
A new transmission model of Zika virus with three transmission routes including human transmission by mosquito bites, sexual transmission between males and females and vertical transmission is established. The basic reproduction number R0 is derived. When R0<1, it is proved that the disease-free equilibrium is globally stable. Furthermore, the optimal control and mitigation methods for transmission of Zika virus are deduced and explored. The MCMC method is used to estimate the parameters and the reasons for the deviation between the actual infection cases and the simulated data are discussed. In addition, different strategies for controlling the spread of Zika virus are simulated and studied. The combination of mosquito control strategies and internal human control strategies is the most effective way in reducing the risk of Zika virus infection.
Citation: Hai-Feng Huo, Tian Fu, Hong Xiang. Dynamics and optimal control of a Zika model with sexual and vertical transmissions[J]. Mathematical Biosciences and Engineering, 2023, 20(5): 8279-8304. doi: 10.3934/mbe.2023361
[1] | Biao Tang, Weike Zhou, Yanni Xiao, Jianhong Wu . Implication of sexual transmission of Zika on dengue and Zika outbreaks. Mathematical Biosciences and Engineering, 2019, 16(5): 5092-5113. doi: 10.3934/mbe.2019256 |
[2] | Minna Shao, Hongyong Zhao . Dynamics and optimal control of a stochastic Zika virus model with spatial diffusion. Mathematical Biosciences and Engineering, 2023, 20(9): 17520-17553. doi: 10.3934/mbe.2023778 |
[3] | Fabio Sanchez, Luis A. Barboza, Paola Vásquez . Parameter estimates of the 2016-2017 Zika outbreak in Costa Rica: An Approximate Bayesian Computation (ABC) approach. Mathematical Biosciences and Engineering, 2019, 16(4): 2738-2755. doi: 10.3934/mbe.2019136 |
[4] | Maghnia Hamou Maamar, Matthias Ehrhardt, Louiza Tabharit . A nonstandard finite difference scheme for a time-fractional model of Zika virus transmission. Mathematical Biosciences and Engineering, 2024, 21(1): 924-962. doi: 10.3934/mbe.2024039 |
[5] | Lan Zou, Jing Chen, Shigui Ruan . Modeling and analyzing the transmission dynamics of visceral leishmaniasis. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1585-1604. doi: 10.3934/mbe.2017082 |
[6] | Chandrani Banerjee, Linda J. S. Allen, Jorge Salazar-Bravo . Models for an arenavirus infection in a rodent population: consequences of horizontal, vertical and sexual transmission. Mathematical Biosciences and Engineering, 2008, 5(4): 617-645. doi: 10.3934/mbe.2008.5.617 |
[7] | Zongmin Yue, Fauzi Mohamed Yusof, Sabarina Shafie . Transmission dynamics of Zika virus incorporating harvesting. Mathematical Biosciences and Engineering, 2020, 17(5): 6181-6202. doi: 10.3934/mbe.2020327 |
[8] | Andrea Pugliese, Abba B. Gumel, Fabio A. Milner, Jorge X. Velasco-Hernandez . Sex-biased prevalence in infections with heterosexual, direct, and vector-mediated transmission: a theoretical analysis. Mathematical Biosciences and Engineering, 2018, 15(1): 125-140. doi: 10.3934/mbe.2018005 |
[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] | Tailei Zhang, Hui Li, Na Xie, Wenhui Fu, Kai Wang, Xiongjie Ding . Mathematical analysis and simulation of a Hepatitis B model with time delay: A case study for Xinjiang, China. Mathematical Biosciences and Engineering, 2020, 17(2): 1757-1775. doi: 10.3934/mbe.2020092 |
A new transmission model of Zika virus with three transmission routes including human transmission by mosquito bites, sexual transmission between males and females and vertical transmission is established. The basic reproduction number R0 is derived. When R0<1, it is proved that the disease-free equilibrium is globally stable. Furthermore, the optimal control and mitigation methods for transmission of Zika virus are deduced and explored. The MCMC method is used to estimate the parameters and the reasons for the deviation between the actual infection cases and the simulated data are discussed. In addition, different strategies for controlling the spread of Zika virus are simulated and studied. The combination of mosquito control strategies and internal human control strategies is the most effective way in reducing the risk of Zika virus infection.
The Zika virus is named after it was first discovered in 1947 in rhesus monkeys in Uganda's Zika forest[1]. The Zika virus is widely distributed in the tropics and subtropics. It is mainly transmitted to humans through the bites of infected female aedes mosquitoes[2]. Many people infected with Zika will experience no symptoms or only mild symptoms. Common Zika symptoms are fever, rash, headache, joint pain, red eyes and muscle pain that can last from a few days to a week[3]. The Zika virus was first detected in Uganda and Tanzania in 1952[4]. Virus activity had been muted and only sporadic cases of Zika virus infection had been reported in equatorial Africa, the Americas, Asia and the Pacific. Since then, there had been a pandemic on the Western Pacific island of Yap in 2007[5]. In 2013 to 2014, there was a larger outbreak in French Polynesia in the Pacific region, which infected about 32,000 people [6]. Zika spread rapidly in Brazil after the first case in South America was detected there in 2015[7]. In October of that year, Brazil reported an increase in microcephaly cases. Pregnant women with Zika risk transmitted the virus to their newborns and gave birth to babies with microcephaly or other birth defects [8]. In addition to mosquito-borne transmission, Zika is transmitted sexually, mainly from men to women. Studies have shown that the Zika virus persists in men's semen longer than other body fluids, up to six months [9]. Zika virus can be transmitted to women several months after a man has recovered. Another mode of transmission that has a high impact is vertical transmission, where the virus is passed from an infected pregnant woman to her newborn. This mode of transmission is likely to result in children being born with microcephaly or other severe fetal brain defects. The virus is also spread in laboratories and through blood transfusions[10].
All sorts of sophisticated models have been built to study the spread of Zika. The possible transmission of Zika virus through sexual transmission was first identified by Foy et al. [11]. Gao et al. [12] established a Zika transmission model with mosquito-borne and sexual transmission of the virus, using the Zika epidemic data from Brazil, Colombia and El Salvador for data simulation. It was concluded that sexual transmission was relatively small contributor to Zika virus transmission but that sexual transmission increases the risk of infection and epidemic size and may prolong outbreaks. He et al. [13] simulated Zika virus infections reported in French Polynesia, Colombia, and the State of Bahia in Brazil. By comparing the simulation results, we would be able to better understand the likely evolution of Zika virus, control the spread of the outbreak and prevent potential transmission. Baca-Carrasco et al. [14] proposed three mathematical models including vector transmission, sexual transmission and population migration. The common conclusion was that the level of endemic disease following Zika virus outbreak was very low and sexual transmission contributed to the extent of the outbreak. Agusto et al. [15] established a transmission model of Zika virus in the absence of disease deaths and performed stability analysis. When expanding the model to include mortality due to Zika virus, different model stability analyses were established and numerical simulations were used to assess the importance of sexual transmission to study the dynamics of the disease. Imran et al. [16] established a model considering the vertical transmission of human and mosquito as well as analyzing its dynamic behavior in detail. Denes et al. [17] established a complex transmission model, which distinguished males and females in the mode of sexual transmission. Data simulations used Zika virus data from Costa Rica and Suriname concluded that mosquito birth and death rates were the most important factors in Zika virus transmission, but sexual transmission also had a significant impact on disease transmission. Ibrahim et al. [18] established a model of Zika virus transmission involving sexually transmitted and asymptomatic carriers. The effects of weather periodicity on mosquito-related parameters were studied. Yuan et al. [19] proposed three modes of transmission of Zika virus, including vector transmission between humans and mosquitoes, transmission by human sexual contact and vertical transmission within mosquitoes. The contribution of each transmission route to the basic reproduction number was analyzed. Numerical simulations confirm that sexual and vertical transmission had different effects on the early and long-term transmission of Zika virus. Busenberg and Cooke [20] earlier discussed the modeling and dynamic analysis of various vertically transmitted diseases. In an S−I−R compartment model, we incorporate vertical transmission into infected category (If,Im) by assuming that a subset of the offspring of infected women (BhλhIfNf) are infectious at birth, while the remaining newborns (Bh−BhλhIfNf) enter the susceptible category. Where Bh represents natural birth rate of humans and λh represents proportion of offspring with congenital infection of infected females. Li et al.[21] used a similar method to consider vertical transmission. Some literatures have conducted detailed studies on the optimal control strategies for the transmission of Zika virus, studied the impact of different control strategy combinations on the transmission of Zika virus in human populations, and obtained some effective strategies for preventing and controlling the transmission of Zika virus disease [22,23].
Few previous studies on Zika transmission have dealt with the vertical transmission in humans. However, the people most affected by Zika infection are newborns. It is of great significance to distinguish the sexes and add neonatal infection for analysis. With numerical simulations, it is clear which control measures are in place to rapidly control the spread of Zika disease, and the changes in actual cases in Colombia show that effective control measures can rapidly control the spread of the disease. Motivated by the above discussion, the goal of this paper is to study the joint influence of mosquito-borne transmission, sexual transmission and vertical transmission in human and mosquitoes on the spread of Zika virus. The sections of this paper are as follows. In Section 2, we introduce a Zika transmission model with three transmission modes: mosquito-borne transmission, sexual contact transmission and vertical transmission. In Section 3, we determine the basic reproduction number of the model and study the local asymptotic stability of the disease-free equilibrium as well as the global stability of the disease-free equilibrium in the case of R0<1. In Section 4, we derive and explore the expression of the optimal control policy. In Section 5, we provide parameter estimates using actual transmission cases in Colombia, explaining why the actual transmission cases do not fit well with the fitted curve. The number of infected cases is simulated under different control strategies. The paper ends with a discussion in last section.
The total human population Nh(t) consists of female human and male human. Female human consists of three compartments: Sf(t), If(t), Rf(t). Sf(t) represents the number of female susceptible individual, If(t) represents the number of female infected individual, Rf(t) represents the number of female recovered individual. Male human consists of four compartments: Sm(t), Im(t), Irm(t), Rm(t). Sm(t) represents the number of male susceptible individuals, Im(t) represents the number of male infected individuals, Irm(t) represents the number of male convalescent individuals who have recovered from the disease but can still transmit it through sexual contact. Rm(t) represents the number of male recovered individual. The total vector population Nv(t) consists of two compartments: Sv(t), Iv(t). Sv(t) represents the number of susceptible mosquitos, Iv(t) represents the number of infected mosquitos. The transmission diagram of the model is shown in Figure 1.
The total human population and the total mosquito population are given by:
Nf(t)=Sf(t)+If(t)+Rf(t),Nm(t)=Sm(t)+Im(t)+Irm(t)+Rm(t),Nv(t)=Sv(t)+Iv(t),Nh(t)=Nf(t)+Nm(t). |
Then the model can be built as follows:
{dSfdt=Bh2−Bh2Y1(t)−Y2(t)Sf−Y3(t)Sf−μhSf,dIfdt=Bh2Y1(t)+Y2(t)Sf+Y3(t)Sf−γIf−μhIf,dRfdt=γIf−μhRf,dSmdt=Bh2−Bh2Y1(t)−Y3(t)Sm−μhSm,dImdt=Bh2Y1(t)+Y3(t)Sm−γIm−μhIm,dIrmdt=γIm−γrIrm−μhIrm,dRmdt=γrIrm−μhRm,dSvdt=Bv−Y5(t)Sv−BvY4(t)−μvSv,dIvdt=Y5(t)Sv+BvY4(t)−μvIv. | (2.1) |
We define the functions Y1(t),Y2(t),Y3(t),Y4(t),Y5(t) by
Y1(t)=λhIfNf,Y2(t)=βIm+krIrmNf,Y3(t)=αhIvNh,Y4(t)=θIvNv,Y5(t)=αvIf+ImNh. |
The model assumes that the population has a constant natural birth rate (Bh) and that the number of males and females is the same, denoting the natural birth rate of males and females by Bh2. When vertical transmission is considered, if a woman is infected with Zika virus during pregnancy, then the individual newborn may also be infected and proportion of congenital infected offspring in infected females is assumed to be λh (see, e.g., [21]). IfNf is the proportion of infected women, so Bh2λhIfNf represents the number of infected newborns, and Bh2−Bh2λhIfNf indicates susceptible female newborns. The same is true of male representations. Male susceptible individuals (Sm) are infected with Zika virus due to the bite of mosquitos infected with Zika virus, the transmission probability of infected mosquitos biting susceptible human is αh, thus entering the category of infected individuals (Im). Male infections decrease due to natural mortality (μh) and cure rates (γ) of infected persons. The male enters the convalescent phase (Irm), during which the virus will not be present in the male blood, but will be present in the male semen for a long period of time (even up to six months). Sexual contact with others during this period of time is a risk of transmitting the virus to others. Due to metabolism, the virus will be metabolized out of the body and eventually transformed into male recovered individual (Rm) at the recovery rate of γr. Female susceptible individuals (Sf) enter the infected category (If) by acquiring the virus through the bite of infected mosquitos with transmission probability (αh) and through sexual contact with transmission probability (β). Female infections decrease due to natural mortality (μh) and cure rates (γ) of infected persons. Recovered women were transferred to the recovered category (Rf).
Similarly, susceptible mosquitoes (Sv) become infected by biting infected human with transmission probability (αv). The proportion of the offspring of infected mosquitoes (Iv) will be infected and assume that the proportion of infected female offspring is θ. Consider the expression of vertical propagation similar to that of humans. It is assumed that the mosquito will not recover after infection. The parameters description of the Zika model are summarized in Table 1.
Parameter | Description(Units) |
Bh | Natural birth rate of humans (days−1) |
μh | Natural death rate of humans (days−1) |
β | Sexual transmission infection rate from infected humans to susceptible humans (days−1) |
αh | Baseline value of transmission rate from mosquitoes to humans(days−1) |
αv | Baseline value of transmission rate from humans to mosquitoes (days−1) |
kr | Relative human-to-human transmissibility of convalescent to symptomatic humans (none) |
γ | Human recovery rate (days−1) |
γr | Recovery rate of convalescent humans (days−1) |
λh | Proportion of offspring with congenital infection of infected females (none) |
Bv | Baseline value of mosquito birth rate (days−1) |
1μv | Mosquito lifespan(days) |
θ | Proportion of offspring congenital infection of infected female mosquitoes (none) |
By reference [24], Theorems 1.1.8 and 1.1.9 of Wiggins [25] for each nonnegative initial condition there is a unique, non-negative solution.
Lemma 1. Let the initial value W(0)≥0, where W(t)=(Sf,If,Rf,Sm,Im,Irm,Rm,Sv,Iv). Then the solutions W(t) of model (2.1) are non-negative for all time t≥0. Furthermore
lim supt→∞Nh(t)=Bhμh,lim supt→∞Nv(t)=Bvμv. |
Proof. Let t1=sup{t>0:W(t)>0}, by the first equation of model (2.1) that
dSfdt=Bh2−Bh2Y1(t)−Y2(t)Sf−Y3(t)Sf−μhSf, |
we get
Sf(t1)=Sf(0)e−(Y2(u)+Y3(u)+μh)t1+e−(Y2(u)+Y3(u)+μh)t1e∫tf0−(Y2(u)+Y3(u))du+μht1∫tf0(Bh2−Bh2Y1(u))du>0. |
It can similarly be shown that W(t)>0 for all t>0. It can be represented as 0<Sf≤Nh(t),0<If≤Nh(t),0<Rf≤Nh(t),0<Sm≤Nh(t),0<Im≤Nh(t), 0<Irm≤Nh(t),0<Rm≤Nh(t),0<Sv≤Nv(t),0<Iv≤Nv(t).
Supplementary model (2.1) are given as
dNhdt=Bh−μhNh. |
dNvdt=Bv−μvNv. |
Hence,
Bhμh=lim inft→∞Nh≤lim supt→∞Nh=Bhμh, |
Bvμv=lim inft→∞Nv≤lim supt→∞Nv=Bvμv. |
as required.
Model (2.1) is analyzed in a biologically meaningful feasible domain. Consider the feasible region
Wh={Sf,If,Rf,Sm,Im,Irm,Rm.:Nh≤Bhμh}, |
Wv={Sv,Iv:Nv≤Bvμv}. |
Lemma 2. The region W=Wh×Wv⊂R7+×R2+ is positively invariant and attracts all positive orbits in W.
Proof. Following steps to establish positive invariance of W. The rate of change in the population is obtained by supplementary model (2.1) to give
dNhdt=Bh−μhNh, |
dNvdt=Bv−μvNv. |
Therefore, it can be obtained
Nh=Nh(0)e−μht+Bhμh(1−e−μht), |
Nv=Nv(0)e−μvt+Bvμv(1−e−μvt). |
Particularly, Nh≤Bhμh, if Nh(0)≤Bhμh. Similarly, Nv≤Bvμv, if Nv(0)≤Bvμv. Thus, the set W is a positive invariant.
In order to prove that the set W is attractive, note that dNhdt<0 and dNvdt<0 if Nh(0)>Bhμh and Nv(0)>Bvμv respectively. Thus, either the solution enters W in finite time, or Nh(t) and Nv(t) approach Bhμh and Bvμv respectively, and therefore the variables of the infection class If,Im,Irm and Iv tend to 0. Thus, the set W is attractive and all solutions in R9+ eventually enter W.
Obviously, the system (2.1) always has a disease-free equilibrium which is given by:
ε0=(S∗f,I∗f,R∗f,S∗m,I∗m,Ir∗m,R∗m,S∗v,I∗v)=(Bh2μh,0,0,Bh2μh,0,0,0,Bvμv,0). |
Using the next-generation matrix method [24] to prove local stability of disease free equilibrium, we construct the transmission vector F representing the new infections flowing only into the infected compartments given by
F=(Bh2λhIfNf+βIm+krIrmNfSf+αhIvNhSf,Bh2λhIfNf+αhIvNhSm,0,BvθIvNv+αvIf+ImNhSv)T, |
the transition vector V represent the outflow from the infectious compartments in system (2.1) is given by
V=(γIf+μhIf,γIm+μhIm,γrIrm+μhIrm−γIm,μvIv)T |
Substitute the value of the disease-free equilibrium Nh=Bhμh and Nf=Nm=Bh2μh, we compute the Jacobian F from F given by
F=(λhμhββkrαh2λhμh00αh20000αvμhBvBhμvαvμhBvBhμv0θμv), |
and the Jacobian V from V given by
V=(γ+μh0000γ+μh000−γγr+μh0000μv). |
Thus, the characteristic polynomial available from of the generation matrix FV−1 is
λ4−k1θ+μhλhk1λ3−μh(Bhβkrμ2vθλh+Bvk2αhαv)k1Bhμ2vk2λ2+12μhβγrkr(2Bhμ2vθλh−Bvαhαv)k1Bhμ2vk2λ=0, |
where k1=γ+μh,k2=γr+μh. We can solve for four eigenroots, one of which is 0, two complex roots, one real root, and find the absolute value of the largest root.
Hence, the basic reproduction number of model (2.1) is
R0=ρ(FV−1)=161k1Bhμvk23√ψ1+6√−1Bhk23μhψ2+23ψ33√k1μv(ψ1+6√−1Bhk23μhψ2k1)B2hk22+13k1θ+μhλhk1 |
where
ψ1=−72Bhβk21krμhμ3vθγrλh+36Bhβk1krμ2hμ3vγrλ2h+8Bhk31k2μ3vθ3−12Bhk1k2μhμ3vθ2λh−12Bhk1k2μ2hμ3vθλ2h+8Bhk2μ3hμ3vλ3h+54Bvβk21krμhμvαhαvγr+36Bvk21k2μhμvθαhαv+36Bvk1k2μ2hμvαhαvλh,ψ2=16B3hβ3k1k3rμ2hμ6vγ3rλ3h−32B3hβ2k21k2k2rμhμ6vθ2γ2rλ2h+32B2hβ2k1k2k2rμ2hμ6vθγ2rλ3h+4B3hβ2k2k2rμ3hμ6vγ2rλ4h+16B3hβk31k22krμ6vθ4γrλh−32B3hβk21k22krμhμ6vθ3γrλ3h+8B3hβk1k22krμ2hμ6vθ2γrλ3h+8B3hβk22krμ3hμ6vθγrλ4h+4B3hk21k32μhμ6vθ4λ2h−8B3hk1k32μ2hμ6vθ3λ3h+4B3hk22μ3hμ6vθ2λ4h+72B2hBvβ2k21k2k2rμhμ4vθαhαhα6vγ2rλh+12B2hBvβ2k1k2k2rμ2hμ4vαhαvγ2rλ2h−8B2hBvβk31k22krμ4vθ3αhαvγr+92B2hBvβk21k22krμhμ4vθ2αhαvγrλh+4B2hBvβk1k22krμ2hμ4vθαhαvγrλ2h−8B2hBvk21k32μhμ4vθ3αhαvλh+32B2hBvk1k32μ2hμ4vθ2αhαvλ2h−8B2hBvk32μ3hμ4vθαhαvλ3h−27BhB2vβ2k21k2k2rμhμ2vα2hα2vγ2r−36BhB2vβk21k22krμhμ2vθα2hα2vγr+12BhB2vβk1k22krμ2hμ2vγrλh+4BhB2vk21k32μhμ2vθ2α2hα2v−40BhB2vk1k32μ2hμ2vθα2hα2vλh+4BhB2vk32μ3hμ2vα2hα2vλ2h+16B3vk1k32μ2hα3hα3v,ψ3=2Bhβk1krμhμ2vγrλh+Bhk21k2μ2vθ2+Bhk1k2μhμ2vθλh+Bhk2μ2hmu2vλ2h+3Bvk1k2μhαhαv. |
Using Theorem 2 of [24], we obtain the local stability of disease-free equilibrium.
Theorem 3.1.1. The disease-free equilibrium ε0 is locally asymptotically stable for R0<1 and is unstable for R0>1.
Using the method which is applied in [26,27,28,29], we obtain global stability of disease-free equilibrium. In order to make the population size of Zika extinction independent of the initial value, we establish the global stability of the disease-free equilibrium point, considering a feasible region
W1={X∈W:Sf≤S∗f,Sm≤S∗m,Sv≤S∗v}, |
where X=Sf,If,Rf,Sm,Im,Irm,Rm,Sv,Iv
Lemma 3. The W1 is positively invariant for model (2.1).
Proof. For the the first equation of the model (2.1), where S∗f=Bh2μh
dSfdt=Bh2−Bh2Y1(t)−Y2(t)Sf−Y3(t)Sf−μhSf≤Bh2−μhSf≤μh[Bh2μh−Sf]=μh(S∗f−Sf). | (3.1) |
Further,
Sf(t)≤S∗f−(S∗f−Sf(0))e−μht. |
Hence, if Sf(0)≤S∗f for t≥0, then Sf(t)≤S∗f for t≥0. Similarly, it can be obtained if Sm(0)≤S∗m for t≥0, then Sm(t)≤S∗m for t≥0 and if Sv(0)≤S∗v for t≥0, then Sv(t)≤S∗v for t≥0. Therefore, in general the field W1 is a positive invariant set and attracts all solutions of model (2.1) in R9+.
Remark 3.2.1. The W1 is a special field that needs to be satisfied when R0<1, which represents the basin of attraction.
Theorem 3.2.1. The disease-free equilibrium ε0 is globally asymptotically stable for R0<1.
Proof. To prove the global stability of the disease-free equilibrium, let X=(Sf,Rf,Sm,Rm,Sv) and Z=(If,Im,Irm,Iv). Therefore, the grouping system can be expressed as
dXdt=F(X,0),dZdt=G(X,Z). | (3.2) |
where, F(X,0) is the right side of ˙Sf,˙Rf,˙Sm,˙Rm,˙Sv when If=Im=Irm=Iv=0 and G(X,Z) is the right side of ˙If,˙Im,˙Irm,˙Iv.
Next, we consider simplifying the system dXdt=F(X,0):
dSfdt=Bh2−μhSf,dRfdt=−μhRf,dSmdt=Bh2−μhSm,dRmdt=−μhRm,dSvdt=Bv−μvSv.. | (3.3) |
It is easy to get an equilibrium for system (3.3),
X∗=(S∗f,R∗f,S∗m,R∗m,S∗v)=(Bh2μh,0,Bh2μh,0,Bvμv). |
Showing that X∗ is a globally stable equilibrium in W1. For this, by the first and second equations of (3.3) we get
Sf(t)=Bh2μh+(Sf(0)−Bh2μh)e−μht,Rf(t)=Rf(0)e−μht. | (3.4) |
Taking the limits of Sf(t) and Rf(t) at t→∞, we have
limt→∞Sf(t)=Bh2μhandlimt→∞Rf(t)=0. |
Similarly, it can be shown that limt→∞Sm(t)=Bh2μh,limt→∞Rm(t)=0, and limt→∞Sv(t)=Bvμv. These asymptotic dynamics are independent of initial in W. Therefore, the convergence of the solution of (3.3) is global in W1. Further, according to [28] we require G(X,Z) to satisfy the two stated conditions:
(i).G(X,0)=0,(ii).G(X,Z)=DzG(X∗,0)−ˆG(X,Z),ˆG(X,Z)≥0, |
where (X∗,0)=(Bh2μh,0,0,Bh2μh,0,0,0,Bvμv,0) and DzG(X∗,0) is the Jacobian of the G(X,Z) at (X∗,0), which is an M-matrix(the off-diagonal elements are nonnegative).
Thus,
DzG(X∗,0)=(−γ−μh+Bhλh2N∗fβS∗fN∗fβkrS∗fN∗fαhN∗hS∗fBhλh2N∗f−γ−μh0αhN∗hS∗m0γ−γr−μh0αvS∗vN∗hαvS∗vN∗h0BvθN∗v−μv), |
and
ˆG(X,Z)=(ξ1If+ξ2Im+ξ3Irm+ξ4Ivξ1If+ξ5Iv0ξ6Iv+ξ7(If+Im)), |
where, ξ1=Bhλh2N∗f(1−N∗fNf), ξ2=βS∗fN∗f(1−SfN∗fNfS∗f), ξ3=βkrS∗fN∗f(1−SfN∗fNfS∗f), ξ4=αhS∗fN∗h(1−SfN∗hNhS∗f), ξ5=αhS∗mN∗h(1−SmN∗hNhS∗m), ξ6=BvθN∗v(1−N∗vNv). ξ7=αvS∗vN∗h(1−SvN∗hNhS∗v).
Further, S∗f=Bh2μh,S∗m=Bh2μh,S∗v=Bvμv,N∗h=Bhμh. We have in that Sf≤S∗f,Sm≤S∗m,Sv≤S∗v. Hence, if the human population is in a state of equilibrium, we can get (1−N∗fNf)>0,(1−SfN∗fNfS∗f)>0,(1−SfN∗hNhS∗f)>0,(1−SmN∗hNhS∗m)>0,(1−N∗vNv)>0 and (1−SvN∗hNhS∗v)>0. Therefore, ˆG≥0. Then, according to the theorem in[28], the global stability of the disease-free equilibrium can be obtained. This would indicate that Zika virus will die out over time and remain stable globally.
We add four time-dependent control variables to the model corresponding to four mitigation strategies. In order to derive the necessary conditions for the existence of optimal control, we use Pontryagin's maximum principle [30]. Abimbade et al. [31] and Olaniyi et al. [32] refers to the derivation of the optimal control problem.
In model (4.1), four control strategies u1(t),u2(t),u3(t),u4(t) are added to extend model (2.1) to obtain model(4.1). u1(t) denotes the use of mosquito nets and other methods to reduce human exposure to mosquitoes; u2(t) means to improve the media, internet and other publicity efforts to enhance human awareness and reduce the probability of sexual transmission; u3(t) advocates delaying pregnancy and reducing the number of babies born with abnormalities. u4(t) means the use of insecticides and other methods to reduce mosquito populations. Here, we assume that the control set is
U={(u1,u2,u3,u4|ui(t)∈L∞[0,tf],0≤ui(t)≤ci,0≤ci≤1,i=1,...,4}. |
The optimal control model is given as
dSfdt=Bh2−(1−u3(t))Bh2Y1(t)−(1−u2(t))Y2(t)Sf−(1−u1(t))Y3(t)Sf−μhSf,dIfdt=(1−u3(t))Bh2Y1(t)+(1−u2(t))Y2(t)Sf+(1−u1(t))Y3(t)Sf−γIf−μhIf,dRfdt=γIf−μhRf,dSmdt=Bh2−(1−u3(t))Bh2Y1(t)−(1−u1(t))Y3(t)Sm−μhSm,dImdt=(1−u3(t))Bh2Y1(t)+(1−u1(t))Y3(t)Sm−γIm−μhIm,dIrmdt=γIm−γrIrm−μhIrm,dRmdt=γrIrm−μhRm,dSvdt=Bv−BvY4(t)−(1−u1(t))Y5(t)Sv−u4(t)Sv−μvSv,dIvdt=BvY4(t)+(1−u1(t))Y5(t)Sv−u4(t)Iv−μvIv. | (4.1) |
Due to non-negative initial conditions and bounded Lebesgue measurable control, this system has non-negative bounded solutions [33]. We consider an optimal control problem to minimize the objective functional
J=∫tf0[A1If+A2Im+A3Irm+A4Iv+12(η1u21+η2u22+η3u23+η4u24)]dt. | (4.2) |
In Eq (4.2), A1,A2,A3 and A4 represent the weights of human infected and mosquito infected, respectively. The weights η1,η2,η3 and η4 are measures of the costs associated with the control variables u1,u2,u3 and u4, respectively.
Theorem 4.1.1. We consider the objective functional J given by Eq (4.2) with (u1,u2,u3,u4)∈U subject to the control model (4.1) with initial conditions. There exists u∗(t)={u∗1,u∗2,u∗3,u∗4}∈U such that
J(u∗1,u∗2,u∗3,u∗4)=min{J(u1,u2,u3,u4)|(u1,u2,u2,u4)∈U}. |
Proof. We can use the result of [33] to prove the existence of an optimal control problem.
By its definition, we know that the control set U is closed and convex, and the integrand is also convex on U. Obviously these state and control variables are non-negative. The control system is bounded, which means the compactness of optimal control. Furthermore, there exists a constant ζ>1 and positive values z1, and z2, such that
J(u1,u2,u3,u4)≥z1(|u1|2+|u2|2+|u3|2+|u4|2)ζ2−z2, |
which completes the existence of the optimal control. A method to prove the existence of optimal control is proposed in [34].
To find an optimal solution we consider the Lagrangian function of the optimal control problem. The Lagrangian function is
L=A1If+A2Im+A3Irm+A4Iv+12(η1u21+η2u22+η3u23+η4u24). |
For model (4.1), we derive the necessary conditions for optimal control according to the Pontryagin maximum principle. The corresponding Hamiltonian function is
H(Sf,If,Rf,Sm,Im,Irm,Rm,Sv,Iv,u1,u2,u3,u4,λi)=L(If,Im,Irm,Iv)+λ1dSfdt+λ2dIfdt+λ3dRfdt+λ4dSmdt+λ5dImdt+λ6dIrmdt+λ7dRmdt+λ8dSvdt+λ9dIvdt=A1If+A2Im+A3Irm+A4Iv+12(η1u21+η2u22+η3u23+η4u24)+λ1[Bh2−(1−u3(t))Bh2Y1(t)−(1−u2(t))Y2(t)Sf−(1−u1(t))Y3(t)Sf−μhSf]+λ2[(1−u3(t))Bh2Y1(t)+(1−u2(t))Y2(t)Sf+(1−u1(t))Y3(t)Sf−γIf−μhIf]+λ3[γIf−μhRf]+λ4[Bh2−(1−u3(t))Bh2Y1(t)−(1−u1(t))Y3(t)Sm−μhSm]+λ5[(1−u3(t))Bh2Y1(t)+(1−u1(t))Y3(t)Sm−γIm−μhIm]+λ6[γIm−γrIrm−μhIrm]+λ7[γrIrm−μhRm]+λ8[Bv−BvY4(t)−(1−u1(t))Y5(t)Sv−u4(t)Sv−μvSv]+λ9[BvY4(t)+(1−u1(t))Y5(t)Sv−u4(t)Iv−μvIv], |
where λi,i=1,...,9 are adjoint variables.
Theorem 4.2.1. Given an optimal control (u∗1,u∗2,u∗3,u∗4), and let Sf,If,Rf,Sm,Im,Irm,Rm,Sv and Iv be the state solutions for model (4.1). Thus, there exist adjoint variables λi,i=1,...,9 satisfying
dλ1dt=(λ1−λ2)[(1−u2(t))Y2+(1−u1(t))Y3]+λ1μh],dλ2dt=−A1+(λ1−λ2)(1−u3(t))Bhλh2Nf+(λ2−λ3)γ+λ2μh+(λ4−λ5)(1−u3(t))Bhλh2Nf+(λ8−λ9)(1−u1(t))αvNhSv,dλ3dt=λ3μh,dλ4dt=(λ4−λ5)(1−u1(t))Y3+λ4μh,dλ5dt=−A2+(λ1−λ2)(1−u2(t))βNfSf+(λ5−λ6)(γ)+λ5μh+(λ8−λ9)(1−u1(t))αvNhSv,dλ6dt=−A3+(λ1−λ2)(1−u2(t))βkrNfSf+(λ6−λ7)γr+λ6μh,dλ7dt=λ7μh,dλ8dt=(λ8−λ9)(1−u1(t))Y5+λ8(μv+u4(t)),dλ9dt=−A4+(λ1−λ2)(1−u1(t))αhNhSf+(λ4−λ5)(1−u1(t))αhNhSm+(λ8−λ9)BvθNv+λ9(μv+u4(t)). |
The boundary conditions are
λi(tf)=0,i=1,...,9. |
Furthermore, the optimal controls u∗1,u∗2,u∗3 and u∗4 are represented by
u∗1(t)=min{max{uc1,0},1},u∗2(t)=min{max{uc2,0},1},u∗3(t)=min{max{uc3,0},1},u∗4(t)=min{max{uc4,0},1}, | (4.3) |
where
uc1=(λ2−λ1)Y3Sf+(λ5−λ4)Y3Sm+(λ9−λ8)Y5Svη1,uc2=(λ2−λ1)Y2Sfη2,uc3=(λ2−λ1)Bh2Y1+(λ5−λ4)Bh2Y1η3,uc4=λ8Sv+λ9Ivη4. |
Hence,
u∗i={0,ifuci≤0,uci,if0<uci<1,1,ifuci≥1, |
where i=1,...4.
Proof. The result of the adjoint system can be obtained from Pontryagin's principle
dλidt=−∂H∂x, |
where x=Sf,If,Rf,Sm,Im,Irm,Rm,Sv and Iv. The boundary conditions are λ∗i(tf)=0,i=1,...,9.
To derive the characterization of the optimal control given by Eq (4.3), we solve the equations on the interior of the control set,
∂H∂u1=0,∂H∂u2=0,∂H∂u3=0,∂H∂u4=0. |
Plugging the bounds for the controls, we obtain the desired characterization.
In this section, we apply our model to study Zika virus transmission cases in Colombia during 2015 to 2017. Based on the actual data of Zika virus transmission, numerical experiments are carried out using matlab and the parameters of the model is estimated using the Markov Chain Monte Carlo (MCMC) procedure. According to [35] data, we know that the average number of births per day in Colombia is about 1826.81 and the mortality rate is 0.0000368. It has been reported that it will take three to seven days for infected individuals with Zika to recover. Therefore, the recovery rate is set to 1/4. We know from [36,37] that γr is 0.06. We choose a set of values of parameters in Table 2.
Parameter | Range | Value | Source |
Bh | - | 1826.81 | [35] |
μh | - | 0.0000368 | [35] |
β | 0.01-0.1 | 0.046586 | MCMC |
αh | 0.03-0.75 | 0.63722 | MCMC |
αv | 0.09-0.75 | 0.6893 | MCMC |
kr | 0.2-0.8 | 0.4249 | MCMC |
γ | 1/3-1/7 | 1/4 | Assumed |
γr | 0.01-0.07 | 0.06 | [36,37] |
λh | 0.001-0.3 | 0.0054773 | MCMC |
Bv | 200-5000 | 4506 | MCMC |
1μv | 4-35 | 8 | Assumed |
θ | 0-0.004 | 0.0021348 | MCMC |
According to Table 2, we know that transmission rate from infected humans to susceptible humans (β) is 0.046586, baseline value of transmission rate from mosquitoes to humans (αh) is 0.63722 and baseline value of transmission rate from humans to mosquitoes (αv) is 0.6893. About vertical transmission, proportion of offspring congenital infection of infected female mosquitoes (θ) is 0.0021348 and proportion of offspring with congenital infection of infected females (λh) is 0.0050859. The probability is relatively small, but it has caused serious consequences, not only the financial impact on society but also the impact on life and health. Therefore, we must pay attention to it.
Clearly, fitting results Figure 2, the data simulation does not match the actual data very well. According to [7], we found that in June 2016, the World Health Organization developed a plan and implemented measures to address the threat posed by Zika virus, including in Colombia, which also implemented measures to reduce the probability of transmission through sexual contact and human mosquito contact. This is the reason for the rapid decline in the actual number of cases from June 2016. Therefore, we're going to do a piecewise simulation(see Figure 3). Before June 2016, transmission rate from infected humans to susceptible humans (β) is 0.046586, baseline value of transmission rate from mosquitoes to humans (αh) is 0.63722 and baseline value of transmission rate from humans to mosquitoes (αv) is 0.6893. After June 2016, transmission rate from infected humans to susceptible humans (β) is 0.045953, baseline value of transmission rate from mosquitoes to humans (αh) is 0.19573 and baseline value of transmission rate from humans to mosquitoes (αv) is 0.30891. The values of other parameters are shown in Table 2. In conclusion, control measures can quickly bring the disease under control. In the next section, we will discuss the impact of several control measures on the transmission of Zika virus.
In this section, we examine the impact of several control measures on the prevalence of the disease and the change in the number of infected people under optimal control. We take the initial value Sf=30000,If=200,Rf=500,Sm=30000,Im=150,Irm=100, Rm=400,Sv=2000,Iv=100. We choose the parameter values as follows, Bh=1826.81,Bv=4506,μh=3.65∗10−5,μv=18,β=0.046586, αh=0.63722,αv=0.6893,kr=0.4249,γ=14,γr=0.06,λh=0.0054773,θ=0.0021348. In order to compare the effectiveness of control measures, we conduct simulations as follows.
To study the impact of different control measures on Zika virus outbreak, we will simulate cases over time in female susceptible individual, male susceptible individual and male convalescent individuals under different control measures. Figure 4 shows the change of the number of infected persons (If,Im,Irm) over time when only the control measure u1 is added and other control measures are zero. As can be seen from the figure, when no control measures are taken, the peak of infected people will be brought to a high degree and then the number will decrease. Reduction of human mosquito exposure rate (u1=0.25) will greatly reduce the peak of infection with slight control, but its impact on Zika virus will be reduced in later stages. Under moderate control of u1 the disease will not break out or become extinct and will remain at a very low incidence. Figure 5 shows the change of the number of infected persons (If,Im,Irm) over time when control measure u4 is added and other control measures are zero. As can be seen from the figure, only mild implementation of control measures (u4=0.2) is required to prevent the outbreak of disease and gradually eliminate the disease over time. Figure 6 shows the change in the number of infected persons (If,Im,Irm) when only u2 is added and other control measures are zero. As can be seen from the figure, when u2 control measures are moderate, the peak value of infected females (If) is slightly reduced and the effect on infected females (Im) and convalescent men (Irm) is not significant. Comparing Figures 4–6, it is clear that controlling mosquito transmission to humans is more effective than controlling sexual transmission.
Figure 7 shows the change in the number of infected persons (If,Im,Irm) when only control measure u3 is added and other control measures are zero. As can be seen from the figure, due to the large number of infected people in the outbreak of the disease, only implementing the control strategy of delaying pregnancy (u3=0.5) has little impact on the number of infected people. Figure 8 shows the change in the number of infected persons (If,Im,Irm) over time when all four control measures were implemented. As can be seen from the figure, only minor control measures are needed to quickly control the outbreak and wipe out the disease in a short time. Figure 9 shows the change of the number of infected persons (If,Im,Irm) over time under the optimal control measures. We know that under optimal control, the disease rarely breaks out and is most effectively controlled.
Reducing the mosquito population is the most effective way to control the spread of the disease by comparing several control measures. In conclusion, rapid control of a Zika outbreak requires a combination of mosquito-borne control strategies and internal human control strategies. The strategy of delaying pregnancy may be effective because of the greater adverse effects of Zika virus infection in newborns and the possibility that delaying pregnancy may reduce the number of infections in newborns.
We develop a Zika virus transmission model with mosquito-borne transmission, sexual transmission and vertical transmission. Because sexual transmission of Zika is mainly male to female, we have differentiated the genders to more accurately represent the method of sexual transmission. Vertical transmission distinguishes between vertical transmission by mosquitoes and vertical transmission by humans. The basic reproduction number is obtained by the next generation matrix method. The global stability of the disease-free equilibrium is derived. The existence and mathematical expression of optimal control are obtained by using Pontryagin's maximum principle.
Based on the data of Zika virus in Colombia from 2015 to 2017, the unknown parameters of the model are estimated by MCMC algorithm. The reason why the fitted curve was not consistent with the actual cases was that the corresponding control measures adopted in Columbia in June 2016 led to a rapid decline in the number of infections after June 2016. A comparison of the four control strategies revealed that a combination of mosquito vector control and internal human control is necessary to bring the disease under control quickly. So in order to quickly control the spread of the Zika virus, we need to not only control the mosquito population and avoid contact with mosquitoes, but also increase awareness of disease transmission in humans and reduce sexual transmission. The Zika virus can cause severe brain defects in newborns and control strategies to delay pregnancy may reduce the number of infections in newborns. However, the change in the number of neonatal infected persons is not clearly indicated in this paper, which will likely be related to follow-up work in this respect.
This work is supported by the National Natural Science Foundation of China (11861044), the NSF of Gansu of China (21JR7RA212 and 21JR7RA535).
The authors declare that they have no conflict of interest.
[1] |
G. W. Dick, S. F. Kitchen, A. J. Haddow, Zika virus (Ⅰ). Isolations and serological specificity, Trans. R. Soc. Trop. Med. Hyg., 46 (1952), 509–520. https://doi.org/10.1016/0035-9203(52)90042-4 doi: 10.1016/0035-9203(52)90042-4
![]() |
[2] |
L. R. Petersen, D. J. Jamieson, A. M. Powers, M. A. Honein, Zika virus, N. Engl. J. Med., 374 (2016), 1552–1563. https://doi.org/10.1056/nejmra1602113 doi: 10.1056/nejmra1602113
![]() |
[3] | Centers for Disease Control and Prevention, About Zika, Overview, Available from: https://www.cdc.gov/zika/about/overview.html. |
[4] |
K. Smithburn, Neutralizing antibodies against certain recently isolated viruses in the sera of human beings residing in East Africa, J. Immunol., 69 (1952), 223–234. https://doi.org/10.4049/jimmunol.69.2.223 doi: 10.4049/jimmunol.69.2.223
![]() |
[5] |
M. R. Duffy, T. H. Chen, W. T. Hancock, A. M. Powers, J. L. Kool, R. S. Lanciotti, et al., Zika virus outbreak on Yap Island, federated states of Micronesia, N. Engl. J. Med., 360 (2009), 2536–2543. https://doi.org/10.1056/NEJMoa0805715 doi: 10.1056/NEJMoa0805715
![]() |
[6] |
D. Musso, E. Nilles, V. M. Cao Lormeau, Rapid spread of emerging Zika virus in the Pacific area, Clin. Microbiol. Infect., 20 (2014), O595–O596. https://doi.org/10.1111/1469-0691.12707 doi: 10.1111/1469-0691.12707
![]() |
[7] | World Health Organization, Zika virus outbreak global response-Interim report May 2016. Available from: https://www.who.int/publications/i/item/zika-virus-outbreak-global-response. |
[8] |
J. Mlakar, M. Korva, N. Tul, M. Popović, M. Poljšak-Prijatelj, J. Mraz, et al., Zika virus associated with microcephaly, N. Engl. J. Med., 374 (2016), 951–958. https://doi.org/10.1056/NEJMoa1600651 doi: 10.1056/NEJMoa1600651
![]() |
[9] |
P. S. Mead, N. K. Duggal, S. A. Hook, M. Delorey, M. Fischer, D. O. McGuire, et al., Zika virus shedding in semen of symptomatic infected men N. Engl. J. Med., 378 (2018), 1377–1385. https://doi.org/10.1056/NEJMoa1711038 doi: 10.1056/NEJMoa1711038
![]() |
[10] | Centers for Disease Control and Prevention, Questions About Zika, Available from: https://www.who.int/publications/i/item/zika-virus-outbreak-global-response. |
[11] |
B. D. Foy, K. C. Kobylinski, J. L. C. Foy, B. J. Blitvich, A. T. da Rosa, A. D. Haddow, et al., Probable non¨Cvector-borne transmission of Zika virus, Colorado, USA, Emerging Infect. Dis., 17 (2011), 880–882. https://doi.org/10.3201/eid1705.101939 doi: 10.3201/eid1705.101939
![]() |
[12] |
D. Gao, Y. Lou, D. He, T. C. Porco, Y. Kuang, G. Chowell, et al., Prevention and control of Zika as a mosquito-borne and sexually transmitted disease: a mathematical modeling analysis, Sci. Rep., 6 (2016), 1–10. https://doi.org/10.1038/srep28070 doi: 10.1038/srep28070
![]() |
[13] |
D. He, D. Gao, Y. Lou, S. Zhao, S. Ruan, A comparison study of Zika virus outbreaks in French Polynesia, Colombia and the State of Bahia in Brazil, Sci. Rep., 7 (2017), 273. https://doi.org/10.1038/s41598-017-00253-1 doi: 10.1038/s41598-017-00253-1
![]() |
[14] |
D. Baca-Carrasco, J. X. Velasco-Hernández, Sex, mosquitoes and epidemics: An evaluation of Zika disease dynamics, Bull. Math. Biol., 78 (2016), 2228–2242. https://doi.org/10.1007/s11538-016-0219-4 doi: 10.1007/s11538-016-0219-4
![]() |
[15] |
F. Agusto, S. Bewick, W. Fagan, Mathematical model for Zika virus dynamics with sexual transmission route, Ecol. Complex, 29 (2017), 61–81. https://doi.org/10.1016/j.ecocom.2016.12.007 doi: 10.1016/j.ecocom.2016.12.007
![]() |
[16] |
M. Imran, M. Usman, M. Dur-e Ahmad, A. Khan, Transmission dynamics of Zika fever: A SEIR based model, Differ. Equtions Dyn. Syst., 29 (2021), 463–486. https://doi.org/10.1007/s12591-017-0374-6 doi: 10.1007/s12591-017-0374-6
![]() |
[17] |
A. D¡äenes, M. A. Ibrahim, L. Oluoch, M. Tekeli, T. Tekeli, Impact of weather seasonality and sexual transmission on the spread of Zika fever, Sci. Rep., 9 (2019), 17055. https://doi.org/10.1038/s41598-019-53062-z doi: 10.1038/s41598-019-53062-z
![]() |
[18] |
M. A. Ibrahim, A. Dénes, Threshold dynamics in a model for Zika virus disease with seasonality, Bull. Math. Biol., 83 (2021), 27. https://doi.org/10.1007/s11538-020-00844-6 doi: 10.1007/s11538-020-00844-6
![]() |
[19] |
X. Yuan, Y. Lou, D. He, J. Wang, D. Gao, A Zika endemic model for the contribution of multiple transmission routes, Bull. Math. Biol., 83 (2021), 111. https://doi.org/10.1007/s11538-021-00945-w doi: 10.1007/s11538-021-00945-w
![]() |
[20] | S. Busenberg, K. Cooke, Vertically Transmitted Diseases: Models and Dynamics, Biomathematics 23, Springer-Verlag, Berlin Heidelberg, 1993. |
[21] |
M. Y. Li, H. L. Smith, L. Wang, Global dynamics of an SEIR epidemic model with vertical transmission, SIAM J. Appl. Math., 62 (2001), 58–69. https://doi.org/10.1137/S0036139999359860 doi: 10.1137/S0036139999359860
![]() |
[22] |
S. Olaniyi, Dynamics of Zika virus model with nonlinear incidence and optimal control strategies, Appl. Math. Inf. Sci., 12 (2018), 969–982. https://doi.org/10.18576/amis/120510 doi: 10.18576/amis/120510
![]() |
[23] |
E. Okyere, S. Olaniyi, E. Bonyah, Analysis of zika virus dynamics with sexual transmission route using multiple optimal controls, Sci. Afr., 9 (2020), e00532. https://doi.org/10.1016/j.sciaf.2020.e00532 doi: 10.1016/j.sciaf.2020.e00532
![]() |
[24] |
P. Van den Driessche, J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002), 29–48. https://doi.org/10.1016/S0025-5564(02)00108-6 doi: 10.1016/S0025-5564(02)00108-6
![]() |
[25] | S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, Springer, Berlin, 1990. |
[26] |
F. B. Agusto, S. Bewick, W. Fagan, Mathematical model of Zika virus with vertical transmission, Infect. Dis. Modell., 2 (2017), 244–267. https://doi.org/10.1016/j.idm.2017.05.003 doi: 10.1016/j.idm.2017.05.003
![]() |
[27] |
K. Blayneh, Y. Cao, H. D. Kwon, Optimal control of vector-borne diseases: Treatment and prevention, Discrete Contin. Dyn. Syst. Ser. B, 11 (2009), 587–611. https://doi.org/10.3934/DCDSB.2009.11.587 doi: 10.3934/DCDSB.2009.11.587
![]() |
[28] | C. Castillo-Chavez, S. Blower, P. Van den Driessche, D. Kirschner, A. A. Yakubu, Mathematical Approaches for Emerging and Reemerging Infectious Diseases, Springer-Verlag, New York, 2002. |
[29] |
C. Bhunu, W. Garira, Z. Mukandavire, Modeling HIV/AIDS and tuberculosis coinfection, Bull. Math. Biol., 71 (2009), 1745–1780. https://doi.org/10.1007/s11538-009-9423-9 doi: 10.1007/s11538-009-9423-9
![]() |
[30] | L. Pontryagin, V. Boltyanskiy, R. Gamkrelidze, E. Mishchenko, Mathematical Theory of Optimal Processes, Interscience, New York, 1962. |
[31] |
S. Abimbade, S. Olaniyi, O. Ajala, M. Ibrahim, Optimal control analysis of a tuberculosis model with exogenous re-infection and incomplete treatment, Optim. Control Appl. Methods, 41 (2020), 2349–2368. https://doi.org/10.1002/oca.2658 doi: 10.1002/oca.2658
![]() |
[32] |
S. Olaniyi, O. Falowo, K. Okosun, M. Mukamuri, O. Obabiyi, O. Adepoju, Effect of saturated treatment on malaria spread with optimal intervention, Alexandria Eng. J., 65 (2023), 443–459. https://doi.org/10.1016/j.aej.2022.09.024 doi: 10.1016/j.aej.2022.09.024
![]() |
[33] | D. L. Lukes, Differential Equations: Classical to Controlled, Academic Press, 162 (1982), 223–225. https://doi.org/10.2307/2322889 |
[34] |
A. Abidemi, S. Olaniyi, O. A. Adepoju, An explicit note on the existence theorem of optimal control problem, J. Phys. Conf. Ser., 2199 (2022), 012021. https://doi.org/10.1088/1742-6596/2199/1/012021 doi: 10.1088/1742-6596/2199/1/012021
![]() |
[35] | World Health Organization, WHO Global Health Observatory data repository, Crude birth and death rate. Data by country, 2003. Available from: https://apps.who.int/gho/data/node.main.CBDR107?lang = en. |
[36] |
A. C. Gourinat, O. OConnor, E. Calvez, C. Goarant, M. Dupont-Rouzeyrol, Detection of Zika virus in urine, Emerging Infect. Dis., 21 (2015), 84–86. https://doi.org/10.3201/eid2101.140894 doi: 10.3201/eid2101.140894
![]() |
[37] |
D. Musso, C. Roche, E. Robin, T. Nhan, A. Teissier, V. M. Cao-Lormeau, Potential sexual transmission of Zika virus, Emerging Infect. Dis., 21 (2015), 359–361. https://doi.org/10.3201/eid2102.141363 doi: 10.3201/eid2102.141363
![]() |
1. | S. Olaniyi, F. M. Chuma, Lyapunov Stability and Economic Analysis of Monkeypox Dynamics with Vertical Transmission and Vaccination, 2023, 9, 2349-5103, 10.1007/s40819-023-01572-w | |
2. | Anuj Kumar, Study of the impact of information and limited medical resources on Zika prevalence: an optimal control approach, 2023, 138, 2190-5444, 10.1140/epjp/s13360-023-04665-z | |
3. | Jhoana P. Romero-Leiton, Elda K.E. Laison, Rowin Alfaro, E. Jane Parmley, Julien Arino, Kamal R. Acharya, Bouchra Nasri, Exploring Zika's Dynamics: A Scoping Review Journey from Epidemic to Equations Through Mathematical Modelling, 2024, 24680427, 10.1016/j.idm.2024.12.016 | |
4. | Alan Carvalho Leandro, Renan Konig Leal, Millena Fernandes, Rivania da Silva Lira, Edenilson Osinski Francisco, Betine Pinto Moehlecke Iser, Josiane Somariva Prophiro, Analysis of Zika Notifications in Brazil Between 2016 and 2023, 2025, 19, 1981-982X, e012118, 10.24857/rgsa.v19n5-017 |
Parameter | Description(Units) |
Bh | Natural birth rate of humans (days−1) |
μh | Natural death rate of humans (days−1) |
β | Sexual transmission infection rate from infected humans to susceptible humans (days−1) |
αh | Baseline value of transmission rate from mosquitoes to humans(days−1) |
αv | Baseline value of transmission rate from humans to mosquitoes (days−1) |
kr | Relative human-to-human transmissibility of convalescent to symptomatic humans (none) |
γ | Human recovery rate (days−1) |
γr | Recovery rate of convalescent humans (days−1) |
λh | Proportion of offspring with congenital infection of infected females (none) |
Bv | Baseline value of mosquito birth rate (days−1) |
1μv | Mosquito lifespan(days) |
θ | Proportion of offspring congenital infection of infected female mosquitoes (none) |
Parameter | Range | Value | Source |
Bh | - | 1826.81 | [35] |
μh | - | 0.0000368 | [35] |
β | 0.01-0.1 | 0.046586 | MCMC |
αh | 0.03-0.75 | 0.63722 | MCMC |
αv | 0.09-0.75 | 0.6893 | MCMC |
kr | 0.2-0.8 | 0.4249 | MCMC |
γ | 1/3-1/7 | 1/4 | Assumed |
γr | 0.01-0.07 | 0.06 | [36,37] |
λh | 0.001-0.3 | 0.0054773 | MCMC |
Bv | 200-5000 | 4506 | MCMC |
1μv | 4-35 | 8 | Assumed |
θ | 0-0.004 | 0.0021348 | MCMC |
Parameter | Description(Units) |
Bh | Natural birth rate of humans (days−1) |
μh | Natural death rate of humans (days−1) |
β | Sexual transmission infection rate from infected humans to susceptible humans (days−1) |
αh | Baseline value of transmission rate from mosquitoes to humans(days−1) |
αv | Baseline value of transmission rate from humans to mosquitoes (days−1) |
kr | Relative human-to-human transmissibility of convalescent to symptomatic humans (none) |
γ | Human recovery rate (days−1) |
γr | Recovery rate of convalescent humans (days−1) |
λh | Proportion of offspring with congenital infection of infected females (none) |
Bv | Baseline value of mosquito birth rate (days−1) |
1μv | Mosquito lifespan(days) |
θ | Proportion of offspring congenital infection of infected female mosquitoes (none) |
Parameter | Range | Value | Source |
Bh | - | 1826.81 | [35] |
μh | - | 0.0000368 | [35] |
β | 0.01-0.1 | 0.046586 | MCMC |
αh | 0.03-0.75 | 0.63722 | MCMC |
αv | 0.09-0.75 | 0.6893 | MCMC |
kr | 0.2-0.8 | 0.4249 | MCMC |
γ | 1/3-1/7 | 1/4 | Assumed |
γr | 0.01-0.07 | 0.06 | [36,37] |
λh | 0.001-0.3 | 0.0054773 | MCMC |
Bv | 200-5000 | 4506 | MCMC |
1μv | 4-35 | 8 | Assumed |
θ | 0-0.004 | 0.0021348 | MCMC |