Skip to main content

ORIGINAL RESEARCH article

Front. Phys., 08 May 2024
Sec. Social Physics

Cost–benefit analysis of the COVID-19 vaccination model incorporating different infectivity reductions

  • 1Complex Systems Research Center, Shanxi University, Taiyuan, China
  • 2Department of Mathematics, Saveetha School of Engineering SIMATS, Chennai, India
  • 3Department of Mathematics, Kwame Nkrumah University of Science and Technology, Kumasi, Ghana

The spread and control of coronavirus disease 2019 (COVID-19) present a worldwide economic and medical burden to public health. It is imperative to probe the effect of vaccination and infectivity reductions in minimizing the impact of COVID-19. Therefore, we analyze a mathematical model incorporating different infectivity reductions. This work provides the most economical and effective control methods for reducing the impact of COVID-19. Using data from Ghana as a sample size, we study the sensitivity of the parameters to estimate the contributions of the transmission routes to the effective reproduction number Re. We also devise optimal interventions with cost–benefit analysis that aim to maximize outcomes while minimizing COVID-19 incidences by deploying cost-effectiveness and optimization techniques. The outcomes of this work contribute to a better understanding of COVID-19 epidemiology and provide insights into implementing interventions needed to minimize the COVID-19 burden in similar settings worldwide.

1 Introduction

Coronavirus disease 2019 (COVID-19) is one of the infectious diseases that has caused a global medical and financial burden. Due to its high transmission rate, the World Health Organization (WHO) officially announced the prevalence of it as a global pandemic on 8th March 2020. The coronavirus spreads directly or indirectly from one infectious living host cell to another by replicating and causing infectious disease in the host (humans and/or animals) [1]. Therefore, it is imperative to investigate how the virus spreads in the community and devise strategies to halt the disease’s transmission.

Globally, COVID-19 has caused over 6 million deaths, and by March 2022, approximately 480 million incidences had been reported, according to [2]. However, the COVID-19 reinfection scenario using the SIR model shows that transmission dynamics could arise as a result of immunity waning, even in cases where the force of reinfection is relatively weak [3]. As mentioned in [4], the COVID-19 model, which entails infection through objects contaminated with SARS-CoV-2, is suggested to be made public knowledge.

COVID-19 spreads quickly, threatening global health and igniting a pandemic. This pandemic has had a major effect on other sectors, especially the socioeconomic sector [5]. As of 30 April 2022, there were 512,466,045 reported incidences and 6,257,512 COVID-19 fatalities worldwide, according to the Worldometer. With 6,046,467 COVID-19 cases reported overall, Indonesia ranks seventh in Asia and 18th globally [6].

These findings highlight the need to investigate the spread of COVID-19 in order to minimize its transmission. Using a deterministic model is one way that mathematics is crucial to simulating the epidemic phenomenon of the disease’s spread. The analysis of the COVID-19 model with declining immunity has advanced significantly. For the latter, the natural immunity period is defined by the vaccine efficacy level, which dictates when to start the mass vaccination strategy based on models involving symptomatic and asymptomatic infected populations, as explained in [6].

Mathematical models have been an important tool in epidemiology since the 18th century and can be used to determine the level of spread of infectious disease, aside from medical and biological research studies and strategies for controlling diseases. A modified susceptible–exposed–infectious–removed (SEIR) model with vaccination, quarantine, and isolated SVEQIMR is used to analyze the spread of COVID-19 disease in [7]. In this article, compartmental models were used to demonstrate the pattern of vaccination and control strategies for infectious diseases.

Furthermore, the compartmental model is used to study the behavior of COVID-19 after the introduction of vaccines in [8]. In this article, the population is divided into distinct compartments to demonstrate the control efficacy of infectious diseases in India. Again, a parameterized nonlinear SEIHR model to analyze the transmission of coronavirus disease in Indonesia using the compartmental model is analyzed in [9].

The transmission dynamics, basic reproduction number, and control measures of COVID-19 are analyzed using the compartmental model in [10]. In this article, effective public health interventions were proposed to control the spread of the disease in Wuhan, China. The cost-effective analysis, global stability, and control strategy for the spread of COVID-19 in Ghana are analyzed using the compartmental model in [11]. In this article, control measures were outlined by formulating the SEAIRV model to investigate human–environment–human transmission. However, these studies do not capture the effects of reducing the infectivity rate on the disease’s transmission.

The aforementioned issues have driven this study to explore the effects of vaccination and infectivity reductions on the spread of COVID-19, devise optimal control interventions, and analyze the cost–benefits of implementing the interventions. The subsequent sections are as follows: in Section 2, we present the design of the epidemiological model together with the definition of the parameters and variables in the model. In Section 3, we analyze the positivity of the model’s solutions, computation of the models’ reproduction numbers, and stability of the model. In Section 4, we present the model parameter estimation and sensitivity analysis. In Section 5, we present the numerical simulations and model analysis. In Section 6, we present the optimal control strategies and analyze the cost–benefit analysis. We finally present the concluding remarks of this study in Section 7.

2 SVEQIMR model formulation

The flowchart below illustrates the mechanism of the model. We denote Nt as the total population divided into seven different compartments: susceptible individuals at a given time St, vaccinated individuals at a given time Vt, exposed individuals at a given time Et, individuals under quarantine at a given time Qt, infected individuals at a given time It, isolated individuals at a given time Mt, and individuals recovered at a given time Rt. We assume that the vaccinated individuals become susceptible again due to the vaccine’s inefficacy. Again, a portion of susceptible individuals undergo self-quarantine, while others enter the exposed class. Some of the exposed individuals become infected and then recover naturally without any special treatment. The remaining proportion of individuals either go through quarantine or become infected. It is also assumed that the individuals in the susceptible and vaccinated compartments would come into direct or indirect contact with the individuals in the exposed, quarantined, and isolated compartments. Therefore, we introduce infectivity reductions re,rq, and rm attributed to the exposed, quarantined, and isolated compartments, respectively. The introduction of these different infectivity reductions is to minimize transmission from the potential carriers of the virus. Again, we assume that the quarantined and isolated individuals are not part of the active population at the time of incidence, so the total active population for the standard incidence at any given time is N1=NtQtMt. Therefore, the forces for infection of susceptible and vaccinated classes are given as ρ1=βsI+reE+rqQ+rmMN1 and ρ2=βvI+reE+rqQ+rmMN1, respectively. Recovered individuals become susceptible again once they come into contact with the possible carriers of the COVID-19 virus. The description of the model’s parameters is given in Table 1. The optimal control interventions κpt,κvt, and κit are explained in the subsequent section.

Table 1
www.frontiersin.org

Table 1. Definition of model parameters.

The total population Nt in Eq. (1) is defined based on Figure 1 as follows:

Nt=St+Vt+Et+Qt+It+Mt+Rt.(1)

Figure 1
www.frontiersin.org

Figure 1. SVEQIMR model flowchart.

The following nonlinear ordinary differential equations illustrate the model.

dSdt=Λ+σV+δ1Q+τRρ1Sω1S,dVdt=ϑSρ2Vω2V,dEdt=ρ1S+ρ2Vω3E,dQdt=ŋS+φEω4Q,dIdt=γEω5I,dMdt=δ2Q+αIω6M,dRdt=εeE+εiI+εmMω7R.(2)

where ρ1=βsI+reE+rqQ+rmMN1,ρ2=βvI+reE+rqQ+rmMN1,ω1=μ+ϑ+ŋ,ω2=μ+σ,ω3=μ+φ+γ+εe,ω4=μ+δ1+δ2,ω5=μ+α+εi, ω6=μ+εm, and ω7=μ+τ.

With initial conditions S00,V00,E00,Q00,I00,M00,and R00, all initial parameters of model (2) are nonnegative.

3 SVEQIMR model mathematical analysis

The detailed analysis of the model is conducted mathematically to show the positivity of solutions and their boundedness within a specific domain.

3.1 Positivity and boundedness of solutions

The variables and parameters of model (2) are nonnegative since it is based on a population with COVID-19.

Theorem 1. Define Wt=St,Vt,Et,Qt,It,Mt,Rt; if W00, then Wt0, and its solutions and initial values are nonnegative for t>0 and bounded in the region R+7.

Proof: Let us consider the following instance where there exists an initial time ti such that

minWti>0 and minWt>0 for all t0,ti.

Here, Wt=St,Vt,Et,Qt,It,Mt,Rt. Without the loss of generalization, minWti=Sti.

Therefore, Sti=0,Vti>0,Qti>0,Rti>0, and St>0, for all t0,ti. However,

dStidt=Λ+σVti+δ1Qti+τRti>0,Λ0,Sti>S00.

This contradicts the claim Sti=0. Therefore, St>0 for all t0. This shows that all the solutions are positive for t0 in all other cases.

Theorem 2. To prove the boundedness of model (2), define a positive invariant set as K=S0,V0,E0,Q0,I0,M0,R0R+7:NtΛμ and attract positive solutions.

Proof: Considering Eq. 1.

The rate of change in the total population is given as

Nt=St+Vt+Qt+It+Mt+Rt,
Nt=ΛμSt+Vt+Et+Qt+It+Mt+Rt,
N=ΛμN.(3)

From Eq. (3), it follows that,

NtΛμ+N0eμt.

Then, 0<NtΛμ. Thus, Nt is bounded, and all solutions in the K approach enter or remain in K. If t,0Nt shows that Nt is a set of positive invariant and is in the region R+7. This theorem proves the existence of COVID-19 at a given time in an area that was not infected with COVID-19 disease, and all the initial state variables are positive. This completes the proof.

3.2 Disease-free equilibrium and effective reproduction number

The disease equilibrium point, ξ0, of the model is achieved by equating system (2) to 0. The result is defined as follows:

Λ+σV+δ1Q+τRρ1Sω1S=0,ϑSρ2Vω2V=0,ρ1S+ρ2Vω3E=0,ŋS+φEω4Q=0,γEω5I=0,δ2Q+αIω6M=0,εeE+εiI+εmMω7R=0.(4)

At disease-free equilibrium (DFE), we set E=Q=I=M=R=0, and the following results are obtained from Eq. (4):

Λ+σVω1S=0,
ϑS=ω2V,
S0=Λμ+σμ+ϑ+ŋμ+σσϑ,
V0=Λϑμ+ϑ+ŋμ+σσϑ.

At disease-free equilibrium,

ξ0=Λμ+σμ+ϑ+ŋμ+σσϑ,Λϑμ+ϑ+ŋμ+σσϑ,0,0,0,0,0.(5)

The corresponding Jacobian matrix of the system (2) evaluated at ξ0=S0,V0,0,0,0,0,0 (see, Eq. (5)) to obtain the disease-free equilibrium Jacobian matrix Jξ0 is given as follows:

Jξ0=ω1σβsS0N0reδ1βsS0N0rqβsS0N0βsS0N0rm0ϑω2βvV0N0reβvV0N0rqβvV0N0βvV0N0rm000βsS0+βvV0N0reω3βsS0+βvV0N0rqβsS0+βvV0N0βsS0+βvV0N0rm0ŋ0φω400000γ0ω500000δ2αω6000εe0εiεmμ,(6)

where N0=NQM using ξ0; hence, N0=S0+V0.

Using Jξ0, Eq. (6) the following matrices can be deduced to evaluate the effective reproduction number Re. We define the next-generation matrix G=FV1 as the square matrix, which consists of matrix F representing new infections and matrix V.

F=βsS0+βvV0N0reβsS0+βvV0N0βsS0+βvV0N0rqβsS0+βvV0N0rm000000000000,V=ω3000φω400γ0ω500δ2αω6,
V1=1ω3000φω3ω41ω400γω3ω501ω50αγω4+φδ2ω5ω3ω4ω5ω6δ2ω4ω6αω5ω61ω6,G=FV1=g11g12g13g14000000000000,(7)

where g11=βsS0+βvV0N0ω3re+βsS0+βvV0γN0ω3ω5+βsS0+βvV0φN0ω3ω4rq+βsS0+βvV0ω4αγ+ω5φδ2N0ω3ω4ω5ω6rm,g12=βsS0+βvV0N0ω4rq+βsS0+βvV0δ2N0ω4ω6rm,g13=βsS0+βvV0N0ω5+βsS0+βvV0αN0ω5ω6rm,g14=βsS0+βvV0N0ω6rm.

The effective reproduction number, Re, of the model is evaluated as the spectral radius of matrix G; that is, PFV1 given as

Re=ReE+ReI+ReQ+ReM,

where ReE=βsS0+βvV0N0ω3re,ReI=βsS0+βvV0γN0ω3ω5,ReQ=βsS0+βvV0φN0ω3ω4rq, and ReM=βsS0+βvV0ω4αγ+ω5φδ2N0ω3ω4ω5ω6rm.

Substituting S0,V0, and N0 into Eq. (7), gives

Re=βsμ+σ+βvϑμ+σ+ϑreω3+γω3ω5+φrqω3ω4+ω4αγ+ω5φδ2rmω3ω4ω5ω6.(8)

Therefore, there is a unique equilibrium in the model, which implies that Re in Eq. (8) is unique. The effective reproduction number, Re, indicates the number of secondary infections that one infected person can produce if they come into contact with people living in a safe zone. Re can be summarized as follows: the exposed class contributed to a significant secondary infection, which is the first term of Re, where a proportion of γω3 individuals entered the infected class. The infected class generated a significant secondary infection, which is the second term of Re. The quarantined class contributed to a significant secondary infection, which is the third term of Re. The isolated class contributed to a significant secondary infection, which is the last term of Re. We consider the dynamics of vaccination and infectivity reductions on Re as follows in Eqs 911 respectively.

• In the absence of vaccination, the reproduction number, Rav, is given as

Rav=βsreω3+γω3ω5+φrqω3ω4+ω4αγ+ω5φδ2rmω3ω4ω5ω6.(9)

• In the absence of infectivity reductions, the reproduction number, Rai, is given as

Rai=βsμ+σ+βvϑμ+σ+ϑ1ω3+γω3ω5+φω3ω4+ω4αγ+ω5φδ2ω3ω4ω5ω6.(10)

• In the absence of vaccination and infectivity reductions, the reproduction number, Ran, is given as

Ran=βs1ω3+γω3ω5+φω3ω4+ω4αγ+ω5φδ2ω3ω4ω5ω6.(11)

3.3 Stability analysis

Next, since the system has unique equilibrium points, we check its stability.

3.4 Local stability of disease-free equilibrium ξ0

The Jacobian matrix, Jξ0, is given in Equation 7. It can be seen from (6) that λ1=ω1,λ2=ω2,andλ3=ω7 are the three eigenvalues of Jξ0. The other eigenvalues are derived from the reduced matrix in Eq. (12):

J1ξ0λI4=Aω3λA1A2A3φω4λ00γ0ω5λ00δ2αω6λ,(12)

where A=βsS0+βvV0N0re,A1=βsS0+βvV0N0rq,A2=βsS0+βvV0N0, and A3=βsS0+βvV0N0rm.

Therefore, the remaining eigenvalues are the roots of the following characteristic polynomial:

Γ4λ=λ4+H3λ3+H2λ2+H1λ+H0=0,(13)

where

H3=ω3+ω4+ω5+ω6A,
H2=ω3ω4+ω5+ω6+ω4ω5+ω6+ω5ω6Aω4+ω5+ω6A2γ,
H1=ω3ω4ω5+ω4ω6+ω5ω6+ω4ω5ω6Aω4ω5+ω4ω6+ω5ω6Bδrta3+a4A1φω5+ω6A2γω4+ω6A3αγ+φδ2,
H0=ω3ω4ω5ω6Aω4ω5ω6A1φω5ω6A2γω4ω6A3αγω4+φδ2ω5,
H0=1Aω3+A2γω3ω5+A1φω4ω5+A3αγω4+φδ2ω5ω3ω4ω5ω6,
H0=1ReE+ReI+ReQ+ReM,
H0=1Re.

Therefore, using the Routh–Hurwitz stability conditions, the roots of Γ4λ in Eq. (13) of the reduced matrix have negative real parts if the following conditions hold: H3>0,H2>0,H1>0, and H0>0. It is obvious that the condition holds if Re<1. This proves that in model (2), disease-free equilibrium, ξ0, is locally stable if Re<1 and unstable if Re>1.

3.5 Existence of endemic equilibrium of the model

Let us consider P*=S*,V*,E*,Q*,I*,M*,R* as the endemic equilibrium for system (2) and equate the derivative to zero (0). The following results are obtained:

Λ+σV*+δ1Q*+τR*ZβsS*ω1S*=0,ϑS*ZβvV*ω2V*=0,ZβsS*+βvV*ω3E*=0,ŋS*+φE*ω4Q*=0,γE*ω5I*=0,δ2Q*+αI*ω6M*=0,εeE*+εiI*+εmM*ω7R*=0,(14)

where Z=I*+reE*+rqQ*+rmM*N1*, for N1*=N*Q*M*.

Then, P*=S*,V*,E*,Q*,I*,M*,R*, and solving Eq. 14 simultaneously gives the following

S*=ϑω4ω6Zβv+ω2Λ+y1I*y4,V*=ϑω4ω6Λ+y1I*y4,E*=ω5γI*,Q*=1ω4γŋϑδ2ω4ω6Zβv+ω2Λ+y1I*+φω5y3I*γy3,M*=1ω6γŋϑω4ω6Zβv+ω2Λ+y1I*+φω5+αγy3I*γy4,R*=1ω7εeω5+γεiI*γ+εmω6γŋϑδ2ω4ω6Zβv+ω2Λ+y1I*+φω5+αγy3I*γy4.(15)

Substituting the expressions for S*,V*,and E* in Eq. 15 into the fifth equation of Eq. 14 and simplifying, we obtain the following equation for I*:

ω4ω6Λ+y1I*Zβv+ω2+ϑy4ω3ω5I*γ=0.

Simplifying the above equation gives Eq. (16)

I*=Λγω4ω6Zβv+ω2+ϑω3ω4y4γω4ω6y1Zβv+ω2+ϑ,(16)

where y1=φδ1ω5γω4+τω7εeω5γ+εi+εmω6φδ2ω5γ+α,y2=σϑω4ω6+εmŋδ2Zβv+ω2,y3=σϑω4ω6+εmŋδ2Zβv+ω2ω4ω6Zβvμ+ŋZ2βsβvω1ω2,andy4=ZKy3Re1.

Note that, in evaluating S* and V*,

y4=Zβs+ω2)(βv+ω1ω4ω6y2,
y4=Z2βsβv+Zβsω2+Zβvω1+ω1ω2ω4ω6y2.

Substituting ω1 and ω2 and simplifying gives

y4=Zβsμ+σ+Zβvϑ+Zβvμ+ŋ+Z2βsβv+μ+ϑ+ŋμ+ϑy2,
y4=Zβsμ+ϑ+Zβvϑy3.

Let K=1μ+σ+ϑreω3+γω3ω5+φrqω3ω4+ω4αγ+ω5φδ2rmω3ω4ω5ω6.

It follows that

y4=ZKβsμ+σ+βvϑKy3,
y4=Zβsμ+σ+βvϑμ+σ+ϑreω3+γω3ω5+φrqω3ω4+ω4αγ+ω5φδ2rmω3ω4ω5ω6Ky3,
y4=ZKy3Re1.(17)

The expressions for S* and V* can be rewritten as follows

S*=ϑω4ω6Zβv+ω2Λ+y1I*ZKy3Re1,
V*=ϑω4ω6Λ+y1I*ZKy3Re1.

Let us consider the denominators of S* and V* in the following cases:

(a) If Re>1, then ZKy3Re1=ZKy3Re1>0. This results in the endemic equilibrium point K*=S*,V*,E*,Q*,I*,M*,R* since the expressions for S* and V* are nonnegative, which implies I*>0.

(b) If Re<1, then since ZKy3Re1=ZKy3Re1<0, which implies I*<0, endemic equilibrium does not exist.

Using Eq. (17), for I*>0 when Re>1, it implies that the endemic equilibrium P* has a positive and unique equilibrium point when Re>1.

3.6 Global stability of endemic equilibrium

Let us consider a globally positively definite and unbounded function Lx with a globally negative time derivative. For Lx<0 for all xx*, then at equilibrium, x* is globally stable for the autonomous system x=fx and Lx is the Lyapunov function.

Theorem 3. The system (14) has an endemic equilibrium P*=S*,V*,E*,Q*,I*,M* that satisfies S*>0,V*>0,E*>0,Q*>0,I*>0,and M*>0. If Re>1, the endemic equilibrium point globally asymptotically stable in a positive region R+7 on H with

Λ=ZβsS*+ω1S*σV*δ1Q*,ω2=ϑS*KβvV*V*,ω3=ZβsS*+βvV*E*,ω4=ŋS*+φE*Q*,ω5=γE*I*,ω6=δ2Q*+αI*M*.(18)

Proof. We define the Lyapunov function L and its endemic equilibrium as follows:

L=kiSS*S*lnSS*+VV*V*lnVV*+EE*E*lnEE*+QQ*Q*lnQQ*+II*I*lnII*+MM*M*lnMM*,(19)

from Eq. (19), we have ki>0 and i=1,2,3,4,5,6.

dLdt=ki1S*SS+1V*VV+1E*EE+1Q*QQ+1I*II+1M*MM.(20)

Substituting (2) and (18) into (20) gives

dLdt=k11S*SZβsS*+ω1S*σV*δ1Q*+σV+δ1Qρ1Sω1S+k21V*VϑSZβvVϑS*ZβvV*V*V+k31E*EZβsS+ZβvVZβsS*+βvV*E*E+k41Q*QŋS+φEŋS*+φE*Q*Q+k51I*IγEγE*I*I+k61M*Mδ2Q+αIδ2Q*+αI*M*M.(21)

Solving the (21) value gives

dLdt=k1ω12SS*S*SS*+k1ZβsS*3S*SEE*SE*S*E+k4ŋS*3SS*QQ*S*Q*SQ+k1σV*3SS*V*VS*VSV*+k1δ1Q*3SS*Q*QS*QSQ*+k2k3ZβvV*3V*VEE*VE*V*E+k4φE*3E*EQQ*EQ*E*Q+k6αI*3I*IMM*IM*I*M+k5γE*3E*EII*EI*E*I+k6δ2Q*3Q*QMM*QM*Q*M.(22)

Let f1=SS*,f2=VV*,f3=EE*,f4=QQ*,f5=II*,and f6=MM*.

k1,k2,k3,k4,k5,and k6 are obtained by setting the coefficients of f1,f2,f3,f4,f1f4 and f2f4 equal to 0, which after solving gives

k1=k3=1,k2=σV*ϑS*ZβvV*,k4=δ1Q*ŋS*+φE*,k5=ZβsS*+βvV*γE*,and k6=δ1Q*δ2Q*+αI*.

Hence, the result is

dLdt=ω12f11f1S*+ZβsS*31f1f3f1f3+σV*3f11f2f2f1+δ1Q*3f11f4f4f1+k2ZβvV*31f2f3f2f3+k4φE*31f3f4f3f4+k5γE*31f3f5f3f5+k6αI*31f5f6f5f6+k6δ2Q*31f5f6f5f6.(23)

From the above Eqs (22, 23), it can be realized that if the arithmetic mean is greater than or equal to their geometric mean, then dLdt0.dLdt=0 only holds if f1=f2=f3=f4=f5=f6. It implies that S=S*,V=V*,E=E*,Q=Q*,and I=I*=M=M*in H; therefore, the largest invariant set is S,V,E,Q,I,MH:dLdt=0. The endemic equilibrium P* is globally asymptotically stable in the positive region R+7 when Re>1 based on the Lyapunov–LaSalle stability theorem.

4 Model parameterization and sensitivity analysis

We evaluate the reproduction number to find the transmission routes of the pandemic and the proportion of individuals. A sensitivity analysis was carried out to evaluate the contribution of each parameter to the reproduction number.

4.1 Estimation of model parameters

We evaluate the reproduction numbers using the values in Table 3. The effective reproduction number Re=0.85575 and a proportion of 0.19998 of the exposed individuals move to the infected class at any given time. The contribution from each transmission route is shown in the table below.

Table 2 reveals that the infectious and exposed classes contribute significantly to the transmission of the disease. This implies that there must be admissible measures to halt the transmission of the disease.

Table 2
www.frontiersin.org

Table 2. Contribution of transmission routes to the effective reproduction number Re.

4.2 Sensitivity analysis

We illustrate the arithmetic behavior of model (2) parameters’ sensitivity. The sensitivity of the parameters in Re for the model is defined in Eq. 24 as

ƤpRe=Rep.pR0%Re%p.(24)
ƤΛRe=1,ƤβsRe=βsμ+σa1,ƤβvRe=βvϑa1,ƤreRe=ReEa2,ƤrqRe=ReQa2,ƤrmRe=ReMa2,ƤεeRe=εeReEω3+ReIω3+ReQω3,ƤεiRe=εiReIω5+ReQω5φδ2rmReM,ƤεmRe=ReMεmω6ReE+ReI+ReM+φrqω3ω4,Ƥδ1Re=δ1ReQω4+ReMω4αγrma2,Ƥδ2Re=δ2ReQω4+ReMω4αγ+φω5rma2,ƤαRe=αReIω5+R0Mω5ReMa2,ƤγRe=γReEω5+ReIω3ω5+ReQω32ω5+ReM1βsω2a1,ƤμRe=μa1ReEω3+R0Iω3+γω3ω52+φrqω32ω4+φrqω3ω42βsa2,ƤσRe=βsσReE+ReI+ReQ+ReMa32ReMa32,ƤϑRe=βvϑReE+ReI+ReQ+ReMa31a2,and ƤφRe=φReEω3+ReIω3+ReQω32ω4+ReMω3a3rqω3ω5δ2rmω3a2,

where a1=βsω2+βvϑ,a2=ReE+ReI+ReQ+ReM,anda3=μ+σ+ϑ.

From the above analysis, it is observed that Re will increase proportionally as the following parameters βs,βv,re,rq,rm increase; on the other hand, R0 decreases proportionally as the following parameters α,γ,σ,φ,μ,ϑ,εe,εi,εm,δ1,δ2 increase since sensitivity indices are sign determined. This can be used to show the numerical importance of the various parameters in Re. Table 3 illustrates the numerical results of the sensitivity index of the various parameters in Re.

Table 3
www.frontiersin.org

Table 3. Values of the sensitivity index of parameters in Re.

Based on the numerical illustrations shown in Table 3, the effective reproduction number Re will change as the parameters change. Sensitivity analysis is interpreted based on the sign associated with the particular parameter. The effective reproduction number Re will decrease as the parameter values with the negative sign increase, while it increases when they decrease. Considering that βs=0.8192, it indicates that βs will increase Re by 81% whenever there is a 1% increment in the transmission rate of the susceptible individuals. This depicts the explanation for all the parameter values with a positive sign. On the other hand, εi=0.1546 indicates that εi will decrease Re by 15% whenever there is a 1% increment in the recovery rate of the infected individuals. This depicts the explanation for all the parameter values with a negative sign.

Now, we examine the relationship between the parameters of effective reproduction number Re by checking the partial rank correlation coefficients (PRCCs) of the parameters. The following figure illustrates the behavioral pattern of each parameter in the transmission dynamics of the disease.

From Figure 2, one could realize that βs,βv,re,rq,γ,φ have a high positive effect on Re, while μ,ϑ,εe have a high negative effect on Re. One could realize that all these parameters are associated with vaccination and interventions to mitigate the infectivity rate in the population.

Figure 2
www.frontiersin.org

Figure 2. Partial rank correlation coefficient (PRCC) of effective reproduction number Re parameters.

4.3 Effects of vaccination and infectivity reductions on the effective reproduction number Re

This subsection explores the dynamics of Re with respect to vaccination and infectivity reduction parameters ϑ,σ,re,rq,rm. The following figure demonstrates the changing effects of vaccination and infectivity reduction parameters on Re.

Figure 3A demonstrates the influence of vaccination rate ϑ and vaccine inefficacy σ. One can realize that the effective reproduction number Re increases as the vaccination rate ϑ decreases, while Re decreases as vaccine inefficacy σ decreases. Figure 3Bg show the effects of vaccination rate ϑ, vaccine inefficacy σ, and infectivity reductions. It is observed that to mitigate the disease’s transmission, there should be effective control measures to minimize the infectivity levels and maximize the vaccination rate to halt the disease’s transmission. Figures 3H–J show the effect of the infectivity reductions on Re. It could be observed that an increase in re at any given time results in an increase in Re (Figure 3H; Figure 3I). It can be seen that re and rq have a significant effect on Re, as demonstrated in Figure 2. Based on the above graphical representation, it is important to minimize the infectivity levels of the exposed class re and quarantined class rq to halt the disease’s transmission.

Figure 3
www.frontiersin.org

Figure 3. Effect of vaccination and infectivity reductions on the effective reproduction number Re. (A) a plot of σ and θ, (B) re and θ, (C) rq and θ, (D) rm and θ, (E) re and σ, (F) rq and σ, (G) rm and σ, (H) rq and re (I) rm and re (J) rm and rq. The meaning of the respective parameters is in Table 1.

5 Numerical simulations and discussions

In this section, the real data are observed to depict the situation of the transmission. The real data are compared with the model’s solution accuracy. The proposed model is applied to explore the transmission of COVID-19 disease in Ghana using data from the WHO [17]. This includes the daily number of infections and the total number of daily infections after the introduction of vaccines. The data are analyzed and compared to other literature works for the simulations. The daily infections and the total number of infections are simulated to analyze the behavioral pattern of the transmission of the disease. In particular, we illustrate the changing effects of the vaccination rate ϑ, vaccine inefficacy σ , and infectivity reductions on the size of infectious individuals.

5.1 Application of the SVEQIMR model

Once the model is formulated, it is necessary to compare it with data to check its validity. Here, we want to check the accuracy and authenticity of the model by verifying the extent to which the model can represent the real situation, as described in [18]. The simulation consists of the application of the data from Ghana to illustrate the transmission of COVID-19 for the period 1st January 2022 to 1st March 2022 when the individuals were vaccinated. The following state variables are considered using data from Ghana for the period 1st January 2022 to 1st March 2022 [17]. The estimated total population of Ghana is 31732129 [19]; hence, N0=31732129, and the assumed initial values are as follows: S0=200000,V0=120000,E0=150000,I0=997,Q0=1000,M0=800,and R0=500. All the parameters used for the simulations are shown in Table 3.

The SVEQIMR model (fitted) depicts the pattern of the real situation (real data), as shown in Figure 4A, which represents the plot of the daily number of infections, and Figure 4B represents the total number of infections for the period of 1st January 2022 to 1st March 2022. The plot of the model results indicates the pattern of the real situation. That is, the fitted model and real situation agree with each other and illustrate the transmission dynamics of the spread of the pandemic.

Figure 4
www.frontiersin.org

Figure 4. Plot of (A) the daily cases and (B) the total number of cases of the fitted model verses the real data from 1st January to 1st March 2022.

5.2 Effects of vaccination and infectivity reductions on COVID-19 incidences

Here, we analyze the influence of vaccination rate and vaccine inefficacy on the spread of COVID-19 disease using the same dataset. Let us consider the following parameters: vaccination rates ϑ=0,0.15,0.25,0.35,0.45 and vaccine inefficacy σ=0,0.02,0.05,0.07,0.09 for the infected class.

Again, we demonstrate the changing effects of the different infectivity reductions on the behavioral pattern of the transmission of COVID-19 disease by considering the following arbitrary values 0.05,0.10,0.15,0.20,0.25 for the infectivity reduction of exposed individuals re, quarantined individuals rq, and isolated individuals rm. The results are presented in Figure 5.

Figure 5
www.frontiersin.org

Figure 5. Plot of (A) the impact of vaccination rate ϑ on COVID-19 daily cases (B) and the impact of vaccine inefficacy σ, (C) infectivity reduction of exposed individuals re, (D) infectivity reduction of quarantined individuals rq, and (E) infectivity reduction of isolated individuals rm on COVID-19 daily cases.

From Figure 5A, it is observed that an increase in the vaccination rate ϑ results in a significant decrease in the daily number of reported cases. On the contrary, an increase in vaccine inefficacy σ results in an increase in the daily number of reported cases, as shown in Figure 5B. It could also be observed that the disease dies out gradually with time as vaccine inefficacy σ approaches zero (0); however, a high vaccination rate ϑ reduces the number of infections. Therefore, there is a significant effect of different values of vaccination rate ϑ and vaccine inefficacy σ on the disease’s spread. The above illustration shows that vaccination against the spread of the coronavirus is very important and must be adhered to.

From Figure 5C, it is realized that an increase in the infectivity reduction of exposed individuals re results in a significant increase and higher peaks in the daily reported cases. Furthermore, an increase in the infectivity reduction of the quarantined class rq results in a moderate increase in the daily reported cases compared to the peaks of re, as shown in Figure 5D. The peaks of daily reported cases are lower than those of the exposed class because the individuals in this group are confined and monitored. In Figure 5E, there is a significant decline in the daily reported cases as the infectivity reduction of the isolated individuals rm decreases, which has fewer peaks than the others. From the above graphical representation, the infectivity reduction of the exposed individuals re should be controlled carefully because the higher the infectivity reduction, the higher the number of daily reported cases and vice versa. This graphical representation can be confirmed by the numerical illustration of the contribution of the transmission route in Table 2. The exposed compartment contributed significantly to the effective reproduction number Re, with a proportion of 0.19998 of the individuals moving to the infected class. This means that approximately 20% of individuals in the exposed compartment become infected at a given time. The above diagram depicts the order of contribution of the transmission route to the effective reproduction number Re.

6 Optimal control problem and cost–benefit analysis

We modify Equation 2 with the following optimal control variables: κpt, representing the public awareness of the prevalence of COVID-19 and related infections where κpt0,1, which reduces the forces of infection ρ1 and ρ2, by 1κpt; κvt, denoting the control effort to intensify COVID-19 vaccination; and κit, denoting the control effort for infectivity reduction, which reduces infectivity by 1κit. All these efforts denote the control interventions in minimizing the transmission of COVID-19 and its reinfections. The modified Eq. (2) is given in Eq. (25) as follows:

dSdt=dSdt=Λ+σV+δ1Q+τR1κpρ1Sμ+κvϑ+ŋS,dVdt=κvϑS1κpρ2Vμ+σV,dEdt=1κpρ1S+ρ2Vμ+1κiφ+1κiγ+εeE,dQdt=ŋS+1κiφEμ+δ1+1κiδ2Q,dIdt=1κiγEμ+α+εiI,dMdt=1κiδ2Q+αIμ+εmM,dRdt=εeE+εiI+εmMμ+τR,(25)

where ρ1 and ρ2 remain as in Section 2 with the initial conditions given in the model (2).

6.1 Objective functional

We now formulate the optimal trajectories that show the effect of the control efforts κpt,κvt,andκit subjected to (36); the objective functional Q is given as

Qκp,κv,κi=0tfh1E+h2I+h3Q+h4M+12z1κp2t+12z2κv2t+12z3κi2tdt.(26)

We focus on minimizing the cost function (26), and the total cost of implementing the optimal control is given as

Z=0tf12z1κp2t+12z2κv2t+12z3κi2tdt.(27)

The parameters z1,z2,and z3 in Eq. 27 are the balancing cost factors for κpt,κvt,andκit, respectively. All the control efforts κpt,κvt,and κit are assumed to be bounded by Lebesgue measurable time-dependent functions on the interval 0,tf, where tf is the final time with the control effort set defined in Eq. 28 as

Γ=κp,κv,κi for 0κp,κv,κi1,,0ttf.(28)

Now, we establish point-wise Hamiltonian H through Pontryagin’s maximum principle to transform the optimal control system (25) and its associated objective functional (26). The following optimal solution is achieved.

H=h1E+h2I+h3Q+h4M+12z1κp2t+12z2κv2t+12z3κi2t+λSΛ+σV+δ1Q+τR1κpρ1Sμ+κvϑ+ŋS+λVκvϑS1κpρ2Vμ+σV+λE1κpρ1S+ρ2Vμ+1κiφ+γ+εeE+λQŋS+1κiφEμ+δ1+1κiδ2Q+λIγEμ+α+εiI+λM1κiδ2Q+αIμ+εmM+λRεeE+εiI+εmMμ+τR,(29)

where λS,λV,λE,λQ,λI,λM,and λR in Eq. 29 are the co-state variables with respect to the state variables, S,V,E,Q,I,M,and R.

Theorem 4. Given κp*t,κv*t,and κi*t as the optimal controls and the corresponding solutions S0,V0,E0,Q0,I0,M0,and R0 of the system (25), which minimizes Zκpt,κvt,κit over Γ, then there exist co-state variables λS,λV,λE,λQ,λI,λM,andλR that satisfy

dλjdt=Hj,(30)

with conditions λjtf=0, where j=S,V,E,Q,I,M,R. Then, the optimality conditions that minimize the Hamiltonian, H, of (29) with respect to the controls are given as

κp*t=minκpmax,max0,λEλSρ10S0+λEλVρ20V0z1,κv*t=minκvmax,max0,λSλVϑS0z2κi*t=minκimax,max0,λQλEφE0+λIλEγE0+λMλQδ2Q0z3.(31)

Proof: We formulate the adjoint equation for the optimal system by taking the partial derivative of Eq. 30 as follows in Eq. (32):

dλSdS=1κpλSλEρ1+λVλSκvϑ+λSλVσV+λQλSŋ+λSλQδ1Q+λSλRτR+μλS,dλVdV=1κpλEλVρ2+λVλSκvϑS+λSλVσ+μλV,dλEdE=h1+1κpλEλSρ1S+λEλVρ2V+λQλE1κiφ+λIλE1κiγ+λRλEεe+μλE,dλQdQ=h2+λQλSŋS+λSλQδ1+λQλE1κiφE+λMλQ1κiδ2+μλQ,dλIdI=h3+1κiλIλEγE+λMλIα+λRλIεi+μλI,dλMdM=h4+1κiλMλQδ2Q+λMλIαI+λRλMεm+μλM,dλRdR=λRλEεeE+λRλIεiI+λRλMεmM+λSλRτ+μλR.(32)

The control set illustrates the co-state system with the optimal conditions.

Hκp=z1κp+λSλEρ10S0+λVλEρ20V0,Hκv=z2κv+λVλSϑS0,Hκi=z3κi+λEλQφE0+λEλIγE0+λQλMδ2Q0.(33)

We solve for κpt,κvt, and κit as κp*t,κv*t, and κi*t of Eq. (33), and the results confirms the expression in Eq. (31) are as follows:

κp*t=minκpmax,max0,λEλSρ10S0+λEλVρ20V0z1,κv*t=minκvmax,max0,λSλVϑS0z2κi*t=minκimax,max0,λQλEφE0+λIλEγE0+λMλQδ2Q0z3.(34)

Therefore, using the bounds of the controls κpt,κvt, and κit, the control efforts are in the compact form given by the optimal condition of the system in Eq. (34); hence, the proof is complete.

6.2 Optimal control strategies

Here, our aim is to determine the number of infections after deploying the optimal control interventions. We explore the effects of implementing the interventions; therefore, the optimality system (36) is solved forward in time and the adjoint system backward in time with the corresponding lower and upper bounds of the controls. We used the population of Ghana to study the behavioral pattern of COVID-19. The estimated total population of Ghana is 31732129 [19]; hence, N0=31732129, and the assumed initial values are as follows: S0=200000,V0=120000,E0=150000,I0=997,Q0=1000,M0=800,and R0=500, together with Λ = 1,364, τ=0.2, and ŋ=0.21 and parameter values illustrated in Table 3. The balance costs associated with the objective functional are assumed to be z1=5,z2=10,and z3=20, and weight hi=100, where i=1,2,3,4. The lower bound (LB) and upper bound (UB) are assumed to be LB1=0,UB1=1,LB2=0,UB2=1,LB3=0,and UB3=1. The results are illustrated according to the strategies to implement the control strategies.

6.2.1 Strategy 1: implementation of public awareness (κp)

The optimal solutions illustrated in Figure 6 account for the observations when the control effort κp is applied accordingly.

Figure 6
www.frontiersin.org

Figure 6. Optimal solutions of implementing strategy 1. Plot (A) blue dotted line is the optimal solution for implementing strategy 1. (B) Optimal control profile for strategy 1.

The optimal solutions illustrated above depict the following observations when public education is only applied:

(a) Figure 6A represents the effect of the control effort κp on the infectious individuals. It implies that the number of individuals will decrease if the control effort is optimally implemented in halting the disease’s transmission. Conversely, it will increase significantly.

(b) Figure 6B represents the profile of the control effort for public awareness of COVID-19. It implies that education on COVID-19 should reach 50% of the population throughout the implementation to halt COVID-19 transmission.

6.2.2 Strategy 2: implementation of vaccination (κv)

The optimal solutions illustrated in Figure 7 account for the observations when the control effort κv is applied accordingly.

Figure 7
www.frontiersin.org

Figure 7. Optimal solutions of implementing strategy 2. Plot (A) blue dotted line is the optimal solution for implementing strategy 2. (B) Optimal control profile for strategy 2.

The optimal solutions illustrated above depict the following observations when vaccination is only applied:

(a) Figure 7A represents the effect of the control effort κv on the infectious individuals. It implies that the number of individuals will decrease if the control effort is optimally implemented in halting the disease’s transmission. Conversely, it will increase significantly.

(b) Figure 7B represents the profile of the control effort for vaccination to prevent COVID-19. It implies that approximately 25% of the population should be vaccinated within 80 days and intensified further up to 75% in the subsequent days throughout the implementation to halt COVID-19 transmission.

6.2.3 Strategy 3: implementation of infectivity treatment (κi)

The optimal solutions illustrated in Figure 8 account for the observations when the control effort κi is applied accordingly.

Figure 8
www.frontiersin.org

Figure 8. Optimal solutions of implementing strategy 3. Plot (A) blue dotted line is the optimal solution for implementing strategy 3. (B) Optimal control profile for strategy 3.

The optimal solutions illustrated above depict the following observations when infectivity treatment is only applied:

(a) Figure 8A represents the effect of the control effort κi on the infectious individuals. It implies that the number of individuals will decrease if the control effort is optimally implemented in halting the disease’s transmission. Conversely, it will increase significantly.

(b) Figure 8B represents the profile of the control effort for COVID-19 reinfection. It implies that approximately 75% of the individuals suspected to be carriers of the virus should be treated/monitored throughout the implementation to halt COVID-19 transmission.

6.2.4 Strategy 4: implementation of public awareness and vaccination (κp,κv)

The optimal solutions illustrated in Figure 9 account for the observations when the control efforts κp and κv are applied accordingly.

Figure 9
www.frontiersin.org

Figure 9. Optimal solutions of implementing strategy 4. Plot (A) blue dotted line is the optimal solution for implementing strategy 4. (B) Optimal control profile for strategy 4.

The optimal solutions illustrated above depict the following observations when control efforts for public awareness and vaccination are applied:

(a) Figure 9A represents the effect of the control efforts κp and κv on the infectious individuals. It implies that the number of individuals will decrease if the control effort is optimally implemented in halting the disease’s transmission. Conversely, it will increase significantly.

(b) Figure 9B represents the profile of the control efforts for public awareness and vaccination against COVID-19. It implies that approximately 80% of the population should be educated on COVID-19 within 22 days, which can be relaxed to approximately 25%, and 50% of the population should be vaccinated throughout the implementation of these interventions to halt COVID-19 transmission.

6.2.5 Strategy 5: implementation of public awareness and infectivity treatment (κp,κi)

The optimal solutions illustrated in Figure 10 account for the observations when all the control efforts κp and κi are applied accordingly.

Figure 10
www.frontiersin.org

Figure 10. Optimal solutions of implementing strategy 5. Plot (A) blue dotted line is the optimal solution for implementing strategy 5. (B) Optimal control profiles for strategy 5.

The optimal solutions illustrated above depict the following observations when all the control efforts for public awareness and infectivity treatment are applied:

(a) Figure 10A represents the effect of the control efforts κp and κi on the infectious individuals. It implies that the number of individuals will decrease to a minimum within 30 days if the control effort is optimally implemented in halting the disease’s transmission. Conversely, it will increase significantly.

(b) Figure 10B represents the profile of the control efforts for public awareness and infectivity treatment. It implies that approximately 80% of the population should be educated on COVID-19, and 50% of suspected carriers of the virus should be treated/monitored throughout the implementation of these interventions to halt COVID-19 transmission.

6.2.6 Strategy 6: implementation of vaccination and infectivity treatment (κv,κi)

The optimal solutions illustrated in Figure 11 account for the observations when the control efforts κv and κi are applied accordingly.

Figure 11
www.frontiersin.org

Figure 11. Optimal solutions of implementing strategy 6. Plot (A) blue dotted line is the optimal solution for implementing strategy 6. (B) Optimal control profiles for strategy 6.

The optimal solutions illustrated above depict the following observations when all the control efforts for vaccination infectivity reduction are applied:

(a) Figure 11A represents the effect of the control efforts κv and κi on the infectious individuals. It implies that the number of individuals will decrease to a minimum within 20 days if the control effort is optimally implemented in halting the disease’s transmission. Conversely, it will increase significantly.

(b) Figure 11B represents the profile of the control efforts for public education and vaccination against COVID-19. It implies that all the interventions U1 and U2 should be implemented at levels higher than 25% from the start of implementation throughout the subsequent days to halt COVID-19 transmission.

6.2.7 Strategy 7: implementation of all controls (κp,κv,κi)

The optimal solutions illustrated in Figure 12 account for the observations when the control efforts κp,κv,and κi are applied accordingly.

Figure 12
www.frontiersin.org

Figure 12. Optimal solutions of implementing strategy 7. Plot (A) blue dotted line is the optimal solution for implementing strategy 7. (B) Optimal control profiles for strategy 7.

The optimal solutions illustrated above depict the following observations when all the control efforts are applied:

(a) Figure 12A represents the effect of the control efforts κp,κv,and κi on the infectious individuals. It implies that the number of individuals will decrease to a minimum within 10 days if the control effort is optimally implemented in halting the disease’s transmission. Conversely, it will increase significantly.

(b) Figure 12B represents the profile of all control efforts. It implies that approximately 80% of the population should be educated on COVID-19, approximately 50% of the population should be vaccinated, and 50% of suspected carriers of the virus should also be treated/monitored throughout the implementation period of these interventions to halt COVID-19 transmission.

6.3 Cost–benefit analysis

Once the strategies are given, it is imperative to know the cost associated with implementing such intervention(s). Therefore, we explore the cost associated with each control strategy to check their effectiveness. We outline some cost-effectiveness approaches to further understand the control strategies.

We consider two procedures, namely, average cost-effectiveness ratio (ACER) and incremental cost-effectiveness ratio (ICER), which have been explained in [2023], to carry out epidemiological studies.

6.4 Average cost-effectiveness ratio

We define the ACER of implementing a strategy as

ACER=OverallcostgeneratedbyapplyingthestrategyOverallinfectionavertedbyapplyingthestrategy.(35)

The overall cost Z stated in (27) would be used to evaluate the total cost that the intervention would generate in Eq. 36. We then compare the ACER values of each strategy, and the one with the lowest value is the most cost-effective, saving costs. Therefore, the cost-effective intervention is considered the strategy with the least ACER value. The expression in Eq. 35 is illustrated as follows.

From Table 4, control strategy 1, which involves the implementation of public education only, has the least value of ACER, indicating cost savings. However, relying solely on this metric is not enough to choose a strategy; hence, we further explore other approaches.

Table 4
www.frontiersin.org

Table 4. Strategies’ ACER values with their overall infection averted and cost incurred.

6.5 Incremental cost-effectiveness ratio

We define the ICER of implementing a strategy as

ICER=ThecostdifferencegeneratedbystrategiesxandyDifferenceintheoverallinfectionavertedinstrategiesxandy.(36)

The total cost function Z stated in (27) would be used to estimate the overall cost that the intervention would generate in Eq. (36). It is worth knowing that the averted total number of infections is the difference between the initial values of Ex and Ix, without control(s) and with controls. The outcomes are tabulated in increasing order of infection averted.

The ICER in Table 5 is calculated as

ICER2=1.2513×10302.0184×1070=6.1995×105,
ICER1=1.4064×1031.2513×1031.5269×1082.0184×107=1.1705×106,
ICER4=1.5279×1031.4064×1031.5279×1081.5269×108=0.0012,
ICER3=5.6206×1031.5279×1033.9351×1083.9351×108=1.7002×105,
ICER5=6.2463×1035.6206×1033.9388×1082.3231×108=0.0017,
ICER6=6.8719×1036.2463×1035.2611×1083.9388×108=4.7312×106,
ICER7=7.4975×1036.8719×1035.2648×1085.2611×108=0.0017.

Table 5
www.frontiersin.org

Table 5. Strategies’ ICER values with their overall infection averted and cost incurred.

Assessing strategies 2 and 1 in Table 5, it is noticed from the ICER that strategy 2 is expensive to deploy in a resource-limited setting; therefore, strategy 2 is removed from the list of possible controls, and the ICER is calculated again. This is presented in Table 6.

Table 6
www.frontiersin.org

Table 6. Strategies’ ICER values with their overall infection averted and cost incurred.

Assessing strategies 1 and 4 in Table 6, it is noticed from the ICER that strategy 4 is expensive to deploy in a resource-limited setting; therefore, strategy 4 is removed from the list of possible controls, and the ICER is calculated again. This is presented in Table 7.

Table 7
www.frontiersin.org

Table 7. Strategies’ ICER values with their overall infection averted and cost incurred.

Assessing strategies 1 and 3 in Table 7, it is noticed from the ICER that strategy 3 is expensive to deploy in a resource-limited setting; therefore, strategy 3 is removed from the list of possible controls, and the ICER is calculated again. This is presented in Table 8.

Table 8
www.frontiersin.org

Table 8. Strategies’ ICER values with their overall infection averted and cost incurred.

Assessing strategies 1 and 5 in Table 8, it is noticed from the ICER that strategy 5 is expensive to deploy in a resource-limited setting; therefore, strategy 5 is removed from the list of possible controls, and the ICER is calculated again. This is presented in Table 9.

Table 9
www.frontiersin.org

Table 9. Strategies’ ICER values with their overall infection averted and cost involved.

Assessing strategies 1 and 6 in Table 9, it is noticed from the ICER that strategy 6 is expensive to deploy in a resource-limited setting; therefore, strategy 6 is removed from the list of possible controls, and the ICER is calculated again. This is presented in Table 10.

Table 10
www.frontiersin.org

Table 10. Strategies’ ICER values with their overall infection averted and cost incurred.

Finally, assessing strategies 1 and 7 in Table 10, it is noticed from the ICER that strategy 7 is expensive to deploy in a resource-limited setting; therefore, strategy 7 is removed from the list of possible controls. Therefore, we conclude that strategy 1 is the most cost-effective strategy to use among the several strategies under study here. From the above analysis, it is obvious that strategy 1, which involves public education, is the intervention that saves cost.

7 Conclusion

We have presented a work that analyzes the changing effects of vaccination and infectivity reductions on the transmission of COVID-19 using data from Ghana. We have estimated the model’s parameters and analyzed their effects on disease transmission through numerical and graphical illustrations. Again, we have exhibited the threshold dynamics of the effective reproduction number Re together with the contributions from the transmission routes (Table 2). We have demonstrated the sensitivity of the model’s parameters to study their effects on the effective reproduction number Re (Table 3; Figure 2; Figure 3).

The aim of this work is to study the effect of vaccination and infectivity reductions in controlling COVID-19 transmission and devise control interventions that save cost to mitigate the transmission; therefore, we have formulated optimal control strategies together with the cost–benefit analysis that consider control measures involving both pharmaceutical and non-pharmaceutical interventions to control COVID-19. We implemented the strategies (Figure 6AFigure 12A), and it was realized that public education, vaccination, and infectivity reductions to prevent COVID-19 should be intensified and reach approximately 25% of the population from the beginning and intensified in the subsequent days (Figure 6BFigure 12B).

It is also worth knowing that public education saves cost as per the cost–benefit analysis compared to the other strategies raised in this work. This intervention can minimize COVID-19, as illustrated in Figure 6A. This intervention should reach approximately 50% of the population throughout the period of its implementation in order to realize the results of strategy 1 (Figure 6). Although strategy 1 saves cost, other strategies elaborated in this work can also be applied, but one has to consider the cost involved in implementing the strategy. The cost involved in applying the optimal control strategies is presented in Table 4Table 10.

The outcomes of the findings imply that both pharmaceutical and non-pharmaceutical measures are very important in controlling the transmission of COVID-19. These control measures should always be vigorously implemented to create public awareness on COVID-19 and its reinfection, as illustrated in Figure 6BFigure 12B, in order to reduce the effective contact rates and rates of acquiring COVID-19, as illustrated in Figure 3.

Although we have demonstrated the dynamics of COVID-19 transmission with vaccination and different infectivity reductions, this work is focused on the homogeneity of the population, and we hope to extend this study to explore the transmission dynamics of COVID-19 reinfection by considering heterogeneity of the population such as age and sex. We encourage individuals to adhere to personal hygiene and be aware of COVID-19 reinfection.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary Material; further inquiries can be directed to the corresponding authors.

Author contributions

RA: conceptualization, data curation, formal analysis, investigation, methodology, writing–original draft, and writing–review and editing. ZJ: funding acquisition, supervision, and writing–review and editing. JY: formal analysis, supervision, and writing–review and editing. JA: formal analysis, investigation, and writing–review and editing.

Funding

The author(s) declare that financial support was received for the research, authorship, and/or publication of this article. This research was supported by the National Natural Science Foundation of China grants 12231012 and 61873154.

Acknowledgments

The authors appreciate the support of Complex Systems Research Center, Shanxi University.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

1. World Health Organization. Coronavirus disease 2019(COVID-19), situation report -51, data as reported by 8 March 2020 (2019). Available at: https://www.who.int/emergencies/disease/novel-coronavirus-2019/situation-reports (Accessed January 10, 2024).

Google Scholar

2. Chen Y, Wang Y, Fleming J, Yu Y, Gu Y, Liu C, et al. Active or latent tuberculosis increases susceptibility to COVID-19 and disease severity (2020). Available at: https://www.medrxiv.org/ (Accessed January 10, 2024).

Google Scholar

3. Salman AM, Ahmed I, Mohd MH, Jamiluddin MS, Dheyab MA. Scenario analysis of COVID-19 transmission dynamics in Malaysia with the possibility of reinfection and limited medical resources scenarios. Comput Biol Med (2021) 133:104372. doi:10.1016/j.compbiomed.2021.104372

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Zamir M, Nadeem F, Alqudah MA, Abdeljawad T. Future implications of COVID-19 through Mathematical modeling. Results Phys (2022) 33:105097. doi:10.1016/j.rinp.2021.105097

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Worldometers. Reported cases and deaths by country or territory (2022). Available at: https://www.worldometers.info/coronavirus/#countries (Accessed April 30, 2022).

Google Scholar

6. Zegarra MAA, Infante SD, Carrasco DB, Liceaga DO. COVID-19 optimal vaccination policies: a modeling study on efficacy, natural and vaccine-induced immunity responses. Math Biosci (2021) 337:108614. doi:10.1016/j.mbs.2021.108614

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Adachi A. Grand challenge in human/animal virology: unseen, smallest replicative entities shape thewholeglobe. FrontMicrobiol (2019) 11:431. doi:10.3389/fmicb.2020.00431

PubMed Abstract | CrossRef Full Text | Google Scholar

8. 1mg. Coronavirus No-panic help guide (2020). Available at: https://smef.org.uk/wpcontent/uploads/2020/03/Corona-Ebook.pdf. (Accessed January 10, 2024)

Google Scholar

9. Annas S, Pratama MI, Rifandi M, Sanusi W, Side S, Stability analysis and numerical simulation of SEIR model for pandemic COVID-19 spread in Indonesia. Chaos, Solitons and Fractal (2020) 139:110072. doi:10.1016/j.chaos.2020.110072

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Pang L, Liu S, Zhang X, Tian T, Zhao Z, Transmission dynamics and control strategies of COVID-19 in Wuhan, China. J Biol Syst (2020) 28:543–60. doi:10.1142/s0218339020500096

CrossRef Full Text | Google Scholar

11. Asamoah JKK, Owusu MA, Jin Z, Oduro F, Abidemi A, Gyasi EO, Global stability and cost-effectiveness analysis of COVID-19 considering the impact of the environment: using data from Ghana. Chaos, Solitons and Fractals (2020) 140:110103. doi:10.1016/j.chaos.2020.110103

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Tchoumi S, Diagne M, Rwezaura H, Tchuenche J, Malaria and COVID-19 co-dynamics: a mathematical model and optimal control. Appl Math Model (2021) 99:294–327. e327. doi:10.1016/j.apm.2021.06.016

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Yan X, Zou Y, Optimal and sub-optimal quarantine and isolation control in SARS epidemics. Math Comp Model (2008) 47(1-2):235–45. doi:10.1016/j.mcm.2007.04.003

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Khan MA, Atangana A, Mathematical modeling and analysis of COVID-19: a study of new variant omicron. Physica A: Stat Mech Its Appl (2022) 599:127452. Article 127452. doi:10.1016/j.physa.2022.127452

CrossRef Full Text | Google Scholar

15. Rwezaura H, Diagne M, Omame A, de Espindola A, Tchuenche J, Mathematical modeling and optimal control of SARS-CoV-2 and tuberculosis co-infection: a case study of Indonesia. Model Earth Syst Environ (2022) 8(4):5493–520. e5520. doi:10.1007/s40808-022-01430-6

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Omame A, Abbas M, Onyenegecha C, A fractional-order model for COVID-19 and tuberculosis co-infection using Atangana–Baleanu derivative. Chaos, Solitons and Fractals (2021) 153:111486. Article 111486. doi:10.1016/j.chaos.2021.111486

PubMed Abstract | CrossRef Full Text | Google Scholar

17. World Health Organization. World health organization covid– 19 report (2024). Available at: https://covid19.who.int/region/afro/country/gh. (Accessed January 10, 2024).

Google Scholar

18. Dietz K. The estimation of the basic reproduction number for infectious diseases. Stat Methods Med Res (1993) 2:23–41. doi:10.1177/096228029300200103

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Osei E, Amu H, Kye-Duodu G, Kwabla MP, Danso E, Binka FN, et al. Impact of COVID-19 pandemic on Tuberculosis and HIV services in Ghana: an interrupted time series analysis. PLoS ONE (2023) 18(9):e0291808. doi:10.1371/journal.pone.0291808

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Asamoah JKK, Owusu MA, Jin Z, Oduro FT, Abidemi A, Gyasi EO, Global stability and cost-effectiveness analysis of COVID-19 considering the impact of the environment: using data from Ghana. Chaos, Solitons and Fractals (2020) 140:110103. doi:10.1016/j.chaos.2020.110103

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Agusto F, Leite M, Optimal control and cost-effective analysis of the 2017 meningitis outbreak in Nigeria. Infect Dis Model (2017) 4(2019):161–87. doi:10.1016/j.idm.2019.05.003

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Asamoah JKK, Okyere E, Abidemi A, Moore SE, Sun G, Jin Z, et al. Optimal control and comprehensive cost-effectiveness analysis for COVID-19. Results Phys (2022) 33:105177. doi:10.1016/j.rinp.2022.105177

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Asamoah JKK, Jin Z, Sun G, Non-seasonal and seasonal relapse model for Q fever disease with comprehensive cost-effectiveness analysis. Results Phys (2021) 22:103889. doi:10.1016/j.rinp.2021.103889

CrossRef Full Text | Google Scholar

Keywords: COVID-19, vaccination, infectivity reductions, optimal control, cost–benefit analysis

Citation: Appiah RF, Jin Z, Yang J and Asamoah JKK (2024) Cost–benefit analysis of the COVID-19 vaccination model incorporating different infectivity reductions. Front. Phys. 12:1383357. doi: 10.3389/fphy.2024.1383357

Received: 07 February 2024; Accepted: 19 March 2024;
Published: 08 May 2024.

Edited by:

Xiaoke Xu, Dalian Nationalities University, China

Reviewed by:

Xiaofeng Luo, North University of China, China
Segun Oke, Ohio University, United States

Copyright © 2024 Appiah, Jin, Yang and Asamoah. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Zhen Jin, jinzhn@263.net; Junyuan Yang, yjyang66@sxu.edu.cn

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.