5+x6)(μv+u3)x10. (3.6)

    Next, we choose βh as the bifurcation parameter. Therefore, solving R20=1 with respect to βh (an argument similar to the proof of Theorem 4), we obtain the following:

    βh=(α2+δ2+μh+u2)(α1+δ1+μh)(γ1+μh)(μh+ωh+u1)μh(γ2+μh)(μv+u3)2Z0γ1(μh+ωh)(γ2+μh)(α2+δ2+μh+u2)βvΛvΛh, (3.7)

    where Z0=u1(γ1+μh)(μh+(δ1δ2u2+α1α2)p+α2+δ2+u2)γ2βvΛvΛhξ. Linearization of system (3.6) at E2 with βh=βh gives

    J=[μhu1ωh0000000a110u1μhωh0000000a21000γ1μh000000a310000γ2μh00000a41000γ1(p1)γ2a5500000000pγ20a66000000000u2a770000000δ1δ2δ3μh000000βvΛvμv+u3βvΛvμv+u300a9900000βvΛvμv+u3βvΛvμv+u3000u3+μv],

    where

    a110=(μh+ωh)(α2+δ2+μh+u2)(α1+δ1+μh)(γ1+μh)(γ2+μh)(μv+u3)2Aγ2γ1μh(μh+ωh)(α2+δ2+μh+u2))βvΛv,a210=ξ(α2+δ2+μh+u2)(α1+δ1+μh)(γ1+μh)(γ2+μh)(μv+u3)2u1Aγ2γ1μh(μh+ωh)(α2+δ2+μh+u2))βvΛv,a310=(μh+ωh)(α2+δ2+μh+u2)(α1+δ1+μh)(γ1+μh)(γ2+μh)(μv+u3)2Aγ2γ1μh(μh+ωh)(α2+δ2+μh+u2))βvΛv,a410=ξ(α2+δ2+μh+u2)(α1+δ1+μh)(γ1+μh)(γ2+μh)(μv+u3)2u1Aγ2γ1μh(μh+ωh)(α2+δ2+μh+u2))βvΛv,a55=μhα1δ1,a66=δ2u2μhα2,a77=δ3μhα3,a99=u3μv,

    and A=((μh2+(ξu1α2δ2ωhu2)μh+((δ2+u2α1+α2δ1)pα2δ2u2)u1ξωh(α2+δ2+u2))γ1+(μh+(δ2+u2α1+α2δ1)pδ2u2α2)u1μhξ). From the direct calculation, it can be shown that J has a simple zero eigenvalue, while the other nine eigenvalues are negative. Hence, we can use the center manifold theory to analyze the bifurcation of our yellow fever model.

    Following the application of the Castillo-Song bifurcation theorem, we must calculate the right and left eigenvectors related to the zero eigenvalue. The right eigenvector of J at eigenvalue zero is given as follows:

    w=(w1,w2,w3,w4,w5,w6,w7,w8,w9,w10)T

    where

    w1=(δ3+μh+α3)(α2+δ2+μh+u2)(γ2+μh)(ξωhu1+μh2+2ωhμh+ωh2)μh(μh+ωh+u1)u1,w2=(δ3+μh+α3)(α2+δ2+μh+u2)(γ2+μh)(ξμh+ξu1+μh+ωh)(μh+ωh+u1)μh,w3=(α2+δ2+μh+u2)(μh+ωh)(γ2+μh)(δ3+μh+α3)u1(γ1+μh),w4=(δ3+μh+α3)(α2+δ2+μh+u2)ξ,w5=(α2+δ2+μh+u2)(δ3+μh+α3)(γ1μh2+((γ2+ωh)A1))(γ1+μh)(α1+δ1+μh)u1,w6=(δ3+μh+α3)ξpγ2,w7=ξu2pγ2,w8=(α2+δ2+μh+u2)γ2(δ1(δ3+μh+α3)(α2+δ2+μh+u2)+A2)μh(α1+δ1+μh),w9=Λvβv(μh2+(ξu1+α2+δ2+ωh+u2)μh+u1A3)u1(μv+u3)2(γ1+μh)(α1+δ1+μh),w10=Λvβv(μh2+(ξu1+α2+δ2+ωh+u2)μh+u1A3)u1(μv+u3)2(γ1+μh)(α1+δ1+μh),

    with

    A1=γ1+ξu1γ2(p+1)μh+γ1(ξ(p+1)u1+ωh)γ2,A2=p((δ3+μh+α3)(μh+α1)δ2+(α3α2α2δ3μhα2μhα3α3u2μhδ3μh2μhu2)δ1+u2δ3(μh+α1)),A3=((1p)(u2+α2+δ2)+p(α1+δ1))ξ+ωh(α2+δ2+u2)γ1+(((δ1δ2u2+α1α2)p+μh+α2+δ2+u2)μhu1ξ)γ2+γ1μh(μh+ωh)(α2+δ2+μh+u2)(δ3+μh+α3).

    However, the left eigenvector of J for eigenvalue zero is given as follows:

    v=(v1,v2,v3,v4,v5,v6,v7,v8,v9,v10)

    where

    v1= v2=v7=v8=v9=0,v3= γ1,v4=(γ1+μh)(μh+(1p)(u2+α2+δ2)+p(α1+δ1))γ2(γ2+μh)(α2+δ2+μh+u2),v5=γ1+μh,v6=α1γ1+α1μh+δ1γ1+δ1μh+γ1μh+μh2α2+δ2+μh+u2,v10=(γ1+μh)(α1+δ1+μh)(u3+μv)Λvβv.

    Using the formula as follows:

    A=n=10k,i,j=1vkwiwj2gkxixj(0,0),   B=n=10k,i,j=1vkwi2gkxiβ(0,0). (3.8)

    we have the following:

    A=2(α3+δ3+μh)2(γ1μh3+((ξu1+γ1)γ2+γ1(δ2+ωh+u2+α2))μh2+M0)u12(μv+u3)2(γ1+μh)(α1+δ1+μh)μh(μh+ωh+u1),B=(ξγ2u1(γ1+μh)(δ1δ2u2+α1α2)p+(α2+δ2+μh+u2)2M1)u1(μv+u3)2(γ1+μh)(α1+δ1+μh)μh(μh+ωh+u1)(γ2+μh)(α2+δ2+μh+u2),

    where M1=(ξγ1γ2u1+ξγ2μhu1+γ1γ2μh+γ1γ2ωh+γ1μh2+γ1μhωh)(α3+δ3+μh)ΛvΛhβv while M0>0 cannot be shown in this study because of its long expression. Because we always have A<0 and B>0, we have the following results.

    Theorem 5. The yellow fever model in system (3.1) always exhibits a forward bifurcation at R0=1.

    Based on Theorem 5, we understand that E2 is locally asymptotically stable when R0>1 but is close to one. Additionally, based on our proposed model, we can conclude that a basic reproduction number less than unity will always ensure the extinction of yellow fever from the population.

    In this section, we perform some numerical experiments based on our previous results on Theorems 3, 4, and 5. First, we analyze the elasticity of each parameter in R0 to determine the most influential parameter for changing the size of R0. Based on this result, we continue our simulation by showing the level set of R0 with respect to each control variable to understand the sensitivity of each parameter. The forward bifurcation diagram shows the impact of R0 on the endemic size of the infected mosquito population. Finally, numerical experiments were conducted to demonstrate the dynamic behavior of the system (3.1) under certain scenarios.

    To conduct the elasticity analysis of R0 with respect to any parameter ρ, we use the following equation [42]:

    ϵρR0=R0ρ×ρR0. (3.9)

    To perform these experiments, we use the following parameter values:

    Λh=1000065×365,Λv=2000030,βh=105,βv=105,ξ=0.05,ωh=0.1,μh=165×365,μv=130,γ1=0.31,γ2=0.31,p=0.1,δ1=0.143,δ2=0.0715,δ3=0.10725,α1=0.00035,α2=0.0004375,α3=0.0002625,u1=0.01,u2=0.01,u3=0.01,

    which gives R0=1.504. Using the Eq (3.9), the elasticity of R0 with respect to each parameter in the system (3.1) with the value of the above parameter is shown in Table 2. Note that because α3 and δ3 do not appear in the expression of R0, ϵα3R0 and ϵδ3R0 are both zero. The positive sign of ϵρR0 means that R0 increases when ρ increases. However, when ϵρR0 is negative, R0 decreases as ρ increases. Hence, all control variables can reduce the value of R0 whenever these control parameters increase. The value of ϵρR0 represents the percentage of increase or decrease of R0 for a 1% increase or decrease in ρ. For example, because ϵu3R0=0.2308, a 1% increase in u3 will reduce R0 by 0.2308%. Additionally, note that |ϵu3R0|=max{|ϵu1R0|,|ϵu2R0|,|ϵu3R0|}. Hence, we conclude that fumigation is more essential in reducing R0 compared to vaccination and hospitalization. These results are in accordance with the recommendations of the WHO regarding fumigation as one of the most recommended efforts to overcome yellow fever [3].

    Table 2.  Elasticity of R0 with respect to each parameter in system (3.1).
    ϵΛhR0 ϵΛvR0 ϵβhR0 ϵβvR0 ϵγ1R0 ϵγ2R0
    0.5 0.5 0.5 0.5 6.7×105 3.64×107
    ϵα1R0 ϵα2R0 ϵδ1R0 ϵδ2R0 ϵωhR0 ϵξR0
    -0.0012 2.32×106 -0.49 -0.00038 0.043 0.0027
    ϵμhR0 ϵμvR0 ϵpR0 ϵu1R0 ϵu2R0 ϵu3R0
    -0.5002 -0.77 0.000186 -0.043 -0.000053 -0.2308

     | Show Table
    DownLoad: CSV

    In addition to understanding the impact of all parameters in Eq (2.1) to R0, we must understand the impact of these parameters on the dynamics of all variables in Eq (2.1). Hence, we conducted a sensitivity analysis using three methods: non-normalization, half normalization, and full normalization using the SimBiology toolbox for MATLAB [43,44].

    Suppose that the yellow fever model in Eq (2.1) has 10 compartments, xi in i=1,2,...,10 represents S,V,E1,E2,I,T1,T2,R,U, and W, respectively. Furthermore, let kj for j=1,2,...,20 represent all the parameters in system (2.1). We assume that the model-balanced equations are expressed as a system of differential equations as follows:

    dxidt=fi(x,k), (3.10)

    where xRn, kRm, and fi(x,k) is the right-hand side of the system (2.1). The non-normalization sensitivities are given as follows:

    Sxikj=xikj, (3.11)

    where Sxikj is the sensitivity coefficient of each xi with respect to each parameter kj. The half-normalization sensitivities are also given by the following equation:

    Sxikj=(1xi)(xikj). (3.12)

    Furthermore, the full-normalization sensitivities are defined as follows:

    Sxikj=(kjxi)(xikj). (3.13)

    The results of the sensitivity analysis without normalization are shown in Figure 3. The results shown in Figure 3 show that the infection parameters βh and βv are very dominant in influencing changes in the populations of S and E1 in humans, and the populations of U and W on mosquitoes. This agrees with the results of the analysis in Table 2 regarding the effect of βh and βv, which are also dominant on R0. Furthermore, it was also observed that the fumigation intervention (u3) was more dominant in suppressing the number of infected mosquitoes, whereas vaccination (u2) was dominant in determining the dynamics of healthy and vaccinated human populations. The simulation results in Figure 3 also confirm that hospitalization is the least effective among all control parameters.

    Figure 3.  Local sensitivity analysis with non–normalization technique of all variables with respect to (a) all parameters in computational simulations and (b) all parameters except βh and βv.

    We perform a half-normalization of our model in Figure 4. Note that βh and βv are still dominant in determining the dynamics of the (3.1) model compared to other parameters. Interestingly, fumigation is not only dominant in regulating population changes in mosquitoes but also in infected human populations. Clearly, the presence of fumigation has a significant impact on the population of E1. This is consistent with the fact that the expected effect of fumigation is a reduction in the number of new infections in the human population. Figure 5 shows the sensitivity analysis using the full normalization approach. The results indicate the same result as using the non-and half-normalization approaches, where fumigation is more dominant compared to other control parameters.

    Figure 4.  Local sensitivity analysis with half–normalization technique of all variables with respect to (a) all parameters in computational simulations and (b) all parameters except βh and βv.
    Figure 5.  Local sensitivity analysis with full–normalization technique of all variables with respect to all parameters in computational simulations.

    Next, we show the bifurcation diagram of system (3.1) using the same parameter values as in previous numerical experiments. We use our results in Theorem 5, which indicates that our proposed model in system (3.1) always exhibits a forward bifurcation at R0=1. The bifurcation diagram is shown in Figure 6. Note that the yellow fever-free equilibrium is always locally asymptotically stable when R0<1. At R0=1, the yellow fever-free equilibrium becomes unstable, while the yellow fever endemic equilibrium starts to arise and is locally asymptotically stable. Note that a larger value of R0 will increase the size of infected mosquitoes in an endemic equilibrium.

    Figure 6.  Bifurcation diagram of system (3.1) with respect to R0 and W.

    To illustrate the dynamics of infected compartments with respect to the values of R0, we apply the Runge-Kutta method to solve system (3.1). In the first simulation we set the values of all parameter similar to the previous simulation, except u1=0.03,u2=0.01,u3=0.03. This setting gives us R0<1. In Figure 7, we can see that all trajectories tend to the yellow fever free equilibrium point asymptotically, which in this case is given by

    (S,V,E1,E2,I,T1,T2,R,U,W)=(7694,2306,0,0,0,0,0,0,10527,0).
    Figure 7.  Dynamics of total infected human (left) and mosquito (right) for short (top) and long (bottom) time period which tend to the yellow fever free equilibrium for various initial conditions.

    To simulate R0>1, we choose the same parameter value as before, except βh=2×105,βv=2×105,u1=0,u2=0,u3=0. In Figure 8, we can see that all solutions tend to the endemic equilibrium point, which is given by

    (S,V,E1,E2,I,T1,T2,R,U,W)=(599,0,1.3,0,3,0,0,9375,19967,33).
    Figure 8.  Dynamics of total infected human (left) and mosquito (right) for short (top) and long (bottom) time period which tend to the yellow fever endemic equilibrium for various initial conditions.

    Optimal control has been used in several epidemiological models to analyze how the best form of intervention should be chosen to reduce the rate of spread of the disease, where the intervention is treated as a time-dependent variable [45,46,47,48]. In these studies, the optimal condition for the control variables concludes that the optimal dynamics of the controls depend on the population involved in the model. These various studies concluded that time-dependent interventions provide more optimal results in suppressing the spread of disease but at a more optimal cost.

    Before we characterize our optimal control problem, we recall all the control variables introduced in Section 2 as follows:

    u1(t): Vaccination rate

    u2(t): Hospitalization rate

    u3(t): Fumigation rate.

    The objective function is expressed using Eq (2.2). Our aim is to minimize the number of infected individuals with optimal interventions, which are read as follows:

    J(ˆu1,ˆu2,ˆu3)=minu1,u2,u3{J(u1,u2,u3)|uiU}, (4.1)

    where U is the admissible control set.

    We use the Pontryagin maximum principle (PMP) [49] to determine the necessary condition for the existence of a triplet of our optimal controls. We define the Hamiltonian of the problem as follows:

    H=ω3E1+ω4E2+ω5I+ω6T1+ω7T2+ω10W+φ1u21+φ2u22+φ3u23+λ1(Λh+ωhVβhSWu1SμhS)+λ2(u1SξβhVWωhV)+λ3(βhSWγ1E1μhE1)+λ4(ξβhVWξVμhV)+λ5(γ1E1+(1p)γ2E2δ1I(μh+α1)I)+λ6(pγ2E2δ2T1u2T1(μh+α2)T1)+λ7(u2T1δ3T2(μh+α3)T2)+λ8(δ1I+δ2T1+δ3T2μhR)+λ9(ΛvβvU(I+T1)(μv+u3)U)+λ10(βvU(I+T1)(μv+u3)W), (4.2)

    where λi for i=1,2,10 are the adjoint variables for the state variables S,V,E1,E2,I,T1,T2,R,U, and W, respectively. Hence, the necessary condition of our optimal control problem is given by the following theorem:

    Theorem 6. For the optimal control variables ˆu1,ˆu2,, and ˆu3, and the optimal solutions ˆS,ˆV,^E1,^E2,ˆI,^T1, ^T2,ˆR,ˆU, and ˆW of system (2.1), the adjoint variables λi for i=1,2,10 that satisfy the following:

    ˙λ1=HS(t)=λ1(βhW+u1+μh)λ2u1λ3βhIv,˙λ2=HV(t)=λ1ωh+λ2(ξβhW+ωh+μh)λ4ξβhW,˙λ3=HE1(t)=ω3+λ3(γ1+μh)λ5γ1,˙λ4=HE2(t)=ω4+λ4(γ2+μh)λ5(1p)γ2λ6pγ2,˙λ5=HI(t)=ω5+λ5(δ1+μh+α1)λ8δ1+λ9βvUλ10βvU,˙λ6=HT1(t)=ω6+λ6(δ2+u2+μh+α2)λ7u2λ8δ2+λ9βvUλ10βvU,˙λ7=HT2(t)=ω7+λ7(δ3+μh+α3)λ8δ3,˙λ8=HR(t)=λ8μh,˙λ9=HU(t)=λ9(βvI+βvT1+μv+u3)λ10(βvI+βvT1),˙λ10=HW(t)=λ10+λ1βhS+λ2ξβhVλ3βhSλ4ξβhV+λ10(μv+u3), (4.3)

    with transversality conditions λi(T)=0 for i=1,2,10. Additionally, the controls ^ui for i=1,2,3 are given by the following:

    ˆu1(t)={0;S(λ1λ2)2φ10,1;S(λ1λ2)2φ11,S(λ1λ2)2φ1;others.ˆu2(t)={0;T1(λ6λ7)2φ20,1;T1(λ6λ7)2φ21,T1(λ6λ7)2φ2;others.ˆu3(t)={0;U(λ9+λ10)2φ30,1;U(λ9+λ10)2φ31,U(λ9+λ10)2φ3;others. (4.4)

    Proof. The adjoint system in Eq (4.3) is taken from the following formula

    ˙λ1=HS,˙λ2=HV,˙λ3=HE1,˙λ4=HE2,˙λ5=HI,˙λ6=HT1,˙λ7=HT2,˙λ8=HR,˙λ9=HU,˙λ10=HW, (4.5)

    with the transversality condition λi(T)=0 for i=1,210. To obtain the control characterization, we first take the derivative of H with respect to each control variable. Hence, we obtain the following:

    Hu1(t)=2φ1u1λ1S+λ2S,Hu2(t)=2φ2u2λ6T1+λ7T1,Hu3(t)=2φ2u3λ9Uλ10I.

    Solve Hui=0 for each control variables, we have the following:

    u1(t)=S(t)(λ1λ2)2φ1,u2(t)=T1(t)(λ6λ7)2φ2,u3(t)=λ9U(t)+λ10U(t)2φ3.

    Hence, based on the standard argument that ^ui should be bounded between umini and umaxi (in this study, we choose umini=0,umaxi=1 for each control variable), we have Eq (4.4).

    In this section, we perform numerical simulations for the optimal control problem. To summarize, we aim to minimize the cost function in Eq (2.2) subject to the yellow fever model in Eq (2.1) as the state system, and the adjoint system in Eq (4.3) with the terminal condition given, and the optimality condition in Eq (4.4). We performed numerical simulations using the forward-backward sweep method [50]. To perform these simulation, we chose the following initial conditions:

    S(0)=8500,V(0)=0,E1(0)=800,E2(0)=0,I(0)=500,T1(0)=0,T2(0)=0,R(0)=200,U(0)=17000,V(0)=3000. (4.6)

    Additionally, to balance each component in the cost function Eq (2.2), we choose ω3=ω4=ω5=ω6=ω7=1, ω10=0.1,φ1=2500,φ2=5000, and φ3=1500. We use the same parameter values as in the previous section, which gives R0=1.504. Hence, without further improvement in the intervention, yellow fever will be endemic to the population. The autonomous simulation of system (3.1) without any intervention (u1(t)=u2(t)=u3(t)=0) is shown as a red curve in Figures 913. We called this simulation as Strategy 0.

    Figure 9.  Dynamics of total infected human, infected mosquitoes, and control variables for Strategy 1.

    For optimal control experiments, we divide our experiments based on the combination of each form of intervention as follows:

    ● Strategy 1: Vaccination only (u1(t)0,u2(t)=u3(t)=0).

    ● Strategy 2: Fumigation only (u1(t)=u2(t)=0,u30).

    ● Strategy 3: Vaccination and hospitalization only (u1(t)0,u2(t)0,u3(t)=0).

    ● Strategy 4: Vaccination and fumigation only (u1(t)0,u2(t)=0,u3(t)0).

    ● Strategy 5: All interventions implemented (u1(t)0,u2(t)0,u3(t)0).

    Strategy 1

    In the first simulation, the optimal control simulation used vaccination as the only means of intervention. The results are shown in Figure 9.

    The implementation of vaccination with the greatest effort should be done at the beginning of simulation. It will reduce the number of infected people, reducing both direct contact between vulnerable mosquitoes and infected people and the number of infected mosquitoes. Reducing the level of intervention will cause the number of infected people and mosquitoes to increase near the end of the simulation. Yellow fever vaccines can cause side effects in rare cases; therefore, the total number of infected people rose immediately after vaccination before slowly decreasing toward the end.

    Based on maximum vaccination at the start, this strategy has a cost of 632,377 and prevents 444,868 infections, leading to a total of 2,981,365 recovered individuals.

    Strategy 2

    The optimal control strategy for this simulation only uses fumigation to control mosquito populations with the aim of breaking the yellow fever infection cycle. The results are shown in Figure 10.

    Figure 10.  Dynamic of total infected human, infected mosquitoes, and control variables for Strategy 2.

    This strategy involves maximum fumigation for nearly three months, before drastically decreasing the rate of fumigation toward the end owing to the rapidly increasing number of infected mosquitoes since the start. This significantly reduced the number of infected people and mosquitoes. Fumigation helps control the number of infected mosquitoes that can come into direct contact with susceptible individuals.

    Although this strategy has a cost of 333,345, which is almost half that of the first strategy, it prevents 784,547 infections, far more than the first strategy. With this strategy, 1,819,666 people will recover by the end of the period.

    Strategy 3

    The third optimal control simulation combines vaccination and treatment. People in T1 receive treatment to aid in the recovery process. Treatment is given because yellow fever vaccines may cause side effects in rare cases. The results are shown in Figure 11.

    Figure 11.  Dynamic of total infected human, infected mosquitoes, and control variables for Strategy 3.

    The implementation of vaccination with the greatest effort should be done at the beginning of simulation. It would be accompanied by the side effects of the yellow fever vaccine. Treatment must counteract this effect. Treatment continued until day 14. This reduces the number of infected individuals, which without any intervention will increase until the 30th day. Vaccination is reduced because of the declining number of infected people.

    This strategy has a higher cost than the previous two strategies (664,808) but is only slightly better at preventing infections (458,867) and yields a similar number of recovered people (2,949,196).

    Strategy 4

    The fourth optimal control simulation uses both vaccination and fumigation to combat yellow fever. The results are shown in Figure 12.

    Figure 12.  Dynamic of total infected human, infected mosquitoes, and control variables for Strategy 4.

    Maximum fumigation and vaccination are given simultaneously for the first 60 days to slow the rise in the number of infected people and mosquitoes.

    Maximum intervention throughout the period is impossible because of cost constraints. Therefore, vaccination is reduced after the 65th day, and fumigation is reduced after the 81st day. This strategy causes the number of infected mosquitoes and infected humans to approach zero on the 14th and the 38th days, respectively.

    This strategy has the highest cost of all four strategies (1,933,426) but yields the same number of prevented infections as the second strategy (815,668) and fewer recovered individuals than the second strategy (1,551,319).

    Strategy 5

    In the last simulation, the optimal control simulation simultaneously involved all interventions (vaccination, hospitalization, and fumigation) to control the spread of yellow fever. The simulation results are shown in Figure 13. Note that all interventions must be given to the fullest extent possible from the start of the simulation. Therefore, the number of infected humans and mosquitoes has decreased significantly since the first day. When the infected population in both human and mosquito populations stops increasing, the vaccination, hospitalization, and fumigation interventions will gradually decrease until they reach zero on the final day of the simulation.

    Figure 13.  Dynamic of total infected human, infected mosquitoes, and control variables for Strategy 5.

    With maximum intervention from the beginning of the simulation, the cost function is worth 2694.146, which is much larger than the cost function value in Strategy 4. Nonetheless, when compared to Strategy 4, the number of infected humans successfully avoided was not significantly different, that is, 816 individuals. Similarly, the number of successfully recovered people is not significantly different from Strategy 4, which is 1551 individuals.

    It is critical to determine the most cost-effective strategy among the five proposed strategies based on the results of our optimal control simulations. To accomplish this, we use three methods: the average cost-effectiveness ratio (ACER), infected averted ratio (IAR), and incremental cost-effectiveness ratio (ICER) [51,52,53].

    Average cost-effectiveness ratio (ACER)

    The ACER method calculates the average cost that should be spent for each number of avoided infected individuals. The ACER formula is as follows:

    ACERStrategy i=Total cost for intervention with Strategy iTotal number of infection averted with Strategy i. (4.7)

    Therefore, a smaller ACER indicates a better strategy. The results of the ACER values for each strategy are shown in Figure 14. Note that the best strategy based on the ACER index implemented fumigation as a single form of intervention (Strategy 2), followed by Strategies 1, 3, 4, and 5. This means that using fumigation as a single form of intervention is the cheapest way, on average, to avoid many newly infected individuals. An interesting result shows that in our numerical experiment, vaccination is not cost-effective in reducing the expected number of newly infected individuals. This is because in our simulation, the number of infected individuals is already high (13% in the human population and 15% in the mosquito population). Furthermore, the simultaneous implementation of all strategies is not cost-effective because we must increase the already high hospitalization cost further.

    Figure 14.  The average cost-effectiveness ratio (ACER) of Strategies 1–5.

    Infected averted ratio (IAR)

    The IAR method calculates the ratio between the number of infected averted individuals and the number of recovered individuals. The equation for IAR is as follows:

    IARStrategy i=Total number of infection averted with Strategy iTotal recovered individuals with Strategy i. (4.8)

    Based on the above equation, a larger IAR indicates a better strategy. Figure 15 shows that a combination of all means of interventions (Strategy 5) is the best strategy in terms of IAR, but it is only slightly different from Strategy 4. This is because when vaccination, hospitalization, and fumigation are implemented together, all means of intervention aim to reduce the endemic situation and avoid newly infected individuals. However, as we already discussed in the previous section, Strategy 5 comes at a high cost. The second-best strategy in terms of IAR is Strategy 4, followed by Strategies 2, 3, and 1. Note that implementing vaccination as a single single form of intervention is the least cost-effective strategy in the terms of IAR. This is because this type of intervention is more concerned with prevention rather than endemic reduction.

    Figure 15.  The infected averted ratio (IAR) of Strategies 1–5.

    Incremental cost-effectiveness ratio (ICER)

    The ICER was the final cost-effectiveness analysis method. The ICER compares the cost differences between two strategies and the number of infected averts. Hence, the equation for ICER is as follows:

    ICERi–j=Difference of cost between Strategy i and Strategy jDifference of number of infection averted between Strategy i and Strategy j. (4.9)

    Based on the numerical simulation, we rank all strategies in an increasing order based on the total number of infections averted in Table 3. Strategy 1 (intervention of vaccination only) has the lowest number of infections averted, while Strategy 5 (all interventions implemented) has the largest number of infections averted.

    Table 3.  Strategies 1–5 in increasing order based on the number of infections averted.
    Strategy Total infections averted Total cost ICER
    Strategy 1 444.868 632.377 1.4215
    Strategy 3 458.867 664.808 2.3167
    Strategy 2 784.547 333.345 -1.0178
    Strategy 4 815.668 1933.426 51.415
    Strategy 5 816 2694.146 2285.8173

     | Show Table
    DownLoad: CSV

    To compare strategies 1 and 3, the ICER is calculated as follows:

    ICER1=632.377444.868=1.4215,ICER3=664.808632.377458.867444.868=2.3167,ICER2=333.345664.808784.547458.867=1.0178,ICER4=1933.426333.345815.668784.547=51.415,ICER5=2694.1461933.426816815.668=2285.8173.

    Note that ICER-3 > ICER-1. Hence, we can exclude ICER-3 from the next calculation because it is more costly than ICER-3. Next, we continue our calculation by comparing the ICER between Strategies 1 and 2. The result of calculating ICER using the same method as before is shown in Table 4.

    Table 4.  Comparison between Strategies 1 and 2.
    Strategy Total infection averted Total cost ICER
    Strategy 1 444.868 632.377 1.4215
    Strategy 2 784.547 333.345 -.88038
    Strategy 4 815.668 1933.426 51.415
    Strategy 5 816 2694.146 2285.8173

     | Show Table
    DownLoad: CSV

    Table 4 shows that ICER-1>ICER-2, which means that Strategy 1 is more costly than Strategy 2. Hence, we exclude Strategy 1 from the next calculation. Up to this step, we have three last candidates for the best strategies, i.e., Strategies 2, 4, and 5. Next, we compare Strategies 2 and 4, the results of which are shown in Table 5.

    Table 5.  Comparison between Strategies 2 and 4.
    Strategy Total infection averted Total cost ICER
    Strategy 2 784.547 333.345 0.4249
    Strategy 4 815.668 1933.426 51.415
    Strategy 5 816 2694.146 2285.8173

     | Show Table
    DownLoad: CSV

    Table 5 shows that Strategy 4 is more costly compared to Strategy 2 because ICER-4 > ICER-2. Hence, we exclude Strategy 4 from the next comparison. The results are presented in Table 6.

    Table 6.  Comparison between Strategies 2 and 5.
    Strategy Total infection averted Total cost ICER
    Strategy 2 784.547 333.345 0.4249
    Strategy 5 816 2694.146 75.0564

     | Show Table
    DownLoad: CSV

    Table 6 indicates that ICER-5 > ICER-2, which means that Strategy 5 (all interventions implemented) is more costly compared to Strategy 2 (only fumigation strategy implemented) in reducing the spread of yellow fever. Additionally, Strategy 5 is dominated by Strategy 2, which means that Strategy 5 is less effective compared to Strategy 2. Hence, we can conclude that fumigation, as a single intervention to reduce the spread of yellow fever, is the most cost-effective strategy compared to other possible strategies.

    Yellow fever is an acute viral hemorrhagic fever caused by the yellow fever virus of the Flavivirus genus. It is spread by the bite of several types of female mosquitoes (Aedes, Haemagogus) [54]. It is estimated that more than 100,000 cases occur each year, with Africa accounting for more than 60% of all cases. There are three types of yellow fever based on species that interact during the transmission process. The first type is the jungle transmission, where the transmission process involves monkeys and mosquitoes, while the second type, namely, the urban type, involves mosquitoes and humans. The last type, namely the intermediate type, which commonly occurs in Africa, involves mosquitoes, monkeys, and humans in the transmission process [55]. Yellow fever has an incubation period of three to six days, and the symptoms vary, from headache, backache, vomiting, bleeding, and yellow eyes to death [56]. Several means of interventions have been implemented to control the spread of yellow fever, such as vaccination and fumigation. Vaccination for yellow fever is strongly recommended for individuals planning to travel to the yellow fever endemic area. However, there are a few cases where yellow fever vaccination has a side effect on some vulnerable individuals, such as the elderly, pregnant women, etc.[25]. Infected individuals who already have severe symptoms should be hospitalized.

    This work introduces a novel optimal control of the yellow fever model based on a ten-dimensional system of ordinary differential equations with a compartmental model approach. Our model includes three possible form of interventions: vaccination, hospitalization, and fumigation. We show that our model is mathematically and biologically well-defined. Our mathematical analysis shows that our model has two possible equilibriums: the yellow fever-free equilibrium point and the yellow fever endemic equilibrium point. We show that the basic reproduction number of our model plays an important role in determining the existence and stability of our equilibrium points. We discovered that the yellow fever-free equilibrium is always locally asymptotically stable when the basic reproduction number is less than one and unstable when it is greater than one. However, the yellow fever endemic equilibrium never exists when the basic reproduction number is less than one and exists (and is stable) when the basic reproduction number is greater than one. Additionally, we found that the endemic size in human and mosquito populations increased rapidly when the basic reproduction number increased. Using the Castillo-Song bifurcation theorem [41], we demonstrated that our model exhibits a forward bifurcation phenomenon at a basic reproduction number of one. Hence, we can conclude that our basic reproduction number becomes a necessary threshold to determine whether yellow fever will be endemic or disappear from the population.

    An optimal control problem is then developed and characterized using the Pontryagin maximum principle [49]. We treated our form of interventions (vaccination, hospitalization, and fumigation) as time-dependent variables. We ran our simulations using a forward-backward method [50] for five types of strategies, depending on the possible combination of strategies. The simulation results demonstrated that all means of intervention can eradicate yellow fever. Our sensitivity analysis found that fumigation was the most elastic parameter for determining the basic reproduction number. A cost-effectiveness analysis was conducted to determine the best strategies to prevent the spread of yellow fever. Three types of cost-effectiveness indicators were used: ACER, IAR, and ICER. We found that vaccination intervention showed the best ACER results, implying that the vaccination strategy as a single intervention provides the lowest average cost required for an averted infection individual. On the other hand, the strategy of combining vaccination, hospitalization, and fumigation showed the best IAR results. Additionally, we found that using fumigation as a single form of intervention shows the best ICER result, which agrees with the ACER result. Therefore, the results of this study can be used by authorities to determine the best strategy to implement in the field to control the spread of yellow fever.

    This study considers only three form of interventions: fumigation, vaccination, and hospitalizations. However, some other means of intervention can be chosen as an option, such as using bed nets or mosquito repellents to reduce the mosquito bite. Furthermore, massive vector control such as fumigation, is a high risk because it can trigger mosquitoes' resistance to fumigants. Hence, a mathematical model concerning more means of intervention and the possibility of mosquito resistance could be considered in future work. Last but not least, it is essential to fit our model to the real-life situation by estimating the best fit parameter. Hence, incidence data of yellow fever is essential for conducting parameter estimation into our model.

    The author would like to thank the anonymous reviewers for his/her valuable comments. This research was funded by the Indonesian RistekBRIN with PDUPT research grant scheme (ID Number: NKB-159/UN2.RST/HKP.05.00/2021), 2021.

    The authors declare that they have no conflict of interest.

    From yellow fever model in Eq (3.1), with a non-negative initial conditions on each variables, we have the following result:

    dSdt|S>0,V0,E10,E20,I0,T10,T20,R0,U>0,V0=Λh+ωhV>0,dVdt|S>0,V=0,E10,E20,I0,T10,T20,R0,U>0,V0=u1S0,dE1dt|S>0,V0,E1=0,E20,I0,T10,T20,R0,U>0,V0=βhSW0,dE2dt|S>0,V0,E10,E2=0,I0,T10,T20,R0,U>0,V0=ξβhVW0,dIdt|S>0,V0,E10,E20,I=0,T10,T20,R0,U>0,V0=γ1E1+(1p)γ2E20,dT1dt|S>0,V0,E10,E20,I0,T1=0,T20,R0,U>0,V0=pγ2E20,dT2dt|S>0,V0,E10,E20,I0,T10,T2=0,R0,U>0,V0=u2T10,dRdt|S>0,V0,E10,E20,I0,T10,T20,R=0,U>0,V0=δ1I+δ2T1+δ3T20,dUdt|S>0,V0,E10,E20,I0,T10,T20,R0,U=0,V0=Λv>0,dWdt|S>0,V0,E10,E20,I0,T10,T20,R0,U>0,V=0=βvU(I+T1)0. (A.1)

    The above calculation at the boundary of each variable indicates that the rates of each variable are non-negative on the boundary planes of R10+. Hence, we can conclude that all vector fields in the boundary are always pointed inward. Therefore, whenever the initial condition is non-negative in R10+, the solution will always remain non-negative for all times t>0. Hence, the proof is complete.

    Because dNdt=ΛhIα1T1α2T2α3μhN, we have

    dNdt=ΛhIα1T1α2T2α3Nμh,ΛhμhN. (B.1)

    Solve the above equation with respect to N(t), we have the following:

    N(t)N(0)eμht+Λhμh(1eμht).

    Therefore, for t, N(t) tends to Λhμh. A similar approach can be used to demonstrate that the total mosquito population (M(t)) is eventually bounded by Λvμv. Hence, the proof is complete.

    Taking only the infected compartment in system (3.1), the transmission (T) and the transition (Σ) matrix, which are evaluated in E1, are given by the following:

    T=[00000(μh+ωh)Λhβhμh(μh+ωh+u1)00000ξβhΛhu1μh(μh+ωh+u1)00000000000000000000βvΛvμv+u3βvΛvμv+u300],

    and

    Σ=[γ1μh000000a220000γ1(1p)γ2μhα1δ10000pγ20δ2u2μhα200000u2δ3μhα3000000μvu3],

    where a22=(1p)γ2pγ2μh. Using the formula in [33], the next generation matrix of system (3.1) is given by K=EtTΣ1E, where

    E=[100010000000000001],Et=[100000010000000001].

    Hence, the basic reproduction number of system (3.1) as the spectral radius of K is given by the following:

    R0=ρ(K)=μh(α2+δ2+μh+u2)(γ1+μh)M1μh(μh+ωh+u1)(γ2+μh)(μv+u3)(γ1+μh)(α2+δ2+μh+u2)(α1+δ1+μh), (C.1)

    with

    M1=(γ1μ3h+((ξu1+γ1)γ2+γ1(δ2+ωh+u2+α2))μ2h+(((ξu1+α2+δ2+ωh+u2)γ1+u1(pα1+pδ1+(α2+δ2+u2)(1p))ξ)γ2+γ1ωh(α2+δ2+u2))μh+γ1γ2(u1(pα1+pδ1+(α2+δ2+u2)(1p))ξ+ωh(α2+δ2+u2)))Λvβv(μh+ωh+u1)(γ2+μh)βh(α1+δ1+μh)Λh.

    Let us simplify the notion of E1,E2,I,T1,T2,W,R,S,V,U as xi for i=1,2,10, respectively. We redefine system (3.1) as follows:

    f1:=dE1dt=βhSWγ1E1μhE1,f2:=dE2dt=ξβhVW(1p)γ2E2pγ2E2μhE2,f3:=dIdt=γ1E1+(1p)γ2E2(μh+α1)Iδ1I,f4:=dT1dt=pγ2E2δ2T1u2T1(μh+α2)T1,f5:=dT2dt=u2T1δ3T2(μh+α3)T2,f6:=dWdt=βvU(I+T1)(μv+u3)W,f7:=dRdt=δ1I+δ2T1+δ3T2μhR,f8:=dSdt=Λh+ωhVβhSWu1SμhS,f9:=dVdt=u1SξβhVWμhVωhV,f10:=dUdt=ΛvβvU(I+T1)(μv+u3)U. (D.1)

    Now, we define our yellow-fever free equilibrium as

    Xs=(x1,x2,x3,x4,x5,x6,x7,x8,x9,x10),=(0,0,0,0,0,0,0,λh(μh+ωh)μh(μh+ωh+u1),Λhu1μh(μh+ωh+u1),Λvμv+u3). (D.2)

    To apply Theorem 2 in [34], we have to show that the system (D.1) satisfies five axioms in [34]. Therefore, we need to separate the infection (F) and the non-infection (V) term in the symstem (D.1) as follows:

    F=[WSβhξβhVW000βvU(I+T1)0000],V=[(γ1+μh)E1(1p)γ2E2+(pγ2+μh)E2γ1E1(1p)γ2E2+(μh+α1)I+δ1Ipγ2E2+(δ2+u2)T1+(μh+α2)T1u2T1+δ3T2+(μh+α3)T2(μv+u3)WμhRδ1Iδ2T1δ3T2ΛhωhV+βhWS+(u1+μh)Su1S+ξβhVW+(μh+ωh)VΛv+βvU(I+T1)+(μv+u3)U]. (D.3)

    Note that Vi can be decomposed as Vi=ViV+i, where Vi and V+i represent the out and in flow in compartment-i, respectively. Now, we are ready to proof the five axioms in [34].

    1. If all compartment (xi, i=1,2,10) are non-negative, then Fi,Vi, and V+i are always non-negative. By substituting xi0 into Fi,Vi, and V+i, we can verify that Fi,Vi, and V+i are always non-negative.

    2. If all compartments are empty, then there is no out-flow rate from each compartment. It is easy to verify that, Vi=0 for i=1,2,10 when we set xi=0 for i=1,2,10.

    3. For all non infected compartment (xi, i=7,8,9,10), we have that Fi=0. This axiom satisfies directly by the expression of F at (D.3).(?)

    4. If xiXs, then Fi=0=V+i=0 for i=1,2,6. By substituting Xs into Fi, i=1,2,6, we have that Fi=0 and V+i=0, i=1,2,6.

    5. If F(x)=0, then all the eigenvalues of Df(Xs) have a negative real part, where Df(Xs) is the Jacobian matrix of (D.1) evaluated at Xs. By substituting Xs into system (D.1), we obtain :

    f1=dE1dt=γ1E1μhE1,f2=dE2dt=(1p)γ2E2pγ2E2μhE2,f3=dIdt=γ1E1+(1p)γ2E2(μh+α1)Iδ1I,f4=dT1dt=γ2E2δ2T1u2T1(μh+α2)T1,f5=dT2dt=u2T1δ3T2(μh+α3)T2,f6=dWdt=(μv+u3)W,f7=dRdt=δ1I+δ2T1+δ3T2μhR,f8=dSdt=Λh+ωhVβhSWu1SμhS,f9=dVdt=u1SξβhVWμhVωhV,f10=dUdt=ΛvβvU(I+T1)(μv+u3)U. (D.4)

    Then, we can evaluate Df(Xs) of system (D.4). By standard calculation, we obtain the eigenvalues of Df(Xs) as follows:

    λ1=μh,λ2=μh,λ3=(μv+u3),λ4=(μv+u3),λ5=(γ2+μh),λ6=(γ1+μh),λ7=(α3+δ3+μh),λ8=(u2+α2+δ2+μh),λ9=(ωh+u1+μh),λ10=(α1+δ1+μh).

    Since all parameter values are non-negative, we can see that all eigenvalues of Df(Xs) are negative.

    Hence, the proof is completed.

    From the expression of F(W) in Eq (3.5), we can see that a2 is always positive, whereas a0<0R20>1. Using the formula for the multiplication of roots in a second-degree polynomial, we have W1×W2=a0a2, where W1 and W2 is the root of F(W). Hence, if R20>1, we have W1×W2<0, which indicates that there exists one positive root, and the other one is negative. Therefore, we have a unique yellow fever endemic equilibrium point when R20>1R0>1.

    Now, we analyze the possibility of the existence of E2 when R20<1. To accomplish this, we use the properties of the second-degree polynomial to have two positive roots, that is, W1×W2=a0a2>0, W1+W2=a1a2>0, and a214a2a00. However, because the expressions of a2,a1, and a0 is not simple, the expression of a214a2a0 is not easy to analyze. Hence, we use another approach to analyze the existence of E2 when R0<1. To accomplish this, we use a gradient analysis of W in R0=1 and W=0. If we can find a condition such that WR0<0, we can have at least one positive root of F(W) when R0<1. Because the expression of R0 contains a square root, and the condition of R0=1 is represented by R20=1, we will analyze the gradient of W with respect to R20 at R20=1 instead of R0=1.

    Solving the expression of R20=1 in Eq (3.3) with respect to βh, and substituting it with F(W), we have a2,a1, and a0, which is an expression of R20. Taking the implicit derivative of W with respect to R20 from F(W), we have the following:

    a2R20(W)2+a2(2W)WR20+a1R20W+a1WR20+a0R20=0. (E.1)

    substitute W=0, and solve it with respect to WR20, we have the following:

    WR20=a0R20a1>0,

    where a0R20=2μh(μv+u3)2(γ2+μh)(γ1+μh)(α2+δ2+μh+u2)(μh+ωh+u1)(α1+δ1+μh)<0 and a1>0. Hence, the gradient of W at R20=1 is always positive without any condition. Therefore, we can conclude that there is no yellow fever endemic equilibrium when R0<1. Hence, the proof is complete.



    [1] J. Florczak-Wyspianska, E. Nawotczynska, W. Kozubski, Yellow fever vaccine-associated neurotropic disease (yel-and)–a case report, Neurol. Neurochir., 51 (2017), 101–105. doi: 10.1016/j.pjnns.2016.09.002. doi: 10.1016/j.pjnns.2016.09.002
    [2] United Nations, Prevention of Yellow Fever, 2019. Available from: https://www.cdc.gov/yellowfever/prevention/index.html.
    [3] United Nations, Yellow Fever, 2019. Available from: https://www.who.int/news-room/fact-sheets/detail/yellow-fever.
    [4] J. E. Staples, A. D. Barrett, A. Wilder-Smith, J. Hombach, Review of data and knowledge gaps regarding yellow fever vaccine-induced immunity and duration of protection, Vaccines, 5 (2020), 1–7. doi: 10.1038/s41541-020-0205-6. doi: 10.1038/s41541-020-0205-6
    [5] D. Aldila, Analyzing the impact of the media campaign and rapid testing for covid-19 as an optimal control problem in east java, Indonesia, Chaos, Solitons Fractals, 141 (2020), 110364. doi: 10.1016/j.chaos.2020.110364. doi: 10.1016/j.chaos.2020.110364
    [6] D. Aldila, Optimal control for dengue eradication program under the media awareness effect, Int. J. Nonlinear Sci. Numer. Simul., 2021. doi: 10.1515/ijnsns-2020-0142. doi: 10.1515/ijnsns-2020-0142
    [7] D. Aldila, Mathematical model for HIV spreads control program with ART treatment, in Journal of physics: Conference series, 974 (2018), 012035.
    [8] C. A. G. Engelhard, A. P. Hodgkins, E. E. Pearl, P. K. Spears, J. Rychtar, D. Taylor, A mathematical model of Guinea worm disease in Chad with fish as intermediate transport hosts, J. Theor. Biol., 521 (2021), 110683. doi: 10.1016/j.jtbi.2021.110683. doi: 10.1016/j.jtbi.2021.110683
    [9] Z. Guo, G. Sun, Z. Wang, Z. Jin, L. Li, C. Li, Spatial dynamics of an epidemic model with nonlocal infection, Appl. Math. Comput., 377 (2020), 125158. doi: 10.1016/j.amc.2020.125158. doi: 10.1016/j.amc.2020.125158
    [10] G. Sun, M. Li, J. Zhang, W. Zhang, X. Pei, Z. Jin, Transmission dynamics of brucellosis: Mathematical modelling and applications in China, Comput. Struct. Biotechnol. J., 18 (2020), 3843–3860. doi: 10.1016/j.csbj.2020.11.014. doi: 10.1016/j.csbj.2020.11.014
    [11] M. Chapwanya, A. Matusse, Y. Dumont, On synergistic co-infection in crop diseases. The case of the Maize Lethal Necrosis Disease, Appl. Math. Modell., 90 (2021), 912–942. doi: 10.1016/j.apm.2020.09.036. doi: 10.1016/j.apm.2020.09.036
    [12] Y. Belgaid, M. Helal, A. Lakmeche, E. Venturino, On the stability of periodic solutions of an impulsive system arising in the control of agroecosystems, in International Symposium on Mathematical and Computational Biology, (2020), 183–199.
    [13] M. Kung'aro, L. Luboobi, F. Shahada, Modelling and stability analysis of SVEIRS yellow fever two host model, Gulf J. Math., 3 (2015), 106–129. doi: 10.1016/j.ces.2015.02.038. doi: 10.1016/j.ces.2015.02.038
    [14] S. Raimundo, M. Amaku, E. Massad, Equilibrium analysis of a yellow fever dynamical model with vaccination, Comput. Math. Methods Med., 2015 (2015), 482091. doi: 10.1155/2015/482091. doi: 10.1155/2015/482091
    [15] U. Danbaba, S. Garba, Stability analysis and optimal control for yellow fever model with vertical transmission, Int. J. Appl. Comput. Math, 6 (2020), 1–34. doi: 10.1007/s40819-020-00860-z. doi: 10.1007/s40819-020-00860-z
    [16] S. Zhao, L. Stone, D. Gao, D. He, Modelling the large-scale yellow fever outbreak in luanda, angola, and the impact of vaccination, PLoS Negl. Trop. Dis., 12 (2018), e0006158. doi: 10.1371/journal.pntd.0006158. doi: 10.1371/journal.pntd.0006158
    [17] F. Agusto, M. Leite, Optimal control and cost-effective analysis of the 2017 meningitis outbreak in Nigeria, Infect. Dis. Modell., 4 (2019), 161–187. doi: 10.1016/j.idm.2019.05.003. doi: 10.1016/j.idm.2019.05.003
    [18] M. Bruyand, M. Receveur, T. Pistone, C. Verdiere, R. Thiebaut, D. Malvy, Yellow fever vaccination in non-immunocompetent patients, Med. Mal. Infect., 38 (2008), 524–532. doi: 10.1016/j.medmal.2008.06.031. doi: 10.1016/j.medmal.2008.06.031
    [19] S. M. Raimundo, M. Amaku, E. Massad, Equilibrium analysis of a yellow fever dynamical model with vaccination, Comput. Math. Methods Med., 2015 (2015), 482091. doi: 10.1155/2015/482091. doi: 10.1155/2015/482091
    [20] Pan American Health Organization/World Health Organization, Epidemiological update: yellow fever, 2018. Available from: https://reliefweb.int/report/brazil/epidemiological-update-yellow-fever-20-march-2018.
    [21] F. M. Shearer, C. L. Moyes, D. M. Pigott, O. J. Brady, F. Marinho, A. Deshpande, et al., Global yellow fever vaccination coverage from 1970 to 2016: an adjusted retrospective analysis, Lancet Infect. Dis., 17 (2017), 1209–1217. doi: 10.1016/S1473-3099(17)30419-X. doi: 10.1016/S1473-3099(17)30419-X
    [22] K. W. Blayneh, A. B. Gumel, S. Lenhart, T. Clayton, Backward bifurcation and optimal control in transmission dynamics of West Nile virus, Bull. Math. Biol., 72 (2010), 1006–1028. doi: 10.1007/s11538-009-9480-0. doi: 10.1007/s11538-009-9480-0
    [23] B. Buonomo, R. D. Marca, Optimal bed net use for a dengue disease model with mosquito seasonal pattern, Math. Methods Appl. Sci., 41 (2018), 573–592. doi: 10.1002/mma.4629. doi: 10.1002/mma.4629
    [24] M. Andraud, N. Hens, C. Marais, P. Beutels, Dynamic epidemiological models for dengue transmission: a systematic review of structural approaches, PLoS ONE, 7 (2012), e49085. doi: 10.1371/journal.pone.0049085. doi: 10.1371/journal.pone.0049085
    [25] P. Cottin, M. Niedrig, C. Domingo, Safety profile of the yellow fever vaccine stamaril: a 17-year review, Expert Rev. Vaccines, 12 (2013), 1351–1368. doi: 10.1586/14760584.2013.836320. doi: 10.1586/14760584.2013.836320
    [26] D. Aldila, T. Götz, E. Soewono, An optimal control problem arising from a dengue disease transmission model, Math. Biosci., 242 (2013), 9–16. doi: 10.1016/j.mbs.2012.11.014. doi: 10.1016/j.mbs.2012.11.014
    [27] T. T. Yusuf, D. O. Daniel, Mathematical modelling of yellow fever transmission dynamics with multiple control measures, Asian Res. J. Math., 13 (2019), 1–15. doi: 10.9734/arjom/2019/v13i430112. doi: 10.9734/arjom/2019/v13i430112
    [28] D. Aldila, M. Angelina, Optimal control problem and backward bifurcation on malaria transmission with vector bias, Heliyon, 7 (2021), e06824. doi: 10.1016/j.heliyon.2021.e06824. doi: 10.1016/j.heliyon.2021.e06824
    [29] B. Handari, F. Vitra, R. Ahya, T. N. S, D. Aldila, Optimal control in a malaria model: intervention of fumigation and bed nets, Adv. Differ. Equations, 2019 (2019), 497. doi: 10.1186/s13662-019-2424-6. doi: 10.1186/s13662-019-2424-6
    [30] M. S. Indriyono Tantoro, Pedoman Pencegahan Penyakit Yellow Fever, in Kementerian Kesehatan Republik Indonesia Direktorat Jenderal Pencegahan dan Pengendalian Penyakit, 2017.
    [31] D. Aldila, B. M. Samiadji, G. M. Simorangkir, S. H. A. Khosnaw, M. Shahzad, Impact of early detection and vaccination strategy in covid-19 eradication program in jakarta, indonesia, BMC Res. Notes, 14 (2021), 132. doi: 10.1186/s13104–021–05540–9. doi: 10.1186/s13104–021–05540–9
    [32] I. M. Wangaria, S. Davisa, L. Stone, Backward bifurcation in epidemic models: Problems arisingwith aggregated bifurcation parameters, Appl. Math. Modell., 40 (2016), 1669–1675. doi: 10.1016/j.apm.2015.07.022. doi: 10.1016/j.apm.2015.07.022
    [33] O. Diekmann, J. A. P. Heesterbeek, M. G. Roberts, The construction of next-generation matrices for compartmental epidemic models, J. R. Soc. Interface, 7 (2010), 873–885. doi: 10.1098/rsif.2009.0386. doi: 10.1098/rsif.2009.0386
    [34] P. van den Driessche, J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002), 29–48. doi: 10.1016/S0025-5564(02)00108-6. doi: 10.1016/S0025-5564(02)00108-6
    [35] G. Simorangkir, D. Aldila, A. Rizka, H. Tasman, E. Nugraha, Mathematical model of tuberculosis considering observed treatment and vaccination interventions, J. Interdiscip. Math., 24 (2021), 1717–1737. doi: 10.1080/09720502.2021.1958515. doi: 10.1080/09720502.2021.1958515
    [36] D. Aldila, K. Rasyiqah, G. Ardaneswari, H. Tasman, A mathematical model of zika disease by considering transition from the asymptomatic to symptomatic phase, in Journal of Physics: Conference Series, 1821 (2021), 012001. doi: 10.1088/1742-6596/1821/1/012001.
    [37] D. Aldila, B. Handari, Effect of healthy life campaigns on controlling obesity transmission: A mathematical study, in Journal of Physics: Conference Series, 1747 (2021), 012003. doi: 10.1088/1742-6596/1747/1/012003.
    [38] J. Li, Y. Zhao, S. Li, Fast and slow dynamics of malaria model with relapse, Math. Biosci., 246 (2013), 94–104. doi: 10.1098/rspb.2016.0048. doi: 10.1098/rspb.2016.0048
    [39] K. Nudee, S. Chinviriyasit, W. Chinviriyasit, The effect of backward bifurcation in controlling measles transmission by vaccination, Chaos Solitons Fractals, 123 (2018), 400–412. doi: 10.1016/j.chaos.2019.04.026. doi: 10.1016/j.chaos.2019.04.026
    [40] O. Sharomi, C. Podder, A. Gumel, E. Elbasha, J. Watmough, Role of incidence function in vaccine-induced backward bifurcation in some HIV models, Math. Biosci., 210 (2007), 436–463. doi: 10.1016/j.mbs.2007.05.012. doi: 10.1016/j.mbs.2007.05.012
    [41] C. Castillo–Chavez, B. Song, Dynamical models of tuberculosis and their applications, Math. Biosci. Eng., 1 (2014), 361–404. doi: 10.3934/mbe.2004.1.361. doi: 10.3934/mbe.2004.1.361
    [42] N. Chitnis, J. Hyman, J. Cushing, Determining important parameters in the spread of malaria through the sensitivity analysis of a mathematical model, Bull. Math. Biol., 70 (2008), 1272–1296. doi: 10.1007/s11538-008-9299-0. doi: 10.1007/s11538-008-9299-0
    [43] S. H. A. Khosnaw, M. Shahzad, M. Ali, F. Sultan, A quantitative and qualitative analysis of the covid–19 pandemic model, Chaos, Solitons Fractals, 138 (2020), 109932. doi: 10.1016/j.chaos.2020.109932. doi: 10.1016/j.chaos.2020.109932
    [44] S. H. A. Khosnaw, M. Shahzad, M. Ali, F. Sultan, Mathematical modelling for coronavirus disease (covid-19) in predicting future behaviours and sensitivity analysis, Math. Model. Nat. Phenom., 15 (2020), 33. doi: 10.1016/j.chaos.2020.109932. doi: 10.1016/j.chaos.2020.109932
    [45] A. Abidemi, N. Aziz, Optimal control strategies for dengue fever spread in Johor, Malaysia, Comput. Methods Programs Biomed., 196 (2020), 105585. doi: 10.1016/j.cmpb.2020.105585. doi: 10.1016/j.cmpb.2020.105585
    [46] D. Aldila, M. Ndii, B. Samiadji, Optimal control on covid-19 eradication program in Indonesia under the effect of community awareness, Math. Biosci. Eng., 17 (2021), 6355–6389. doi: 10.3934/mbe.2020335. doi: 10.3934/mbe.2020335
    [47] N. Sharma, R. Singh, J. Singh, E. Castillo, Modeling assumptions, optimal control strategies and mitigation through vaccination to zika virus, Chaos, Solitons Fractals, 150 (2021), 111137. doi: 10.1016/j.chaos.2021.111137. doi: 10.1016/j.chaos.2021.111137
    [48] N. Sweilam, S. Al-Mekhlafi, Optimal control for a fractional tuberculosis infection model including the impact of diabetes and resistant strains, J. Adv. Res., 17 (2019), 125–137. doi: 10.3934/mbe.2020335. doi: 10.3934/mbe.2020335
    [49] L. Pontryagin, V. Boltyanskii, R. Gamkrelidze, E. Mishchenko, The Mathematical Theory of Optimal Processes, Interscience Publishers John Wiley & Sons, Inc., New York-London.
    [50] S. Lenhart, J. Workman, Optimal Control Applied to Biological Models, Chapman and Hall, London/Boca Raton, 2007.
    [51] K. Okosun, O. Rachid, N. Marcus, Optimal control strategies and cost-effectiveness analysis of a malaria model, BioSystems, 111 (2013), 83–101. doi: 0.1016/j.biosystems.2012.09.008.
    [52] F. Agusto, I. ELmojtaba, Optimal control and cost-effective analysis of malaria/visceral leishmaniasis co–infection, PLoS ONE, 12 (2017), 1–31. doi: 10.1371/journal.pone.0171102. doi: 10.1371/journal.pone.0171102
    [53] J. Akanni, F. Akinpelu, S. Olaniyi, A. Oladipo, A. Ogunsola, Modelling financial crime population dynamics: optimal control and cost-effectiveness analysis, Int. J. Dyn. Control, 2019 (2019), 1–14. doi: 10.1007/s40435-019-00572-3. doi: 10.1007/s40435-019-00572-3
    [54] E. Q. Lima, M. L. Nogueira, Viral hemorrhagic fever-induced acute kidney injury, in Seminars in Nephrology, 28 (2008), 409–415. doi: 10.1016/j.semnephrol.2008.04.009.
    [55] R. Klitting, E. A. Gould, C. Paupy, X. De Lamballerie, What does the future hold for yellow fever virus?, Genes, 9 (2018), 291. doi: 10.3390/genes9060291. doi: 10.3390/genes9060291
    [56] I. McGuinness, J. D. Beckham, K. L. Tyler, D. M. Pastula, An overview of yellow fever virus disease, Neurohospitalist, 7 (2017), 157. doi: 10.1177/1941874417708129. doi: 10.1177/1941874417708129
  • This article has been cited by:

    1. Dipo Aldila, Bapan Ghosh, Dynamical Analysis on a Malaria Model with Relapse Preventive Treatment and Saturated Fumigation, 2022, 2022, 1748-6718, 1, 10.1155/2022/1135452
    2. Dipo Aldila, Meksianis Z. Ndii, Nursanti Anggriani, Hengki Tasman, Bevina D. Handari, Impact of social awareness, case detection, and hospital capacity on dengue eradication in Jakarta: A mathematical model approach, 2023, 64, 11100168, 691, 10.1016/j.aej.2022.11.032
    3. Dipo Aldila, Nadya Awdinda, Faishal F. Herdicho, Meksianis Z. Ndii, Chidozie W. Chukwu, Optimal control of pneumonia transmission model with seasonal factor: Learning from Jakarta incidence data, 2023, 9, 24058440, e18096, 10.1016/j.heliyon.2023.e18096
    4. S. Olaniyi, F. M. Chuma, Lyapunov Stability and Economic Analysis of Monkeypox Dynamics with Vertical Transmission and Vaccination, 2023, 9, 2349-5103, 10.1007/s40819-023-01572-w
    5. Dipo Aldila, Chita Aulia Puspadani, Rahmi Rusin, Mathematical analysis of the impact of community ignorance on the population dynamics of dengue, 2023, 9, 2297-4687, 10.3389/fams.2023.1094971
    6. Nawaf L. Alsowait, Mohammed M. Al-Shomrani, Ismail Abdulrashid, Salihu S. Musa, Synergistic Impact of Active Case Detection and Early Hospitalization for Controlling the Spread of Yellow Fever Outbreak in Nigeria: An Epidemiological Modeling and Optimal Control Analysis, 2024, 12, 2227-7390, 3817, 10.3390/math12233817
  • Reader Comments
  • © 2022 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(3856) PDF downloads(196) Cited by(6)

Figures and Tables

Figures(15)  /  Tables(6)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog