Research article Special Issues

A deterministic transmission model for analytics-driven optimization of COVID-19 post-pandemic vaccination and quarantine strategies


  • This study developed a deterministic transmission model for the coronavirus disease of 2019 (COVID-19), considering various factors such as vaccination, awareness, quarantine, and treatment resource limitations for infected individuals in quarantine facilities. The proposed model comprised five compartments: susceptible, vaccinated, quarantined, infected, and recovery. It also considered awareness and limited resources by using a saturated function. Dynamic analyses, including equilibrium points, control reproduction numbers, and bifurcation analyses, were conducted in this research, employing analytics to derive insights. Our results indicated the possibility of an endemic equilibrium even if the reproduction number for control was less than one. Using incidence data from West Java, Indonesia, we estimated our model parameter values to calibrate them with the real situation in the field. Elasticity analysis highlighted the crucial role of contact restrictions in reducing the spread of COVID-19, especially when combined with community awareness. This emphasized the analytics-driven nature of our approach. We transformed our model into an optimal control framework due to budget constraints. Leveraging Pontriagin's maximum principle, we meticulously formulated and solved our optimal control problem using the forward-backward sweep method. Our experiments underscored the pivotal role of vaccination in infection containment. Vaccination effectively reduces the risk of infection among vaccinated individuals, leading to a lower overall infection rate. However, combining vaccination and quarantine measures yields even more promising results than vaccination alone. A second crucial finding emphasized the need for early intervention during outbreaks rather than delayed responses. Early interventions significantly reduce the number of preventable infections, underscoring their importance.

    Citation: C. K. Mahadhika, Dipo Aldila. A deterministic transmission model for analytics-driven optimization of COVID-19 post-pandemic vaccination and quarantine strategies[J]. Mathematical Biosciences and Engineering, 2024, 21(4): 4956-4988. doi: 10.3934/mbe.2024219

    Related Papers:

    [1] Pannathon Kreabkhontho, Watchara Teparos, Thitiya Theparod . Potential for eliminating COVID-19 in Thailand through third-dose vaccination: A modeling approach. Mathematical Biosciences and Engineering, 2024, 21(8): 6807-6828. doi: 10.3934/mbe.2024298
    [2] Fang Wang, Lianying Cao, Xiaoji Song . Mathematical modeling of mutated COVID-19 transmission with quarantine, isolation and vaccination. Mathematical Biosciences and Engineering, 2022, 19(8): 8035-8056. doi: 10.3934/mbe.2022376
    [3] Shina D. Oloniiju, Olumuyiwa Otegbeye, Absalom E. Ezugwu . Investigating the impact of vaccination and non-pharmaceutical measures in curbing COVID-19 spread: A South Africa perspective. Mathematical Biosciences and Engineering, 2022, 19(1): 1058-1077. doi: 10.3934/mbe.2022049
    [4] Giulia Luebben, Gilberto González-Parra, Bishop Cervantes . Study of optimal vaccination strategies for early COVID-19 pandemic using an age-structured mathematical model: A case study of the USA. Mathematical Biosciences and Engineering, 2023, 20(6): 10828-10865. doi: 10.3934/mbe.2023481
    [5] Durward Cator, Qimin Huang, Anirban Mondal, Martial Ndeffo-Mbah, David Gurarie . Individual-based modeling of COVID-19 transmission in college communities. Mathematical Biosciences and Engineering, 2022, 19(12): 13861-13877. doi: 10.3934/mbe.2022646
    [6] Damilola Olabode, Jordan Culp, Allison Fisher, Angela Tower, Dylan Hull-Nye, Xueying Wang . Deterministic and stochastic models for the epidemic dynamics of COVID-19 in Wuhan, China. Mathematical Biosciences and Engineering, 2021, 18(1): 950-967. doi: 10.3934/mbe.2021050
    [7] 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
    [8] Akhil Kumar Srivastav, Pankaj Kumar Tiwari, Prashant K Srivastava, Mini Ghosh, Yun Kang . A mathematical model for the impacts of face mask, hospitalization and quarantine on the dynamics of COVID-19 in India: deterministic vs. stochastic. Mathematical Biosciences and Engineering, 2021, 18(1): 182-213. doi: 10.3934/mbe.2021010
    [9] Chenwei Song, Rui Xu, Ning Bai, Xiaohong Tian, Jiazhe Lin . Global dynamics and optimal control of a cholera transmission model with vaccination strategy and multiple pathways. Mathematical Biosciences and Engineering, 2020, 17(4): 4210-4224. doi: 10.3934/mbe.2020233
    [10] Maryam Al-Yahyai, Fatma Al-Musalhi, Ibrahim Elmojtaba, Nasser Al-Salti . Mathematical analysis of a COVID-19 model with different types of quarantine and isolation. Mathematical Biosciences and Engineering, 2023, 20(1): 1344-1375. doi: 10.3934/mbe.2023061
  • This study developed a deterministic transmission model for the coronavirus disease of 2019 (COVID-19), considering various factors such as vaccination, awareness, quarantine, and treatment resource limitations for infected individuals in quarantine facilities. The proposed model comprised five compartments: susceptible, vaccinated, quarantined, infected, and recovery. It also considered awareness and limited resources by using a saturated function. Dynamic analyses, including equilibrium points, control reproduction numbers, and bifurcation analyses, were conducted in this research, employing analytics to derive insights. Our results indicated the possibility of an endemic equilibrium even if the reproduction number for control was less than one. Using incidence data from West Java, Indonesia, we estimated our model parameter values to calibrate them with the real situation in the field. Elasticity analysis highlighted the crucial role of contact restrictions in reducing the spread of COVID-19, especially when combined with community awareness. This emphasized the analytics-driven nature of our approach. We transformed our model into an optimal control framework due to budget constraints. Leveraging Pontriagin's maximum principle, we meticulously formulated and solved our optimal control problem using the forward-backward sweep method. Our experiments underscored the pivotal role of vaccination in infection containment. Vaccination effectively reduces the risk of infection among vaccinated individuals, leading to a lower overall infection rate. However, combining vaccination and quarantine measures yields even more promising results than vaccination alone. A second crucial finding emphasized the need for early intervention during outbreaks rather than delayed responses. Early interventions significantly reduce the number of preventable infections, underscoring their importance.



    The coronavirus disease of 2019 (COVID-19) is a disease caused by the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) virus and was declared a pandemic by The World Health Organization (WHO) on March 11, 2020 [1]. Symptoms of COVID-19 include fever, cough, shortness of breath, fatigue, headache, nausea, vomiting, sore throat, and diarrhea. These symptoms typically appear about two to 14 days after the infection [2]. As of July 5, 2023, a total of 6,811,818 people in Indonesia have been reported infected with this virus, and 161,867 people have died from the infection [3]. Even though Indonesia has transitioned from the pandemic phase to the endemic phase of COVID-19 [4], it is important to note that the symptoms of COVID-19 can be exacerbated by the presence of comorbidities or underlying health conditions [5], and there are still reported cases of mortality due to COVID-19 [6]. Therefore, it remains crucial to continue practicing preventive measures against COVID-19.

    Since the advent of COVID-19, which began affecting various parts of the world in late 2019, numerous scientific approaches have been implemented to comprehend the spread and control of the virus. These approaches encompass medical and public health [7,8] and social [9,10], mathematical [11,12,13,14,15], and many other perspectives. In the field of mathematics, the approaches employed by the authors are quite diverse. These range from ordinary differential equations (ODEs) [12,13,14,15,16] to fractional differential equations [11,17,18], partial differential equations [19,20], stochastic models [21,22], and the utilization of big data analysis approaches [23,24].

    Mathematical modeling is a crucial tool for examining and understanding the dynamics of COVID-19 transmission, particularly in the context of specific interventions. The most popular intervention in the early spread of COVID-19 is the implementation of social distancing. Some authors have introduced their mathematical model to analyze the impact of social distancing on controlling COVID-19. Aldila et al. [25] developed a model to assess the impact of social distancing policies in the city of Jakarta, Indonesia. They found that the intervention can reduce the spread of COVID-19, but the outbreak of COVID-19 will arise soon when social distancing is relaxed. Another model introduced by Gevertz et al. [26] in their paper considers nonsocially distant individuals. They found that implementing relaxation policies at regular intervals could effectively reduce the peak infectious load. However, it is crucial to note that this schedule is highly sensitive to both parameter values and the frequency of the schedule. Recently, Arik et al. [27] used a fractional-order differential equation to analyze the impact of the environment and social distancing on the spread of COVID-19.

    Another intervention that many authors considered in their model is the vaccination strategy [11,13,14,15]. Saharan and Tee [28] used a simple susceptible-exposed-infectious-recovered (SEIR) model to analyze the effectiveness of a vaccine in controlling the spread of COVID-19 in Malaysia and Thailand. They used the Holt-Winters method to forecast their model dynamics. A fractional-order model was introduced by the authors in [29] to examine the impact of vaccines on the spread of COVID-19. A diffusion framework was used in their model to analyze the spatial spread of COVID-19. They found from their numerical experiments that it is imperative to implement strategic measures to ensure the timely and efficient distribution of vaccines. A mathematical model was introduced by the author in [30] to analyze the optimal allocation of COVID-19 vaccines in the Philippines. Using their model and data from the Philippines, they found that it is crucial to prioritize the primary vaccination series for the pediatric population (aged 5 to 11) within the first three months. Additionally, it is advisable to administer initial booster shots to individuals aged 12 and older during this timeframe. Please see [12,31,32,33] for more examples of the implementation of a mathematical model to assess the impact of vaccines on the spread of COVID-19.

    Other models aim to demonstrate the significance of quarantine in suppressing the spread of COVID-19 [34,35,36,37,38]. These models emphasize the complementary role of quarantine, alongside vaccination, in effectively mitigating the transmission of the virus. Authors in [34,35] have highlighted that while quarantine alone can reduce the spread of COVID-19 in the long-term, it cannot completely eradicate the virus. Nevertheless, the quarantine program can be combined with other initiatives, such as vaccination [36,37] and interventions to enhance population awareness [38].

    Based on this explanation, we can deduce that the long-term elimination of COVID-19 cannot be accomplished solely through vaccination or quarantine programs when executed in isolation from other factors. An essential element that can enhance the effectiveness of both vaccination and quarantine efforts is heightened public awareness. Consequently, certain models attempted to underscore the role of public awareness in the spread of COVID-19 [38,39,40,41]. Collectively, these models highlight that an increase in public awareness can effectively control the spread of COVID-19.

    Hospital care and capacity play a crucial role in controlling COVID-19. Adequate hospital resources, such as beds, ventilators, medical staff, and equipment, are essential for managing and treating severe cases of COVID-19. Hospitals provide care for individuals with more severe symptoms, and having sufficient capacity is crucial to ensuring that everyone who needs medical attention can receive it promptly. Adequate and timely medical treatment provided in hospitals plays a crucial role in reducing the impact and transmission of diseases, as evidenced in countries such as the USA, Brazil, India [42] and Indonesia [43]. The availability and proper utilization of resources, such as masks and ventilators, also significantly contribute to the recovery of COVID-19 patients [44]. Moreover, several models highlight that hospitalization, when combined with other interventions like vaccination, can positively contribute to COVID-19 recovery [12]. Some models also indicate limited resources in treatment using a saturated function, leading to the conclusion that treatment resources are an important factor in COVID-19 recovery [45,46,47,48,49].

    From the above explanation, it can be broadly seen that maximizing the intensity of vaccine administration, implementing quarantine measures, enforcing social restrictions, and improving the quality of hospitals will increase the possibility of eliminating COVID-19 from the community. However, this comes with the consequence of the high cost associated with the required interventions. Therefore, an optimal intervention is needed to minimize the spread of COVID-19 in the population. One approach that can be taken is to tailor COVID-19 interventions based on field needs—in this case, depending on time. One mathematical approach that can be used is to construct the problem as an optimal control problem. The basic idea of this approach is to treat the intervention u as a time-dependent variable u(t). The aim of this implementation is to minimize the number of infected individuals at the lowest possible cost for interventions. Many authors have introduced mathematical models with optimal control for COVID-19 interventions. The authors in [50] developed an optimal control model to analyze the impact of virus protection on the dynamics of COVID-19. The optimal control model for how COVID-19 spread, considering lockdown measures and asymptomatic cases, is discussed in [51]. They found that implementing a lockdown is essential in the early stages of an outbreak. On the other hand, combining vaccination and treatment is effective in controlling and preventing the spread of COVID-19. A more complex model, introduced by the authors in [52], accommodates the low-high-risk population and three types of interventions, namely isolation, detection, and treatment. Their parameter values were estimated using infection incidence data from Shanghai (March–April 2022). They found that when the budget is limited, implementing isolation and detection is more effective. On the other hand, if the budget is sufficient, implementing treatment alone is more effective compared to other possible scenarios. For more examples of optimal strategies for COVID-19 control, refer to [53,54].

    From the above explanation, we understand that although many references discuss the implementation of vaccination and quarantine as strategies to control COVID-19, there is still a lack of papers that consider the combined impact of population awareness on the effectiveness of these interventions. Therefore, the objective of this paper is to develop a deterministic model that incorporates vaccination and quarantine compartments while also considering the effects of saturation functions that represent awareness and resource limitations. Given that individuals who have contracted COVID-19 develop immunity to the disease for at least six months [55], this paper considers individuals with prior COVID-19 infections to be immune to the disease. An optimal control approach is also considered in our model to assess the optimal scenario for suppressing the spread of COVID-19 with a low cost of intervention through a combination of vaccination and quarantine.

    This paper is organized as follows: The first section discusses the state of the art of the research. In Section 2, we encompass the construction of the model, where we delve into the assumptions employed, the differential equation system, and the transmission diagram. We engage in a comprehensive discussion on model analysis, which includes positivity analysis and the parameterization of model parameters in Section 3. In Section 4, we construct the nondimensionalized version of our model to examine the COVID-19-free equilibrium points, the endemic equilibrium, local stability analysis, and bifurcation analysis. Section 5 discusses a series of numerical experiments incorporating elasticity analysis, sensitivity analysis, and autonomous simulation. We define our optimal control model in Section 6. Several numerical experiments for various scenarios are conducted in this section. Lastly, some discussion and conclusions are provided in Section 7.

    Before constructing the model, it is necessary to determine the assumptions or limitations of the model. In this paper, the following assumptions are made: We assume that the total population remains constant, there are no deaths caused by COVID-19, there is no waning immunity, and vaccination does not provide perfect protection against infection for vaccinated individuals. We assume that a vaccinated individual has better protection against contact with other individuals. Furthermore, we include medical interventions to accelerate the recovery rate of infected individuals. The human population is divided into five compartments based on their health status, namely susceptible (S), vaccinated (V), infected (I), quarantined (Q), and recovered (R).

    The model construction process is based on the transmission diagram shown in Figure 1. The modeling process is as follows: In our model, we incorporate the intervention of vaccination with a specific rate denoted as u1. Under this intervention, vaccinated individuals are segregated into a separate class, denoted as V. It is important to note that, based on many reports [12], COVID-19 vaccines do not offer complete protection against infection. Therefore, vaccinated individuals can still potentially contract COVID-19, albeit at a reduced rate, represented by the parameter β2. This rate is smaller than the infection rate for susceptible individuals in the S class (β1). In several reports [56,57,58], it has been found that individuals who have received vaccination tend to exhibit a heightened awareness toward the spread of COVID-19. This is attributed to their enhanced knowledge of the dangers of COVID-19, whether acquired through media, social interactions, or other means. An increase in the number of COVID-19 infections tends to heighten public awareness of the disease spread. As the number of infected individuals rises, there is often greater media coverage, discussions, and public discourse surrounding the disease. This increased attention can contribute to a heightened awareness among the general population about the importance of preventive measures, the severity of the situation, and the potential risks associated with the disease [59]. Hence, we consider that the probability of infection should be a monotonically decreasing function of I. Hence, the infection term within the vaccinated class is modeled by a saturated function expressed as β2VI1+b1I.

    Figure 1.  Transmission diagram of compartments.

    Additionally, we assume that individuals who become infected can be detected at a rate denoted as u2. Once detected, individuals receive medical treatment either in a hospital setting or are monitored from home, which we categorize as a quarantine intervention. As a result, their recovery rate increases. Drawing from various reports [60,61], it has been observed that the escalating number of COVID-19 infection cases during the pandemic may disrupt the quality of hospital services. This phenomenon is attributed, among other factors, such as hospital staff burnout and the presence of resource constraints, as well as challenges in adaptation and mitigation. Public health authorities and healthcare providers frequently collaborate to deploy strategic interventions aimed at effectively managing case surges and mitigating the adverse effects on the standard of healthcare. Based on these facts, we assumed that the recovery rate of quarantined individuals could become saturated due to the high number of individuals in quarantine in the hospital. Therefore, we model the recovery rate of quarantined individuals as a a saturated function γ0+γ11+b2Q. With this definition of the recovery function, we can see that the recovery rate of quarantined individuals reaches its maximum when Q is relatively small (Q0), i.e., γ0+γ1. On the other hand, when the number of infected individuals is approximately large, the recovery rate for quarantined individuals reaches its minimum level, i.e., γ0.

    Based on the assumptions outlined above, the mathematical model for the spread of COVID-19 is formulated as follows:

    dSdt=Λhu1Sβ1SIμS, (2.1a)
    dVdt=u1Sβ21+b1IVIμV, (2.1b)
    dIdt=β1SI+β21+b1IVIu2Iγ0IμI, (2.1c)
    dQdt=u2I(γ0+γ11+b2Q)QμQ, (2.1d)
    dRdt=γ0I+(γ0+γ11+b2Q)QμR, (2.1e)

    with the parameter description given in Table 1. Since the total population N is S+V+I+Q+R, then we have that the dynamic of total population is given by:

    dNdt=Λhμ(S+V+I+Q+R)=ΛhμN.
    Table 1.  Table of Parameter in Model (2.1).
    Par Interpretation Unit Interval Ref. Value Note
    N Total of human population in West Java people 4.9×107 [62] 4.9×106 Assumed
    Λh Natural birth rate peopleday [0,μN] Assumed 499400071.85365 Assumed
    u1 Rate of vaccination intervention 1day [0, 1] Assumed 0.0344 Estimated
    u2 Rate of quarantine intervention 1day [0, 1] Assumed 0.1540 Estimated
    β1 Contact between the susceptible population and the infected population 1peopleday [1×109, 9.7×107] [63,64] 1.4017107 Estimated
    μ Natural death rate 1day [168.79×365, 177.18×365] [65] 171.85365 Assumed
    β2 Contact between the vaccinated population and the infected population 1peopleday [1×109, 9.7×107] [63,64] 8.4101108 Estimated
    γ0 Natural recovery rate 1day [114,17] [64,66] 0.0751 Estimated
    γ1 Maximum additional recovery rate due to quarantine 1day [114,17] [64,66] 0.0476 Estimated
    b1 Saturation coefficient due to the number of infected individuals 1people [104,0.0179] [25,48] 0.0019 Estimated
    b2 Saturation coefficient due to the number of vaccinated individual 1people [104,0.0376] [25,48] 0.0224 Estimated

     | Show Table
    DownLoad: CSV

    If the total of population is approximately constant, then we can approximate the value of Λh by assuming dNdt=0, which lead us to Λh=μN.

    In this section, a nonnegativity analysis of the solution obtained from Model (1) will be performed.

    Theorem 1. Region Ω which is defined by:

    Ω={(S,V,I,Q,R)R5+|0<S,V,I,Q,R,S+I+V+Q+Rmax{N(0),Λhμ}},

    is positively invariant with respect to Model (2.1).

    Proof. Model (2.1) can be written as follows:

    dXdt=CX+D,

    where:

    X=[S,V,I,Q,R]T,D=[Λh,0,0,0,0]T,andC=[d10000u1d2000β1Id3d40000u2d5000γ0d6μ],

    while d1=u1μβ1I, d2=β11+b1IIμ, d3=β21+b1II, d4=u2μγ0, d5=(γ0+γ11+b2Q)μ, and d6=γ0+γ11+b2Q. Due to the nonnegative entries outside the main diagonal, matrix C is classified as a Metzler matrix. Since D0, then Model (2.1) possesses a positive invariant in R5+ [67]. Hence, all trajectories of solution of Model (2.1), which originate from an initial state in R5+, confine forever.

    Summing up all equations in Model (2.1), we have:

    dNdt=ΛhμN.

    Then the solution is N(t)=Λhμ+(N(0)Λhμ)eμt. If N(0)=Λhμ, then we have N(t)=Λhμ for all time t. If N(0)<Λhμ, then N(t) will be monotonically increasing and tends to Λhμ for t. On the other hand, if N(0)>Λhμ, then N(t) will be monotonically decreasing and tends to Λhμ for t. Therefore, all feasible solutions of the Model (2.1) enter the region Ω, implying that the region is an attracting set.

    To estimate the parameters for Model (2.1), a data fitting process was conducted for the number of COVID-19 infections in West Java, Indonesia, from July 1, 2022, to October 1, 2022. In the context of our model, this data represents the dynamics of I(t). Our aim is to fit the dynamic output of I(t) from our model with the available data. Mathematically, this task can be expressed as minimizing the Euclidean distance between COVID-19 incidence data (denoted by Idata) and the solution of I(t) from Model (2.1), using the best-fit parameters β1,β2,u1,u2,b1,b2,γ0,γ1, and the best initial conditions S(0),V(0),I(0),Q(0),R(0). Hence, we define the following objective function:

    C=T0(IdataI(t))2dt, (3.1)

    where T is the maximum number of existing data points. Our aim is to find the optimal parameters:

    Ψ={β1,β2,u1,u2,b1,b2,γ0,γ1,S(0),V(0),I(0),Q(0),R(0)},

    such that:

    C(Ψ)=minΘC(Ψ), (3.2)

    where Θ is the admissible value for the estimated parameters (fourth column of Table 1). The following constraints need to be mentioned:

    ● The infection rate of vaccinated individuals (β2) is smaller than that of susceptible individuals (β1). Hence, we have β2<β1.

    ● Vaccination and quarantine is a non-stop intervention by the government. Hence, u1>0,u2>0.

    ● For biological interpretation purposes, we have γ0,γ1,b1,b2, and all initial conditions are positive.

    We approach our parameter estimation challenge as the task of minimizing a constrained, nonlinear, multivariable function. Specifically, our objective is to minimize C subject to the constraints imposed by the ODE model (2.1), while ensuring that the Model parameters and initial conditions adhere to biologically reasonable values. We use the fmincon toolbox in MATLAB to solve our problem. The result is shown in Figure 2. The best-fit parameter is provided in Table 1 (sixth column), while the best fit initial condition is as follows:

    S(0)=3460000,V(0)=1870000,I(0)=465,Q(0)=40,R(0)=49.
    Figure 2.  Parameter estimation result using the best-fit parameters. The black dot represents the incidence data (Idata), the blue dot represents the model results for compartment I, and the red dot represents the model forecasting result (I(t)) until t=100.

    For more parameter estimation methods, readers may read [68,69,70,71] for further references.

    From the results of parameter estimation in Table 1, we can see that the control reproduction number is still greater than one, which indicates that the number of infected individuals (I(t)) will continue to exist as t (see Section 4 for further analytical results on the COVID-19 endemic equilibrium point). We found that u1=0.0344 and u2=0.1540, which resulted in the findings shown in Figure 2. If the control interventions were stopped (u1=u2=0), then we would obtain R0=9.316. Any relaxation of the control variables will increase Rc. On the other hand, if we increase the intensity of vaccination and quarantine, then Rc can be reduced. Figure 3 shows the impact of relaxation/improvement of vaccination and quarantine interventions on the dynamics of infected individuals I.

    Figure 3.  Forecasting results on the dynamics of infected individuals with vaccination and quarantine interventions, showing (a) relaxed measures and (b) improved measures.

    Let us assume the total population is constant. Hence we have:

    x1=SN,x2=VN,x3=IN,x4=QN,x5=RN,

    which represents the proportion of each class to the total human population. Assuming τ=tγ0, m=μγ0,a1=β1Nγ0,a2=β2γ0,g1=γ1γ0,c1=b1N,c2=b2N,e1=u1γ0,e2=u2γ0, and reducing the susceptible compartments, a nondimensionalized version of the Model (2.1) is obtained as follows:

    dx1dτ=me1x1a1x1x3mx1,dx2dτ=e1x1a21+c1x3x2x3mx3,dx3dτ=a1x1x3+a21+c1x3x2x3e2x3x3mx3,dx4dτ=e2x3(1+g11+c2x4)x4mx4, (4.1)

    where x5=1x1x2x3x4.

    In this section, we initiate our analysis by determining the form of the COVID-19-free equilibrium, followed by the calculation of the control reproduction number, denoted as Rc. Reproduction numbers have been utilized by various authors in epidemiological models [11,49] to assess the dynamic behavior of the proposed model, including the existence and stability of equilibrium points.

    Theorem 2. The COVID-19-free equilibrium of system (4.1) is given by:

    E0=(x1,x2,x3,x4)=(mm+e1,e1m+e1,0,0),

    which always exists unconditionally.

    Proof. The equilibrium points associated with Model (1) are obtained when the population changes in each compartment remain constant, namely, when dx1dτ=0,dx2dτ=0,dx3dτ=0, and dx4dτ=0. As the COVID-19-free equilibrium point represents a condition without COVID-19 in the population, we have x3=0. By direct calculation, substituting x3=0 into dx1dt,dx2dt, and dx4dt, we obtain x1=mm+e1, x2=e1m+e1, and x4=0. From the expressions of x1,x2,x3, and x4, it is clear that they always exist in R40 without any existence criteria. Hence, the proof is complete.

    From the direct consequences of the expression of E0, we have the following corollary.

    Corollary 1. Without the intervention of vaccination, the COVID-19-free equilibrium of system (4.1) is given by:

    Corollary 1. Without the intervention of vaccination, the COVID-19-free equilibrium of system (4.1) is given by:

    E0=(x1,x2,x3,x4)=(1,0,0,0).

    Theorem 4.2. The control reproduction number from system (4.1) is Rc=a1m+a2e1(e2+1+m)(m+e1).

    Proof. We use a next-generation matrix approach [72] to determine the control reproduction number of the system (4.1). At first, we determine the Jacobian matrix of the subsystem of (4.1) that only involves the infected compartments x3 and x4. The Jacobian matrix evaluated at E0 is given by:

    J=[a1mm+e1+a2e1(m+e1)e21m0e21g1m]. (4.2)

    The matrix J can be written as the sum of the transmission matrix T and the transition matrix Σ, where:

    T=[a1mm+e1+a2e1(m+e1)000]andΣ=[1me20e21g1m].

    Since the second row of matrix T consists entirely of zeros, the next-generation matrix can be derived as:

    NGM=ETΣ1E=[a1m+a2e1(e2+1+m)(m+e1)],

    where E=[10] and E is the transpose of E. Since the spectral radius of the matrix above is given by a1m+a2e1(e2+1+m)(m+e1), we can conclude that Rc=a1m+a2e1(e2+1+m)(m+e1). Hence, the proof is complete.

    Theorem 4. Let Rc be Rc that satisfies k214k0k2=0 and K=(e2+1+m)2(m+e1)a2e1(e2+1+m)+a2e1(mc1+a2), then, system (4.1) has:

    1) One endemic equilibrium if Rc>1,

    2) No endemic equilibrium for Rc<1 if K0,

    3) No endemic equilibrium for Rc<Rc<1 if K<0,

    4) Two endemic equilibrium for RcRc<1 if K<0.

    Proof. The endemic equilibrium represents a condition in which the disease persists within the system [73], specifically at the point (x1,x2,x3,x4) with x30. As there is no change in the system when the endemic equilibrium point is reached, we set dx1dτ=0,dx2dτ=0,dx3dτ=0, and dx4dτ=0 with x30. By direct calculation, we can find:

    E1=(x1,x2,x3,x4), (4.3)

    where x1=ma1x3+m+e1,x2=e1m(c1x3+1)(a1x3+m+e1)(mc1x3+m+a2x3), x4 is the positive root of e2x3(1+g1c1x4+1)x4mx4=0, and x3* is taken from the positive roots of the following polynomial:

    f(x3)=k2x32+k1x3+k0=0, (4.4)

    where:

    k2=(e2+1+m)(mc1+a2)(Rc(e2+1+m)(m+e1)a2e1)m,k1=((e2+1+m)(c1(Rc1)m2Rc(e2+1+m)2(m+e1)+(c1e1+a2)(Rc1)m+Rca2e1)a2e1(mc1+a2)),k0=m(e2+1+m)(m+e1)(Rc1).

    To analyze the existence of the endemic equilibrium, it is necessary to guarantee the existence of a positive root in the polynomial (4.4). If the roots are positive, then x1,x2, and x4 will also be positive. From Descartes' rule of signs, it can be concluded that when Rc>1, the system will have a unique endemic equilibrium point. Furthermore, for k0<0, it is possible for the system to have zero or two positive roots. Hence, further analysis is needed. Gradient analysis is conducted at Rc=1,x3=0. Using implicit derivative on polynomial (4.4), we have:

    f(x3)Rc|Rc=1,x3=0=m(e2+1+m)(m+e1)(e2+1+m)2(m+e1)(e2+1+m)a2e1+[a2e1(mc1+a2)]. (4.5)

    Hence, if K=(e2+1+m)2(m+e1)(e2+1+m)a2e1+[a2e1(mc1+a2)]<0, then we will have a negative gradient of x3 at Rc=1,x3=0. This indicates the existence of positive roots x3 for some intervals when Rc<1. Since f(x3) is a second-degree polynomial, then f(x3) will have a turning point at Rc=Rc, where Rc is the value of Rc that satisfies the discriminant of f(x3) equal to zero, which is k214k0k2=0. Hence, we have proved case 3 of the theorem. If Rc<Rc, then we will have no positive root (case 4 of the theorem proved). Lastly, if K0, then we will have a positive gradient, which confirms case 2 in the proven theorem. Furthermore, we use Descartes' rule of signs to determine the possible number of positive roots of the polynomial f(x3). The results are shown in Table 2. It can be observed that f(x3) may have 0, 1, or 2 positive roots. Additionally, it is possible to have an endemic equilibrium even when R0<1.

    Table 2.  The maximum number of roots for x3 can be determined using Descartes' rule of signs for the equation in (4.4).
    Case k2 k1 k0 Rc Change of Sign Number of Possible Roots
    1 - + - Rc<1 0 0 or 2
    2 - + + Rc>1 1 1
    3 - - - Rc<1 2 0
    4 - - + Rc>1 1 1

     | Show Table
    DownLoad: CSV

    In the previous section, we have shown that the COVID-19-free equilibrium always exists, while the endemic equilibrium exists depending on the conditions of Rc and K. In this section, we show how the stability of the COVID-19-free equilibrium can only be achieved if Rc<1.

    Theorem 5. For e10 and e20, the point E0 in the system (1) is locally asymptotically stable if Rc<1 and unstable if Rc>1.

    Proof. To demonstrate the stability of the COVID-19-free equilibrium point, we linearize system (4.1) around the COVID-19-free equilibrium point [73]. The Jacobian matrix is formulated as follows:

    Jac|DFE=[me10a1mm+e10e1ma2e1(m+e1)000a1mm+e11+a2e1(m+e1)e2m000e2(m1)g1]. (4.6)

    The eigenvalues of the matrix above can be obtained as follows:

    λ1=m,λ2=(m+e1),λ3=(m+1+g1),λ4=m2ma1+me1+me2+e1e2+m+e1a2e1(m+e1),=(Rc1)(e2+1+m).

    E0 is locally asymptotically stable if all eigenvalues of matrix (4.6) have negative values. Since λ1 to λ3 are already negative and λ4<0 for Rc<1, E0 is locally asymptotically stable for Rc<1. Moreover, E0 is unstable if any eigenvalue of matrix (4.6) is positive. Therefore, E0 is unstable for Rc>1.

    Here, we analyze the local stability of the COVID-19 endemic equilibrium point near the condition Rc=1 using the Castillo-Song bifurcation theorem [74].

    Theorem 6. The COVID-19 model in system (4.1) undergoes backward bifurcation at Rc=1 if K<0, and forward bifurcation when K>0.

    Proof. Using the Castillo-Song theorem [74], we examine bifurcation from system (2.1), which has already been nondimensionalized to become system (4.1). System (4.1) can be represented as follows:

    f1=me1x1a1x1x3mx1,f2=e1x1a21+c1x3x2x3mx3,f3=a1x1x3+a21+c1x3x2x3e2x3x3mx3,f4=e2x3(1+g11+c2x4)x4mx4.

    For Rc=1, we take the bifurcation parameter a1=a1 where a1 represents the contact rate β1. We have:

    a1=(e2+1+m)(m+e1)a2e1m.

    Furthermore, the linearization matrix around the COVID-19-free equilibrium point for a1=a1 is obtained as follows:

    J=[me10(e2+1+m)(m+e1)+a2e1(m+e1)0e1ma2e1(m+e1)0000000e2(m1)g1].

    Subsequently, we computed the right and left eigenvectors of the zero eigenvalues of A, denoted as w=(w1,w2,w3,w4)T and v=(v1,v2,v3,v4), respectively, yielding the following expression:

    w1=((m+1)+g1)((e2+1+m)(m+e1)a2e1)e2(m+e1)2,w2=((e2+1+m)(m+e1)+ma2)((m+1)+g1)e1e2(m+e1)2m,w3=m+1+g1e2,w4=w4,v1=0,v2=0, v3=v3,v4=0.

    Using the theorem [74] and computing the nonzero partial derivatives of f(x) at the disease-free equilibrium, the associated bifurcation coefficients are defined as follows:

    A=4k, i, j = 1vkwiwj2fkxixj(0,0),=2((e2+1+m)2(m+e1)a2e1(e2+1+m)+a2e1(mc1+a2))((m+1)+g1)2(m+e1)me22,=2K((m+1)+g1)2(m+e1)me22, (4.7)
    B=4k, i = 1vkwi2fkxia1(0,0),=m(m+1+g1)(m+e1)e2. (4.8)

    Since B>0, according to [74], backward bifurcation occurs when A>0K<0, and forward bifurcation occurs when A<0K>0. Hence, the theorem is proved.

    In this section, we present the results of numerical experiments conducted on model (2.1) regarding the elasticity analysis, bifurcation diagram, and autonomous simulations. The control reproduction number of the original Model in (2.1) is given by:

    Rc=Λ(μβ1+β2u1)(μ+u1)μ(u2+γ0+μ).

    Hence, without any intervention, the basic reproduction number of the original model in (2.1) is given by:

    R0=Λβ1μ(μ+γ0).

    All model parameters that have been used in this section are given in Table 1, which shows that Rc=1.827>1. Without any improvement in the control variables, COVID-19 will continue to exist in West Java.

    At this stage, an analysis will be carried out to examine the sensitivity of the control reproduction number (Rc) to various parameters that affect its value, as well as the relationship between these parameters. By considering different values for these parameters, we can observe how the control reproduction number of the system (2.1) varies while keeping the values of the other parameters constant. To evaluate the sensitivity, we employ the forward sensitivity index of Rc, which is denoted by:

    EσRc=Rcσ×σRc, (5.1)

    where σ is a parameter in the system (2.1).

    We calculate the elasticity indices of Rc with respect to all parameter values to observe the effects of changes in parameters on Rc. Using the formula and parameter values in Table 1, we obtained the following results:

    EμRc=2μ3β1((β1+3β2)u1+β1(u2+γ0))μ22u1β2(u1+u2+γ0)μu12β2(u2+γ0)(μβ1+β2u1)(μ+u1)(u2+γ0+μ)=0.99,Eβ1Rc=μβ1μβ1+β2u1=0.0018,Eβ2Rc=β2u1μβ1+β2u1=0.998,Eu2Rc=u2u2+γ0+μ=0.673,Eγ0Rc=γ0u2+γ0+μ=0.326,Eγ1Rc=0,Eu1Rc=u1(β1β2)μ(μβ1+β2u1)(μ+u1)=0.00073,Eb1Rc=0,Eb2Rc=0.

    To analyze the impact of changes in vaccination intensity and quarantine measures, we recalculate the elasticity index by adjusting the values of u1 and u2. The parameters used are the same as those shown in Figure 3. It is evident that reducing the intensity of vaccination and quarantine increases the elasticity of vaccine intervention while decreasing the elasticity of quarantine. Conversely, when we increase the intensity of vaccination and quarantine, the effectiveness of vaccination decreases while the elasticity of quarantine increases. Additionally, we observe that the quarantine intervention is more elastic in affecting R0 compared to vaccination. Another noteworthy point from Table 3 is that the infection rate of vaccinated individuals is much more sensitive compared to that of non-vaccinated individuals. Hence, it is essential to develop a high-quality vaccine that can provide maximum protection against the COVID-19 infection.

    Table 3.  Changes of elasticity index of Rc with respect to u1,u2,γ0,β1, and β2 when u1 and u2 are relaxed (reduced) and when it is improved (increased).
    Parameter u1 and u2 relaxed u1 and u2 improved
    25% 50% 75% 85% 50% 100% 150% 200%
    u1 -0.0007 -0.0015 -0.003 -0.004 -0.00049 -0.00036 -0.00029 -0.00024
    u2 -0.607 -0.507 -0.34 -0.236 -0.755 -0.804 -0.837 -0.861
    γ0 -0.392 -0.492 -0.659 -0.763 -0.244 -0.195 -0.162 -0.139
    β1 0.0024 0.0036 0.007 0.012 0.001 0.00092 0.00073 0.0006
    β2 0.997 0.996 0.992 0.987 0.998 0.999 0.999 0.999

     | Show Table
    DownLoad: CSV

    In this section, we draw the bifurcation diagram using the best-fit parameter values obtained from the previous section. The result is presented in Figure 4, where the forward bifurcation occurs at Rc=1. The blue and red curves represent the equilibrium states of COVID-19-free and endemic conditions, respectively. The solid and dotted curves represent the stable and unstable equilibrium points. BP denotes the branching point at Rc=1. It is clear to see that when Rc<1, the COVID-19-free equilibrium is stable, but it becomes unstable when it surpasses the branching point BP. At the same time, at the branching point BP, the endemic equilibrium starts to emerge as Rc increases.

    Figure 4.  Forward bifurcation of system (4.1). Pi for i=1,2,3,4 is the sample point of Rc at 0.5, 3.7, 6 and 8, respectively. BP is the branching point at Rc=1.

    To illustrate the dynamics of the Model (2.1), we use four sample points in Figure 4, denoted as P1,P2,P3 and P4. We observe that P1 lies in the region where Rc<1, and the autonomous simulation is shown in Figure 5, panel (a), for the original model in the Model (2.1). It is evident that the system tends toward a stable COVID-19-free equilibrium. On the other hand, for sample points P2,P3 and P4, when Rc>1, the system tends toward an endemic equilibrium. The numerical simulation results are given in panels (b)–(d), respectively.

    Figure 5.  The impact of Rc(β1,β2) on the dynamics of each compartment from Model (2.1). Panels (a)–(d) represent the sample points P1 to P4, as shown in Figure 4. Specific parameter values used in the simulation are given in Table 4.
    Table 4.  Numerical results on the stable equilibrium points of Model (2.1) depend on Rc.
    Point β1 β2 Rc S V I Q R
    P1 2.79×108 1.11×108 0.5 5197 4.98×106 0 0 0
    P2 1.4×107 8.4×108 3.76 5194 3.76×106 203 391 1.22×106
    P3 6.87×107 2.75×107 6 5144 1.67×106 550 1101 3.3×106
    P4 9.16×107 3.66×107 8 5119 1.32×106 610 1223 3.6×106

     | Show Table
    DownLoad: CSV

    Model (2.1) considers two control variables: u1, representing vaccination intervention, and u2, representing quarantine intervention. Trivially, from our sensitivity analysis in the previous section, we understand that larger values of u1 and u2 will yield better results for eradicating COVID-19. However, it will come at a high cost. Hence, our main objective is to minimize the number of infected individuals, I and Q, with the lowest possible cost of u1 and u2. To address this problem, we constructed a cost function as follows:

    J=T0(ω1I+ω2Q+φ1u21+φ2u22)dt, (6.1)

    which represents the total cost due to the high number of infected individuals and the level of control intensity. Note that ω1 and ω2 represent the weight parameters for I and Q, respectively. On the other hand, φ1 and φ2 represent the weight costs for u1 and u2. Next, by applying Pontryagin's maximum principle [75], we obtain the Hamiltonian function as follows:

    H=ω1I+ω2Q+φ1u21+φ2u22+λ1dSdt+λ2dVdt+λ3dIdt+λ4dQdt+λ5dRdt, (6.2)

    where λ1 for i=1,2,3,4,5 represents the adjoint variables for S,V,I,Q, and R, respectively. Thus, taking the partial derivatives of H with respect to each of the state variables yields the adjoint system, as shown below:

    dλ1dt=HS=β1Iλ3(β1Iμu1)λ1u1λ2,=β1I(λ2λ3)+u1(λ1λ2)+μλ1,dλ2dt=HV=β2Inλ3Ib1+1(β2InIb1+1μ)λ2,=β2I1+b1I(λ2λ3)+μλ2,dλ3dt=HI=ω1(β1Sβ2VIb1(Ib1+1)2+β2VIb1+1u2γ0μ)λ3u2λ4λ5γ0+β1Sλ1(β2VIb1(Ib1+1)2β2VIb1+1)λ2,=ω1+β1S(λ1λ3)+β2V1+b1I(1b1I1+b1I)(λ2λ3)++u2(λ3λ4)+γ0(λ3λ5)+μλ3,dλ4dt=HQ=ω2(γ1b2Q(b2Q+1)2γ0γ1b2Q+1μ)λ4(γ1b2Q(b2Q+1)2+γ0+γ1b2Q+1)λ5,=ω2+γ11+b2Q(b2Q1+b2Q1)(λ5λ4)+γ0(λ4λ5)+μλ4,dλ5dt=HR=μλ5, (6.3)

    completed with the transversality condition λi(t=T)=0 for i=1,2,3,4,5. Furthermore, the optimal controls (u1,u2) are given by:

    u1=min{max{umin1,S(λ1λ2)2φ1},umax1},u2=min{max{umin2,I(λ3λ4)2φ2},umax2}, (6.4)

    where umini and umaxi, for i=1,2, represent the lower and upper bounds of u1 and u2, respectively.

    The solution of our optimal control problem in Eq (6.4) through analytical methods is unattainable due to the high dimensionality and nonlinearity of the ordinary differential equations system for the state variables in Model (2.1) and the adjoint variable in (6.3). Consequently, we employ a forward-backward sweep method to solve our problem numerically [73]. Additional examples demonstrating the implementation of this method can be found in [76,77,78,79]. The initial step involves providing an initial guess for the control variables u1 and u2 for all t[0,T]. By combining this guess with an initial condition for the state variables, we solve Model (2.1) in the time interval t[0,T] forward in time. Subsequently, utilizing the obtained results, we solve the adjoint system in (6.3) backward in time, subject to the transversality condition λi(T)=0. With the solutions for the state and adjoint variables in hand, we update the control variables using the formula in (6.4). This iterative process is repeated until the convergence criteria is met, with convergence defined as |Jiteration-k+1Jiteration-k|105.

    We have divided the numerical experiments in this section into three different categories for three different purposes. The first simulation is conducted to determine the most cost-effective scenario based on the combination of controls. The second simulation aims to assess the impact of the initial population condition on the dynamics of control. The last simulation is conducted to evaluate the influence of the initial basic reproduction number on the effectiveness of the COVID-19 control strategy.

    In this section, we conducted a numerical experiment to address our optimal control problem, exploring various combinations of control interventions. The first scenario involves the simultaneous implementation of vaccination and quarantine measures to contain the spread of COVID-19. The outcomes are presented in Figure 6. It is evident that when both of these interventions are employed concurrently, there is a notable reduction in the number of infected individuals. The dynamics of vaccination control necessitate an initial high rate of application during the early stages of the simulation, followed by a gradual decrease over time. Conversely, the quarantine intervention exhibits a decreasing trend in response to a heightened number of infected individuals being placed under quarantine. The total cost of intervention for this scenario is 3.51×106.

    Figure 6.  Optimal control results when vaccination and quarantine are implemented together. Panels (a)–(f) represent the dynamics of S, V, I, Q, I+Q and control variables, respectively. The total infected averted with this scenario is 4.014×107, with a total cost for intervention of 3.47×106.

    The subsequent simulation aims to elucidate the population's response when vaccination is implemented as the sole intervention. The results are presented in Figure 7. It is evident that without quarantine measures, vaccination intervention necessitates a more intensive application, as indicated in panel (f). Consequently, the vaccinated population approaches nearly the entire total population, while the number of infected individuals, although exhibiting an increase, remains significantly lower in comparison to scenarios without any control measures. The overall cost of intervention in this scenario is only marginally higher than that of the first scenario, where vaccination and quarantine are implemented concurrently.

    Figure 7.  Optimal control results when vaccination and quarantine are implemented together. Panels (a)–(f) represent the dynamics of S, V, I, Q, I+Q and control variables, respectively. The total infected averted with this scenario is 4.02×107, with a total cost for intervention of 3.51×106.

    The final simulation was conducted to assess the impact of quarantine intervention when employed as the sole measure to mitigate COVID-19. The results are depicted in Figure 8. It is evident that without vaccination intervention, quarantine measures do not have a significant effect on curtailing the number of infected individuals, unlike the two preceding scenarios. To counterbalance the high number of infections, intensive vaccination interventions must be administered in the initial stages (refer to panel (f)). However, since our model assumes that the recovery of quarantined individuals follows a saturation function, it becomes increasingly challenging to increase the recovery rate as more individuals are placed under quarantine. This is because there are a high number of individuals who require treatment. Therefore, this intervention must be scaled back over time to minimize intervention costs. Consequently, quarantine interventions do not yield significant results in reducing infection numbers, as observed in the two previous scenarios.

    Figure 8.  Optimal control results when vaccination and quarantine are implemented together. Panels (a)–(f) represent the dynamics of S, V, I, Q, I+Q and control variables, respectively. The total infected averted with this scenario is 3.12×105, with a total cost for intervention of 3.51×106.

    In contrast to the simulations in the previous subsection, which focused on intervention combinations, this subsection focuses on scenarios driven by distinct initial conditions. These simulations are conducted under the assumption of very low initial infection counts, and this scenario can be termed an endemic prevention scenario, where intervention is implemented before rapid COVID-19 transmission occurs. The initial condition is given by:

    S(0)=5330000,V(0)=0,I(0)=100,Q(0)=0,R(0)=0.

    The simulation outcomes are illustrated in Figure 9. It is evident that when the number of infected individuals at the initial time is extremely low, the intervention administered is also relatively modest (refer to panel (f) in the blue curve). Vaccination intervention begins to escalate as the number of infected individuals increases (see panels (c)–(e) highlighted in red). On the other hand, the intensity of quarantine intervention diminishes over time, reflecting the reduced effectiveness of treatment intervention for individuals who are quarantined. This is due to the increasing number of individuals who are subject to quarantine. The total cost incurred for this scenario is 5.33×106, which is greater than when the initial condition involved a higher number of infected individuals. However, the number of new infections successfully averted is substantially larger, amounting to 6.51×107.

    Figure 9.  Optimal control results when vaccination and quarantine are implemented together but when COVID-19 has just started to spread (indicated by the small number of infected individuals at t=0). Panels (a)–(f) represent the dynamics of S, V, I, Q, I+Q and control variables, respectively. The total infected averted with this scenario is 6.51×107, with a total cost for intervention of 5.33×106.

    The final simulation in this chapter was conducted to assess the influence of the endemic level, as indicated by the magnitude of Rc in the population, on the intensity of control measures required to manage COVID-19. To achieve this, the parameters used were consistent with those in Table 1, except for β1 and β2, which were reduced by 50%. As a result, we obtained Rc=4.658, which is 50% smaller than the estimate from the previous chapter. The results are displayed in Figure 10. Due to the lower value of Rc, the intensity of vaccination control shown in Figure 10 (panel (f)) does not need to be as high as indicated in Figure 6 (panel (f)). Nonetheless, interventions of this nature have proven effective in significantly reducing the number of infected individuals (see panel (e)). The total cost incurred in this scenario is 9.43×106, which is substantially smaller than the scenario in Figure 6. The number of infections successfully averted amounts to 3.4×106.

    Figure 10.  Optimal control results when the initial basic reproduction number is smaller, i.e., Rc=4.658, which is 50% smaller than in scenario one in Section 6.2.1. Panels (a)–(f) represent the dynamics of S, V, I, Q, I+Q and control variables, respectively. The total infected averted with this scenario is 9.43×106, with a total cost for intervention of 3.4×106.

    In this paper, we delved into an analysis of the dynamic characteristics of the COVID-19 transmission model, specifically the susceptible-vaccinated-infected-quarantine-recovered (SVIQR) model. Our investigation considers critical factors such as human awareness and resource constraints for treating quarantined individuals. To estimate model parameters, we utilized incidence data sourced from West Java, Indonesia. Our exploration encompassed mathematical analyses of equilibrium points, their local stability, and the control reproduction number (Rc), all of which are presented analytically.

    Our findings revealed a significant insight: The model may display forward or backward bifurcations at Rc=1. This suggests that achieving a reproductive number below one alone is not always sufficient to ensure the complete eradication of COVID-19 from a population. This result is consistent with another COVID-19 model introduced in [33,76]. In practice, backward bifurcation commonly occurs due to several possible factors, such as reinfection [80], relapse, imperfect vaccination [81,82], or limited treatment capacity [83]. In the context of our study on COVID-19, the existence of the backward bifurcation suggests the potential for decision-making challenges, as the expectation that COVID-19 will always disappear when Rc<1 is not consistently observed in real-world scenarios. The implications of this backward bifurcation highlight the possibility of a significant shift from a stable COVID-19-free state to a large endemic state, emphasizing the complex dynamics of disease control in public health. Another important result of our research is the sensitivity analysis. Sensitivity analysis underscores the potential effectiveness of minimizing contact between infected individuals who are not in quarantine and vaccinated individuals as a means of controlling the disease's spread.

    To address budget limitations, we reconstructed our model as an optimal control model, introducing vaccination and quarantine interventions as time-dependent parameters. By employing Pontriagin's maximum principle, we rigorously formulated our optimal control problem and utilized the forward-backward sweep method for numerical solutions. Our simulations, conducted across various scenarios that may manifest in real-world situations, yielded noteworthy insights.

    To begin, our numerical experiments underscore the paramount importance of vaccination as the primary intervention in curtailing infections. Vaccination has been proven effective in reducing the likelihood of vaccinated individuals contracting the virus, thereby minimizing the number of infected individuals. However, combining vaccination with quarantine measures still yields more promising results than relying solely on vaccination. A second critical finding emphasizes the necessity of implementing interventions at the onset of an outbreak rather than waiting for it to escalate. This early intervention strategy leads to a significantly higher number of preventable infections. {The results of our optimal control simulation study underscore the importance of considering vaccination and quarantine strategies in controlling the spread of COVID-19. Both interventions play distinct roles: Vaccination is highly effective in preventing the initial spread of COVID-19 during the early stages of the pandemic, while quarantine measures are best focused as the pandemic begins to escalate. The combination of both strategies is essential for effectively mitigating the outbreak of COVID-19. Implementing these strategies based on needs will yield optimal results, resulting in a reduced budget for intervention and preventing a larger number of COVID-19 infections.

    While our study sheds considerable light on the significance of the COVID-19 vaccination, there are several possible future directions in COVID-19 research from this article and other references from other disciplines. In many instances, there exists a gap between genetics, environmental science, and epidemiology in the context of COVID-19. Potential novel insights lie in understanding the determinants of COVID-19 susceptibility and informing targeted interventions and public health strategies to mitigate the impact of the pandemic. Readers can explore a mathematical approach for this scenario in [84]. Another promising avenue in COVID-19 modeling involves bridging rumor spreading models with COVID-19 dynamics, leveraging insights from the Maki-Thompson model [85]. Readers can adapt the Maki-Thompson model to incorporate elements relevant to COVID-19 transmission dynamics, such as contact rates, transmission probabilities, and population demographics. Furthermore, readers can delve into research directions by exploring the concept of strategic decision-making and resource allocation in the context of pandemic response, utilizing a stochastic control approach [86]. Through this approach, we can gain a better understanding of the effectiveness of strategies for managing public health crises and ultimately contribute to improved preparedness and resilience in the face of future pandemics. In conclusion, by embracing interdisciplinary approaches and leveraging mathematical modeling techniques, we can pave the way for innovative solutions to combat COVID-19 and enhance our preparedness for future pandemics.

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

    This research is funded by Universitas Indonesia through the PUTI Research Grant Scheme, 2023 (ID Number: NKB-469/UN2.RST/HKP.05.00/2023).

    The authors declare there is no conflict of interest.



    [1] The World Health Organization (WHO), Coronavirus disease (covid-19) pandemic, 2023. Available from: https://www.who.int/europe/emergencies/situations/covid-19.
    [2] Centers for Disease Control and Prevention, Symptoms of covid-19, 2023. Available from: https://www.cdc.gov/coronavirus/2019-ncov/symptoms-testing/symptoms.html.
    [3] The World Health Organization (WHO), Indonesia situation of covid-19, 2023. Available from: https://covid19.who.int/region/searo/country/id.
    [4] Ministry of State Apparatus Utilization and Bureaucratic Reform, Indonesia (KEMENPANRI), Indonesia telah bergerak menuju endemi covid-199, 2023. Available from: https://www.menpan.go.id/site/berita-terkini/berita-daerah/indonesia-telah-bergerak-menuju-endemi-covid-19.
    [5] Communication Team of the National Committee for Handling Corona Virus Disease 2019 (Covid-19) and National Economic Recovery, Indonesia, Waspadai komorbid, salah satu faktor risiko yang memperparah gejala covid-19, 2022. Available from: https://covid19.go.id/artikel/2022/02/15/waspadai-komorbid-salah-satu-faktor-risiko-yang-memperparah-gejala-covid-19.
    [6] The Ministry of Health Republic Indonesia (KEMENKES RI), Covid 19 update, 2023. Available from: https://www.mayoclinic.org/diseases conditions/coronavirus/in-depth/herd-immunity-and coronavirus.
    [7] K. Cardwell, B. Clyne, N. Broderick, B. Tyner, G. Masukume, L. Larkin, et al., Lessons learnt from the covid-19 pandemic in selected countries to inform strengthening of public health systems: a qualitative study, Public Health, 225 (2023), 343–352. https://doi.org/10.1016/j.puhe.2023.10.024 doi: 10.1016/j.puhe.2023.10.024
    [8] F. M. Ekawati, M. Muchlis, N. G. Iturrieta-Guaita, D. A. D. Putri, Recommendations for improving maternal health services in indonesian primary care under the covid-19 pandemic: Results of a systematic review and appraisal of international guidelines, Sex. Reprod. Healthcare, 35 (2023), 100811. https://doi.org/10.1016/j.srhc.2023.100811 doi: 10.1016/j.srhc.2023.100811
    [9] A. Rupp, P. Limpaphayom, Benefits of corporate social responsibility during a pandemic: Evidence from stock price reaction to covid-19 related news, Res. Int. Bus. Finance, 68 (2024), 102169. https://doi.org/10.1016/j.ribaf.2023.102169 doi: 10.1016/j.ribaf.2023.102169
    [10] I. D. Selvi, Online learning and child abuse: the covid-19 pandemic impact on work and school from home in indonesia, Heliyon, 8 (2022), e08790. https://doi.org/10.1016/j.heliyon.2022.e08790 doi: 10.1016/j.heliyon.2022.e08790
    [11] R. Banerjee, R. K. Biswas, Fractional optimal control of compartmental sir model of covid-19: Showing the impact of effective vaccination, IFAC-PapersOnLine, 55 (2022), 616–622. https://doi.org/10.1016/j.ifacol.2022.04.101 doi: 10.1016/j.ifacol.2022.04.101
    [12] M. L. Diagne, H. Rwezaura, S. Y. Tchoumis, J. M. Tchuenche, A mathematical model of covid-19 with vaccination and treatment, Comput. Math. Methods Med., 2021 (2021), 1250129. https://doi.org/10.1155/2021/1250129 doi: 10.1155/2021/1250129
    [13] J. N. Paul, I. S. Mbalawata, S. S. Mirau, L. Masandawa, Mathematical modeling of vaccination as a control measure of stress to fight covid-19 infections, Chaos, Solitons Fractals, 166 (2023), 112920. https://doi.org/10.1016/j.chaos.2022.112920 doi: 10.1016/j.chaos.2022.112920
    [14] B. Yang, Z. Yu, Y. Cai, The impact of vaccination on the spread of covid-19: Studying by a mathematical model, Physica A, 590 (2022), 126717. https://doi.org/10.1016/j.physa.2021.126717 doi: 10.1016/j.physa.2021.126717
    [15] A. Ayalew, M. Yezbalew, T. Tilahun, T. Tesfa, Mathematical model and analysis on the impacts of vaccination and treatment in the control of the covid-19 pandemic with optimal control, J. Appl. Math., 2023 (2023), 8570311. https://doi.org/10.1155/2023/8570311 doi: 10.1155/2023/8570311
    [16] C. W. Chukwu, R. T. Alqahtani, F. F. Herdicho, A pontryagin's maximum principle and optimal control model with cost-effectiveness analysis of the covid-19 epidemic, Decis. Anal. J., 8 (2023), 100273. https://doi.org/10.1016/j.dajour.2023.100273 doi: 10.1016/j.dajour.2023.100273
    [17] S. Bhatter, K. Jangid, A. Abidemi, K. M. Owolabi, S. D. Purohit, A new fractional mathematical model to study the impact of vaccination on covid-19 outbreaks, Decis. Anal. J., 6 (2023), 100156. https://doi.org/10.1016/j.dajour.2022.100156 doi: 10.1016/j.dajour.2022.100156
    [18] C. Xu, Y. Yu, G. Ren, Y. Sun, X. Si, Stability analysis and optimal control of a fractional-order generalized seir model for the covid-19 pandemic, Appl. Math. Comput., 457 (2023), 128210. https://doi.org/10.1016/j.amc.2023.128210 doi: 10.1016/j.amc.2023.128210
    [19] C. M. Wachira, G. O. Lawi, L. O. Omondi, Travelling wave analysis of a diffusive covid-19 model, J. Appl. Math., 2022 (2022), 60522274. https://doi.org/10.1155/2022/6052274 doi: 10.1155/2022/6052274
    [20] B. Barnes, I. Takyi, B. E. Owusu, F. Ohene Boateng, A. Saahene, E. Saarah Baidoo, et al., Mathematical modelling of the spatial epidemiology of covid-19 with different diffusion coefficients, Int. J. Differ. Equations, 2022, 7563111. https://doi.org/10.1155/2022/7563111 doi: 10.1155/2022/7563111
    [21] A. El Koufi, N. El Koufi, Stochastic differential equation model of covid-19: Case study of pakistan, Results Phys., 34 (2022), 105218. https://doi.org/10.1016/j.rinp.2022.105218 doi: 10.1016/j.rinp.2022.105218
    [22] M. Pajaro, N. M. Fajar, A. A. Alonso, I. Otero-Muras, Stochastic sir model predicts the evolution of covid-19 epidemics from public health and wastewater data in small and medium-sized municipalities: A one year study, Chaos, Solitons Fractals, 164 (2022), 112671. https://doi.org/10.1016/j.chaos.2022.112671 doi: 10.1016/j.chaos.2022.112671
    [23] V. V. Khanna, K. Chadaga, N. Sampathila, R. Chadaga, A machine learning and explainable artificial intelligence triage-prediction system for covid-19, Decis. Anal. J., 7 (2023), 100246. https://doi.org/10.1016/j.dajour.2023.100246 doi: 10.1016/j.dajour.2023.100246
    [24] K. Moulaei, M. Shanbehzadeh, Z. Mohammadi-Taghiabad, H. Kazemi-Arpanahi, Comparing machine learning algorithms for predicting covid-19 mortality, BMC Med. Inf. Decis. Making, 22 (2022), 1–12. https://doi.org/10.1186/s12911-021-01742-0 doi: 10.1186/s12911-021-01742-0
    [25] D. Aldila, S. H. Khoshnaw, E. Safitri, Y. R. Anwar, A. R. Bakry, B. M. Samiadji, et al., A mathematical study on the spread of covid-19 considering social distancing and rapid assessment: the case of jakarta, indonesia, Chaos Solitons Fractals, 139 (2020), 110042. https://doi.org/10.1016/j.chaos.2020.110042 doi: 10.1016/j.chaos.2020.110042
    [26] J. L. Gevertz, J. M. Greene, C. H. Sanchez-Tapia, E. D. Sontag, A novel covid-19 epidemiological model with explicit susceptible and asymptomatic isolation compartments reveals unexpected consequences of timing social distancing, J. Theor. Biol., 510 (2021), 110539. https://doi.org/10.1016/j.jtbi.2020.110539 doi: 10.1016/j.jtbi.2020.110539
    [27] I. A. Arik, H. K. Sari, S. I. Araz, Numerical simulation of covid-19 model with integer and non-integer order: The effect of environment and social distancing, Results Phys., 51 (2023), 106725. https://doi.org/10.1016/j.rinp.2023.106725 doi: 10.1016/j.rinp.2023.106725
    [28] S. Saharan, C. Tee, A covid-19 vaccine effectiveness model using the susceptible-exposed-infectious-recovered model, Healthcare Anal., 4 (2023), 100269. https://doi.org/10.1016/j.health.2023.10026 doi: 10.1016/j.health.2023.10026
    [29] A. I. Alaje, M. O. Olayiwola, A fractional-order mathematical model for examining the spatiotemporal spread of covid-19 in the presence of vaccine distribution, Healthcare Anal., 4 (2023), 100230. https://doi.org/10.1016/j.health.2023.100230 doi: 10.1016/j.health.2023.100230
    [30] R. Pino, V. M. Mendoza, E. A. Enriques, A. C. Velasco, R. Mendoza, An optimization model with simulation for optimal regional allocation of covid-19 vaccines, Healthcare Anal., 4 (2023), 100244. https://doi.org/10.1016/j.health.2023.100244 doi: 10.1016/j.health.2023.100244
    [31] O. J. Watson, G. Barnsley, J. Toor, A. B. Hogan, P. Winskill, A. C. Ghani, Global impact of the first year of covid-19 vaccination: a mathematical modelling study, Lancet Infect. Dis., 22 (2022), 1293–1302. https://doi.org/10.1016/S1473-3099(22)00320-6 doi: 10.1016/S1473-3099(22)00320-6
    [32] A. K. Paul, M. A. Kuddus, Mathematical analysis of a covid-19 model with double dose vaccination in bangladesh, Results Phys., 35 (2022), 105392. https://doi.org/10.1016/j.rinp.2022.105392 doi: 10.1016/j.rinp.2022.105392
    [33] O. I. Idisi, T. T. Yusuf, K. M. Owolabi, B. A. Ojokoh, A bifurcation analysis and model of covid-19 transmission dynamics with post-vaccination infection impact, Healthcare Anal., 3 (2023), 100157. https://doi.org/10.1016/j.health.2023.100157 doi: 10.1016/j.health.2023.100157
    [34] I. Ul Haq, N. Ullah, N. Ali, K. S. Nisar, A new mathematical model of covid-19 with quarantine and vaccination, Mathematics, 11 (2022), 142. https://doi.org/10.3390/math11010142 doi: 10.3390/math11010142
    [35] D. S. A. A. Reza, M. N. Billah, S. S. Shanta, Effect of quarantine and vaccination in a pandemic situation: A mathematical modelling approach, J. Math. Anal. Model., 2 (2021), 77–81. https://doi.org/10.48185/jmam.v2i3.318 doi: 10.48185/jmam.v2i3.318
    [36] Y. Gu, S. Ullah, M. A. Khan, Mathematical modeling and stability analysis of the covid-19 with quarantine and isolation, Results Phys., 34 (2022), 105284. https://doi.org/10.1016/j.rinp.2022.105284 doi: 10.1016/j.rinp.2022.105284
    [37] F. Wu, X. Liang, J. Lein, Modelling covid-19 epidemic with confirmed cases-driven contact tracing quarantine, Infect. Dis. Modell., 8 (2023), 415–426. https://doi.org/10.1016/j.idm.2023.04.001 doi: 10.1016/j.idm.2023.04.001
    [38] A. K. Saha, S. Saha, C. N. Podder, Effect of awareness, quarantine and vaccination as control strategies on covid-19 with co-morbidity and re-infection, Infect. Dis. Modell., 7 (2022), 660–689. https://doi.org/10.1016/j.idm.2022.09.004 doi: 10.1016/j.idm.2022.09.004
    [39] S. S. Musa, S. Queshi, S. Zhao, A. Yusuf, U. T. Mustapha, D. He, Mathematical modeling of covid-19 epidemic with effect of awareness programs, Infect. Dis. Modell., 6 (2021), 448–460. https://doi.org/10.1016/j.idm.2021.01.012 doi: 10.1016/j.idm.2021.01.012
    [40] A. A. Anteneh, Y. M. Bazazaw, S. Palanisam, Mathematical model and analysis on the impact of awareness campaign and asymptomatic human immigrants in the transmission of covid-19, Biomed Res. Int., 2022 (2022), 6260262. https://doi.org/10.1155/2022/6260262 doi: 10.1155/2022/6260262
    [41] M. A. Balya, B. O. Dewi, F. I. Lestari, G. Ratu, H. Rosuliyana, T. Windyhani, et al., Investigating the impact of social awareness and rapid test on a covid-19 transmission model, Commun. Biomathematical Sci., 4 (2021), 46–64. https://doi.org/10.5614/cbms.2021.4.1.5 doi: 10.5614/cbms.2021.4.1.5
    [42] A. K. Srivastav, M. Gosh, S. S. Bandekar, Modeling of covid-19 with limited public health resources: a comparative study of three most affected countries, Eur. Phys. J. Plus, 136 (2021), 359. https://doi.org/10.1140/epjp/s13360-021-01333-y doi: 10.1140/epjp/s13360-021-01333-y
    [43] U. M. Rifanti, A. R. Dewi, N. Nurlaili, S. T. Hapsari, Model matematika covid-19 dengan sumber daya pengobatan yang terbatas, J. Math. Appl., 18 (2021), 23–36. https://doi.org/10.12962/limits.v18i1.8207 doi: 10.12962/limits.v18i1.8207
    [44] S. Cakan, Dynamic analysis of a mathematical model with health care capacity for covid-19 pandemic, Chaos, Solitons, Fractals, 139 (2020), 110033. https://doi.org/10.1016/j.chaos.2020.110033 doi: 10.1016/j.chaos.2020.110033
    [45] I. I. Oke, Y. T. Oyebo, O. F. Fakoya, V. S. Benson, Y. T. Tunde, A mathematical model for covid-19 disease transmission dynamics with impact of saturated treatment: Modeling, analysis and simulation, Open Access Lib. J., 8 (2021), 1–20. https://doi.org/10.4236/oalib.1107332 doi: 10.4236/oalib.1107332
    [46] M. Elhia, L. Boujallal, M. Alkama, O. Balatif, M. Rachik, Set-valued control approach applied to a covid-19 model with screening and saturated treatment function, Complexity, 2020 (2020), 9501028. https://doi.org/10.1155/2020/9501028 doi: 10.1155/2020/9501028
    [47] J. K. Ghosh, S. K. Bhiswas, S. Sarkar, U. Ghosh, Mathematical modelling of covid-19: A case study of italy, Math. Comput. Simul., 194 (2022), 1–18. https://doi.org/10.1016/j.matcom.2021.11.008 doi: 10.1016/j.matcom.2021.11.008
    [48] R. T. Alqahtani, A. Ajbar, Study of dynamics of a covid-19 model for saudi arabia with vaccination rate, saturated treatment function and saturated incidence rate, Mathematics, 9 (2021), 3134. https://doi.org/10.3390/math9233134 doi: 10.3390/math9233134
    [49] I. Ali, S. U. Khan, Dynamics and simulations of stochastic covid 19 epidemic model using legendre spectral collocation method, AIMS Math., 8 (2023), 4220–4236. https://doi.org/10.3934/math.2023210 doi: 10.3934/math.2023210
    [50] S. S. Chaharborj, S. S. Chaharborj, J. H. Asl, P. S. Phang, Controlling pandemic covid-19 using optimal control theory, Results Phys., 26 (2021), 104311. https://doi.org/10.1016/j.rinp.2021.104311 doi: 10.1016/j.rinp.2021.104311
    [51] R. P. Kumar, S. Basu, P. K. Santra, D. Ghosh, G. S. Mahapatra, Optimal control design incorporating vaccination and treatment on six compartment pandemic dynamical system, Results Control Optim., 7 (2022), 100115. https://doi.org/10.1016/j.rico.2022.100115 doi: 10.1016/j.rico.2022.100115
    [52] T. Li, Y. Guo, Optimal control and cost-effectiveness analysis of a new covid-19 model for omicron strain, Physica A, 606 (2022), 128134. https://doi.org/10.1016/j.physa.2022.128134 doi: 10.1016/j.physa.2022.128134
    [53] A. Bilgram, P. G. Jensen, K. Y. Jørgensen, K. G. Larsen, M. Mikučionis, M. Muñiz, et al., An investigation of safe and near-optimal strategies for prevention of covid-19 exposure using stochastic hybrid models and machine learning, Decis. Anal. J., 5 (2022), 100141. https://doi.org/10.1016/j.dajour.2022.100141 doi: 10.1016/j.dajour.2022.100141
    [54] M. S. Khatun, S. Das, P. Das, Dynamics and control of an sitr covid-19 model with awareness and hospital bed dependency, Chaos, Solitons Fractals, 175 (2023), 114010. https://doi.org/10.1016/j.chaos.2023.114010 doi: 10.1016/j.chaos.2023.114010
    [55] Mayo Clinic, Herd immunity and covid-19: What you need to know, 2023. Available from: https://www.mayoclinic.org/diseases conditions/coronavirus/in-depth/herd-immunity-and coronavirus.
    [56] D. K. A. Mannan, K. M. Farhana, Knowledge, attitude and acceptance of a covid‐19 vaccine: a global cross‐sectional study, Int. Res. J. Bus. Social Sci., 6 (2020), 1–23.
    [57] S. M. Saeied, M. M. Saeied, I. A. Kabbash, S. Abdo, Vaccine hesitancy: beliefs and barriers associated with covid‐19 vaccination among egyptian medical students, J. Med. Virol., 93 (2021), 4280–4291. https://doi.org/10.1002/jmv.26910 doi: 10.1002/jmv.26910
    [58] M. O. Elgendy, M. E. A. Abdelrahim, Public awareness about coronavirus vaccine, vaccine acceptance, and hesitancy, J. Med. Virol., 93.(2021), 6535–6543. https://doi.org/10.1002/jmv.27199 doi: 10.1002/jmv.27199
    [59] S. Funk, E. Gilad, C. Watkins, V. A. A. Jansen, The spread of awareness and its impact on epidemic outbreaks, Proc. Natl. Acad. Sci., 106 (2009), 6872–6877. https://doi.org/10.1073/pnas.0810762106 doi: 10.1073/pnas.0810762106
    [60] G. Kelly, S. Petti, N. Noah, Covid-19, non-covid-19 and excess mortality rates not comparable across countries, Epidemiol. Infect., 149 (2021), e176. https://doi.org/10.1017/S0950268821001850 doi: 10.1017/S0950268821001850
    [61] S. Nourazari, S. R. Davis, R. Granovsky, R. Austin, D. J. Straff, J. W. Joseph, L. D. Sanchez, Decreased hospital admissions through emergency departments during the covid-19 pandemic, Am. J. Emerg. Med., 42 (2021), 203–210. https://doi.org/10.1016/j.ajem.2020.11.029 doi: 10.1016/j.ajem.2020.11.029
    [62] Statistics Center of West Java Province (BPS), Jumlah penduduk menurut kabupaten/kota (jiwa), 2018-2020, 2023. Available from: https://jabar.bps.go.id/indicator/12/133/1/jumlah-penduduk-menurut-kabupaten-kota.html.
    [63] A. Abidemi, J. O. Akanni, O. D. Makinde, A non-linear mathematical model for analysing the impact of covid-19 disease on higher education in developing countries, Healthcare Anal., 3 (2023), 100193. https://doi.org/10.1016/j.health.2023.100193 doi: 10.1016/j.health.2023.100193
    [64] E. A. Iboi, O. Sharomi, C. N. Ngonghala, A. B. Gumel, Mathematical modeling and analysis of covid-19 pandemic in nigeria, Math. Biosci. Eng., 17 (2020), 7192–7220. https://doi.org/10.3934/mbe.2020369 doi: 10.3934/mbe.2020369
    [65] Statistics Center of West Java Province (BPS), [komponen ipg] usia harapan hidup 2020-2022, 2023. Available from: https://jabar.bps.go.id/indicator/40/185/1/-komponen-ipg-usia-harapan-hidup-.html.
    [66] S. Olaniyi, O. S. Obabiyi, K. O. Okosun, A. T. Oladipo, S. O. Adewale, Athematical modelling and optimal cost-effective control of covid-19 transmission dynamics, Eur. Phys. J. Plus, 135 (2021), 1–20. https://doi.org/10.1140/epjp/s13360-020-00954-z doi: 10.1140/epjp/s13360-020-00954-z
    [67] A. Abate, A. Tiwari, S. Sastrys, Box invariance in biologically-inspired dynamical systems, Automatica, 45 (2009), 1601–1610. https://doi.org/10.1016/j.automatica.2009.02.028 doi: 10.1016/j.automatica.2009.02.028
    [68] P. Taylan, G. W. Weber, L. Liu, F. Yerlika-Ozkurt, On the foundations of parameter estimation for generalized partial linear models with b-splines and continuous optimization, Comput. Math. Appl., 60 (2010), 134–143. https://doi.org/10.1016/j.camwa.2010.04.040 doi: 10.1016/j.camwa.2010.04.040
    [69] A. A. Tappe, M. Schulze, R. Schenkendorf, Neural odes and differential flatness for total least squares parameter estimation, IFAC-PapersOnLine, 55 (2020), 421–426. https://doi.org/10.1016/j.ifacol.2022.09.131 doi: 10.1016/j.ifacol.2022.09.131
    [70] O. Aydogmuz, T. Ali Hakan, A modified multiple shooting algorithm for parameter estimation in odes using adjoint sensitivity analysis, Appl. Math. Comput., 390 (2021), 125644. https://doi.org/10.1016/j.amc.2020.125644 doi: 10.1016/j.amc.2020.125644
    [71] S. Khalilpourazari, H. Hashemi Doulabi, A. O. Ciftcioglu, G. W. Weber, Gradient-based grey wolf optimizer with gaussian walk: Application in modelling and prediction of the covid-19 pandemic, Expert Syst. Appl., 177 (2021), 114920. https://doi.org/10.1016/j.eswa.2021.114920 doi: 10.1016/j.eswa.2021.114920
    [72] P. Driesche, J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmissions, Math. Biosci., 180 (2002), 29–48. https://doi.org/10.1016/S0025-5564(02)00108-6 doi: 10.1016/S0025-5564(02)00108-6
    [73] M. Martcheva, An Introduction to Mathematical Epidemiology, Springer, Berlin, Germany, 2015.
    [74] C. Castillo–Chavez, B. Song, Dynamical models of tuberculosis and their applications, Math. Biosci. Eng., 1 (2024), 361–404. https://doi.org/10.3934/mbe.2004.1.361 doi: 10.3934/mbe.2004.1.361
    [75] S. Leinhart, J. T. Workman, Optimal Control Applied to Biological Models, Taylor & Francis Group, CRC press, Boca Raton, Florida, USA, 2007.
    [76] D. Aldila, Cost-effectiveness and backward bifurcation analysis on covid-19 transmission model considering direct and indirect transmission, Commun. Math. Biol. Neurosci., 2020 (2020), 49. https://doi.org/10.28919/cmbn/4779 doi: 10.28919/cmbn/4779
    [77] D. Aldila, A. Nadya, Fatmawati, F. F. Herdicho, M. Z. Ndii, C. W. Chukwu, Optimal control of pneumonia transmission model with seasonal factor: Learning from jakarta incidence data, Heliyon, 9 (2023), e18096. https://doi.org/10.1016/j.heliyon.2023.e18096 doi: 10.1016/j.heliyon.2023.e18096
    [78] D. Aldila, M. Angelina, Optimal control problem and backward bifurcation on malaria transmission with vector bias, Heliyon, 7 (2021), e06824. https://doi.org/10.1016/j.heliyon.2021.e06824 doi: 10.1016/j.heliyon.2021.e06824
    [79] D. Aldila, Optimal control for dengue eradication program under the media awareness effect, Int. J. Nonlinear Sci. Numer. Simul., 24 (2023), 95–122. https://doi.org/10.1515/ijnsns-2020-0142 doi: 10.1515/ijnsns-2020-0142
    [80] O. Sharomi, C. N. Podder, A. B, Gumel, B. Song, Mathematical analysis of the transmission dynamics of hiv/tb co-infection in the presence of treatment, Math. Biosci. Eng., 5 (2008), 145–174. https://doi.org/10.3934/mbe.2008.5.145 doi: 10.3934/mbe.2008.5.145
    [81] E. H. Elbasha, C. N. Podder, A. B. Gumel, Analyzing the dynamics of an sirs vaccination model with waning natural and vaccine-induced immunity, Nonlinear Anal. Real World Appl., 12 (2011), 2692–2705. https://doi.org/10.1016/j.nonrwa.2011.03.015 doi: 10.1016/j.nonrwa.2011.03.015
    [82] A. B. Gumel, Causes of backward bifurcations in some epidemiological models, J. Math. Anal. Appl., 395 (2012), 355–365. https://doi.org/10.1016/j.jmaa.2012.04.077 doi: 10.1016/j.jmaa.2012.04.077
    [83] D. Aldila, J. P. Chavez, K. P. Wijaya, N. C. Ganegoda, G. M. Simorangkir, H. Tasman, et al., A tuberculosis epidemic model as a proxy for the assessment of the novel m72/as01e vaccine, Commun. Nonlinear Sci. Numer. Simul., 120 (2023), 107162. https://doi.org/10.1016/j.cnsns.2023.107162 doi: 10.1016/j.cnsns.2023.107162
    [84] K. Erik, W. Gerhard-Wilhelm, T. E. Babaee, Foundations of semialgebraic gene-environment networks, J. Dyn. Games, 7 (2020), 253–268. https://doi.org/10.3934/jdg.2020018 doi: 10.3934/jdg.2020018
    [85] S. Belen, E. Kropat, W. Gerhard-Wilhelm, On the classical maki–thompson rumour model in continuous time, Cent. Eur. Oper. Res., 19 (2011), 1–17. https://doi.org/10.1016/j.jmaa.2015.06.054 doi: 10.1016/j.jmaa.2015.06.054
    [86] E. Savku, A stochastic control approach for constrained stochastic differential games with jumps and regimes, preprint, arXiv: 2301.12921. https://doi.org/10.48550/arXiv.2301.12921
  • This article has been cited by:

    1. Muhammad Abdurrahman Rois, Cicik Alfiniyah, Santi Martini, Dipo Aldila, Farai Nyabadza, Modeling and optimal control of COVID-19 with comorbidity and three-dose vaccination in Indonesia, 2024, 6, 25889338, 181, 10.1016/j.jobb.2024.06.004
    2. Olumuyiwa James Peter, Dipo Aldila, Tawakalt Abosede Ayoola, Ghaniyyat Bolanle Balogun, Festus Abiodun Oguntolu, Modeling tuberculosis dynamics with vaccination and treatment strategies, 2025, 28, 24682276, e02647, 10.1016/j.sciaf.2025.e02647
    3. Dipo Aldila, Abdullah Hasan Hassan, Mohamad Hifzhudin Noor Aziz, Putri Zahra Kamalia, An analytical transmission model for evaluating pneumonia vaccination and control strategies, 2025, 7, 27724425, 100394, 10.1016/j.health.2025.100394
  • 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(1952) PDF downloads(204) Cited by(3)

Figures and Tables

Figures(10)  /  Tables(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog