424
Views
0
CrossRef citations to date
0
Altmetric
Special Issue in Memory of Abdul-Aziz Yakubu

Mathematical model of Ehrlichia chaffeensis transmission dynamics in dogs

, &
Article: 2287082 | Received 07 Jun 2023, Accepted 17 Nov 2023, Published online: 11 Dec 2023

Abstract

Ehrlichia chaffeensis is a tick-borne disease transmitted by ticks to dogs. Few studies have mathematical modelled such tick-borne disease in dogs, and none have developed models that incorporate different ticks' developmental stages (discrete variable) as well as the duration of infection (continuous variable). In this study, we develop and analyze a model that considers these two structural variables using integrated semigroups theory. We address the well-posedness of the model and investigate the existence of steady states. The model exhibits a disease-free equilibrium and an endemic equilibrium. We calculate the reproduction number (T0). We establish a necessary and sufficient condition for the bifurcation of an endemic equilibrium. Specifically, we demonstrate that a bifurcation, either backward or forward, can occur at T0=1, leading to the existence, or not, of an endemic equilibrium even when T0<1. Finally, numerical simulations are employed to illustrate these theoretical findings.

MATHEMATICS SUBJECT CLASSIFICATIONs:

1. Introduction

Ehrlichiosis are tick-borne diseases caused by obligate intracellular rickettsia bacteria in the genera Ehrlichia [Citation1–5]. These bacteria are classified within the group of the α-proteobacteria, order Rickettsiales, family Anaplasmataceae, genus Ehrlichia [Citation3,Citation6]. This genus consists of obligate intracellular Gram-negative bacteria [Citation1,Citation4,Citation6] that mainly infect leukocytes (such as monocytes, macrophages, granulocytes and neutrophils), and endothelial cells in mammals, and salivary glands, intestinal epithelium, and hemolymph cells of ticks [Citation1,Citation7–12]. The genus Ehrlichia comprises of six recognized tick-transmitted species: E. canis, E. muris, E. chaffeensis, E. ewingii, E. minasensis, and E. ruminantium [Citation3,Citation11], and in recent times, other Ehrlichia species have been reported [Citation7,Citation8,Citation11]. Going by our current knowledge, a large number of Ehrlichia species might not have been described [Citation11].

The reservoir hosts for Ehrlichia species include numerous wild animals, as well as some domesticated species and livestock [Citation6]. Some of these Ehrlichia species affect animals including pets such as cats and dogs [Citation6,Citation13], and a limited number have been know to infect humans [Citation2,Citation4,Citation6,Citation13–16]. Specifically, Ehrlichia canis, E. chaffeensis, E. ewingii, E. muris, and E. ruminantium are members of the genus Ehrlichia known to naturally infect various mammalian hosts such as cats, dogs, ruminants, and mice, and are responsible for emerging zoonoses in humans [Citation7,Citation8,Citation10,Citation17].

Ehrlichia are tick-borne diseases transmitted by ticks in the family Ixodidae. Rhipicephalus sanguineus (brown dog tick) and Ixodes ricinus are vectors for Ehrlichia canis [Citation5,Citation6]. E. canis can also be transmitted by Amblyomma cajennense, and experimentally by Dermacentor variabilis (American dog tick) [Citation6]. In North America A. americanum (Lone Star tick) is the primary vector for both E. chaffeensis and E. ewingii [Citation4,Citation6,Citation18,Citation19]. ‘Outside North America, E. chaffeensis has been found in ticks in the genera Amblyomma, Haemaphysalis, Dermacentor, and Ixodes in Asia, in R. sanguineus in Cameroon, and in A. parvum in Argentina’ [Citation6]. Haemaphysalis flava and Ixodes persulcatus complex ticks transmits E. muris [Citation6]. E. ruminantium is the only known Ehrlichia species that infects cattle and it is found on the continent of Africa and a couple of Caribbean islands [Citation8].

Ticks life-cycle span between 2-to-3 years depending on the species [Citation20,Citation21]. They go through four life stages namely egg, six-legged larva, eight-legged nymph, and adult. To survive each life stage after hatching from eggs, the ticks must take a blood meal from a host; but, most ticks will die if they are unable to find a host [Citation20]. Ticks feed once on a host, then fall off and develop into the next stage. This feeding pattern creates pathways for diseases to be transmitted from hosts to hosts [Citation22]. Ixodes scapularis (black-legged tick) life cycle generally lasts two years, while the life cycle of Amblyomma americanum (the lone star tick) is around three years long [Citation21]. Some species like Rhipicephalus sanguineus (the brown dog tick) whose primary host is the domestic dog, prefer to feed on the same host during all its life stages [Citation20,Citation22]. Since R. sanguineus are endophilic, they are found inside houses and dog kennels [Citation22].

Dogs are susceptible to infection with multiple Ehrlichia spp., including E. chaffeensis, E. ewingii, and E. canis [Citation2,Citation12,Citation19,Citation23,Citation24]. In 2019, over 200,000 dogs tested positive for antibodies against Ehrlichia spp. within the United States out of 7,056,709 dogs tested, while over 1000 dogs tested positive out of 168,216 dogs tested in Canada [Citation25]. According to Gettings et al. [Citation25] the distribution of infected dogs follows the distribution of the related tick vectors. For instance, Amblyomma americanum is commonly found on dogs and people in the southeastern and southcentral United States [Citation2]. In Beall et al. [Citation2] the overall seroprevalence of E. canis, E. chaffeensis, and E. ewingii across the United States in 2012 was 0.8%, 2.8%, and 5.1%, respectively. The highest E. canis seroprevalence of 2.3% was found in Arkansas, Louisiana, Oklahoma, Tennessee, and Texas. E. chaffeensis seroreactivity was 6.6% in Arkansas, Kansas, Missouri, and Oklahoma (the central region), and 4.6% in Georgia, Maryland, North Carolina, South Carolina, Tennessee and Virginia (the southeast region). Seroreactivity of E. ewingii was highest in the central region with 14.6% value while the southeast region had a seroreactivity value of 5.9%.

Ehrlichia in dogs was discovered in the 1970s when military dogs were returning from the Vietnam war. They found this disease to be extremely severe in German Shepards, Doberman Pinschers, Belgium Malinois, and Siberian Huskies. Several studies including experimental researches and serological surveys have been carried out to understand Ehrlichiosis transmission in canine including Ehrlichia Chaffeensis [Citation1,Citation15,Citation26–28]. We only found two quantitative studies modelling Ehrlichiosis in human [Citation29,Citation30]. Thus, in this study, we developed a mathematical model for Amblyomma americanum in the Great Plains and dogs infected with Ehrlichia Chaffeensis with the aim of understanding the quantitative properties of the model including the asymptotic dynamics of the disease in dogs across the Plains.

The rest of the paper is organized as follows: Section 2 gives the description of the model including the interactions between the ticks and dogs; Section 3 describes the results of the study. These results include the existence and uniqueness of solutions, the disease invasion process, and the bifurcation analysis. More precisely, we show that, depending on the sign of a constant Cbif – referred to as the bifurcation parameter – which depends on model's parameters, a bifurcation occurs at T0=1 that is either a forward or backward bifurcation. In a forward bifurcation, which occurs when Cbif<0, there exists a unique endemic equilibrium if and only if T0>1. However, in a backward bifurcation, which occurs when Cbif>0, no endemic equilibrium exists for T01 small enough; a unique endemic equilibrium exists if T0>1 while multiple equilibria exist when 0T0<1 close enough to 1. Finally in Section 4 we numerically illustrate such bifurcation results.

2. Model formulation

The transmission model of Ehrlichia Chaffeensis incorporates two subgroups: dogs and ticks. At any time t, the dog population is divided into susceptible SD(t), infected iD(t,a), and chronically infected cD(t,a), recovered RD(t). Here, the variable a represents the time since infection. Thus, the total number of dogs at time t is quantified by ND(t)=SD(t)+0iD(t,a)da+0cD(t,a)da+RD(t).The ticks population is structured into several stages: eggs (E), larva (L), nymph (N), and adult (A). We denote by S={E,L,N,A} the set of ticks stages. Let STk(t) be the number of susceptible ticks of stage KS at time t. We also denote by iTk(t,a) the number of infected ticks of stage KS{E} at time t and which are infected since time a. The model proposed assumes that there are no infected eggs. Therefore the total number of ticks of stage kS{E} at time t is given by (1) NTk(t)=STk(t)+0iTk(t,a)da.(1) The dogs-ticks infection life cycle is shown in Figure .

Figure 1. Flow chart of the model.

Figure 1. Flow chart of the model.

Dogs' population dynamic. At any time t, infected ticks (which are infected since time a) induce an infection within the dogs' population through the force of infection λT(t,a), such that (2) λT(t,a)=βT(a)KS{E}iTk(t,a),(2) where βT denotes the infectivity of infected ticks. Therefore, newly infected dogs are given by SD(t)ND(t)0λT(t,a)da and ϵRD(t)ND(t)0λT(t,a)da, where ϵ is constant parameter accounting for the relative disease transmission to previously recovered dogs. Note that ϵ can be considered age-dependent (ϵ=ϵ(a)) with no more difficulties in the analysis proposed here. Thus, the dogs population dynamic is then described by the below system (3) {dSD(t)dt=θDSD(t)ND(t)0λT(t,a)daμDSD(t),iD(t,0)=SD(t)+ϵRD(t)ND(t)0λT(t,a)da,iD(t,a)t+iD(t,a)a=(μD+γ(a)+δD(a)+νD(a))iD(t,a),cD(t,a)t+cD(t,a)a=(μD+δC(a))cD(t,a)cD(t,0)=0νD(a)iD(t,a)da,dRD(t)dt=0γ(a)iD(t,a)daϵRD(t)ND(t)0λT(t,a)daμDRD(t).(3) In the above model, susceptible dogs are recruited at rate θD and all dogs die naturally at rate μD. Infected and chronically infected dogs have a disease-induced mortality rate δD and δC. Infected dogs progress to a chronic infection at rate νD. Finally, only non-chronic infections are assumed to recover from the infection at rate γ.

Ticks' population dynamic. At time t, infected dogs (chronic or not) induce an infection within the ticks population through the force of infection λD(t), such that λD(t)=0[βD1(a)iD(t,a)+βD2(a)cD(t,a)]da,where βDj(a) denotes the infectivity of infected and chronically infected dogs a-time post infection. Therefore, newly infected ticks of stage kS{E} are given by STk(t)ND(t)λD(t). The ticks population dynamic is then described by the system below (4) {dSTE(t)dt=rE(1STE(t)K)NTA(t)(αE+μE)STE(t),dSTL(t)dt=αESTE(t)STL(t)ND(t)λD(t)(αL+μL)STL(t),dSTN(t)dt=αLSTL(t)STN(t)ND(t)λD(t)(αN+μN)STN(t),dSTA(t)dt=αNSTN(t)STA(t)ND(t)λD(t)μASTA(t),and for kS{E},iTk(t,0)=STk(t)ND(t)λD(t),iTk(t,a)t+iTk(t,a)a=(μk+αk)iTk(t,a)(4) with αA=0. In the above system of the ticks dynamic, the eggs' production rate at time t is rE(1STE(t)K)NTA(t), where rE is the number of eggs produced by adult ticks NTA, and the parameter K is the eggs' carrying capacity. Parameters αk,s are ticks progression rates from the eggs' to adult' stage. The death rate of ticks at each k-stage is μk. The notations of all variables and parameters are summarized in Table .

Table 1. State variables and parameters used for simulations.

3. Main results

This section is devoted to the main results of this paper. Overall, our results are based on the following assumption on the parameters of system (Equation3)–(Equation4). More precisely, we assume that

Assumption 3.1

(1)

The recruitment rate θD of susceptible dogs and the natural death rate μD are positive constants. The ticks' stage transition rates αk,s and death rates μk,s are positive constants. The parameter ϵ, rE and the carrying capacity K are positive constants;

(2)

Parameters γ,δD,νD,δC and the transmission rates βT,βDk, k = 1, 2, belong in L+((0,+),R){0L};

(3)

The initial condition is such that SD(0)=SD0>0, RD(0)=RD00, iD(0,)=iD0L+1((0,+),R), cD(0,)=cD0L+1((0,+),R), STk0(0)=STk00, and iTk(0,)=iTk0L+1((0,+),R) for kS.

3.1. Existence and uniqueness of nonnegative solution

In this section, we state the results concerning the existence of a globally defined nonnegative solution to system (Equation3)–(Equation4).

In order to state our main results, we will rewrite the system in a more appropriate equivalent form. To do this, denote by ek,k{1,2,3} and vk, k{1,2} the canonical basis of R3 and R2, respectively. Thus, we can consider the states variables in a vector form by setting for every t0 and a0 {ST(t)=STL(t)e1+STN(t)e2+STA(t)e3,iT(t,a)=iTL(t,a)e1+iTN(t,a)e2+iTA(t,a)e3,iD(t,a)=iD(t,a)v1+cD(t,a)v2.Moreover, we define the vector βD(a) of transmission rates with components βD1(a) and βD2(a), i.e. βD(a)=βD1(a)v1+βD2(a)v2,and define the vector 1=e1+e2+e3. Let M be ticks stage progression matrix from the larval stage to the adult stage, i.e. (5) M:=(μL+αL00αLμN+αN00αNμA).(5) Therefore, using the notation , to denotes the inner product in R2 and R3 and setting for all a0 (6) θ(a):=(γ(a)+δD(a)+νD(a)00δC(a)),(6) Model (Equation3)–(Equation4) takes the following form (7) {dSD(t)dt=θDSD(t)ND(t)0βT(a)1,iT(t,a)daμDSD(t),iD(t,a)t+iD(t,a)a=(μD+θ(a))iD(t,a),iD(t,0)=SD(t)+ϵRD(t)ND(t)0βT(a)1,iT(t,a)dav1+0νD(a)v1,iD(t,a)dav2,dRD(t)dt=0γ(a)v1,iD(t,a)daϵRD(t)ND(t)0βT(a)1,iT(t,a)daμDRD(t),(7) and (8) {dSTE(t)dt=rE(1STE(t)K)(ST(t),e3+0iT(t,a),e3da)(αE+μE)STE(t),dST(t)dt=αESTE(t)e1MST(t)ST(t)ND(t)0βD(a),iD(t,a)da,iT(t,a)t+iT(t,a)a=MiT(t,a),iT(t,0)=ST(t)ND(t)0βD(a),iD(t,a)da,(8) subject to the initial condition SD(0)=SD0>0, iD(0,)=iD0L+1((0,+),R2), RD(0)=RD0R+, STE(0)=STE0R+, iD(0,)=iD0L+1((0,+),R2), ST(0)=ST0R+3, and iT(0,)=iT0L+1((0,+),R3). The following result then concerns the existence and uniqueness of nonnegative solutions to (Equation7)–(Equation8).

Theorem 3.2

Let Assumption 3.1 be satisfied. Then there exists a unique continuous globally defined integrated solution to (Equation7)–(Equation8). Moreover, if we set (9) Π(a,τ):=eτa(μD+θ(l))dl,aτ0.(9) then the solution satisfies, for all t0 and a0, {dSD(t)dt=θDSD(t)ND(t)0βT(a)1,iT(t,a)daμDSD(t),dRD(t)dt=0γ(a)iD(t,a)daϵRD(t)ND(t)0βT(a)1,iT(t,a)daμDRD(t),dSTE(t)dt=rE(1STE(t)K)(ST(t),e3+0iT(t,a),e3da)(αE+μE)STE(t),dST(t)dt=αESTE(t)e1MST(t)ST(t)ND(t)0βD(a),iD(t,a)da,iD(t,a)={Π(a,at)iD0(at),ifatSD(ta)+ϵRD(ta)ND(ta)0βT(a)1,iT(ta,τ)dτΠ(a,0)v1+0νD(τ)v1,iD(ta,τ)dτΠ(a,0)v2,ifa<tand iT(t,a)={eMtiT0(at)ifateMaST(ta)ND(ta)0βD(τ),iD(ta,τ)dτifa<t.Moreover, the total population of Dogs, ND(t), at time t is the unique solution to (10) {dND(t)dt=θDμDND(t)0δD(a)iD(t,a)da0δC(a)cD(t,a)da,t>0ND(0)=STE0+STL0+STN0+STA0+0[iTL0(a)+iTN0(a)+iTA0(a)]da.(10)

Proof.

The existence and positiveness of solutions of system (Equation3)–(Equation4) can be addressed using an integrated semigroup approach and Volterra integral formulation. More precisely, a similar approach as in [Citation37] can be applied for detailed proof of Theorem 3.2. However, here we give a brief sketch of such an approach. For all tR, let us set u(t)=(SD(t),RD(t),0R2,iD(t,),STE(t),ST(t),0R3,iT(t,))Y,where Y is the space Y=R×R×R2×L1((0,),R2)×R3×R3×R3×L1((0,),R3).Such a Banach space Y is endow with the usual product norm Y.

Let M : D(M)YY the linear operator defined by D(M)=R×R×{0R2}×W1,1((0,),R2)×R3×R3×{0R3}×W1,1((0,),R3),and M(SDRD0R2iD()STEST0R3iT())=(μDSDμDRDiD(0)iD(μD+θ)iD(αE+μE)STEMSTiT(0)iTMiT),as well as the map F:YY such that F(SDRD0R2iD()STEST0R3iT())=(θDSDND0βT(a)1,iT(a)da0γ(a)v1,iD(a)daϵRDND0βT(a)1,iT(a)daSD+ϵRDND0βT(a)1,iT(a)dav1+0νD(a)v1,iD(a)dav20rE(1STEK)(ST,e3+0iT(a),e3da)αESTEe1STND0βD(a),iD(a)daSTND0βD(a),iD(a)da0).Observe that the nonlinear map F is not well defined on D(M)¯ due to the term ND and is then not locally Lipschitz continuous. Furthermore, for any ζ>0, let us introduce the space Yζ={uD(M)¯:κ(u)ζ}D(M)¯,where κ:YY is the operator defined by κ(u)=SD+0iD(a)da+0cD(a)da+RD.We can now define the nonlinear operator Fζ:YζY by FζF. Therefore, system (Equation7)–(Equation8) can be rewritten as the following non-densely defined abstract Cauchy problem: (11) ddtu(t)=Mu(t)+Fζ(u(t)),t>0,u(0)YζY+.(11) We then address the existence and uniqueness of bounded solutions to (Equation11) in a manner similar to the approach in [Citation37].

3.2. Disease invasion process

System (Equation7)–(Equation8) exhibits two disease-free equilibria. More precisely, in an infection-free environment, we have a disease-free equilibrium either in the presence of dogs and ticks or only in the presence of dogs. But, only the former is interesting within this context because the disease is transmitted from dogs to ticks and vice versa. Such an equilibrium is named after the disease-free equilibrium of system (Equation3)–(Equation4). Let us introduce the following threshold parameter (12) RST0=rEμAkSαkαk+μk=rEαEαE+μEαLαL+μLαNαN+μN1μA.(12) The parameter RST0 accounts for the ticks' reproduction number and is such that, αEαE+μE represents the fraction of eggs that progress to the larval stage, αLαL+μL is the fraction of larval that progress to the nymphal stage, αNαN+μN is the fraction of nymph that progress to the adult stage, 1μA is the life expectancy of adult ticks and rE is the ticks egg laying rate. Furthermore, when RST0>1, the disease-free equilibrium of system (Equation3)–(Equation4) is given by (13) E0:=(S¯D,0L1((0,+),R2),0,S¯TE,S¯T,0L1((0,+),R3)),(13) where (14) {S¯D=θDμD,S¯TE=K(11RST0),S¯T=αES¯TEk=13ekj=1kαj1μj+αj,withα0=1,α3=0.(14) In the last equality, the correspondence L1, N2, and A3 is used to facilitate the notations. See Section 6 for a detailed computation of E0.

Next, we introduce the following threshold (15) T0=T0DT×T0TD,(15) where T0DT quantifies the transmission capability from dogs to ticks and T0TD the transmission capability from ticks to dogs. More precisely, we have T0DT=0βD1(a)πD(a)daTransmission from infected dogs to ticks+0βD2(a)πC(a)da0νD(a)πD(a)daTransmission from chronically infected dogs to ticks,and (16) T0TD=kS{E}S¯kjS{E}S¯jProportion ticks of stage $k$×T0kTD,(16) where T0kTD denotes the vectorial capacity of ticks population of stage k and is explicitly given by (17) T0kTD=jS{E}S¯jS¯DTicks/Dogs ratio×0βT(a)1,eMaekda,Transmission from ticks at stage k to Dogs(17) with (18) {πD(a)=e0a(μD+δD(l)+νD(l))dl=Π(a,0)v1,πC(a)=e0a(μD+δC(l))dl=Π(a,0)v2.(18) The map aπD(a) (resp. aπC(a)) describes the probability to still be infected (resp. chronically infected) a-time post-infection. Based on the above notations, we now state the invasion dynamics in terms of the threshold T0 defined in (Equation15).

Theorem 3.3

Let Assumption 3.1 be satisfied. Assume in addition that the ticks' reproduction number satisfies RST0>1. Then the following properties hold true:

(i)

If T0<1 then the disease-free equilibrium E0 is locally asymptotically stable.

(ii)

If T0>1 then the disease-free equilibrium E0 is unstable.

The proof of Theorem 3.3 is given in Section 7.

3.3. Existence of an endemic equilibrium and bifurcation

In this section, we state our main result concerning necessary and sufficient conditions for the existence of an endemic equilibrium to system (Equation7)–(Equation8) and forward (resp. backward) bifurcation at T0=1. Denote by E an endemic equilibrium to system (Equation7)–(Equation8) that is a time-independent solution to system (Equation7)–(Equation8). (19) E=(STE,ST,iT,SD,iD,RD)(19) with (20) {ND=STE+01,iT(a)da+1,ST+SD+0v1,iD(a)da+0v2,iD(a)da+RD0=θDμDND0δD(a)v1,iD(a)da0δC(a)v2,iD(a)da(20) such that STE>0, STint(R+3), iT0L1((0,+),R3), SD>0, iD0L1((0,+),R2), and RD>0. By means of scaling and a suitable change of variables, we prove in Section 8 that the existence of an endemic equilibrium E can be reduced to the existence of a positive solution of one equation with one unknown. More precisely, we prove that there exists an endemic equilibrium E if and only if there exists K>0 satisfying Δ(T0,K)=1with Δ(T0,K):=1ID(T0,K)(1ϵ)RD(T0,K)ND(K)f(KT0)1T0TDT0.The maps KND(K), KRD(T0,K), and KID(T0,K) are given by {ND(K):=θDμD+T0TDK(0δD(a)πD(a)da+0δC(a)πC(a)da)RD(T0,K):=T0TDKND(K)ϵKT0f(KT0)+μDND(K)0γ(a)πD(a)daID(T0,K)=KT0TD0πD(a)da+KT0TD0νD(a)πD(a)da0πC(a)daand the function xf(x) is defined by f(x):=θDμDk=13(j=1kμj+αj(μj+αj)+x)S¯kjS{E}S¯jT0kTDwith T0TD (resp. T0kTD) is the vectorial (k's ticks stage vectorial) capacity given in (Equation16) (resp. in (Equation17)). Recall that in the above sum, we have used the notations L1, N2, and A3 with α3=0. Furthermore, if we consider the bifurcation parameter Cbif1:=1μDT0TD(0δD(a)πD(a)da+0δC(a)πC(a)da)k=13(j=1k1μj+αj)S¯kjS{E}S¯jT0kTDT0TD(1ϵ)1μDT0TD0γ(a)πD(a)daT0TD0πD(a)daT0TD0νD(a)πD(a)da0πC(a)dathen we have the following results whose proof is given in Section 8.

Theorem 3.4

Let Assumption 3.1 be satisfied. Assume in addition that the ticks' reproduction number satisfies RST0>1. Then the following properties hold true

(i)

If Cbif>0 then we have a backward bifurcation at T0=1, that is there exists an endemic equilibrium for T0<1 close to 1;

(ii)

If Cbif<0 then we have a forward bifurcation at T0=1, that is there exists an endemic equilibrium for T0>1 close to 1 and no endemic equilibrium for T0<1 close to 1.

4. Numerical simulations

In this section, we present numerical simulations to illustrate the forward and backward bifurcation results of Model (Equation3)–(Equation4). These simulations were conducted using finite volume numerical schemes implemented with the R software (http://www.r-project.org/). The values of the parameters used are given in Table . In all the figures, we have used the initial conditions SD(0)=10, iD(0,a)=0.03amax(0,100a), cD(0,a)=0.01amax(0,100a), RD(0)=0, STE(0)=1, STL(0)=10, STN(0)=0, STA(0)=2, iTL(0,a)=0.01, iTN(0,a)=0.3, iTA(0,a)=1. Given the parameter values in Table , the ticks' reproduction number RST066.193.

A forward bifurcation occurs at T0=1 (Theorem 3.4), which means that whenever T0<1, then the disease-free equilibrium E0 is locally asymptotically stable (Theorem 3.3) and no endemic equilibrium exists. Asymptotically, the disease go extinct (Figure (b), where T00.992). However, if T0>1, then E0 is unstable (Theorem 3.3) and an endemic equilibrium exists if Cbif<0 (Theorem 3.4). The disease is asymptotically persistent and the solution converges to the endemic equilibrium (Figure (a), where T01.103 and Cbif9.4×104).

Figure 2. Forward bifurcation diagram with convergence either to (a) the endemic equilibrium when T01.103>1; (b) the disease-free equilibrium when T00.992<1. Here rE=1, δD(a)=1.735×104, δC(a)=3.47×104, βD2(a)=0.038, βT(a)=0.15504, and Cbif=9.4×104 are fixed. The probability of infection βD1(a) is considered constant with: (a) βD1(a)=0.076 and (b) βD1(a)=0.0506. The other parameters are given by Table .

Figure 2. Forward bifurcation diagram with convergence either to (a) the endemic equilibrium when T0≈1.103>1; (b) the disease-free equilibrium when T0≈0.992<1. Here rE=1, δD(a)=1.735×10−4, δC(a)=3.47×10−4, βD2(a)=0.038, βT(a)=0.15504, and Cbif=−9.4×10−4 are fixed. The probability of infection βD1(a) is considered constant with: (a) βD1(a)=0.076 and (b) βD1(a)=0.0506. The other parameters are given by Table 1.

A backward bifurcation occurs at T0=1. This means that whenever T0>1, the disease-free equilibrium is unstable and there exists a unique endemic equilibrium. In such a situation, the solutions converge asymptotically to this endemic equilibrium (Figures (a)), where T01.318). By contrast to the forward bifurcation (Figures and (a)), when T0<1 and Cbif5.21×104>0 there exists an endemic equilibrium (Theorem 3.4). Furthermore, there exists a threshold T00.422<1 such that (i) for T0<T0, there is no endemic equilibrium and the solutions converge to the disease-free equilibrium (Figures (b) and (b), where T00.415<T0) and (ii) for T0<T0<1, there exist an endemic equilibrium such that, depending on the initial condition, the solution can converge to this endemic equilibrium (Figures (c) and (b), where T0<T00.843).

Figure 3. Backward bifurcation diagram with convergence either to (a) the endemic equilibrium when T01.318>1; (b) the disease-free equilibrium when T00.415<T0<1; (c) the endemic equilibrium when T0<T00.843<1. Here rE=1, δD(a)=0.1735, δC(a)=0.347, βD2(a)=0.608, βT(a)=0.15504, and Cbif=5.21×104 are fixed. The probability of infection βD1(a) is considered constant with: (a) βD1(a)=0.76, (b) βD1(a)=0.304, and (c) βD1(a)=0.456. The other parameters are given by Table .

Figure 3. Backward bifurcation diagram with convergence either to (a) the endemic equilibrium when T0≈1.318>1; (b) the disease-free equilibrium when T0≈0.415<T0∗<1; (c) the endemic equilibrium when T0∗<T0≈0.843<1. Here rE=1, δD(a)=0.1735, δC(a)=0.347, βD2(a)=0.608, βT(a)=0.15504, and Cbif=5.21×10−4 are fixed. The probability of infection βD1(a) is considered constant with: (a) βD1(a)=0.76, (b) βD1(a)=0.304, and (c) βD1(a)=0.456. The other parameters are given by Table 1.

Figure 4. Forward and backward bifurcation diagram. (a) Forward bifurcation diagram with Cbif=9.4×104 where all the parameters are the same as in Figure . (b) Backward bifurcation diagram with Cbif=5.21×104 and the parameters are the same as in Figure . Note that Cbif does not depend on the parameter βD1.

Figure 4. Forward and backward bifurcation diagram. (a) Forward bifurcation diagram with Cbif=−9.4×10−4 where all the parameters are the same as in Figure 2. (b) Backward bifurcation diagram with Cbif=5.21×10−4 and the parameters are the same as in Figure 3. Note that Cbif does not depend on the parameter βD1.

5. Discussion and conclusions

5.1. Discussion

In this work, we have developed a novel partial differential equation (PDE) model that extends classical epidemiological models proposed for tick-borne diseases (see [Citation29,Citation30]). The model proposed here incorporates the ability to accurately account for the different ticks' developmental stages (discrete variable) as well as the variations in infectiousness over the course of infection (continuous variable). In our study, we established the mathematical well-posedness of the model using the integrated semigroups theory. However, the presence of singularity in the force of infection whenever the total dogs' population is zero introduces a complication to the analysis. Note that this is a mathematical construct, since without dogs the model would not hold nor make sense.

We derived an explicit formula for the reproduction number T0, extending the classical formula. We identified two possible behaviours around T0=1. The first scenario involves a forward bifurcation, indicating that an epidemic can only occur if T0>1. In the second scenario, a backward bifurcation is observed, where an epidemic can arise if T0<1, provided that T0 is sufficiently close to 1. These findings have significant epidemiological implications, especially in an endemic region where controlling the epidemic is of importance. In the forward bifurcation case, simply reducing the reproduction number T0 below 1 is sufficient to halt the epidemic. However, in the backward bifurcation scenario, the number must be reduced below a second threshold, denoted as T0.

The epidemiological implication of the backward bifurcation phenomenon is that the classical requirement of having the basic reproduction number (R0) less than one is no longer sufficient to ensure effective disease eradication or elimination [Citation38]. This implies that disease can invade to a relatively high endemic level once the reproduction number R0 is more than one; decreasing R0 below one do not necessarily make the disease disappear, see Figure (b) and Figure in [Citation38]. Backward bifurcation has been shown to occur in several vector-borne disease models [Citation37,Citation39–45]. For instance Garba et al. [Citation44] showed for a dengue model the possibility of backward bifurcation where a locally stable disease-free equilibrium coexists with a locally stable endemic equilibrium. Also, the age-structure malaria and Chikungunya models in [Citation37,Citation39,Citation42,Citation43] were shown to exhibit backward bifurcation when R0 is less than one. This phenomenon is equally possible in the absence of any age-structure [Citation39,Citation42,Citation43]. Backward bifurcation in these vector-borne disease models is induced by disease induced mortality. Furthermore, backward bifurcation can equally be induced by several other factors like vaccination [Citation46–49], exogenous re-infection which are known to occur in tuberculosis models [Citation50–52], and cross-immunity for instance in a two strain influenza transmission model [Citation53]. Other epidemiological mechanism like differential susceptibility in risk-structured models could also cause backward bifurcation in disease transmission models [Citation38]. We should note that the age-structured Ehrlichia chaffeensis model (Equation3)–(Equation4) include disease induce death (δD and δC) in the dog population and hence the source of the backward bifurcation.

5.2. Conclusion

In conclusion, we have developed a novel partial differential equation model of Ehrlichia chaffeensis transmission dynamics in dogs. The model incorporates the different developmental life stages of ticks (discrete variable) as well as the duration of infection (continuous variable). The following results were obtained from our theoretical analysis and numerical simulations:

  1. The developed model is well-posed;

  2. The model always exhibits a disease-free equilibrium along with an endemic equilibrium;

  3. The model has a reproduction number, denoted as T0;

  4. A necessary and sufficient condition for the bifurcation of an endemic equilibrium was established using semigroup approach;

  5. A bifurcation (forward or backward), can occur at T0=1.

6. The disease-free equilibrium of (3)–(4)

The dynamic of ticks in the absence of infection is governed by the following system of equations (21) {dSTE(t)dt=rE(1STE(t)K)STA(t)(μE+αE)STE(t)dSTL(t)dt=αESTE(t)(μL+αL)STL(t)dSTN(t)dt=αLSTL(t)(μN+αN)STN(t)dSTA(t)dt=αNSTN(t)μASTA(t)(21) with initial condition STE(0)=STE00, STL(0)=STL00, STN(0)=STN00, and STA(0)=STA00. It is easy to see that {0}×R+3 is invariant with respect to the ordinary differential equation so that if STE0=0 then the larval, nymphal, and adult stages exponentially goes to 0. From this, we have that a necessary condition for the ticks to persist is STE0>0. Moreover, by straightforward computations, we find that (Equation21) has a unique positive stationary solution if and only if the ticks reproduction number RST0 defined by (Equation12) satisfies RST0>1. By straightforward computations, we have that the equilibrium of system (Equation21) is given by (22) {S¯TE=KRST01RST0,S¯TL=αEμL+αLS¯TE,S¯TN=αLμN+αNαEμL+αLS¯TE,S¯TA=αNμAαLμN+αNαEμL+αLS¯TE.(22) Note that in accordance with the compact formulation (Equation7)–(Equation8), the above expressions of RST0 and S¯Tk, with kS, can be rewritten as (23) RST0:=rEαEαE+μEM1e1,e3,(23) and (24) {S¯TE=K(11RST0),S¯T=αES¯TEk=13ekj=1kαj1μj+αj,with α0=1,α3=0,(24) and using the correspondence L1, N2, and A3.

7. Threshold and proof of Theorem 3.3

In order to obtain the threshold of (Equation7)–(Equation8) that determine disease invasion dynamics in a completely susceptible population of ticks and dogs, we consider the linearized equation to (Equation7)–(Equation8) at the disease-free equilibrium E0. More precisely, we linearize the infective compartments of (Equation7)–(Equation8) around E0 to obtain the following system (25) {iD(t,a)t+iD(t,a)a=(μD+θ(a))iD(t,a)iD(t,0)=0βT(a)1,iT(t,a)dav1+0νD(a)v1,iD(t,a)dav2(25) and (26) {iT(t,a)t+iT(t,a)a=MiT(t,a)iT(t,0)=S¯TS¯D0βD(a),iD(t,a)da(26) whose initial conditions satisfy iD(0,)L1((0,+),R2) and iD(0,)L1((0,+),R3). In order to define the threshold and study the local asymptotic stability of the disease-free equilibrium, we will make use of the semigroup approach by reformulating (Equation25) and (Equation26) as an abstract Cauchy problem. For this purpose, we consider the Banach spaces X:=R2×L1((0,+),R2), Y=R3×L1((0,+),R3), X0:={0R2}×L1((0,+),R2) and Y0:={0R3}×L1((0,+),R3). Let AD:D(AD)X0X and AT:D(AT)Y0Y be the linear operators defined by AD(0R2φ)=(φ(0)φ(μD+θD()φ))andAT(0R3ϕ)=(ϕ(0)ϕMϕ)with D(AD)={0R2}×W1,1((0,+),R2) and D(AT)={0R3}×W1,1((0,+),R3).Let A:D(A)Y0×X0Y×X with D(A)=D(AD)×D(AT)andA(φϕ)=(AD[φ]AT[ϕ]).Let B:X0×Y0X×Y be the bounded linear operator defined by (27) B(φϕ)=((v1PT[ϕ]+v2QD[φ]0L1((0,+),R2))S¯TS¯D(WD[φ]0L1((0,+),R3)))(27) where we have set for each φ=(0R2φ)X0 {WD[φ]:=0βD(a),φ(a)daPD[φ]:=0γ(a)v1,φ(a)daQD[φ]:=0νD(a)v1,φ(a)daand each ϕ=(0R3ϕ)Y0 {WT[ϕ]:=0e3,ϕ(a)daPT[ϕ]:=0βT(a)1,ϕ(a)da.Hence, setting for all t>0 φD(t):=(0R2iD(t,)) and ϕT(t):=(0R3iT(t,))and φD0:=(0R2iD(0,)) and ϕT0:=(0R3iT(0,))the system (Equation25)–(Equation26) can be rewritten as the following abstract Cauchy problem (28) {ddt(φD(t)ϕT(t))=A(φD(t)ϕT(t))+B(φD(t)ϕT(t)),t>0(φD(0)ϕT(0))=(φD0ϕT0)X0×Y0.(28) In order to define the threshold that determines the local asymptotic stability of the disease-free equilibrium, we need first to prove the following lemma. Before proceeding, let us set X+:=R+2×L1((0,+),R2), Y+:=R+2×L1((0,+),R2), X0+:=X+X0, and Y0+:=Y+Y0.

Lemma 7.1

Let Assumption 3.1 be satisfied. Then the following properties hold true

(i)

A is resolvent positive i.e. (λA)1 maps X+×Y+ into itself for all large λ in the resolvent set ρ(A) of A;

(ii)

The spectral bound s(A) i.e. (29) s(A)=sup{(λ):λσ(A)}(29) satisfies s(A)<0 and (s(A),+)ρ(AD)ρ(AT).

Proof.

In order to obtain the conclusion of the lemma, we first prove that A is resolvent positive. To do this, let us note that if λρ(AD)ρ(AT) then λρ(A) and (λA)1(φ~ϕ~)=((λAD)1[φ~](λAT)1[ϕ~])so that the resolvent of A is obtained by determining the resolvent of AD and AT. By standard computations, we have for each λρ(AD) and φ~=(xφ~)Y (30) (λAD)1(xφ~)=(0R2φ)φ(a)=eλaΠ(a,0)x+0aeλ(aτ)Π(a,τ)φ~(τ)dτ,a0(30) (where Π(a,0) is defined in (Equation9)) while for each λρ(AT) and ϕ~=(yϕ~X we have (31) (λAT)1(yϕ)=(0R3ϕ)ϕ(a)=eλaeMay+0aeλ(aτ)eM(aτ)ϕ~(τ)dτ,a0.(31) Next, recalling the definition of M and θ(a) respectively in (Equation5) and (Equation6) it follows that if (32) {λ>min(μL+αL,μN+αN,μA)=:μ1,<0andλ>μDmin(essinfR+θ11(a),essinfR+θ22(a))=:μ2,<0(32) then λρ(AD)ρ(AT). Next, recalling from [Citation54,Citation55] if A is resolvent positive then (33) {s(A)=inf{λρ(A):(λA)1(X+×Y+)(X+×Y+)}(s(A),+)ρ(A)(33) and we infer from (Equation30) to (Equation33) that s(A)<0. The proof is completed.

Thanks to Lemma 7.1 and the positiveness of the bounded linear operator B (i.e. it maps X0+×Y0+ into X+×Y+), one can use the theory developed in [Citation56] to define a threshold which determines the sign of s(A+B) i.e. the spectral bound of A+B given by s(A+B)=sup{(λ):λσ(A+B)}.More precisely, we set (34) T^0:=r(B(A)1)(34) where r(B(A)1) is the spectral radius of the bounded linear operator B(A)1. The following lemma will allow us to obtain the relationship between the sign of T^01 and the growth bound of the C0-semigroup generated by (A+B)0 i.e. the part of A+B in X0×Y0. Note that the latter threshold T^0 does not exhibit explicitly the parameters of the model but this will be resolved after proving the local stability properties.

Lemma 7.2

Let Assumption 3.1 be satisfied. Then s(A+B) and T^01 have the same sign. Moreover, we have the following properties

(i)

s((A+B)0)=s(A+B)=ω0((A+B)0) with ω0((A+B)0) the growth bound of the C0-semigroup generated by (A+B)0

(ii)

ω0,ess((A+B)0)ω0(A0)=s(A0)=s(A) with ω0,ess((A+B)0) the essential growth bound of the C0-semigroup generated by (A+B)0 and A0 is the part of A in X0×Y0.

Proof.

We first note that X×Y is an AL-space [Citation56, Theorem 3.14] with positive cone X+×Y+ that is normal and generating (see [Citation56]) so that (35) s(A0)=ω0(A0)ands((A+B)0)=ω0((A+B)0)(35) with ω0(A0) the growth bound of the C0-semigroup generated by A0. The first assertion of the lemma is a direct application of [Citation56, Theorem 3.5]. Next, we prove properties (i) and (ii). Since ρ(A)=ρ(A0) and ρ(A+B)=ρ((A+B)0) (see [Citation57]) it follows that s(A0)=s(A) and s((A+B)0)=s(A+B). This proves (i) and the equality ω0(A0)=s(A0)=s(A). Next, note that B is compact since B(X0×Y0) is a finite-dimensional space. Therefore, we infer from [Citation58, Theorem 1.2] that ω0,ess((A+B)0)ω0(A0) and the proof is completed.

Lemma 7.3

Let Assumption 3.1 be satisfied. Assume in addition that the ticks' reproduction number satisfies RST0>1. Then the disease-free equilibrium is locally asymptotically stable if T^0<1 and unstable if T^0>1.

Proof.

If T^0<1 then by Lemma 7.3 we have ω0((A+B)0)<0 resulting to the local asymptotic stability of the disease-free. If T^0>1 then ω0((A+B)0)>0 and since ω0,ess((A+B)0)<0 it follows that ω0,ess((A+B)0)<ω0((A+B)0). Therefore, using [Citation58, Theorem 3.2.] one knows that (A+B)0 has an isolated positive eigenvalue. Thus we refer to [Citation57, Proposition 5.7.4]) that the disease-free equilibrium is unstable.

The threshold T0^ defined in (Equation34) is somehow abstract and does not allow the interpretation of the stability of the disease-free to be made in terms of the parameters of the model. Therefore, we will give in the following and equivalent threshold which incorporate explicitly the parameters of the model. To this end, we first make a remark that allows us to simplify the determination of T0. Let us set Y1=R2×L1((0,+),R2), and X1=R3×L1((0,+),R3). Next, we observe that Y1×X1 is finite-dimensional and B(Y0×X0)Y1×X1 so that the spectral bound of B(A)1 coincides with the spectral bound of B(A)|Y1×X11. The advantage of considering B(A)|Y1×X11 is that we can obtain the spectral radius by solving an eigenvalue problem of a matrix.

Proposition 7.4

Let Assumption 3.1 be satisfied. Assume in addition that the ticks' reproduction number satisfies RST0>1 and set (36) T0=1S¯D0βT(a)1,eMaS¯Tda[0βD1(a)πD(a)da+0νD(a)πD(a)da0βD2(a)πC(a)].(36) Then T^01 and T01 have the same sign.

Proof.

We first give the explicit form of the linear operator B(A)|X1×Y11. Let (φ1ϕ1)X1×Y1 be given with φ1:=x0L1((0,+),R2)) and ϕ1:=(y0L1((0,+),R3)). Then using the resolvent formula of AD and AT respectively in (Equation30) and (Equation31) we obtain (37) (AD)1[φ1]=(0R2Π(,0)x)=:φ0(37) and (38) (AT)1[ϕ1]=(0R3eMy)=:ϕ0.(38) Hence using the explicit form of B in (Equation27) it comes (39) B(A)1(φ1ϕ1)=B(φ0ϕ0)=((v1PT[ϕ0]+v2QD[φ0]0L1((0,+),R2))S¯TS¯D(WD[φ0]0L1((0,+),R3)))(39) with {v1PT[ϕ0]+v2QD[φ0]=v10βT(a)1,eMayda+v20νD(a)v1,Π(a,0)xdaWD[φ0]=0βD(a),Π(a,0)xdaso that the spectral radius of B(A)1 is given by the spectral radius of the linear operator C:R2×R3R2×R3 given by C(xy)=(v10βT(a)1,eMayda+v20νD(a)v1,Π(a,0)xdaS¯TS¯D0βD(a),Π(a,0)xda)Since C is a positive linear operator of finite-dimensional spaces, its spectral radius T^0 is an eigenvalue. Moreover, let (xy)R2×R3 be a non zero vector such that C(xy)=T^0(xy). Then we have the following system of equations (40) {v10βT(a)1,eMayda+v20νD(a)v1,Π(a,0)xda=T^0xS¯TS¯D0βD(a),Π(a,0)xda=T^0y(40) from where we obtain the equality (41) y=1T^0S¯TS¯D0βD(a),Π(a,0)xda.(41) Hence, plugging (Equation41) into the first equation of (Equation40) gives v11T^01S¯D0βT(a)1,eMaS¯Tda0βD(a),Π(a,0)xda+v20νD(a)v1,Π(a,0)xda=T^0x.Thus, setting x=x1v1+x2v2 and recalling that βD(a),Π(a,0)v1=βD1(a)πD(a), βD(a),Π(a,0)v2=βD2(a)πC(a), and v1,Π(a,0)v1=πD(a) we obtain by identification {1S¯D0βT(a)1,eMaS¯Tda0[x1βD1(a)πD(a)+x2βD2(a)πC(a)]da=T^02x1x10νD(a)πD(a)da=T^0x2so that (42) 1S¯D0βT(a)1,eMaS¯Tda[x10βD1(a)πD(a)da+1T^0x10νD(a)πD(a)da0βD2(a)πC(a)da]=T^02x1.(42) From the above computations, we see that (xy) is the null vector if and only if x1=0. Therefore, the spectral radius of C satisfies (43) T^02=1T^0T0,2+T0,3(43) with {T0,2:=1S¯D0βT(a)1,eMaS¯Tda0βD1(a)πD(a)daT0,3:=1S¯D0βT(a)1,eMaS¯Tda0νD(a)πD(a)da0βD2(a)πC(a)da.Next, observe that T0 defined in (Equation15) is also given by T0=T0,2+T0,3. Moreover, by similar computations, one obtains that the non-zero eigenvalues of C satisfy (44) λ3=T0,2+λT0,3.(44) To prove that T^01 and T01 have the same sign, we will make use of (Equation43) and (Equation44). Indeed if T^0=1 then (Equation43) implies that T0=1. If T0=1 then (Equation44) becomes λ3=(1T0,3)+λT0,3 with 1T0,30 so that the non-zero solution of (Equation44) have modulus less than one. This proves that T^0=1 whenever T0=1. To complete the proof, we observe that if T0>1 then (Equation43) implies that 1<T02=1T0T0,2+T0,3<T0,2+T0,3=R0. Similarly if T0<1 then R0<1.

8. Proof of Theorem 3.4

In order to determine the endemic equilibrium of the model (Equation7)–(Equation8), we proceed in several steps. In the first step, we give an equivalent system to (Equation45), the second step is concerned with the solvability of (Equation46), and the last step concerns the bifurcation properties. In this section, we will always suppose that Assumption 3.1 is satisfied and RST0>1. Let us note that using the notation (Equation18) together with Theorem 3.2, one knows that an endemic equilibrium E defined in (Equation19) satisfies (45) {0=rE(1STEK)(ST,e3+0iT(a),e3da)(αE+μE)STEiT(a)=1ND0βD(a),iD(a)daeMaST0R3=αESTEe11ND0βD(a),iD(a)daSTMST,(45) and (46) {0=θDSDND0βT(a)1,iT(a)daμDSDiD(a)=SD+ϵRDND0βT(a)1,iT(a)daπD(a)v1+0νD(τ)v1,iD(a)daπC(a)v20=0γ(a)v1,iD(a)daϵRDND0βT(a)1,iT(a)daμDRD0=θDμDND0δD(a)v1,iD(a)da0δC(a)v2,iD(a)da.(46) On the stationary states of ticks: Let us first prove the following lemma which describes the stationary states of the eggs.

Lemma 8.1

The total number of eggs remains unchanged at the endemic and the disease-free equilibrium i.e. S¯TE=S¯TE. Moreover, we always have (47) ST+0iT(a)da=S¯T.(47)

Proof.

Note that, using the third equation of (Equation45) we have (48) 0R3=αESTEM1e11ND0βD(a),iD(a)daM1STST(48) Next, integrating the second equation of (Equation45) it follows that (49) 0iT(a)da=1ND0βD(a),iD(a)da0eMaSTda=1ND0βD(a),iD(a)daM1ST(49) Thus, by combining (Equation48) and (Equation49) we obtain (50) 0iT(a)da+ST=αESTEM1e1=STES¯TES¯T.(50) Therefore, plugging (Equation50) in the first equation of (Equation53) we obtain the following equality (51) 0=rE(1STEK)STES¯TES¯T,e3(αE+μE)STE(51) or equivalently (52) 0=rE(1STEK)S¯T,e3(αE+μE)S¯TE.(52) The right-hand side of (Equation52) is a decreasing function of STE so that the unique solution to (Equation51) is STE=S¯TE. The equality (Equation47) now follows from (Equation50).

Thanks to Lemma 8.1 the system of equations (Equation45) is equivalent to (53) {iT(a)=1ND0βD(a),iD(a)daeMaST0R3=αESTEe11ND0βD(a),iD(a)daSTMST.(53) Note that for each nonnegative constant η00 the matrix M+η0I3 (where I3 is the identity operator on R3) is invertible with inverse (M+η0I3)1 satisfying (54) (M+η0I3)1e1=k=13(j=1kαj1μj+αj+η0)ek,with α0=1,α3=0,(54) and where the correspondence L1, N2, and A3 is used. Therefore, using (Equation53) together with (Equation54), for η0=1ND0βD(a),iD(a)da, it follows that (55) ST=αESTEk=13ekj=1kNDαj1ND(μj+αj)+0βD(a),iD(a)da,(55) with α0=1 and α3=0. Thus, recalling the expression of S¯T in (Equation24) we obtain (56) ST=k=13S¯T,ekekj=1kND(μj+αj)ND(μj+αj)+0βD(a),iD(a)da(56) and the system of equations (Equation53) becomes equivalent to (57a) ST=k=13(j=1kμj+αj(μj+αj)+WD)S¯T,ekek(57a) (57b) iT(a)=WDeMaST(57b) (57c) WD=0βD(a),iD(a)NDda.(57c) From the above comments, it follows that ST and iT are entirely determined by the variables WD and iDND.

On the stationary states of dogs: In order to obtain an equivalent formulation to (Equation46) we introduce the new variables (58) iD(a):=iD(a)ND,SD:=SDNDandRD:=RDND.(58) Next, we set (59) PT:=0βT(a)1,iT(a)da,(59) and (60) {PD:=0γ(a)v1,iD(a)daQD:=0νD(a)v1,iD(a)da,UD:=0δD(a)v1,iD(a)daUC:=0δC(a)v2,iD(a)da(60) With these new variables, system (Equation46) becomes (61) {0=θDSDPTμDNDSDNDiD(a)=(SD+ϵRD)PTπD(a)v1+NDQDπC(a)v20=NDPDϵRDPTμDNDRD0=θDμDNDNDUDNDUC(61) which is equivalent to (62) {ND=θDμD+UD+UCSD=θDμDND+PTRD=NDPDϵPT+μDNDiD(a)=SD+ϵRDNDPTπD(a)v1+QDπC(a)v2.(62) Next, we show the relationship between PT and WD defined respectively in (Equation57c) and (Equation59). To do so, we integrate (Equation57b) to obtain (63) PT=WD0βT(a)1,eMaSTda(63) and by using the expression of ST defined in (Equation57a) we obtain the following more explicit formula (64) PT=WDk=13(j=1kμj+αj(μj+αj)+WD)S¯kjS{E}S¯jT0kTDS¯D,(64) with T0kTD the k's stage vectorial capacity defined in (Equation17).

In the following, we determine the equations for the variables PD, QD, UD, UC and WC defined in (Equation60). To this end, recalling that βD(a)=βD1(a)v1+βD2(a)v1, we use successively the equation of iD(a) in (Equation62) to obtain (65a) WD=SD+ϵRDNDPT0βD1(a)πD(a)da(65a) (65b) +QD0βD2(a)πC(a)da(65b) (65c) PD=SD+ϵRDNDPT0γ(a)πD(a)da(65c) (65d) QD=SD+ϵRDNDPT0νD(a)πD(a)da(65d) (65e) UD=SD+ϵRDNDPT0δD(a)πD(a)da(65e) (65f) UC=SD+ϵRDNDPT0δC(a)πC(a)da.(65f)

Note that, plugging (Equation65d) into (Equation65a) gives the following equality (66) {WD=SD+ϵRDNDPT[0βD1(a)πD(a)da+0νD(a)πD(a)da0βD2(a)πC(a)da].(66) Let us now define the function (67) f(WD):=k=13(j=1kμj+αj(μj+αj)+WD)S¯kjS{E}S¯jT0kTDS¯D(67) and observe that the vectorial capacity T0TD is given by (68) T0TD=f(0)S¯D(68) so that (69) {T0=f(0)S¯D[0βD1(a)πD(a)da+0νD(a)πD(a)da0βD2(a)πC(a)da].(69) In the following, we rewrite our system of equations by using the variable (70) K:=SD+ϵRDNDPTT0TD.(70) To do so, we first observe that (Equation66) and (Equation64) take respectively the following form (71) WD=KT0andPT=WDf(WD)=KT0f(KT0)(71) while (Equation65a) is now given by (72a) WD=SD+ϵRDNDPT0βD1(a)πD(a)da(72a) (72b) +QD0βD2(a)πC(a)da(72b) (72c) PD=KT0TD0γ(a)πD(a)da(72c) (72d) QD=KT0TD0νD(a)πD(a)da(72d) (72e) UD=KT0TD0δD(a)πD(a)da(72e) (72f) UC=KT0TD0δC(a)πC(a)da.(72f) Hence, using the first equation of (Equation62) together with (Equation72e), and (Equation72f) we obtain the following necessary conditions (73) ND(K):=θDμD+KT0TD(0δD(a)πD(a)da+0δC(a)πC(a)da).(73) Furthermore, the second, third and fourth equation of (Equation62) together with (Equation71), (Equation72c) and (Equation72d) provide (74a) RD(T0,K):=T0TDKND(K)ϵKT0f(KT0)+μDND(K)0γ(a)πD(a)da(74a) (74b) SD(T0,K):=θDμDND(K)+KT0f(KT0)(74b) (74c) iD(a):=KT0TDπD(a)v1+KT0TD0νD(a)πD(a)daπC(a)v2.(74c) Moreover, from the equality PT=KT0f(KT0) and (Equation70) it comes that for K>0 we must have the following equality (75) 1=SD(T0,K)+ϵRD(T0,K)ND(K)f(KT0)T0TDT0.(75) Recalling that we have the necessary condition SD+RD+01,iD(a)da=1, we consider the following equation Δ(T0,K)=1with (76) Δ(T0,K):=1(1ϵ)RD(T0,K)ID(T0,K)ND(K)f(KT0)T0TDT0(76) and (77) ID(T0,K):=KT0TD0πD(a)da+KT0TD0νD(a)πD(a)da0πC(a)da.(77)

Lemma 8.2

There exists an endemic equilibrium to (Equation7)–(Equation8) if and only if there exists K>0 such that Δ(T0,K)=1.

Proof.

The necessity of the lemma follows from the preceding arguments. Next, we prove the sufficiency. To this end, assume that there exists K>0 such that Δ(T0,K)=1. Let ND be given by the right-hand side of (Equation73). Let RD and iD be given respectively by the right-hand side of (Equation74a) and (Equation74c). Observe that with these definitions we have (78) ID(T0,K)=01,iDda(78) so that setting (79) SD:=1RDID(T0,K)(79) it comes from (Equation76) (80) 1=SD+ϵRDNDf(KT0)T0TDT0.(80) Note that setting WD=KT0, PT=WDf(WD), with f defined in (Equation67), and multiplying (Equation80) by K it follows that (81) K=SD+ϵRDNDPTT0TDandWD=SD+ϵRDNDPTT0TDT0.(81) Therefore, our definition of RD and iD respectively in (Equation74a) and (Equation74c) together with (Equation81) give us (82) {RD=SD+ϵRDNDPTNDϵPT+μDND0γ(a)πD(a)daiD(a)=SD+ϵRDNDPTπD(a)v1+SD+ϵRDNDPT0νD(a)πD(a)daπC(a)v2.(82) Moreover, using (Equation81) one also note that our definition of ND in (Equation73) leads to (83) μDND=θD(SD+ϵRD)PT0δD(a)πD(a)da(SD+ϵRD)PT0δC(a)πC(a)da.(83) Next, we observe that the RD-equation in (Equation82) is equivalent to (84) μDNDRD=ϵPTRD+(SD+ϵRD)PT0γ(a)πD(a)da.(84) Note that from the above formulas of RD and iD in (Equation82) we have the following identities (85) {0νD(a)v1,iD(a)da=SD+ϵRDNDPT0νD(a)πD(a)da0γ(a)v1,iD(a)da=SD+ϵRDNDPT0γ(a)πD(a)da0δD(a)v1,iD(a)da=SD+ϵRDNDPT0δD(a)πD(a)da0δC(a)v2,iD(a)da=SD+ϵRDNDPT0νD(a)πD(a)da×0δC(a)πC(a)da(85) from where we can rewrite (Equation83), (Equation84) and the iD-equation in (Equation82) as follow (86) {μDND=θDND0δD(a)v1,iD(a)daND0δC(a)v2,iD(a)daμDNDRD=ϵPTRD+ND0γ(a)v1,iD(a)daiD(a)=SD+ϵRDNDPTπD(a)v1+0νD(a)v1,iD(a)daπC(a)v2.(86) Next, observe that (87) {iD(a)=(μD+θ(a))iD(a)iD(0)=SD+ϵRDNDPTv1+0νD(a)v1,iD(a)dav2(87) so that by integrating (Equation87) from 0 to + we obtain SD+ϵRDNDPTv10νD(a)v1,iD(a)dav2=0(μD+θ(a))iD(a)daand by using (Equation85) it follows that (88) μD01,iD(a)da=SD+ϵRDNDPT0(γ(a)+δD(a))v1,iD(a)da0δC(a)v2,iD(a)da.(88) Next, multiply (Equation89) by ND and use the first equation of (Equation86) to obtain (89) μDND01,iD(a)da=(SD+ϵRD)PTθD+μDNDND0γ(a)v1,iD(a)da(89) Hence, summing (Equation89) and the second equation of (Equation86) it follows that (90) SDPTθD+μDND=μDND01,iD(a)da+μDNDRD=μDND(1SD)(90) that is (91) 0=θDSDPTμDNDSDSD=θDμDND+PT.(91) Finally, we infer from the iD-equation of (Equation86), the equality (Equation91) together with our definition of RD, and ND that (Equation62) is satisfied. The proof is completed by using the fact that we have set PT=WDf(WD) with WD=KT0 and f given by (Equation67).

Thanks to Lemma 8.2, the existence of positive equilibrium to (Equation7)–(Equation8) is subjected to the existence of K>0 such that (92) Δ(T0,K)=1(92) where Δ(T0,K) is given by (Equation76) and (Equation77). Before studying the existence of solution to (Equation92), we first observe that from (Equation73), (Equation74a) and (Equation77) we have (93) {ND(0)=θDμD=S¯DRD(T0,0)=ID(T0,0)=0,T00.(93) We also note that for each T0>0 we have (94) limK+RD(T0,K)=l>0andlimK+ID(T0,K)=+(94) with l:=1ϵT0TDθD0γ(a)πD(a)da0δD(a)πD(a)da+0δC(a)πC(a)da.Therefore, using the equality T0TD=f(0)S¯D it follows from (Equation93) and (Equation76) that (95) Δ(T0,0)=T0andlimK+Δ(T0,K)=.(95) It is now clear from (Equation95) that for each T0>0 there exists K¯:=K¯(T0)>0 such that Δ(T0,K¯)=0. In particular, if T0>1 then there exists K0(0,K¯) such that Δ(T0,K0)=1 leading to the existence of an endemic equilibrium.

On the forward and backward bifurcations: In the following, we deal with the existence of forward and backward bifurcation at T0=1. Roughly speaking, we have to prove that there exists a positive map K0 defined in some right neighbourhood (forward bifurcation) or left neighbourhood (backward bifurcation) of T0=1 such that Δ(T0,K0(T0))=1 and K0(1)=0. Since Δ(1,0)=1, one can prove the existence of forward and backward bifurcations using implicit function theorem which is reduced here to the study of KΔ(1,0). This motivates the following lemma.

Lemma 8.3

The following property is satisfied (96) T0Δ(T0,0)=1,T0>0(96) and KΔ(1,0)=1μDT0TD(0δD(a)πD(a)da+0δC(a)πC(a)da)k=13(j=1k1μj+αj)S¯kjS{E}S¯jT0kTDT0TD(1ϵ)1μDT0TD0γ(a)πD(a)daT0TD0πD(a)daT0TD0νD(a)πD(a)da0πC(a)da.

Proof.

Since Δ(T0,0)=T0 for each T0>0 the first equality of the lemma follows. To compute KΔ(1,0), it is enough to take the derivative of Δ(1,K) with respect to K at K = 0. To do so let us first note that by using (Equation76) we have ND(K)Δ(1,K)=f(K)T0TD[(1ϵ)RD(1,K)+ID(1,K)]f(K)T0TD.Recalling that Δ(1,0)=1, ND(0)=θDμD=S¯D, RD(1,0)=ID(1,0)=0 and taking the derivative of ND(K)Δ(1,K) with respect to K at K = 0 it follows that S¯DKΔ(1,0)=dND(0)dK+f(0)T0TDS¯D[(1ϵ)KRD(1,0)+KID(1,0)].By straightforward computations, we obtain from (Equation73) and (Equation77) that dND(0)dK=S¯DμDT0TD(0δD(a)πD(a)da+0δC(a)πC(a)da)and KID(1,0)=T0TD0πD(a)da+T0TD0νD(a)πC(a)da0πC(a)da.To complete the proof it remains to compute KRD(1,0) and f(0). We first compute KRD(1,0). To this end, we compute the derivative of RD(1,K) from (Equation74a) with respect to K and set K = 0 to obtain KRD(1,0)=1μDT0TD0γ(a)πD(a)da.To compute f(0) recall that f(WD):=k=13(j=1kμj+αj(μj+αj)+WD)S¯kjS{E}S¯jT0kTDS¯Dso that setting f^k(WD):=j=1kgj(WD),k=1,2,3,gj(WD):=μj+αj(μj+αj)+WDwe obtain f(WD):=k=13f^k(WD)S¯kjS{E}S¯jT0kTDS¯D.Therefore, computing the derivative of ln(f^k(WD)) one obtains f^k(WD)=f^k(WD)j=1kgj(WD)gj(WD)and since f^k(0)=gk(0)=1 it follows that f^k(0)=j=1kgj(0)=j=1k1μj+αjthat is (97) f(0)=k=13(j=1k1μj+αj)S¯kjS{E}S¯jT0kTDS¯D.(97) The proof is completed.

Thanks to Lemma 8.3 one knows that if KΔ(1,0)=Cbif10 then by the implicit function theorem there exists ξ(0,1) and a smooth map K0:(1ξ,1+ξ)R such that (98) Δ(T0,K0(T0))=1,T0(1ξ,1+ξ)andK0(1)=0.(98) Thus, up to reduce ξ, it follows that the sign of K0 on (1ξ,1+ξ) is given by the sign of dK0(T0)dT0 at T0=1. Differentiating (Equation98) with respect to T0 and using Lemma 8.3 we obtain dK0(1)dT0=T0Δ(1,0)KΔ(1,0)=Cbif

so that

  • If Cbif>0 then, up to reduce ξ, we have K0(T0)>0 for T0(1ξ,1) which corresponds to a backward bifurcation ;

  • If Cbif<0 then, up to reduce ξ, K0(T0)>0 for T0(1,1+ξ) which corresponds to a forward bifurcation.

Disclosure statement

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

Additional information

Funding

FBA is supported by the National Science Foundation under the EPSCOR Track 2 grant number 1920946.

References

  • AV Barrantes-González, AE Jiménez-Rocha, JJ Romero-Zuñiga, et al. Understanding ehrlichia canis infections in dogs of costa rica: hematological findings and indicative clinical signs. Open J Vet Med. 2016;06(11):163–175. doi: 10.4236/ojvm.2016.611020
  • MJ Beall, AR Alleman, EB Breitschwerdt, et al. Seroprevalence of ehrlichia canis, ehrlichia chaffeensis and ehrlichia ewingii in dogs in North America. Parasit Vectors. 2012;5(1):1–11. doi: 10.1186/1756-3305-5-29
  • JS Dumler, AF Barbet, C Bekker, et al. Reorganization of genera in the families rickettsiaceae and anaplasmataceae in the order rickettsiales: unification of some species of ehrlichia with anaplasma, cowdria with ehrlichia and ehrlichia with neorickettsia, descriptions of six new species combinations and designation of ehrlichia equi and'hge agent'as subjective synonyms of ehrlichia phagocytophila. Int J Syst Evol Microbiol. 2001;51(6):2145–2165. doi: 10.1099/00207713-51-6-2145
  • AD Nair, C Cheng, DC Jaworski, et al. Ehrlichia chaffeensis infection in the reservoir host (white-tailed deer) and in an incidental host (dog) is impacted by its prior growth in macrophage and tick cell environments. PLoS ONE. 2014;9(10):e109056. doi: 10.1371/journal.pone.0109056
  • A Procajło, E Skupień, M Bladowski, et al. Monocytic ehrlichiosis in dogs. Pol J Vet Sci 14(3) (2011), pp. 515–520. doi:10.2478/v10181-011-0077-9
  • CM Ehrlichiosis, CH Fever, TC Pancytopenia, et al. Ehrlichiosis and anaplasmosis: Zoonotic species. 2013. The Center for Food Security & Public Health, Iowa State University. Accessed on August 7 2022. https://www.cfsph.iastate.edu/Factsheets/pdfs/ehrlichiosis.pdf
  • DM Aguiar, TF Ziliani, X Zhang, et al. A novel ehrlichia genotype strain distinguished by the trp36 gene naturally infects cattle in brazil and causes clinical manifestations associated with ehrlichiosis. Ticks Tick Borne Dis. 2014;5(5):537–544. doi: 10.1016/j.ttbdis.2014.03.010
  • BA Allsopp. Natural history of ehrlichia ruminantium. Vet Parasitol. 2010;167(2-4):123–135. doi: 10.1016/j.vetpar.2009.09.014
  • P Brouqui, K Matsumoto. Bacteriology and phylogeny of anaplasmataceae. Rickettsial Dis. 2007;191–210.
  • LA Cohn. Ehrlichiosis and related infections. Vet Clin Small Anim Pract. 2003;33(4):863–884. doi: 10.1016/S0195-5616(03)00031-7
  • L Fargnoli, C Fernandez, LD Monje. Novel ehrlichia strain infecting cattle tick amblyomma neumanni, Argentina, 2018. Emerging Infect Dis. 2020;26(5):1027–1030. doi: 10.3201/eid2605.190940
  • M Groves, G Dennis, H Amyx, et al. Transmission of ehrlichia canis to dogs by ticks (rhipicephalus sanguineus). Am J Vet Res. 1975;36(7):937–940.
  • MN Saleh, KE Allen, MW Lineberry, et al. Ticks infesting dogs and cats in north america: biology, geographic distribution, and pathogen transmission. Vet Parasitol. 2021;294:109392. doi: 10.1016/j.vetpar.2021.109392
  • K Mahachi, E Kontowicz, B Anderson, et al. Predominant risk factors for tick-borne co-infections in hunting dogs from the USA. Parasit Vectors. 2020;13(1):1–12. doi: 10.1186/s13071-020-04118-x
  • RFDC Vieira, TSWJ Vieira, DDAG Nascimento, et al. Serological survey of ehrlichia species in dogs, horses and humans: zoonotic scenery in a rural settlement from southern brazil. Rev Inst De Med Trop São Paulo. 2013;55(5):335–340. doi: 10.1590/S0036-46652013000500007
  • CB Yancey, BC Hegarty, BA Qurollo, et al. Regional seroreactivity and vector-borne disease co-exposures in dogs in the united states from 2004–2010: utility of canine surveillance. Vector-Borne Zoonotic Dis. 2014;14(10):724–732. doi: 10.1089/vbz.2014.1592
  • CD Paddock, JE Childs. Ehrlichia chaffeensis: a prototypical emerging pathogen. Clin Microbiol Rev. 2003;16(1):37–64. doi: 10.1128/CMR.16.1.37-64.2003
  • BE Anderson, KG Sims, JG Olson, et al. Amblyomma americanum: a potential vector of human ehrlichiosis. Am J Trop Med Hyg. 1993;49(2):239–244. doi: 10.4269/ajtmh.1993.49.239
  • O Anziani, S Ewing, R Barker. Experimental transmission of a granulocytic form of the tribe ehrlichieae by dermacentor variabilis and amblyomma americanum to dogs. Am J Vet Res. 1990;51(6):929–931.
  • Centers for Disease Control and Prevention (CDC). How ticks spread disease. [accessed 2023 Aug 7]. Available from: https://www.cdc.gov/ticks/life_cycle_and_hosts.html.
  • E Guo, FB Agusto. Baptism of fire: modeling the effects of prescribed fire on lyme disease. Can J Infect Dis Med Microbiol. 2022;2012.
  • J van der Krogt. Ehrlichia canis infections on the island of curaçao an overview of the clinical picture and current diagnostics & therapies. 2010. Accessed August 7 2023. https://studenttheses.uu.nl/bitstream/handle/20.500.12932/4426/Ehrlichia%20canis%20infections%20on%20the%20island%20of%20Curacao%20by%20J.S.%20van%20der%20Krogt.pdf?sequence=1
  • S Ewing, J Dawson, A Kocan, et al. Experimental transmission of ehrlichia chaffeensis (rickettsiales: ehrlichieae) among white-tailed deer by amblyomma americanum (acari: ixodidae). J Med Entomol. 1995;32(3):368–374. doi: 10.1093/jmedent/32.3.368
  • SE Little, TP O'Connor, J Hempstead, et al. Ehrlichia ewingii infection and exposure rates in dogs from the southcentral united states. Vet Parasitol. 2010;172(3-4):355–360. doi: 10.1016/j.vetpar.2010.05.006
  • JR Gettings, SC Self, CS McMahan, et al. Local and regional temporal trends (2013–2019) of canine ehrlichia spp. seroprevalence in the usa. Parasit Vectors. 2020;13(1):1–11. doi: 10.1186/s13071-020-04022-4
  • MU Aziz, S Hussain, B Song, et al. Ehrlichiosis in dogs: A comprehensive review about the pathogen and its vectors with emphasis on South and East Asian countries. Vet Sci. 2023;10(1):21. doi: 10.3390/vetsci10010021
  • BA Qurollo, J Buch, R Chandrashekar, et al. Clinicopathological findings in 41 dogs (2008-2018) naturally infected with ehrlichia ewingii. J Vet Intern Med. 2019;33(2):618–629. doi: 10.1111/jvim.2019.33.issue-2
  • BA Qurollo, R Chandrashekar, BC Hegarty, et al. A serological survey of tick-borne pathogens in dogs in north america and the Caribbean as assessed by anaplasma phagocytophilum, a. platys, ehrlichia canis, e. chaffeensis, e. ewingii, and borrelia burgdorferi species-specific peptides. Infect Ecol Epidemiol. 2014;4(1):24699.
  • H Gaff, L Gross, E Schaefer . Results from a mathematical model for human monocytic ehrlichiosis. Clin Microbiol Infect. 2009;15:15–16. doi: 10.1111/j.1469-0691.2008.02131.x
  • HD Gaff, LJ Gross. Modeling tick-borne disease: a metapopulation model. Bull Math Biol. 2007;69(1):265 288 doi: 10.1007/s11538-006-9125-5
  • North Shore Animal League. Prevent a litter – spay and neuter your pets. [accessed 2022 Aug 7]. Available from: https://www.animalleague.org/wp-content/uploads/2017/06/dogs-multiply-pyramid.pdf.
  • A Burke. How long do dogs live?. [accessed 2022 Aug 7]. Available from: https://www.akc.org/expert-advice/health/how-long-do-dogs-live/.
  • Mar Vista Animal Medical Center. Ehrlichia infection (canine). 2020 [accessed 2022 Aug 7]. Available from: https://www.marvistavet.com/ehrlichia-infection-canine.pml.
  • K Williams. BSc, DVM, CCRP; Ryan Llera, BSc, DVM; Ernest Ward, DVM. Ehrlichiosis in dogs. [accessed 2022 Aug 7]. Available from: https://vcahospitals.com/know-your-pet/ehrlichiosis-in-dogs.
  • C Wheeler. The life cycle of a tick. [accessed 2022 Aug 7]. Available from: https://study.com/learn/lesson/tick-life-cycle-reproduction-eggs.html.
  • GH Antoinette Ludwig, HS Ginsberg, NH Ogden. A dynamic population model to investigate effects of climate and climate-independent factors on the lifecycle of amblyomma americanum. J Med Entomol 53(1) (2015), pp. 99–115.
  • Q Richard, M Choisy, T Lefèvre, et al. Human-vector malaria transmission model structured by age, time since infection and waning immunity. Nonlinear Anal: Real World Appl. 2022;63:103393. doi: 10.1016/j.nonrwa.2021.103393
  • AB Gumel. Causes of backward bifurcations in some epidemiological models. J Math Anal Appl. 2012;395(1):355–365. doi: 10.1016/j.jmaa.2012.04.077
  • F.B. Agusto, S. Easley, K. Freeman, et al. Mathematical model of three age-structured transmission dynamics of chikungunya virus. Comput Math Methods Med. 2016;2016:1–31. doi: 10.1155/2016/4320514
  • KW Blayneh, AB Gumel, S Lenhart, et al. Backward bifurcation and optimal control in transmission dynamics of West Nile virus. Bull Math Biol. 2010;72(4):1006–1028. doi: 10.1007/s11538-009-9480-0
  • J Dushoff, W Huang, C Castillo-Chavez. Backwards bifurcations and catastrophe in simple models of fatal diseases. J Math Biol. 1998;36(3):227–248. doi: 10.1007/s002850050099
  • F Forouzannia, A Gumel. Dynamics of an age-structured two-strain model for malaria transmission. Appl Math Comput. 2015;250:860–886.
  • F Forouzannia, A Gumel. Mathematical analysis of an age-structured model for malaria transmission dynamics. Math Biosci. 2014;247:80–94. doi: 10.1016/j.mbs.2013.10.011
  • SM Garba, AB Gumel, MA Bakar. Backward bifurcations in dengue transmission dynamics. Math Biosci. 2008;215(1):11–25. doi: 10.1016/j.mbs.2008.05.002
  • Z Mukandavire, AB Gumel, W Garira, et al. Mathematical analysis of a model for HIV-malaria co-infection. Mathematical Biosciences and Engineering 6(2) (2009), pp. 333–362. doi:10.3934/mbe.2009.6.333
  • FB Agusto, AB Gumel. Theoretical assessment of avian influenza vaccine. DCDS Ser B. 2010;13(1):1–25. doi: 10.3934/dcdsb.2010.13.1
  • F Brauer. Backward bifurcations in simple vaccination models. J Math Anal Appl. 2004;298(2):418–431. doi: 10.1016/j.jmaa.2004.05.045
  • EH Elbasha, AB Gumel. Theoretical assessment of public health impact of imperfect prophylactic hiv-1 vaccines with therapeutic benefits. Bull Math Biol. 2006;68(3):577–614. doi: 10.1007/s11538-005-9057-5
  • O Sharomi, C Podder, A Gumel, et al. Role of incidence function in vaccine-induced backward bifurcation in some hiv models. Math Biosci. 2007;210(2):436–463. doi: 10.1016/j.mbs.2007.05.012
  • C Castillo-Chavez, B Song. Dynamical models of tuberculosis and their applications. Math Biosci Eng. 2004;1(2):361–404. doi: 10.3934/mbe.2004.1.361
  • Z Feng, C Castillo-Chavez, AF Capurro. A model for tuberculosis with exogenous reinfection. Theor Popul Biol. 2000;57(3):235–247. doi: 10.1006/tpbi.2000.1451
  • O Sharomi, C Podder, A Gumel, et al. Mathematical analysis of the transmission dynamics of hiv/tb coinfection in the presence of treatment. Math Biosci Eng. 2008;5(1):145.174 doi: 10.3934/mbe.2008.5.145
  • SM Garba, MA Safi, AB Gumel. Cross-immunity-induced backward bifurcation for a model of transmission dynamics of two strains of influenza. Nonlinear Anal Real World Appl. 2013;14(3):1384–1403. doi: 10.1016/j.nonrwa.2012.10.003
  • W Arendt. Resolvent positive operators. Proc Lond Math Soc. 1987;s3-54(2):321–349. doi: 10.1112/plms/s3-54.2.321
  • RD Nussbaum. Positive operators and elliptic eigenvalue problems. Math Z. 1984;186(2):247–264.
  • HR Thieme. Spectral bound and reproduction number for infinite-dimensional population structure and time heterogeneity. SIAM J Appl Math. 2009;70(1):188–211. doi: 10.1137/080732870
  • P Magal, S Ruan. Theory and applications of abstract semilinear Cauchy problems. Volume 201 of Applied Mathematical Sciences. Cham: Springer International Publishing; 2018.
  • A Ducrot, Z Liu, P Magal. Essential growth rate for bounded linear perturbation of non-densely defined Cauchy problems. J Math Anal Appl. 2008;341(1):501–518. doi: 10.1016/j.jmaa.2007.09.074