352
Views
0
CrossRef citations to date
0
Altmetric
Applied & Interdisciplinary Mathematics

Uniform persistence and backward bifurcation of vertically transmitted vector-borne diseases

ORCID Icon | (Reviewing editor:) & (Reviewing editor:)
Article: 2264581 | Received 07 Jul 2022, Accepted 26 Sep 2023, Published online: 18 Oct 2023

ABSTRACT

Vertical transmissions of vector-borne diseases such as dengue virus, malaria, West Nile virus and Rift Valley fever are among challenging factors in disease control. In this paper, we used a system of differential equations to study the impacts of vertical transmission (in the vector) and horizontal (vector-to-host) transmissions cycle on the uniform persistence of a vertically transmitted disease. This is given analytically when the secondary reproduction number is greater than unity. Additional numerical results indicate that vertical transmission rates raise the epidemic level. However, we provide analytical results to reveal that vertical transmission of the disease in vectors alone without horizontal transmission, could not make the disease endemic. Furthermore, by transforming the system of differential equations to a nonlinear eigenvalue equation, the existence of a backward bifurcation is established. The directions of bifurcation is forward when the disease-induced death rate of hosts is reduced below a threshold. It is shown that the bifurcation is backward when the disease-induced death rate in the host is above a critical value. Additionally, increased disease transmission parameters also contribute to backward bifurcation. Numerical results are also provided to highlight that, depending on the initial conditions, the disease could be persistent even when the secondary reproduction number is less than unity.

MATHEMATICS SUBJECT CLASSIFICATION:

View correction statement:
Correction

1. Introduction

Vector-borne diseases such as malaria, dengue virus, West Nile virus, Rift-Valley fever and yellow fever are among major public health problems that claim the lives of human beings and domestic animals. These diseases also cause economic challenges in the world, mostly in developing countries. Additionally, the geographic distributions of some of these vector-borne diseases overlap with the developed world, including many parts of the United States (CitationCenters for Disease Control and Prevention) and Europe (Hubálek & Halouzka, Citation1999). Furthermore, many vector-borne diseases are recurring and becoming endemic in some parts of the world owing to deteriorating environmental conditions due to human activities. Some of these are deforestation, poorly managed irrigation systems, dams, and other factors connected to climate variations which influence physiological changes in mosquitoes

Major disease vectors, such as Anophelis, Aedes aegypti and Aedes albopictus mosquito species transmit dengue virus (Adams & Boots, Citation2010; Feng & Velasco-Hernández, Citation1997), West Nile virus (Blayneh et al., Citation2010; Cruz-Pacheco et al., Citation2009; Pu et al., Citation2023), and Rift Valley fever (Chitnis et al., Citation2013; Gaff et al., Citation2007) to humans and domestic animals. These mosquito species also pass the disease causing pathogen to their eggs which contributes to vertical transmission. The survival of some exposed eggs through the winter season is believed to enable the emerging adult mosquitoes to continue with the transmission cycle once conditions are favorable (Adams & Boots, Citation2010).

Vertical transmission of yellow fever is believed to be the cause for overwintering of the disease keeping transmission cycle in Amazon (Francisco, Citation2007). However, although vertical transmission enhances seasonal outbreak of diseases, it alone, could not be enough to make the disease endemic without vector-to-host (also known as horizontal) transmission cycle. Therefore, a more significant challenge comes from the presence of a horizontal transmission. Moreover, according to studies of mosquito-borne diseases, the most effective fights against horizontal transmission encompass mosquito control and personal protections (Blayneh, Citation2017; Blayneh et al., Citation2009, Citation2010; Gordillo, Citation2013). It is clear that several efforts are employed to control mosquitoes, such as releasing sterile mosquito population (Zheng, Citation2022) and effects of larva predators (Zhu et al., Citation2023). The existence of a backward bifurcation phenomena is one indicator that control efforts could fail, because of existence of endemic equilibrium points for secondary reproductive number reduced below one. Contributions to the phenomena of backward bifurcation come from a combination of several factors, primarily, poor personal protections and also fatality of the diseases (Blayneh, Citation2017; Blayneh et al., Citation2010; Chitnis et al., Citation2006; Dushoff et al., Citation1998). Disease transmissions within the host in addition to horizontal (vector-host) transmission cycle could add more challenges to the dynamic of a vector-borne disease, see for example the work of (Wang et al., Citation2020), where the analysis explores age of infection. Our work focuses on a vertically transmitted disease within the vector population, the effects of vertical transmission rate on disease prevalence and other factors which contribute to the persistence of the vector-borne disease.

Persistence of diseases contributes to challenges against control efforts. A biological system exhibits the phenomena of persistence when a solution which is initially positive (componentwise) has no compartment that dies out in the long run (Allen, Citation2007). This means, no boundary equilibrium is the omega limit set of any solution with initial condition in the positive cone. Several articles have investigated the dynamics of biological systems reflecting different forms of persistence. Some focus on models describing ecological systems, such as Lotka-Volterra type applied to food-chain (see (Allen, Citation2007) and references therein) and predator-prey (Allen, Citation2007; Joseph et al., Citation1986). Likewise, the persistence of epidemics problems has been addressed by few authors. Among them, we have (Thieme, Citation2001) on uniform weak-persistence of diseases, Robelo (Robelo et al., Citation2014) on persistence of diseases in a periodic environment. Similarly, the uniform persistence of diseases are modeled by some authors, notable examples include van den Driessche and Yakubu on discrete-time model (van den Driessche & Yakubu, Citation2019), Gourley et al. (Citation2014) on bluetongue dynamic and (Salako, Citation2023) on the impact of population size and movement on two-strain infectious diseases. Clearly, uniform persistence is strong enough to guarantee that any solution with positive initial value remains bounded away from the boundary equilibrium.

The objectives of this work are to prove the existence of a backward bifurcation when the disease-induced death rate in the host population is above a critical value, mosquito population equilibrates to a large number and personal protections are poor; to establish the uniform persistence of vertically transmitted vector-borne diseases and also, to prove that vertical transmission in the vector alone does not cause the disease to be endemic without vector-to-host transmission cycle. These results are established using the model studied in (Blayneh, Citation2017). Specifically, the important roll played by vector-to-host transmission cycle to establish the disease as endemic is proved in Section (3). Furthermore, the existence of a backward bifurcation is established in Section (4). The uniform persistence of the disease, when the secondary reproduction number is greater than unity, is proved in Section (5). This implies that despite the initial value, each compartment of the disease approaches a positive quantity. Our results highlight the importance of reducing the epidemiology threshold (secondary reproduction number) below unity and factors which cause this effort to fail. Additionally, in the presence of a backward bifurcation, depending on the initial values, the disease could equilibrate to an endemic level for some R0 less than unity. This is supported by numerical results presented in Section (5). Discussion and concluding remarks are given in Section (6).

2. The model for vector-host transmission cycle

The model studied in (Blayneh, Citation2017) is extended to address very important problems emerging from vertically transmitted vector-borne diseases, where the vertical transmission takes place in the vector population. First, we present some background steps which lead to the main model. To this end, we use x1,x2,x3 and x4 to describe the number of hosts which are susceptible, exposed, infectious, and recovered, respectively. Similarly, y1,y2 and y3 represent the number of susceptible, exposed and infectious vectors, respectively. Furthermore, N=i=14xiandP=i=13yi, denote the total population sizes of the host and vector, respectively.

As it is given in (Blayneh, Citation2017), the inflow of exposed mosquitoes (also known as vectors) comes from transmission of the disease through bites of susceptible mosquitoes (on exposed or infectious) hosts and also, adults emerging from eggs laid by infectious mosquitoes. More specifically, susceptible mosquitoes become exposed to the disease by a bite from exposed and infectious hosts with transmission efficacy θ1 and θ2, respectively. Thus, the forces of infection for a susceptible vector due to a bite of exposed host is ϕθ1x2N, and due to a bite of infectious host is ϕθ2x3N, where ϕ is the number of bites that a host sustains per unit time. Therefore, the number of exposed mosquitoes due to horizontal (vector-host) transmission per unit time is (ϕθ1x2N+ϕθ2x3N)y1. Additional inflow of exposed mosquitoes comes from vertical transmission given by δ1ζy3, where ζ is the rate of disease transmission from a female mosquito to eggs per unit time and δ1 is the proportion of eggs reaching to adult stage. Thus, the inflow of exposed mosquitoes is (ϕθ1x3N+ϕθ2x2N)y1+δ1ζy3. The average number of mosquito recruitment is δ0 and this quantity is increased by a proportion of eggs from susceptible and exposed mosquitoes reaching to adult stage, plus eggs from infectious mosquitoes which became susceptible adults, which is δ1(1ζ)y3. Thus, the susceptible mosquito population increases by δ0+δ1(y1+y2)+δ1(1ζ)y3 per unit time, which implies dy1dt=δ0+δ1(y1+y2)+δ1(1ζ)y3(ϕθ1x3N+ϕθ2x2N)y1γy1.

The change in exposed mosquito population is dy2dt=ϕθ1x3y1N+ϕθ2x2y1N+δ1ζy3φy2γy2. It should be noted that the rate of disease progression in mosquito is φ, and γ is the natural death rate of vectors. The inflow of exposed hosts is given by ϕβy3x1N and the rate of progress from exposed to infectious level in the host is d, which is also known as an incubation rate of the disease in a host per unit time. Thus, dx2dt=βϕy3x1Ndx2μx2, where µ is natural mortality rate of hosts. Infectious hosts have three ways to exit the compartment: by recovery at a rate of r, by natural death at a rate of µ or due to disease-induced death rate which is α. Based on these facts, the rate of change in x3, infectious hosts isdx3dt=dx2(r+α+μ)x3. Furthermore, a recovered host becomes susceptible to the disease at a rate of ψ. Incoming host population to the community is assumed to be susceptible. Since the disease is not vertically transmitted in the host population, all new borne hosts are susceptible. These results are the bases for the system of differential equation model given by the system in model (Equation1). For the description of all parameters refer to ().

(1) dx1dt=ρ+ΛN+ψx4βϕy3x1Nμx1dx2dt=βϕy3x1Ndx2μx2dx3dt=dx2(r+α+μ)x3dx4dt=rx3(ψ+μ)x4dy1dt=δ0+δ1(y1+y2)+δ1(1ζ)y3ϕθ1x3y1Nϕθ2x2y1Nγy1dy2dt=ϕθ1x3y1N+ϕθ2x2y1N+δ1ζy3φy2γy2dy3dt=φy2γy3,(1)

Table 1. Model parameter and their descriptions

Next, with X=(x1,x2,x3,x4) and Y=(y1,y2,y3), we define the set

(2) Ω={(X,Y)R+4×R+3,i=14xi[0,ρμΛ],i=13yi[0,δ0γδ1]}={XR+4,i=14xi[0,ρμΛ]}×{YR+3,i=13yi[0,δ0γδ1]}(2)

which is biologically feasible and forward invariant under system (Equation1). Specifically, if (X, Y) is a solution of system (Equation1) with nonnegative initial values (X(0),Y(0)), where 0i=14xi(0)ρμΛ and 0i=13yi(0)δ0γδ1, then i=14xi(t)[0,ρμΛ] and i=13yi(t)[0,δ0γδ1] and furthermore, the total population size is eventually attracted to this set (Blayneh, Citation2017).

The disease-free equilibrium of (Equation1) is

(3) E0=(x1,0,0,0,y1,0,0),(3)

where x1=ρμΛ, y1=δ0γδ1. This is obtained by setting the disease-related variables to zero: x2=x3=x4=y2=y3=0, then solving for x1 and y1 from the equilibrium equation of System (Equation1). For the sake of simplifying notations, we introduce ki,i=1,,4.

(4) k1=d+μ,k2=ψ+μ,k3=r+α+μ,k4=φ+γ.(4)

A1: γ>δ1, μ>Λ and ζ[0,1).

The secondary reproduction number which is obtained using the approaches in (van den Driessche & Watmough, Citation2002) is

(5) R0=R0(ζ)=ζδ1φ2γk4+(ζδ1φ2γk4)2+Rh2,(5)

where

(6) Rh=ϕβφ(θ1d+θ2k3)y1k1k3k4γx1.(6)

The theory of next-generation matrix provided by (van den Driessche & Watmough, Citation2002), states that the secondary reproduction number R0 is ρ(FV1), the spectral radius of FV1. F is the Jacobian of new infection function (βϕy3x1N,0,0,ϕθ1x3y1N+ϕθ1x3y1N+ϕθ2x2y1N+δ1ζy3,0) it is proved that the disease-free equilibrium point E0, given by (Equation3), of system (Equation1) is locally asymptotically stable if R0<1 and unstable if R0>1, where R0 is defined by (Equation5) (see (Blayneh, Citation2017)).

In Section 4, we present some conditions under which the disease equilibrate to endemic level even if R0<1. Next, we look into the limitations of vertical transmissions and the importance of horizontal transmission to magnify the dynamic a vertically transmitted vector-borne disease.

3. Can vertical transmission sustain the virus without a horizontal transmission?

Vector-host contact is the base for a horizontal transmission of any vector-borne disease. In vertically transmitted vector-borne diseases, additional transmission occurs due to mosquitoes passing the disease to eggs, which could become infectious adults maintaining the transmission cycle within the mosquito population. The goal of this section is to show that vertical transmission alone can not support the persistence of the disease in the absence of interactions with hosts (no horizontal transmission). A host could be human or animal where female mosquitoes get blood meal. We begin our study by assuming that the host population is susceptible, but there are infectious mosquitoes supporting vertical transmission with no opportunity to pass the disease to hosts. To see this, consider the system (Equation1) with β=ϕ=θ1=θ2=0. In the absence of horizontal transmission, all of the terms describing disease transmission from infectious hosts to susceptible mosquitoes are eliminated from fifth equation in system (Equation1) Furthermore, since the focus is on the disease dynamics in the mosquito population, without contribution of disease from infectious hosts, we consider only the last three equations of the system, which are

(7) dy1dt=δ0+δ1(y1+y2)+δ1(1ζ)y3γy1dy2dt=δ1ζy3(φ+γ)y2dy3dt=φy2γy3.(7)

It should be noted that the parameters and variables used in (Equation7) are the same as those in (Equation1). Accordingly, the only flow of exposed mosquitoes comes from δζy3 which is exposed adults from eggs of infectious mosquito. Exposed vectors develop the disease to be infectious at a rate of φ. Mortality rate of mosquitoes per unit time is γ. Clearly, (Equation7) is a linear system, and the last two equations do not involve y1. Therefore, we first consider the subsystem

(8) dy2dt=δ1ζy3(φ+γ)y2dy3dt=φy2γy3.(8)

The coefficient matrix of this subsystem is

(9) A=((φ+γ)δ1ζφγ).(9)

The trace of A<0 and the determinant of A is γ(φ+γ)φζδ1>0, because 0ζ<1 and δ1<γ. This implies that the eigenvalues of A have negative real parts. Thus, the unique equilibrium (y2,y3)=(0,0) of the subsystem (Equation8) is globally asymptotically stable. Suppose (y1(t),y2(t),y3(t)) is a solution of (Equation7) with positive initial conditions (y1(0),y2(0),y3(0)). It is clear that the first equation in (Equation7) is equivalent to

(10) dy1dt+(γδ1)y1=δ0+δ1y2++δ1(1ζ)y3.(10)

Using integrating factor e(γδ1)t, we write the solution with initial condition y1(0), as

(11) y1(t)=δ0γδ1+y1(0)e(γδ1)tδ0e(γδ1)tγδ1+δ10tB(s)ds,(11)

where

(12) B(s)=(y2(s)+(1ζ)y3(s))e(γδ1)(st).(12)

Furthermore, lim supty2(t)+(1ζ)y3(t)∣=0. Thus, the last term in (Equation11) satisfies the following condition

(13) 00t(y2(s)+(1ζ)y3(s))e(γδ1)(st)ds∣≤δ1γδ1(1e(γδ1)t)lim supty2(t)+(1ζ)y3(t)∣=0.(13)

Therefore, y1(t) given by (Equation11) approaches δ0γδ1. Since (y2,y3) approaches (0,0), we conclude that the equilibrium of system (Equation7), (δ0γδ1,0,0), is globally asymptotically stable. The conclusion from this result is that according to our model, vertical transmission alone does not sustain the disease. Some studies support the importance of epidemic circulation to make a disease endemic (see Adams & Boots (Citation2010) and references therein). Clearly, since the disease is vertically transmitted from mosquito to egg, some surviving infected eggs could become infectious adults. However, this may not be significant enough to produce more infectious adults which could sustain the number of infectious mosquitoes in the environment. Epidemic circulation in an environment, which is primarily supported by vector-host transmission cycle could establish the disease as endemic. On the other hand, many factors such as rainfall and temperature affect the physiology of mosquitoes. The contribution of these factors could amplify the effects of vertical transmission on the prevalence of the disease.

() shows how vertical transmission amplifies disease prevalence.

Figure 1. A time plot reflecting the effects of vertical transmission in the vector population on disease prevalence: ζ = 0.0046 (continuous/green curve), ζ = 0.3046 (dashed/blue curve) and ζ = 0.5046 (continuous/red curve). Other parameters are α=0.0065, γ=1/17, β=0.15,θ1=0.0083,θ2=0.0072, μ=3.91×105, \upLambda=1.48×105,ρ=205,δ0=1150,r=1/7,δ1=0.0397,d=1/6,ψ=0.00243,ϕ=2. For each case R0<1.

Figure 1. A time plot reflecting the effects of vertical transmission in the vector population on disease prevalence: ζ = 0.0046 (continuous/green curve), ζ = 0.3046 (dashed/blue curve) and ζ = 0.5046 (continuous/red curve). Other parameters are α=0.0065, γ=1/17, β=0.15,θ1=0.0083,θ2=0.0072, μ=3.91×10−5, \upLambda=1.48×10−5,ρ=205,δ0=1150,r=1/7,δ1=0.0397,d=1/6,ψ=0.00243,ϕ=2. For each case R0<1.

4. Backward bifurcation and possible causes

In this section, we establish the existence of a backward bifurcation phenomena. Our approach is by building an operator equation for the exposed groups in the host and the vector populations. Then, we define a bifurcation parameter where the critical value of this parameter corresponds to R0=1. We proved that the operator has a geometrically simple characteristic eigenvalue. Then, we use a theorem given in (Cushing, Citation1998) to conclude the existence of a bifurcating branch at the critical point. Similar approaches are used in (Chitnis et al., Citation2006, Citation2013). The bifurcation result given in (Blayneh, Citation2017) involves complex parameter combination, because of which it was not possible to highlight on the disease-induced death rate and related parameters describing disease transmission cycles. Our approach in this work clears that there is no backward bifurcation when the disease-induce death rate is reduced below a threshold level. Additionally, horizontal transmission cycle and the steady state level of mosquitoes play a role in the emergence of a backward bifurcation when the disease-induce death rate is increased. To begin, let E=(x1,x2,x3,x4,y1,y2,y3) be an endemic equilibrium of System (Equation1), then close to the disease-free equilibrium E0=(x1,0,0,0,y1,0,0), x2 and y2 satisfy the following equations

(14) x2=ϕβφγk1y2+ϕβφγk1[H1ρk2k3+dαρk3]x2y2y2=H2δ0γ(μΛ)ρ(γk4δ1ζφ)x2+dαH2(μΛ)δ0γk3ρ2(γk4δ1ζφ)x22(γδ1)k4H2(μΛ)ρ(γk4δ1ζφ)x2y2,(14)

where

(15) H1=ψrdk1k2k3+Λ(k2k3+rd+rk2)andH2=ϕ(θ1d+θ2k3)k3(γδ1).(15)

Note that assumption (A1) in Section (2) implies that the term γk4δ1ζφ>0. Next, select a bifurcation parameter

(16) Δ=βϕδ0(μΛ)ρ(γδ1)=βϕy1x1,(16)

and using Δ, transform EquationEquation (14) to

(17) x2=Δφx1γk1y1y2+Δφx1γk1k2k3y1(H1+dαk2)x2y2y2=Δ(θ1d+θ2k3)γβk3(γk4δ1ζφ)[x2+dαρk3x22k4y1γx2y2].(17)

Then the nonlinear eigenvalue equation becomes

(18) x=ΔLx+h(Δ,x),(18)

where

(19) L=(0φx1γk1y1(θ1d+θ2k3)γβk3(γk4δ1ζφ)0),(19)
(20) h(Δ,x)=(Δφx1γρk1k2k3y1(H1+dαk2)x2y2Δ(θ1d+θ2k3)γβk32(γk4δ1ζφ)[αdρx22k4k3γy1x2y2])andx=(x2y2).(20)

To apply known bifurcation theorem and pull fundamental results, we first need to lay preliminary steps. To this end, the set R+2={(x2,y2),x2>0,y2>0} is a Banach space under the Euclidean norm. The boundary R+2 carries vectors where at least one component is zero. System (Equation1) has only one disease-free equilibrium, therefore, x=(0,0) is unique on R+2. Furthermore, L:R+2R+2 is a compact operator. This follows from the fact that for any bounded subset D of R+2, L(D) lies in a compact subset of R+2. For an open interval ΘR, we define

(21) Σ=Θ×R+2(21)

and verify the existence of a continuum C+Σ which bifurcates from a critical point (Δ0,0), where

(22) Δ0=βy1k1k3(γk4δ1ζφ)x1φ(θ1d+θ2k3)(22)

Here, a pair (Δ,x) on C+ represents a solution x of (Equation18) corresponding to Δ, where x=(x2,y2)τ, (τ is transpose) is a component of an endemic equilibrium E=(x1,x2,x3,x4,y1,y2,y3) for system (Equation1) which is connected to (Equation14). It is clear that the remaining components of E can be solved from the equilibrium equation. Specifically, for (x2,y2) on the bifurcating curve for (Equation14), we go to system (Equation1) and solve for x3 from the third equation, then for x4 from the fourth equation. Also, adding the fifth and sixth equations, we solve for y1. Similarly we solve for y1, adding the first and the second equations. It should be noted that in the presence of a backward bifurcation, there are two endemic equilibrium points for R0<1.

We now proceed to establish the existence of a continuum C+Σ of solutions of (Equation18) and identify bifurcation points for EquationEquation (18). This is provided in the following lemma.

Lemma 4.1.

The nonlinear eigenvalue equation given by (Equation18) has a continuum C+Σ, where (Δ,x)C+{(Δ0,0)} is a solution pair of (Equation18), where Δ0 is defined by (Equation22) and x=(x2,y2). C+ bifurcates at (Δ0,0), where Δ=Δ0 if and only if R0=1.

Proof.

First, the eigenvalues of L are λ+=x1φ(θ1d+θ2k3)βy1k1k3(γk4δ1ζφ) and λ=x1φ(θ1d+θ2k3)βy1k1k3(γk4δ1ζφ).

Let

(23) a12=φx1γk1y1,a21=(θ1d+θ2k3)γβk3(γk4δ1ζφ),(23)

and using these new notations we have λ+=a12a21 and λ=a12a21. Clearly, λ+ and λ are algebraically simple and nonzero. This implies that

(24) Δ0=1λ+=1/a12a21(24)

and Δ01=1λ=1/a12a21 are characteristic values of L. Corresponding to Δ0, the right characteristic vector of L is Δ0(a12a21,a21)τ,andΔ0(a21,a12a21) is the left characteristic vector. However, it should be noted that the characteristic vector corresponding to Δ01 is (1,a12a21a21) and it is not positive componentwise.

According to Rabinowitz (Citation1971) and Cushing (Citation1998), the eigenvalue problem given by (Equation18) has a continuum of nontrivial solutions, (Δ,x), of (Equation17) one which bifurcates from (Δ0,0) and the other from (Δ01,0). Since the left and right characteristic vectors corresponding to Δ0 are both positive componentwise, the continuum bifurcating from (Δ0,0), which we denote by C+ is positive, but the other continuum bifurcating from (Δ01,0) is not positive. Furthermore, since there is no other critical point different from (Δ0,0) and (Δ01,0), the branch C+ remains in the positive quadrant.

Finally, we show that Δ=Δ0 is equivalent to R0=1. To see this, observe that Δ=Δ0 implies βϕy1x1=βk1k3(γk4δ1ζφ)y1(θ1d+θ2k3)x1, then squaring both sides and simplifying terms we get ϕ2βφ(θ1d+θ2k3)y1γk1k3k4x1=1δ1ζφγk4. The left side of this equation is Rh2, which is given by (Equation6), and if we add (δ1ζφ2γk4)2 to both sides, then we get Rh2+(δ1ζφ2γk4)2=(1δ1ζφ2γk4)2 which reduces to δ1ζφ2γk4+(δ1ζφ2γk4)2+Rh2=1.

Next, we study the direction of the bifurcating branch C+. For points on the continuum C+ close to the critical point (Δ0,0), the Lyapunov–Schmidt expansion is given as follows.

(25) x=ϵV+ϵ2U+O(ϵ3)Δ=Δ0+ϵΔ1+O(ϵ2),where(25)
(26) V=(v1v2),U=(u1u2)(26)

are vectors in R+2 and Δ0 is defined by (Equation22). To this end, the direction of bifurcation is determined by the sign of Δ1. We use the expansion given by (Equation25) in (Equation20) from which we get h(Δ,x)=h~1ϵ2+O(ϵ3), where

(27) h~1=(h11h21)=(x1φγρk1k3y1(H1k2+dα)Δ0v1v2(θ1d+θ2k3)γβk3(γk4δ1ζφ)[dαv1ρk3k4v2γy1]Δ0v1).(27)

Since the term without ϵ is zero on both sides of (Equation18), we equate terms with coefficient ϵ, then ϵ2. To this end, from coefficients of ϵ we get V=Δ0LV, which implies that V is a right characteristic vector corresponding to the characteristic value Δ0. That is, V=1λ+(a12a21,a21)τ, which means, v1=1 and v2=a21/a12. Next, equating coefficients of ϵ2, we get U=Δ1LV+Δ0LU+h~1 which is equivalent to

(28) (IΔ0L)U=Δ1LV+h~1,(28)

where h~1=(h11,h21)τ is the coefficient of ϵ2 and is given by (Equation27). Since (IΔ0L) is singular, by Fredholm’s Alternative (Stakgold, Citation1979), EquationEquation (28) is solvable for U if the nonhomogeneous part is orthogonal to the null space of the adjoint operator to IΔ0L, which in this case is the transpose, (IΔ0L)τ. Suppose z is in the null space of (IΔ0L)τ, then (IΔ0L)τz=0 which is equivalent to zτ(IΔ0L)=0. This implies that zτ belongs to the space generated by the left characteristic vector of L corresponding to Δ0 which is (w1,w2)=1a12a21(a21,a12a21), that is,

(29) <(w1,w2),(Δ1LV+h~1)>=0.(29)

Next, we study the sign of Δ1 using the condition given by (Equation29) which is imposed on the forcing function in EquationEquation (28). As it is clear from (Equation25), the direction of bifurcation at (Δ0,0) is determined by the sign of Δ1. Our focus will be on the following equation which follows from (Equation29).

(30) Δ1=(w1h11+w2h21)a12w1v2+a21w2v1.(30)

Clearly, a12w1v2+a21w2v1 is positive, therefore, the sign of Δ1 is determined by the sign of

w1h11+w2h21. To this end,

(31) w1h11+w2h21=Δ0a21ρk3[H1k2+2dαρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1].(31)

We need to check the sign of w1h11+w2h21, where H1 is given by (Equation15). Define a threshold F by

(32) F=(d+μΛ)(r+μ)dμ+Λ(1D),whereD=r(Λ(ψ+μ)+d(ψ+Λ))(r+μ)(d+μΛ)(ψ+μ).(32)

The following proposition can be established by simple algebraic manipulations.

Proposition 4.2.

Suppose D and F are given by (Equation32) and r<d, then

(a)

D<1.

(b)

H1k2+2dα=(d+Λμ)(αF).

(c)

If R0<1, then ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1<ρk3k1k4ϕβφy1.

In this proposition (a) implies that F is positive. If 0α<F, Proposition (4.2) (b) implies that H1k2+2dα<0, thus w1h11+w2h21<0. In this case, Δ1>0 (see (Equation30)). Therefore, we have a forward bifurcation at (Δ0,0). We would like to remark that (Δ,0)=(Δ0,0) is equivalent to (R0,0)=(1,0), where 0 is (0,0)τ. This will clarify that the bifurcation curve emerges at R0=1 following the results from (Equation25).

On the other hand, if α>F, then H1k2+2dα>0. However, this alone does not make w1h11+w2h12 positive (see EquationEquation (31)), we need additional condition, that is,

(33) H1k2+2dα>ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1.(33)

then from (Equation31) we can see that w1h11+w2h21>0. Therefore, Δ1<0, and the bifurcation is backward. We like to remark that from Proposition (4.2) (c), when α>F, the inequality given by (Equation33) could be achieved if

(34) βϕ>ρk1k2k3k4φy1(H1+2dαk2)(34)

It is possible to see that F<1 and if α>F, then from Proposition (4.2) (b), the first term in (Equation31) is positive. Therefore, in this case, the sign of w1h11+w2h12 is determined by comparing the squares of (d+Λμ)(αF) and ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1. Denote the difference of the square of these terms by g(α), that is,

(35) g(α)=[(d+Λμ)(αF)]2[ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1]2=Aα2Bα+C,(35)

where A=(d+Λμ)2ρ2k42k1θ2βφy1x1(γk4δ1ζφ), B=2F(d+Λμ)2+ρ2k42k1(θ1d+2θ2(r+μ))βφy1x1(γk4δ1ζφ) and C=(d+Λμ)2F2ρ2k42k1(r+μ)(θ1d+θ2(r+μ)2)βφy1x1(γk4δ1ζφ).

Clearly g(F)<0, because the first term in g(α) is zero. Furthermore, since g(α) is continuous, there is an interval (F,F+η)(0,1),0<η<1F where g(α)<0. Thus, for these values of α, the bifurcation is forward. On the other hand, if y1 is increased, or βϕ is large enough, then it is possible to see that g(1)>0. Therefore, by intermediate value theorem, there is αc(F,1), where g(α)<0 for α(F,αc) whereas, g(α)>0 for α(αc,1). Thus, when the disease-induced death rate is higher than the threshold F, while there is high mosquito density and increased disease transmission, a backward bifurcation emerges if the disease-induced death rate exceeds a critical value, αc. For example, as it can be seen for parameters used in (), g(0.0013)<0 (the bifurcation is forward) and g(0.0062)>0 (the bifurcation is backward).

Theorem 4.3.

Suppose F is defined by (Equation32). The system given by (Equation1) exhibits

(a)

a backward bifurcation phenomena if

(i)

the disease-induced death rate, α>F and

(ii)

H1k2+2dα>ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1.

(b)

The bifurcation is forward when αF.

Figure 2. Bifurcation curves reflecting the results of theorem (4.3). Unstable branch (red), stable branch (blue), green (extinction); α = 0.0062 (continuous curve) and α = 0.0013 (dashed curve). Other parameters are γ=1/17, β = 0.15, ζ=0.77,θ1=0.0083,θ2=0.0063, μ=3.91×105,\upLambda=1.56×105,ρ=200,δ0=1100,r=1/7,δ1=0.0399,d=1/6,φ=1/8,ψ=0.00233,ϕ=2. based on the data we used in this figure, we have F=0.0015, for α=0.0062, H1k2+2dα=7.8937×104 and ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1=3.4734×105. for α=0.0013, H1k2+2dα=2.7186×105 and ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1=3.3930×105. the bifurcation is backward for α = 0.0062 and forward when α = 0.0013.

Figure 2. Bifurcation curves reflecting the results of theorem (4.3). Unstable branch (red), stable branch (blue), green (extinction); α = 0.0062 (continuous curve) and α = 0.0013 (dashed curve). Other parameters are γ=1/17, β = 0.15, ζ=0.77,θ1=0.0083,θ2=0.0063, μ=3.91×10−5,\upLambda=1.56×10−5,ρ=200,δ0=1100,r=1/7,δ1=0.0399,d=1/6,φ=1/8,ψ=0.00233,ϕ=2. based on the data we used in this figure, we have F=0.0015, for α=0.0062, H1k2+2dα=7.8937×10−4 and ρk3k4y1∗(θ1d+θ2k3)k1y1∗βφk3(γk4−δ1ζφ)x1∗=3.4734×10−5. for α=0.0013, H1k2+2dα=−2.7186×10−5 and ρk3k4y1∗(θ1d+θ2k3)k1y1∗βφk3(γk4−δ1ζφ)x1∗=3.3930×10−5. the bifurcation is backward for α = 0.0062 and forward when α = 0.0013.

The result in () supports what we highlight in Theorem (4.3). For the set of parameters used in () we have F=0.0015, then when α = 0.0062 we have H1k2+2dα=7.8937×104 and ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1=3.4734×105. Therefore, w1h11+w2h21>0 (see EquationEquation (31)) thus, Δ1<0 and the bifurcation is backward. On the other hand, for α = 0.0013 we get H1k2+2dα=2.7186×105, causing w1h11+w2h21 to be negative. Therefore, the bifurcation is forward.

5. Uniform persistence

From an ecological point of view, persistence in a given community results when each species in the community is saved from extinction in the long run. In epidemic models, the analogy is: if a disease is proved to be persistent, then it is endemic, where every state variable (number in a disease compartment) is bounded away from zero. We establish that the disease is uniformly persistent in an environment when the secondary reproduction number is greater than one. To proceed, we first define a threshold η as a function of ɛ (to be used in Lemmas 5.1 and 5.2).

(36) η(ε)=γk1k3k4[1ζδ1φγk4ϕ2βk(1ε)γk1k3k4×(θ1d+θ2k3)(y1ε)(1/x1ε)],(36)

where k1,k3 and k4 are given by (Equation4). Note that

(37) limε0η(ε)=γk1k3k4(1ζδ1φγk4Rh2).(37)

Remark: (1ζδ1φγk4Rh2)<0 for R0>1.

To see this, note that from (Equation5), (R0ζδ1φ2γk4)2=ζδ1φ2γk4+Rh2 which is equivalent to

(38) R02R0ζδ1φγk4=Rh2.(38)

Thus, when R0>1 we have R0ζδ1φ2γk4>1ζδ1φ2γk4 and since ζδ1φ2γk4<1 we have 1ζδ1φ2γk4>0. This implies (R0ζδ1φ2γk4)2>(1ζδ1φ2γk4)2 which is equivalent to R02R0ζδ1φγk4>1ζδ1φγk4. Then using (Equation38), we have Rh2>1ζδ1φγk4. Therefore, 1ζδ1φγk4Rh2<0 when R0>1.

In the following lemma and in Theorem (5.3), intΩ represents the interior of Ω (the set is defined by (Equation2)).

Lemma 5.1.

Suppose z0intΩ such that the orbit Φ(z0,t)=z(t) approaches the disease-free equilibrium E0 as t. Then, there is ɛ > 0 and a large to such that

βϕy3x1Nβϕy3(1ε)

and

ϕθ1x3y1N+ϕθ2x2y1Nϕ(θ1x3+θ2x2)(y1ε)(1x1ε)fortt0.

Proof.

Since the orbit Φ(z0,t)=z(t) approaches E0 as t, we have limt1N(t)=1x1, limtx1(t)N(t)=1 and limty1(t)=y1. Moreover, given ε>0, there are large t1, t2 and t3 such that 1x1ε<1N<ε+1x1 for all t>t3, |x1(t)N(t)1|<ε for all tt1 and also |y1y1|<ε for all tt2, which means 1ε<x1(t)N(t), y1>y1ε for tt0=max{t1,t2,t3}. These results lead to

βϕy3x1Nβϕy3(1ε)

and

ϕθ1x3y1N+ϕθ2x2y1Nϕ(θ1x3+θ2x2)(y1ε)(1x1ε)fortt0.

Next, in preparation to establish the uniform persistence of system (Equation1), we define a matrix

(39) B=(Θ1Θ2Θ3Θ4),(39)

where

Θ1=((d+μ1)00d(r+α+μ1)00r(ψ+μ1)),Θ2=(0βϕ(1ε)0000)
Θ3=(ϕθ2(y1ε)(1/x1ε)ϕθ1(y1ε)(1/x1ε)0000),
Θ4=((φ+γ)δ1ζφγ).

Lemma 5.2.

B has a positive eigenvalue when R0>1.

Proof.

We begin with the characteristic equation of B which is f(h)=(h+k2)(h4+b3h3+b2h2+b1h+b0), where

(40) b3=γ+k1+k3+k4,b2=k3k4+γk1φδ1ζ,b1=k3k4(γ+k1)+γk1(k3+k4)δ1ζ(k1+k3)ϕ2θ2β(1ε)M,b0=γk1k3k4[1ζδ1φγk4ϕ2βφ(1ε)γk1k3k4(θ1d+θ2k3)(y1ε)(1x1ε)].(40)

Note that b0=η(ε), where η(ε) is given by (Equation36).

Thus, since limε0η(ε)=γk1k3k4(1ζδ1φγk4Rh2)<0 and η is continuous as a function of ɛ, there is 0<ε<min{1,1x1} such that η(ε)<0. Therefore, R0>1 ensures b0<0. Clearly the coefficient b3 is positive. Also b2(k3γ1)ζδ1φ+γk1 is positive, which is possible if γ is reduced and α is increased. Regardless of the sign of b1, there is exactly one change of sign in the coefficient sequence {b0,b1,b2,b3}. Therefore, by Descartes’s rule of sign (Anderson et al., Citation1998), f(h)=0 has one positive root when R0>1.

Theorem 5.3.

If R0>1, the system given by (Equation1) is uniformly persistent, that is, the boundary equilibrium E0 is not the ωlimit set of any orbit in intΩ, where Ω is given by (Equation2).

Proof.

Recall that Ω is forward invariant under system (Equation1) (Blayneh, Citation2017). Define a set A={(x1,0,0,0,y1,0,0):0<x1ρμ1λand0<y1δ0γδ1}. Clearly zAω(z)={E0}, where E0=(x1,0,0,0,y1,0,0) is the disease-free equilibrium of (Equation1) and ω(z) is the omega limit set of a solution with initial zA. Moreover, it should be noted that {E0} is isolated compact and invariant in Ω. To prove the uniform persistence of the disease, we need to prove that {E0} is a weak repeller for intΩ (Thieme, Citation2004). The proof goes by contradiction. To begin with, suppose there is an initial point z0=(x10,x20,x30,x40,y10,y20,y30) all positive components and z0intΩ, such that the orbit Φ(z0,t)=z(t) approaches E0 as t. By Lemma (5.1), there is ɛ > 0 and a large t0, such that

βϕy3x1Nβϕy3(1ε)

and

ϕθ1x3y1N+ϕθ2x2y1Nϕ(θ1x3+θ2x2)(y1ε)(1x1ε)fortt0.

Let F(X,Y)=(f1,f2,f3,f4,f5) be the vector field in EquationEquation (1) associated with the components (x2,x3,x4,y2,y3), in that order. Then f1βϕ(1ε)y3(d+μ1)x2 and f4ϕ(θ1x3+θ2x2)(y1ε)(1x1ε)+ζδ1y3(φ+γ)y2. The foregoing discussion leads to the following differential inequality.

(41) dx2dtβϕ(1ε)y3(d+μ1)x2,dx3dtdx2(r+α+μ1)x3dx4dtrx3(ψ+μ1)x4,dy3dtφy2γy3dy2dtϕ(θ1x3+θ2x2)(y1ε)(1x1ε)+ζδ1y3(φ+γ)y2.(41)

Observe that the Jacobian of the new vector field, representing the right side of (Equation41) at (0,0,0,0,0) is B, the matrix given by (Equation39). Now consider the linear system dzdt=Bz. Note that B=(bij),i,j=1,,5 has the following properties.

(i)

For every proper subset I of N={1,2,3,4,5}, there is iI and jNI such that bij0. Thus, B is irreducible,

(ii)

B is quasi-positive that is, bij0,ij and i,j{1,,5}, then (i) and (ii) insure that B has a simple real eigenvalue which dominates the real parts of all eigenvalues of B, that is, λ0>ϑ,ϑσ(B)(note that ϑis the real part ofϑ) with a positive eigenvector v (Theorem A .45 in (Thieme, Citation2004)). Furthermore, this along with the result of Lemma (5.2) ensures that B has exactly one positive eigenvalue when R0>1. Consequently, λ0>0. Now, pick an initial point z(τ)=(x2(τ),x3(τ),x4(τ),y(τ),y3(τ)) on the manifold rv:r>0, specifically xi(τ)>r0vi1,i=2,3,4, y2(τ)>r0v4 and y3(τ)>r0v5. Clearly, Bv=λ0v>0, then since the semiflow of dzdt=Bz is order preserving, the orbit z(t,z(τ)) goes to +. Additionally, (Equation41) implies that the solutions of dzdt=Bz are bounded from above by (x2(t),x3(t),x4(t),y2(t),y3(t)) for all tτ (Hall, Citation1969). This contradicts the fact that Ω is forward invariant under system (Equation1). Thus, {E0} is a weak repeller for intΩ (Thieme, Citation1993). This completes the proof.

6. Discussion and conclusion

One of the main challenges in disease control is failure to use the reduction of secondary reproduction number (R0) below unity as means to predict disease elimination. This happens when a backward bifurcation exists. In this case, two endemic equilibrium points exist for R0<1. We used a system of differential equations to study the dynamics of a vector-borne disease and conditions related to existence of a backward bifurcation. Our approaches implement analytical and simulation techniques. Bifurcation analysis of the model in Section (2) is presented in Section (3), where Theorem (4.3) reveals that when the disease-induced death rate in the host is reduced below a threshold value given by F there is no backward bifurcation, which means control efforts lead to disease elimination. Still, for some values of α>F, the bifurcation is forward however, but if α exceeds a critical value, then the disease becomes endemic. Increased mosquito level and disease transmission between mosquito and human play a role here. () offers numerical results regarding the backward bifurcation and parameters involved to this effect. Other challenges in a vertically transmitted vector-borne disease are related to transmissions of the disease from female mosquito to egg which is referred to as vertical transmission. Graphical results showing the effect of this transmission in raising the peak of epidemic are presented in (). Numerical simulations are also presented where the infectious population from mosquitoes and humans approach endemic equilibrium levels in the long run even if R0<1 as given in ().

Figure 3. Endemic asymptotic level of infectious mosquitoes and humans for R0<1. ζ = 0.77 α=0.038, γ=1/18, β=0.87,θ1=0.0079,θ2=0.0072, μ=3.91×105,\upLambda=1.48×105,ρ=200,δ0=1120,r=1/7,δ1=0.0397,d=1/6,ψ=0.00233,ϕ=2,φ=1/8. for these set of parameters, R0=0.3944<1, but mosquito and human populations equilibrate to endemic levels (see time plots (a) and (c) with initial value (20,40,1500,40,2000,40,10000)). The bifurcation diagrams of mosquitoes and humans as R0 changes given in windows (b) and (d) show backward bifurcations where two endemic equilibria exists for R0<1.

Figure 3. Endemic asymptotic level of infectious mosquitoes and humans for R0<1. ζ = 0.77 α=0.038, γ=1/18, β=0.87,θ1=0.0079,θ2=0.0072, μ=3.91×10−5,\upLambda=1.48×10−5,ρ=200,δ0=1120,r=1/7,δ1=0.0397,d=1/6,ψ=0.00233,ϕ=2,φ=1/8. for these set of parameters, R0=0.3944<1, but mosquito and human populations equilibrate to endemic levels (see time plots (a) and (c) with initial value (20,40,1500,40,2000,40,10000)). The bifurcation diagrams of mosquitoes and humans as R0 changes given in windows (b) and (d) show backward bifurcations where two endemic equilibria exists for R0<1.

Moving forward, we like to clarify that baseline parameters which are widely used in the literature (see ()) are implemented in the numerical results of this article. We list some of them here. Estimation of parameters for mortality rate of mosquitoes such as Ades aegypti is based on the average life of mosquitoes which is from 14 to 17 days. The incubation period of dengue virus in a mosquito is from 7 to 14 days which is the base for estimating φ to be between 114 and 17. Thus, we estimate φ by 18. Also, the recovery from dengue virus is about seven days, therefore, the recovery rate r is considered to be in [0,17]. In most developing countries, human (host) life expectancy (in years) is in the range from 45 to 70. Accordingly, human mortality rate is estimated to be in the interval [170(365),145(365)]. Disease transmission rate from exposed hosts, θ2 is estimated to be 0.0063, and it is less than θ1 which is estimated to be 0.0083, θ1 is transmission from infectious hosts to mosquitoes.

Table 2. Parameter estimation for dengue fever. Rate is per day

() is interpreted as follows: Vertical transmission of the disease in the mosquito population increases epidemic level. Keeping other parameters fixed, we increased the vertical transmission rate ζ. We selected three different values, namely 0.0046, 0.3046 and 0.5046, then observed that an increase of this rate by 30% causes the epidemic level to increase. The pick epidemic level is higher for a 50% increase in vertical transmission rate. These results are represented graphically in (). It should be clear that in each case the secondary reproduction number, R0, is less than unity: here are the corresponding values: R0=0.0673 (ζ=0.0046), R0=0.1662 (ζ=0.3046) and R0=0.2492 (ζ=0.5046). If disease prevention efforts focus on adult as well as larva control, then the effect of this transmission on raising the epidemic level could be diminished.

Simulation results given in () illustrate some of the conclusions drawn from Theorem (4.3). The data and estimation of key parameters are based on dengue virus and supporting references are provided in (). In this theorem, a backward bifurcation is possible when Δ1<0, which is the coefficient of ϵ in the Lyapunov–Smith expansion given by EquationEquation (25). Additionally, the sign of w1h11+w2h12 and Δ1 are opposite (see (Equation30)). Furthermore, the theorem establishes that reduction of the disease-induced death rate, α, below a threshold denoted by F will eliminate backward bifurcation. Backward bifurcation is a phenomena where endemic equilibrium points exist despite managing to reduce the secondary reproduction number below unity. The main bifurcation result in connection with model parameters states that the sign of the right term in EquationEquation (31) will determine the direction of bifurcation. We list what the connection between this theorem and the numerical results in () where estimation of parameters is done based on dengue virus . For the set of parameters used in (), we have F=0.0015. Clearly, for α=0.0013<0.0015, the bifurcation is forward. However, when α=0.0062>F, the bifurcation is backward, as the sign of Δ1<0. For this value of α, H1k2+2dα=7.8937×104 and ρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1=3.4734×105. Some numerical results could support the fact that having α>F may not alone change the bifurcation result. For example, in addition to value of α used in (), at α=0.0016, H1k2+2dαρk3k4y1(θ1d+θ2k3)k1y1βφk3(γk4δ1ζφ)x1=2.2807\break×1053.3980×105=1.1172×106. The bifurcation analysis used here help point the role played by the fatality of the disease as measured by the parameter for disease-induced death rate.

As it is explained early, what is given in () is that when there is backward bifurcation, solutions with some initial value equilbrate to a positive value even if R0<1. The graphs of the infectious populations are plotted. From this one can conclude that the other components, such as exposed also approach an endemic equilibrium levels. Although in Section (5), persistence of the disease is established for R0>1, the result we see in () suggested possibility of persistence of the disease when conditions for backward bifurcation are met. It should be noted that for the set of parameter values used in (), R0=0.3944 which is less than one.

Based on our model, without vector-host transmission cycle, vertical transmission alone, cannot cause the disease to be endemic or create a backward bifurcation phenomena. However, vertical transmission in the vector could increase the prevalence and epidemic level of the disease in the vector and host populations. Therefore, efforts to mitigate the effects of vertical transmissions which primarily should focus on mosquito control as well as improved measures against disease transmission, such as strong personal protection. Furthermore, disease-induced death rate, if kept low may not cause endemic equilibrium, this means as long as the secondary reproduction number, R0, is kept below one, the disease could not establish in a given community. If mosquito level and mosquito-human disease transmission cycles are increased and the disease-induced death rate in hosts passes a critical level, then the disease could be endemic even when R0, is reduced below one. We conjecture that there could be a critical value of this threshold, below which disease elimination is possible. When the secondary reproduction number is increased beyond one, the disease is uniformly persistent, each disease compartment is bounded away from zero.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Adams, B., & Boots, M. (2010). How important is vertical transmission in mosquitoes for the persistence of dengue? Insights from a mathematical model. Epidemics, 2(1), 1–10. https://doi.org/10.1016/j.epidem.2010.01.001
  • Allen, L. (2007). An introduction to mathematical biology. Pearson Prentice Hall.
  • Anderson, B., Jackson, J., & Sitharam, M. (1998). Descarte’s rule of signs revisited. The American Mathematical Monthly, 105(5), 447–451. https://doi.org/10.1080/00029890.1998.12004907
  • Blayneh, K. W. (2017). Vertically transmitted vector-borne diseases and the effects of extreme temperature. International Journal of Apllied Mathematics, 30(2), 177–209. https://doi.org/10.12732/ijam.v30i2.8
  • Blayneh, K., Cao, Y., & Kwon, H.-D. (2009). Optimal control of vector-borne diseases: Treatment and prevention. Discrete and Continuous Dynamical Systems Series B, 11(3), 587–611. https://doi.org/10.3934/dcdsb.2009.11.587
  • Blayneh, K. W., Gumel, A. B., Lenhart, S., & Clayton, T. (2010). Backward bifurcation and optimal control in transmission dynamics of West Nile virus. Bulletin of Mathematical Biology, 72(4), 1006–1028. https://doi.org/10.1007/s11538-009-9480-0
  • Centers for Disease Control and Prevention. http://www.cdc.gov/dengue
  • Centers for Disease Control and Prevention: Malaria. http://www.cdc.gov/Malaria/faq.htm2
  • Chan, M., Johansson, M. A., & Vasilakis, N. (2012). The incubation periods of dengue viruses. PLoS ONE, 7(11), e50972. https://doi.org/10.1371/journal.pone.0050972
  • Chitnis, N., Cushing, J. M., & Hyman, J. M. (2006). Bifurcation analysis of a mathematical model for malaria transmission. SIAM Journal on Applied Mathematics, 67(1), 24–45. https://doi.org/10.1137/050638941
  • Chitnis, N., Hyman, J. M., & Manore, C. A. (2013). Modelling vertical transmission in vector-borne diseases with application to Rift Valley fever. Journal of Biological Dynamics, 7(1), 11–40. https://doi.org/10.1080/17513758.2012.733427
  • Cruz-Pacheco, G., Esteva, L., & Vargas, C. (2009). Seasonality and outbreaks in West Nile varus infection. Bulletin of Mathematical Biology, 71(6), 1378–1393. https://doi.org/10.1007/s11538-009-9406-x
  • Cushing, J. M. (1998). An introduction to structured population dynamics. SIAM.
  • Dushoff, J., Wenzhang, H., & Castillo-Chavez, C. (1998). Backwards bifurcations and catastrophe in simple models of fatal diseases. Journal of Mathematical Biology, 36(3), 227–248. https://doi.org/10.1007/s002850050099
  • Feng, Z., & Velasco-Hernández, J. X. (1997). Competitive exclusion in a vector-host model for the dengue fever. Journal of Mathematical Biology, 35(5), 523–544. https://doi.org/10.1007/s002850050064
  • Francisco, P. M. (2007). Factors affecting the emergence and prevalence of vector borne infections and the role of vertical transmission. Journal of Vector Borne Diseases, 44, 157–163.
  • Gaff, H. D., Hartley, D. M., & Leahy, N. P. (2007). An epidemiological model of Rift Valley fever. Electronic Journal of Differential Equations, 2007(115), 1–12.
  • Gordillo, L. F. (2013). Reducing mosquito-borne disease outbreak size: The relative importance of contact and tranmissibility in a network model. Applied Mathematical Modeling, 37(18–19), 8610–8616. https://doi.org/10.1016/j.apm.2013.03.056
  • Gourley, S. A., Röst, G., & Thieme, H. R. (2014). Uniform persistence in a model for bluetongue dynamics. SIAM Journal on Mathematical Analysis, 46(2), 1160–1184. https://doi.org/10.1137/120878197
  • Hall, J. K. (1969). Ordinary differential equations. Wiley-Interscience.
  • Hubálek, Z., & Halouzka, J. (1999). West Nile fever. A reemerging mosquito-borne viral disease in Europe. Emerging Infectious Diseases, 5(5), 643–650. https://doi.org/10.3201/eid0505.990505
  • Joseph, W., So, H., & Freedman, H. I. (1986). Persistence and global stability in a predator-prey model consisting of three prey genotypes with fertility differences. Bulletin of Mathematical Biology, 48(5), 469–484. https://doi.org/10.1016/S0092-8240(86)90002-9
  • Pu, L., Lin, Z., & Lou, Y. (2023). A West Nile virus nonlocal model with free boundaries and seasonal succession. Journal of Mathematical Biology, 86(2), 25. https://doi.org/10.1007/s00285-022-01860-x
  • Rabinowitz, P. H. (1971). Some global results for nolinear eigenvalue problems. Journal of Functional Analysis, 7(3), 487–513. https://doi.org/10.1016/0022-1236(71)90030-9
  • Robelo, C., Margheri, A., & Bacaër, N. (2014). Persistence in some periodic epidemic models with infection age or constant periods of infection. Discrete and Continuous Dynamical Systems-Series B(DCDS-B), 19(4), 1155–1170. https://doi.org/10.3934/dcdsb.2014.19.1155
  • Salako, R. B. (2023). The impact of population size and movement on the persistence of a two-strain infectious disease. Journal of Mathematical Biology, 86(1), 5. https://doi.org/10.007/s00285-022-01842-z
  • Stakgold, I. (1979). Green’s functions and boundary value problems. Wiley-Interscience.
  • Thieme, H. R. (1993). Persistence under relaxed point-dissipativity (with applications to an epidemic model). SIAM Journal on Mathematical Analysis, 24(2), 407–435. https://doi.org/10.1137/0524026
  • Thieme, H. R. (2001). Disease extinction and disease persistence in age structured epidemic models, H.R. Thieme, nonlinear analysis, theory. Methods & Applications, 47(9), 6181–6194. https://doi.org/10.1016/S0362-546X(01)00677-0
  • Thieme, H. R. (2004). Mathematics in population biology, Princeton series in theoretical and computational Biology. Princeton.
  • van den Driessche, P., & Watmough, J. (2002). Reproduction numbers and sub-threshold endemic equibria for compartmental models of disease transmission. Mathematical Biosciences, 180(1–2), 29–48. https://doi.org/10.1016/S0025-5564(02)00108-6
  • van den Driessche, P., & Yakubu, A.-A. (2019). Disease extinction versus persistence in discrete-time epidemic models. Bulletin of Mathematical Biology, 81(11), 4412–4446. https://doi.org/10.1007/s11538-018-0426-2
  • Wang, X., Chen, Y., & Martcheva, M. (2020). Asymptotic analysis of a vector-borne disease model with the age of infection. Journal of Biological Dynamics, 14(1), 332–367. https://doi.org/10.1080/17513758.2020.1745912
  • WHO. https://who.int/news-room/fact-sheets/detail/dengue-and-sever-dengue.
  • Zheng, B. (2022). Impact of releasing period and magnitude on mosquito population in a sterile release model with delay. Journal of Mathematical Biology, 85(2), 18. https://doi.org/10.1007/s00285-022-01785-5
  • Zhu, Z., Hui, Y., & Hu, L. (2023). The impact of predators of mosquito larvae on wolbachia spreading dynamics. Journal of Biological Dynamics, 17(1), 1. https://doi.org/10.1080/17513758.2023.2249024