
The basic reproduction ratio R0 of more general periodic and time-delayed impulsive model which the period of model coefficients is different from that of fixed impulsive moments, is developed. That R0 is the threshold parameter for the stability of the zero solution of the associated linear system is also shown. The developed theory is further applied to a swine parasitic disease model with pulse therapy. Threshold results on its global dynamics in terms of R0 are obtained. Some numerical simulation results are also given to support our main results.
Citation: Hai-Feng Huo, Fan Wu, Hong Xiang. On threshold dynamics for periodic and time-delayed impulsive systems and application to a periodic disease model[J]. Mathematical Biosciences and Engineering, 2022, 19(1): 836-854. doi: 10.3934/mbe.2022038
[1] | Amar Nath Chatterjee, Fahad Al Basir, Yasuhiro Takeuchi . Effect of DAA therapy in hepatitis C treatment — an impulsive control approach. Mathematical Biosciences and Engineering, 2021, 18(2): 1450-1464. doi: 10.3934/mbe.2021075 |
[2] | Lin Zhao, Zhi-Cheng Wang, Liang Zhang . Threshold dynamics of a time periodic and two–group epidemic model with distributed delay. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1535-1563. doi: 10.3934/mbe.2017080 |
[3] | Jie He, Zhenguo Bai . Global Hopf bifurcation of a cholera model with media coverage. Mathematical Biosciences and Engineering, 2023, 20(10): 18468-18490. doi: 10.3934/mbe.2023820 |
[4] | Hongyong Zhao, Yangyang Shi, Xuebing Zhang . Dynamic analysis of a malaria reaction-diffusion model with periodic delays and vector bias. Mathematical Biosciences and Engineering, 2022, 19(3): 2538-2574. doi: 10.3934/mbe.2022117 |
[5] | Rong Ming, Xiao Yu . Global dynamics of an impulsive vector-borne disease model with time delays. Mathematical Biosciences and Engineering, 2023, 20(12): 20939-20958. doi: 10.3934/mbe.2023926 |
[6] | Zhenguo Bai . Threshold dynamics of a periodic SIR model with delay in an infected compartment. Mathematical Biosciences and Engineering, 2015, 12(3): 555-564. doi: 10.3934/mbe.2015.12.555 |
[7] | Tingting Zhao, Robert J. Smith? . Global dynamical analysis of plant-disease models with nonlinear impulsive cultural control strategy. Mathematical Biosciences and Engineering, 2019, 16(6): 7022-7056. doi: 10.3934/mbe.2019353 |
[8] | Wenhao Chen, Guo Lin, Shuxia Pan . Propagation dynamics in an SIRS model with general incidence functions. Mathematical Biosciences and Engineering, 2023, 20(4): 6751-6775. doi: 10.3934/mbe.2023291 |
[9] | Biwen Li, Qiaoping Huang . Synchronization of time-delay systems with impulsive delay via an average impulsive estimation approach. Mathematical Biosciences and Engineering, 2024, 21(3): 4501-4520. doi: 10.3934/mbe.2024199 |
[10] | Yan Zhang, Sanyang Liu, Zhenguo Bai . Global dynamics of a di usive single species model with periodic delay. Mathematical Biosciences and Engineering, 2019, 16(4): 2293-2304. doi: 10.3934/mbe.2019114 |
The basic reproduction ratio R0 of more general periodic and time-delayed impulsive model which the period of model coefficients is different from that of fixed impulsive moments, is developed. That R0 is the threshold parameter for the stability of the zero solution of the associated linear system is also shown. The developed theory is further applied to a swine parasitic disease model with pulse therapy. Threshold results on its global dynamics in terms of R0 are obtained. Some numerical simulation results are also given to support our main results.
The basic reproduction number (ratio) R0 is an important parameter in epidemiology and within-host pathogen dynamics. R0 is defined as the expected number of secondary cases produced by a single (typical) infection in a completely susceptible population in epidemiology. R0 determines whether an infectious disease will break out usually. If R0>1, it means that an individual can produce more than one individuals on average, then the population persists, if R0<1, an individual can produce less than one individual on average, thus the population becomes extinct. Diekmann et al. [1] first introduced the next generation matrix approach to R0 and Van den Driessche et al. [2] established the theory of R0 for autonomous compartmental epidemic models. These two works have been widely used in various infectious disease models. For epidemic models with periodic coefficients, Bacaër et al. [3] presented a general definition of R0, that is, R0 is the spectral radius of an integral operator on the space of continuous periodic functions. Since then, there has been a lot of research on R0 in epidemic models. Wang et al. [4] proved that R0 is a threshold parameter for the local stability of the disease-free periodic solution of periodic compartmental ODE models. Liang et al. [5] and Zhao [6], proved similar results for a class of time-delay evolution equations, respectively. Bai et al. [7] generalized the above results to impulse differential equations. Further, Thieme [8] characterized the relationship between spectral bound and reproduction number for infinite-dimensional population structure and time heterogeneity. Bacaër et al. [9] obtained a more biological interpretation of R0 for periodic models, and Inaba [10] used the concept of a generation evolution operator to give a new definition of R0 in heterogeneous environments.
In population biology and epidemiology, more and more people adopt impulsive delay differential equations. For instance, Huo et al. [11] studied the global attractivity of positive periodic solutions for an scalar impulsive differential equations. Gourley et al. [12] evaluated the effectiveness of the age-structure culling strategies for controlling the vector-borne diseases by pulse model with two delays. Shen et al. [13] studied the positive periodic solutions of a predator-prey model with impulsive harvest and time delays. Motivated by the recent works of Liang et al. [5], Zhao [6], especially Bai et al. [7], by using a method similar to Bai et al. [7], we show that R0−1 determines the stability of the zero solution of the associated periodic linear impulsive system and there are no special requirements for the period of the model coefficients and pulse time of the model coefficients.
This paper is organized as follows. In Section 2, we introduce the definition of the basic reproduction ratio R0 for impulse delay periodic compartmental model and prove a stability equivalence result (Theorem 2.1). In Section 3, we apply the developed theory to an impulsive model of parasitic diseases in aquaculture with time delay, and establish a threshold type result in terms of R0. A brief discussion section then completes the paper.
Let τ≥0 be a given number, denote
PC([−τ,0],Rm)={ϕ:[−τ,0]→Rm|ϕ(t−)=ϕ(t),∀t∈(−τ,0],ϕ(t+) exists for t∈[−τ,0) and ϕ(t+)=ϕ(t) for all but at most finite number of points t∈[−τ,0)}, |
where ϕ(t+)=lims→t+ϕ(s) and ϕ(t−)=lims→t−ϕ(s). We equip the linear space PC([−τ,0],Rm) with the norm ∥⋅∥τ defined by ∥ϕ∥τ=sup−τ≤θ≤0|ϕ(θ)|, where |⋅| is any convenient norm on Rm (for example the Euclidean norm). For a left continuous function u:R→Rm and u(t+)=u(t) for all but at most finite number of points within a limited interval, define ut∈PC([−τ,0],Rm) by
ut(θ)=u(t+θ),∀θ∈[−τ,0] |
Let F:R→L(PC([−τ,0],Rm),Rm) be a map and V(t) be a continuous m×m matrix function on R. Assume that F(t) and V(t) are ω−periodic in t for some real number ω>0, and for each fixed ϕ, F(t)ϕ is a measurable function of t, continuous in ϕ for each fixed t and there exists M>0 such that
‖F(t)‖≤M, t∈[0,ω]. |
We consider a linear impulsive and periodic functional differential system
u′(t)=F(t)ut−V(t)u(t),a.e. t>0,t≠tn, | (2.1a) |
u(t+)−u(t)=Pnu(t),t=tn,n∈N, | (2.1b) |
where Pn is m×m matrix such that det(Pn+I)≠0 (n∈N), I is the identity matrix, tk:N→R+ and there exists an integer q≥1 such that tn+q=tn+ω, Pn+q=Pn, n∈N.
Let ˜J=[−τ,0)∩{ti+nω:n∈Z and i=1,...,q} (possibly empty), and J=[−τ,0)∖˜J. Denote the space PCJ by
PCJ={ϕ:[−τ,0]→Rm|ϕ(t−)=ϕ(t),∀t∈(−τ,0],ϕ(t+) exists for t∈[−τ,0) and ϕ(t+)=ϕ(t) for t∈J}. |
It is easy to see that PCJ⊂PC([−τ,0],Rm), and PCJ is a Banach space endowed with the norm ∥⋅∥τ. Given ϕ∈PCJ we consider the following initial condition
u(θ)=ϕ(θ),θ∈[−τ,0]. | (2.2) |
A function is said to be a solution of (2.1) with initial conditions (2.2) if the following condition are satisfied:
(i) u(t) is absolutely continuous on each interval (tn,tn+1], n∈N.
(ii) u(t+n), u(t−n) existandu(t−n)=u(tn), n∈N.
(iii) u(t) satisfies (2.1a) for almost everywhere in [0,+∞)∖{tn} and u(t) satisfies (2.1b) for every t=tn, n∈N.
By the general theory of impulsive delay differential equations [14], it follows that for any ϕ∈PCJ, system (2.1) has a unique solution u(t,ϕ) on [0,+∞) with u0=ϕ.
The internal evolution of individuals in the infectious compartments (e.g., natural and disease-induced deaths, movements among compartments and recovery) is governed by the linear impulsive ordinary differential system:
{dudt=−V(t)u(t),t≠tn,u(t+)−u(t)=Pnu(t),t=tn,n∈Z. | (2.3) |
By the theory of impulsive differential equations [15,Sect.1.2], we denote by Un(t,s)(n∈Z,t,s∈(tn−1,tn]) the fundamental matrix for the linear equation
dudt=−V(t)u(t)(tn−1<t≤tn). |
Then the evolution matrices associated with (2.3) can be written as
W(t,s)={Un(t,s) for t,s∈(tn−1,tn],Un+1(t,t+n)n∏j=i+1(Pj+I)Uj(tj,t+j−1)(Pi+I)Ui(ti,s) for ti−1<s≤ti<tn<t≤tn+1 |
A straightforward verification shows that
W(t,t)=IW(tn,t−n)=IW(t+n,s)=(I+Pn)W(tn,s)(s∈(tn−1,tn])∂∂tW(t,s)=−V(t)W(t,s)(t>s,t≠tn) | (2.4) |
In order to introduce the basic reproduction ratio for system (2.1), we need the following assumptions.
(H1) For any given t∈R, operator F(t):PC([−τ,0],Rm)→Rm is positive in the sense that F(t)PC([−τ,0],Rm+)⊆Rm+.
(H2) For any given t∈R, matrix −V(t) is cooperative, and r(W(ω,0))<1,
(H3) Pn+I is nonnegative for every n∈N, which means all entries of Pn+I are nonnegative,
where r(W(ω,0)) represents the spectral radius of W(ω,0). By the theory of impulsive delay differential equations [15,Remark 3.5] with (H2), we can conclude that there exists K≥1 and α>0 such that
‖W(t,s)‖≤Ke−α(t−s)t≥s,s∈R | (2.5) |
In view of the periodic environment, we suppose that v(t), ω−periodic in t, is the initial distribution of infectious individuals. For any t≥s, F(t−s)vt−s is the distribution of newly infected individuals at time t−s, which is produced by the infectious individuals who were introduced over the time interval [t−s−τ,t−s]. Then W(t,t−s)F(t−s)vt−s is the distribution of those infected individuals who were newly infected at time t−s and remain in the infected compartments at time t. It follows that
∫∞0W(t,t−s)F(t−s)vt−sds=∫∞0W(t,t−s)F(t−s)v(t−s+⋅)ds |
is the distribution of accumulative new infections at time t produced by all those infectious individuals introduced at all previous times to t.
By the definition of W(t,s), for any given s≥0, W(t,t−s)vt−s is the distribution of those infectious individuals who were introduced at time t−s and remain in the infected compartments at time t, then w(t):=∫∞0W(t,t−s)vt−sds is the distribution of accumulative infectious individuals who were introduced at all previous times to t and remain in the infected compartments at time t. Thus, the distribution of newly infected individuals at time t is
F(t)wt=F(t)∫∞0W(t+⋅,t−s+⋅)v(t−s+⋅)ds. |
Let the Banach space Xω be given by
Xω={v:R→Rm|v is continuous in (tn,tn+1),n∈Z,v(t−n),v(t+n) exist and v(t−n)=v(tn),v(t+ω)=v(t) for t∈R).} |
with the norm ‖v‖Xω=supt∈[0,ω]|v(t)|, where |⋅| is any convenient norm on Rm and the positive cone X+ω:={v∈Xω:v(t)≥0,∀t∈R}. According to the above analysis, we define two linear operators on Xω by
[Lv](t)=∫∞0W(t,t−s)F(t−s)v(t−s+⋅)ds,t∈R,v∈Xω, |
and
[ˆLv](t)=F(t)∫∞0W(t+⋅,t−s+⋅)v(t−s+⋅)ds,t∈R,v∈Xω. |
Moreover, let A and B be two bounded linear operators on Xω defined by
[Av](t)=∫∞0W(t,t−s)v(t−s+⋅)ds,[Bv](t)=F(t)vt,t∈R,v∈Xω. |
It then follows that L=A∘B and ˆL=B∘A, and hence, L and ˆL have the same spectral radius. Motivated by the concept of next generation operators in [7], we define the spectral radius of L and ˆL as the basic reproduction ratio
R0:=r(L)=r(ˆL) |
for periodic system (2.1).
In view of the previous assumptions and conclusions, one can verify that L and ˆL are well defined. To see this, we introduce the following operator.
For any given λ∈R, let Eλ be a linear operator on PC([−τ,0],Rm) defined by
[Eλϕ]=eλθϕ(θ),θ∈[−τ,0],ϕ∈PC([−τ,0],Rm). |
It is easy to see that ‖Eλ‖≤max{1,e−λτ} for any λ∈R. Then we introduce a family of linear operators Lλ on Xω:
[Lλv](t)=∫∞0e−λsW(t,t−s)F(t−s)Eλv(t−s+⋅)ds,t∈R,v∈Xω. |
Obviously, L0=L, and Lλ is well defined for all λ>−α. Further, we can prove the following assertion.
Lemma 2.1. For each λ>−α,Lλ is positive, continuous and compact on Xω.
Proof. Let λ>−α be given. Clearly, Eλ is a positive linear operator on PC([−τ,0],Rm). By hypothesis (H1), (H2) and (H3), F(t) and W(t,s)(t≥s) are positive linear operators. Then it is easy to see that Lλ is positive on Xω. In view of
∣e−λsW(t,t−s)F(t−s)Eλv(t−s+⋅)∣≤KMe−(λ+α)smax{1,e−λτ}∥v∥Xω |
for any given t∈R, s∈[0,+∞) and v∈Xω, we have
∣[Lλv](t)∣≤∫∞0∣e−λsW(t,t−s)F(t−s)Eλv(t−s+⋅)∣ds≤KMmax{1,e−λτ}∥v∥Xω∫∞0e−(λ+α)sds≤KMmax{1,e−λτ}λ+α∥v∥Xω,t∈R,v∈Xω. |
Thus,
∥Lλv∥Xω=maxt∈[0,ω]∣[Lλv](t)∣≤KMmax{1,e−λτ}λ+α∥v∥Xω, |
That means Lλ is bounded, and hence, continuous on Xω
Let v∈Xω and ∀t_,¯t∈(0,ω] with t_<¯t. Note that
[Lλv](t)=∫t−∞e−λ(t−s)W(t,s)F(s)Eλvsds,t∈R,v∈Xω, |
we have
|[Lλv](ˉt)−[Lλv](t_)∣≤∫ˉt−∞∣e−λ(ˉt−s)W(ˉt,s)F(s)Eλvs−e−λ(t_−s)W(t_,s)F(s)Eλvs|ds+∫ˉtt_e−λ(t_−s)|W(t_,s)F(s)Eλvs|ds≤∫ˉt−∞e−λ(t_−s)|(e−λ(ˉt−t_)W(ˉt,t_)−I)W(t_,s)F(s)Eλvs|+KMmax{1,e−λτ}λ+α∥v∥Xω(e(λ+α)(ˉt−t_)−1). |
Thus |[Lλv](ˉt)−[Lλv](t_)∣→0 as |ˉt−t_|→0, and hence [Lλv](t) is quasiequicontinuous. Therefore, by applying [15,Lemma 2.4], Lλ is compact on Xω.
In order to study the properties of the solution of the equation (2.1), we need the following auxiliary system. For any given λ∈[0,+∞), consider the linear impulsive system:
{u′(t)=λF(t)ut−V(t)u(t),a.e. t>0,t≠tn,u(t+)−u(t)=Pnu(t),t=tn,n∈N, | (2.6) |
For any ϕ∈PC+J, [14,Corollary 3.1] implies that system (2.6) admits a unique solution uλ(t,ϕ) on [0,+∞) satisfying uλ0=ϕ. In addition, the following statements are valid.
Lemma 2.2. (i) For any ϕ∈PC+J, the solution uλ(t,ϕ) of (2.6) with uλ0=ϕ is positive.
(ii) Let uλ1(t,ϕ) and uλ2(t,ϕ) be two solutions of (2.6) with 0≤λ1<λ2. Then uλ1(t,ϕ)≤uλ2(t,ϕ) for all t≥0.
Proof. (i)For ϵ>0, we consider the following small perturbations of system (2.1)
{u′(t)=λF(t)ut−V(t)u(t)+ϵe,a.e. t>0,t≠tn,u(t+)−u(t)=Pnu(t),t=tn.n∈N, | (2.7) |
where e=(1,1,...,1)T∈Rm. For any ψ∈PC+J, we define
f(t,ψ)=λF(t)ψ−V(t)ψ(0), |
fϵ(t,ψ)=λF(t)ψ−V(t)ψ(0)+ϵe. |
Let uλ,ϵ(t,ϕ) be the solution of (2.7) with initial condition uλ,ϵ0=ϕ∈PC+J. Now we show that uλ,ϵ(t,ϕ)≫0 for t>0 whenever ϕ≥0. Due to uλ,ϵ(0,ϕ)=ϕ(0)≥0 and (uλ,ϵ)′+(0,ϕ)=f(0,ϕ)+ϵ>0, where (uλ,ϵ)′+(0,ϕ) is the right derivative with respect to (2.7). It is easy to see that there exist a positive number ℓ>0 such that uλ,ϵ(t,ϕ)≫0 for 0<t<ℓ. Therefore, if the result is false, we can assume that there exists ˆt>0 such that
uλ,ϵ(t,ϕ)≫0 for 0<t<ˆt and uλ,ϵi(ˆt,ϕ)=0 for some i. |
Then we get (uλ,ϵi)′_(ˆt,ϕ)≤0, where (uλ,ϵi)′_(ˆt,ϕ) is the left-hand derivative of uλ,ϵi at t=ˆt. However, uλ,ϵˆt∈PC([−τ,0],Rm+) and by (H1), (H2) and (2.7),
(uλ,ϵi)′_(ˆt,ϕ)=fi(ˆt,uλ,ϵˆt)+ϵ>0, |
which is a contradiction. By (H3) the positivity of uλ,ϵ(t,ϕ) doesn't change at impulsive points, thus uλ,ϵ(t,ϕ)≫0 for t∈(0,+∞). Let ϵ→0, then uλ,ϵ(t,ϕ)→uλ(t,ϕ), and hence uλ(t,ϕ)≥0 for t∈[0,+∞).
(ⅱ) Note that (2.6) is a linear equation, (ⅱ) is a corollary of (ⅰ).
Let U(t,λ) be the solution maps of system (2.6) from PCJ to PC([−τ,0],Rm) that is,
U(t,λ)ϕ=uλt(ϕ), |
where uλ(t,ϕ) is the unique solution of (2.6) with uλ0=ϕ∈PCJ. It is easy to see that U(ω,λ) is an operator from PCJ to PCJ. The basic reproduction number of (2.6) is represented by R(λ):=r(λL)=λR0. For convenience, we write U:=U(ω,1) for λ=1. Then we have the following result.
Lemma 2.3. The existence of fixed point of U in PCJ is equivalent to that of ω-periodic solution of (2.1).
Proof. Suppose that u(t,ϕ∗) is a periodic solution of (2.1), that is, u(t+ω,ϕ∗)=u(t,ϕ∗) for t∈[−τ,+∞). Note that U is the Poincaré map of system (2.6) with λ=1, the proof is obvious.
Conversely, assume that ϕ∗∈PCJ and Uϕ∗=ϕ∗. Let u(t):=u(t,ϕ∗) be the solution of (2.1) with u0=ϕ∗. Let v(t):=u(t+ω),∀t>0. Then Consider the derivative of v(t) we get
v′(t)=u′(t+ω)=F(t+ω)ut+ω−V(t+ω)u(t+ω)=F(t)vt−V(t)v(t),a.e. t>0,t≠tn |
and at the impulsive points v(t+n)=u(t+n+q)=(I+Pn+q)u(tn+q)=(I+Pn)v(tn),n∈N. In addition, v(θ)=u(θ+ω)=uω(θ)=u0(θ)=ϕ∗(θ) for all θ∈[−τ,0]. Thus the uniqueness of solutions of (2.1) imply that u(t+ω)=u(t) for all t≥−τ. Clearly, u(t+n)=u(t+n+q),∀n∈N. Hence, u(t) is a periodic solution of (2.1).
Lemma 2.4. For each n>τω, Un is positive and compact on PCJ.
Proof. Let u(t,ϕ) be the unique solution of (2.1) with u0=ϕ∈PCJ. By Lemma 2.2(ⅰ), Un,n∈N is positive on PCJ. Under the assumption that nω>τ, Unϕ can be written as follows
[Unϕ](θ)=ϕ(0)+∫nω+θ0F(s)us−V(s)u(s)ds+∑k:tk∈[0,nω+θ)Pku(tk,ϕ),−τ≤θ≤0, |
where tk is the impulsive point on [0,nω+θ). Let B be the bounded subset of PCJ. Obviously, for any ϕ∈B, u(s,ϕ) is bounded for s∈[0,nω], and hence, Un(B) is bounded.
Next, we prove that the set Un(B) is quasiequicontinuous whenever nω>τ. Then, with [15,Lemma 2.4], we can deduce that Un is compact for nω>τ.
Note that V(t) is continuous and periodic, then ‖V(t)‖≤N and ‖F(t)‖≤M for t>0, Thus
∫βα∣F(s)us−V(s)u(s,ϕ)∣ds≤(M+N)B′(β−α)=M1(β−α) |
for 0≤α≤β≤nω, where B′≥max0≤s≤nω|u(s,ϕ)| with ϕ∈B. Suppose that the interval (nω−τ,nω) has p impulsive points, denote by θ(n)i,1≤i≤p. Let θ(n)0=nω−τ and θ(n)p+1=nω. Then for any ˉt,t_∈(−τ,0] with nω+ˉt,nω+t_∈(θ(n)i,θ(n)j] for some i,j, we have
|[Unϕ](ˉt)−[Unϕ](t_)|≤∫nω+ˉtnω+t_∣F(s)us−V(s)u(s,ϕ)∣ds+|∑k:tk∈[0,nω+ˉt)Pku(tk,ϕ)−∑k:tk∈[0,nω+t_)Pku(tk,ϕ)|≤M1|ˉt−t_|+|∑k:tk∈[0,nω+ˉt)Pku(tk,ϕ)−∑k:tk∈[0,nω+t_)Pku(tk,ϕ)|. |
which implies that |[Unϕ](ˉt)−[Unϕ](t_)|→0 as |ˉt−t_|→0. Hence, Un(B) is quasiequicontinuous. The proof is completed.
Lemma 2.5. If r(U)=1, then R0≥1.
Proof.Fix an integer n0>0 such that n0ω>τ. Then the operator Un0:=U(n0ω) is compact and positive by Lemma 2.4. The Krein-Rutman theorem [16,Theorem 1.1] implies that there is ϕ∈PC+J∖{0} such that Un0ϕ=r(Un0)ϕ=ϕ. Note that
(Un0−I)ϕ=(U−I)(Un0−1+Un0−2+⋅⋅⋅+U+I)ϕ:=(U−I)ϕ∗=0, |
from the positivity of U, we get that ϕ∗∈PC([−τ,0],Rm0)∖{0} and Uϕ∗=ϕ∗. It follows from Lemma 2.3 that u∗(t):=u(t,ϕ∗) with u∗0=ϕ∗ is an ω-periodic solution of (2.1). We extend u∗(t) to the entire R in the following way
ˆu∗(t)={u∗(t),t≥0,u∗(t+(n+1)ω),t∈(−(n+1)ω,−nω],n∈N, |
It is easy to see that ˆu∗(t) is an ω-periodic solution of system (2.1). By the constant-variation formula,
ˆu∗(t)=W(t,t0)ˆu∗(t0)+∫tt0W(t,s)F(s)ˆu∗sds,t≥t0,t0∈R. | (2.8) |
Letting t0→−∞ in (2.8), together with (2.5) and the boundedness of ˆu∗(t) on R, we have
ˆu∗(t)=∫t−∞W(t,s)F(s)ˆu∗sds,t∈R. |
Thus 1 is an eigenvalue of L on Xω, and hence, R0≥1.
Lemma 2.6. If R0=1, then r(U)≥1.
Proof. If R0=1, then there exists φ∗∈Xω such that
φ∗=∫t−∞W(t,s)F(s)φ∗sds,t∈R. |
It is easy to see that
(φ∗)′=F(t)φ∗t−V(t)φ∗(t),a,e,t∈R. |
At the impulsive points tn, (2.4) indicates that φ∗(t+n)=(I+Pn)φ∗(tn),n∈Z. Thus, u∗(t):=φ∗|t∈[−τ,+∞) is a periodic solution of (2.1). Lemma 2.3 implies that Uu∗0=u∗0, hence, r(U)≥1.
Lemma 2.7. r(U)>1 if and only if R0>1.
Proof.(Necessity) If 0<λ1<λ2, then U(ω,λ1)(ϕ)≤U(ω,λ2)(ϕ) for any ϕ∈PC+J by Lemma 2.2(ⅰ). Note that U(ω,λ) is a positive and bounded linear operator on PCJ, then [17,Theorem 1.1] implies that U(ω,λ) is a nondecreasing function of λ on (0,+∞). By Lemma 2.3, there exists n0∈N such that Un0 is compact. Since r(U)=r(U(ω,1))>1 and limλ→0+r(U(ω,λ))=r(W(ω,0))<1, By the continuity of spectral radius for compact linear operators [18,Theorem 2.1(a)] and U and Un0 have the same spectral radius, we can conclude that there exists λ0∈(0,1) such that r(U(ω,λ0))=1. Thus R(λ0)≥1 by Lemma 2.5. Then R0=R(λ0)λ0>1.
(Sufficiency) It is easy to see that R(1R0)=1 since R(1)=R0>1. Then r(U(ω,R−10))≥1 due to Lemma 2.6. Note that r(U(ω,λ)) is nondecreasing in λ∈[0,+∞), we have r(U(ω,1)≥r(U(ω,R−10)).
Suppose, by contradiction, that r(U(ω,1))=1. Then r(U(ω,λ))=1 for all λ∈[R−10,1]. By the proof of Lemma 2.5, we see that 1 is an eigenvalue of λL with positive eigenvector for all λ∈[R−10,1]. Thus λ−1 is an eigenvalue of L with positive eigenvector for all λ∈[R−10,1]. But this is impossible since the compact linear operator L has only countablely many eigenvalues. This shows that r(U(ω,0))>1. The proof is completed.
In fact, we have proved the main theorem of this section.
Theorem 2.1. R0−1 has the same sign as r(U)−1.
The following result is about the numerical calculation of R0 (see, e.g., [7]). For completeness, here we list the main steps.
Theorem 2.2. If R0>0 then λ=R−10 is the unique solution of r(U(ω,λ))=1.
Proof. The proof is essentially the same as that of Bai [7], we omit it here.
Remark 2.1. With Theorem 2.2, we can numerically compute R0 according to the following three steps:
With Theorem 2.2, we can numerically compute R0 according to the following three steps:
(i) Choose initial function v0∈Int(PC+J).
(ii) For any given λ∈[0,+∞), let an=‖U(ω,λ)vn−1‖τ and vn=U(ω,λ)vn−1‖U(ω,λ)vn−1‖τ for all n≥1. By Liang et al. [5,Lemma 2.5], it follows that
r(U(ω,λ))=limn→∞‖Un(ω,λ)v0‖1n=limn→∞(n∏k=1)1nak. |
Moreover, if limn→∞an exists, then r(U(ω,λ))=limn→+∞an.
(iii) Compute the numerical solution λ0 to r(U(ω,λ))=1 via the bisection algorithm, and hence, R0=1λ0.
Mathematical methods have been used to study infectious diseases for a long time and certain results have been obtained. People have added various factors to the model by analyzing the actual situation to better reflect the actual problem. For example, Sharma et al. [19] considered the incubation period of the disease and used the time-delayed model, Gupta et al. [20] used a time-delay SEIRD model to analyze the spread of COVID-19 infection in a population. Church et al. [21] considered the vaccination situation and used the impulse model, etc.
The non-impulsive system has been studied in [6] and non-delayed system has been studied in [22], In this section, inspired by the above work, we combine a variety of factors and apply the theory developed in the previous section to an impulsive model of parasitic diseases in pork farming with time delay. Parasites can cause a decrease in feed conversion rate and weight gain, and there is a risk of pig-to-human transmission in pork and various organs after slaughter. This has caused huge economic losses to the global pork industry [23]. Concentrated breeding makes the treatment of parasitic diseases easier to achieve. The purpose of treatment is achieved by adding to the feed and spraying the corresponding drugs on the body surface. The timely removal of feces and other potential sources of infection can effectively reduce the infection rate of parasitic diseases. However, there are few researches on the above strategies in mathematics. The following simulation of the above methods by applying the method of infectious disease dynamics has some reference significance for when and how to control parasitic diseases in reality.
Let S(t),E(t) and I(t) be the total numbers at time t of the susceptible, exposed, and infective populations, respectively. For simplicity, we assume that the latent period of the disease is τ, and the incidence rate function f(t,S,I) depends on time t and variables S and I. Let μ(t) be the natural death rate of the population. It then follows that the rate of entry into the infective class from the exposed one at time t is
e−∫tt−τμ(r)drf(t−τ,S(t−τ),I(t−τ)). |
As discussed in [24], E(t) can be represented as
E(t)=∫tt−τe−∫tsμ(r)drf(s,S(s),I(s))ds. |
It is easy to show that
E′(t)=f(t,S(t),I(t))−e−∫tt−τμ(r)drf(t−τ,S(t−τ),I(t−τ))−μ(t)E(t). |
Then, we obtain the following non-autonomous SEI model with impulses:
{S′(t)=δ(t)−f(t,S(t),I(t))−μ(t)S(t),E′(t)=f(t,S(t),I(t))−e−∫tt−τμ(r)drf(t−τ,S(t−τ),I(t−τ))−μ(t)E(t),I′(t)=e−∫tt−τμ(r)drf(t−τ,S(t−τ),I(t−τ))−μ(t)I(t),}a.e.t>0,t≠tn,S(t+)=S(t)+θE(t)+θI(t),E(t+)=E(t)−θE(t),I(t+)=I(t)−θI(t),}t=tn,n∈N. | (3.1) |
Here δ(t) is the recruitment rate, θ represents the treatment rate of parasite drugs.
According to [25], we need to impose the following compatibility condition:
E(0)=∫0−τe−∫0sμ(r)drf(s,S(s),I(s))ds. | (3.2) |
Assume that f(t,S,I) and all these time-dependent coefficients are ω-periodic in t for some real number ω>0. Obviously the function
p(t):=e−∫tt−τμ(r)dr |
is also ω-periodic, and hence, model (3.1) is an ω-periodic and time-delayed system with impulses. To study the dynamic behavior of system (3.1), we make the following assumptions, which are natural considering the biological background of the system (3.1):
(A1) δ(t) and μ(t) are all non-negative and continuous functions with δ(t)>0 and ∫ω0μ(t)dt>0.
(A2) f(t,S,I) is a C1-function with the following properties:
(i) f(t,S,0)≡0,f(t,0,I)≡0, and ∂f(t,S,0)∂I are positive and non-decreasing for all S>0.
(ii) ∂f(t,S,I)∂S≥0 and f(t,S,I)≤∂f(t,S,0)∂II for all (t,S,I)∈R×R2+.
By virtue of (A1), the scalar linear periodic equation
S′(t)=δ(t)−μ(t)S(t) | (3.3) |
has a unique positive ω-periodic solution S∗(t), which is globally stable in R. Linearizing system (3.1) at its disease-free periodic solution (S∗(t),0,0), we obtain the following periodic linear impulsive differential equation for the infective variable I:
{I′(t)=a(t)I(t−τ)−μ(t)I(t),a.e.t>0,t≠tn,I(t+)=(1−θ)I(t),t=tn,n∈N. | (3.4) |
where
a(t)=p(t)∂f(t−τ,S∗(t−τ),0)∂I |
Take m=1, F(t)ϕ=a(t)ϕ(−τ), and V(t)=μ(t). Then system (2.1) becomes system (3.4), and
W(t,s)=e−∫tsμ(r)dr∏k:tk∈[s,t)(1−θ)∀t≥s,s∈R |
[Lv](t)=∫∞0W(t,t−s)F(t−s)v(t−s+⋅)ds=∫∞0W(t,t−s)a(t−s)v(t−s−τ)ds=∫∞τW(t,t−s+τ)a(t−s+τ)v(t−s)ds=∫∞0K(t,s)v(t−s)ds,∀t∈R,v∈Xω, |
where
K(t,s)={W(t,t−s+τ)a(t−s+τ),ifs≥τ0,ifs<τ. |
From the definition of the basic reproduction number, we get R0=r(L) for the system (3.4)
By the general theory of impulsive delay differential equations [14], it follows that for any ϕ∈PCJ, system (3.1) has a unique solution u(t,ϕ)=(S(t),E(t),I(t)) on [0,+∞) with u0=ϕ∈D.
D={ϕ:[−τ,0]→R3+|ϕ(t−)=ϕ(t),∀t∈(−τ,0],ϕ(t+) exists for t∈[−τ,0),ϕ(t+)=ϕ(t) for t∈J, and ϕ2(0)=∫0−τe−∫0sμ(r)drf(s,ϕ1,ϕ3)ds.}. |
Let N(t)=S(t)+E(t)+I(t). Then we have
N′(t)=δ(t)−μ(t)N(t),t>0. | (3.5) |
Thus, the global stability of S∗(t) for (3.3) implies that solutions of system (3.1) with initial data in D exist globally on [0,+∞) and are ultimately bounded.
Just like the procedure in Sect. 2, we take m=1 and let U(t):PCJ→PC([−τ,0],R) be the solution maps of (3.1), that is,
U(t)ϕ=ut(ϕ),t≥0,ϕ∈PCJ, |
where u(t,ϕ) is the unique solution of (3.1) satisfying u0=ϕ∈PCJ. Then U:=U(ω) is the Poincarˊe (period) map associated with (3.1) on PCJ. Let r(U) be the spectral radius of U. Then the following assertion holds true.
Lemma 3.1. Let μ=lnr(U)ω. Then there exists a positive and ω-periodic function v∗(t) such that eμtv∗(t) is a solution of (3.1).
Proof. Since the proof is essentially the same as that of [7,Lemma 8], we omit it here.
Theorem 3.1. Let (A1) and (A2) hold. Then if R0<1, then the disease-free periodic solution (S∗(t),0,0) of (3.1) is globally attractive.
Proof. Let P(t) be the solution maps of the impulsive delay differential equations (3.4) on PCJ, that is, P(t)ϕ=It(ϕ),t≥0, where I(t,ϕ) is the unique solution of (3.4) satisfying I0=ϕ∈PC+J∖{0}. Put P:=P(ω) is the Poincaré (period) map associated with system (3.4). We have sign(R0−1)=sign(r(P)−1) due to Theorem 2.1.
Let ϕ∈PC+J∖{0}. We assert that there exists t∗∈[0,τ] such that I(t∗)>0. If that's not correct, it means I(t)=0 for all t∈[0,τ]. Then the first equation of (3.4) implies
a(t)I(t−τ)=0,t∈[0,τ], |
but a(t)>0, which contradicts the assumption that ϕ>0. In view of
{I′(t)≥−μ(t)I(t),a.e.t>0,t≠tn,I(t+)=(1−θ)I(t),t=tn,n∈N, |
we have
I(t)≥I(t∗)e−∫tt∗μ(s)ds∏k:tk∈[t∗,t)(1−θ),t≥t∗. |
This means that Pn is strongly positive on PCJ whenever nω≥2τ. By a method similar to the proof of Lemma 2.4, we can show that Pn is compact when nω≥2τ. Since Pn=P(nω), [26,Lemma 3.1] implies that r(P) is a simple eigenvalue of P with a strongly positive eigenvector, and the modulus of any other eigenvalue is less than r(P).
Since R0<1, we have r(P)<1. Let Pϵ be the Poincaré map of the following perturbed impulsive delay differential system
{I′(t)=p(t)∂f(t−τ,S∗(t−τ)+ϵ,0)∂II(t−τ)−μ(t)I(t),a.e.t>0,t≠tn,I(t+)=(1−θ)I(t),t=tn,n∈N. | (3.6) |
Note that limϵ→0r(Pϵ)=r(P)<1, we can fix a sufficiently small number ϵ>0 such that r(Pϵ)<1. By Lemma 3.1, there is positive ω-periodic function vϵ(t) such that uϵ(t)=eμϵtvϵ(t) is a positive solution of (3.6), where μϵ=lnr(Pϵ)ω<0. For any given ϕ∈D, let v(t,ϕ)=(S(t),E(t),I(t)). Due to (3.5) and the global stability of S∗(t) for (3.3), there exists a sufficiently large integer n1>0 such that n1ω≥τ and S(t)<S∗+ϵ,∀t≥n1ω−τ. In view of assumption (A2), we have
{I′(t)≤p(t)∂f(t−τ,S∗(t−τ)+ϵ,0)∂II(t−τ)−μ(t)I(t),a.e.t>0,t≠tn,I(t+)=(1−θ)I(t),t=tn,n∈N, |
for all t≥n1ω. Choose a sufficiently large number K>0 such that I(t)≤Kuϵ(t),∀t∈[n1ω−τ,n1ω]. Thus, the same method as in [7,Theorem 3] shows that
I(t)≤Kuϵ(t)=Keμϵtvϵ(t),∀t≥n1ω, |
and hence, limt→∞I(t)=0. Thus, for any given κ∈R, by the continuity of f(t,S,I) and f(t,S,0)=0, there is T1>0 such that f(t,S(t),I(t))−e−∫tt−τμ(t)drf(t−τ,S(t−τ),I(t−τ))<κ,
By the second and fifth equations of (3.1), we have
{E′(t)≤κ−μ(t)E(t),a.e.t>0,t≠tn,E(t+)=(1−θ)E(t),t=tn. |
Consider the following impulsive differential equation
{x′(t)=κ−μ(t)x(t),a.e.t>0,t≠tn,x(t+)=(1−θ)x(t),t=tn. | (3.7) |
By virtue of (A1), we see that the system (3.7) has a unique positive ω-periodic solution x∗ which is globally stable in R.
x∗(t)=ce−∫t0μ(s)ds∏k:tk∈[0,t)(1−θ)+κ∫t0e−∫tsμ(s)ds∏k:tk∈[s,t)(1−θ)ds |
Let κ→0 and t→∞, then x∗(t)→0. Thus the comparison theorem for impulsive differential equations and the global stability of x∗(t) for (3.7) implies that E(t)→0 as t→∞. Note that S(t)+E(t)+I(t)=S1(t), where S1(t) is the solution of (3.1) satisfying S1(0)=ϕ1(0)+ϕ2(0)+ϕ3(0) It is easy to see that limt→∞(S1(t)−S∗(t))=0 then limt→∞(S(t)−S∗(t))=0.
Theorem 3.2. Let (A1) and (A2) hold. If R0>1, then there exists a real number η>0 such that the solution u(t,ϕ)=(S(t),E(t),I(t)) satisfies lim inft→∞I(t)≥η for any ϕ∈D.
Proof. Let Pδ be the Poincaré map of
{I′(t)=p(t)(∂f(t−τ,S∗(t−τ),0)∂I−δ)I(t−τ)−μ(t)I(t),a.e.t>0,t≠tn,I(t+)=(1−θ)I(t),t=tn,n∈N |
on PCJ. Since limδ→0r(Pδ)=r(P)>1, we can fix a small number δ>0 such that r(Pδ)>1. It then follows that there is a small number η0>0 such that
f(t−τ,S∗(t−τ)−η0,I)≥(∂f(t−τ,S∗(t−τ),0)∂I−δ)I(t),∀I∈[0,η0]. |
We claim that lim supt→∞I(t)>η0,∀ϕ∈D. Suppose, by contradiction, that there is a ˆt>0 such that I(t)≤η0 for all t≥ˆt. Then system (3.6) becomes
{I′(t)≥p(t)(∂f(t−τ,S∗(t−τ),0)∂I−δ)I(t−τ)−μ(t)I(t),a.e.t>0,t≠tn,I(t+)=(1−θ)I(t),t=tn,n∈N. |
Choose a sufficiently small real number ˆk such that
I(t)≥ˆkeμδtvδ(t),t∈[ˆt−τ,ˆt]. |
Note that μδ>0, let t→∞,then I(t)→∞, which is a contradiction. Then there are only two possibilities: (ⅰ) for all large t, I(t)≥η0; (ⅱ) I(t) oscillates around η0. Obviously, we only need to consider the second case. Put t_ and ¯t are sufficiently large such that
I(t_)=I(¯t)=η0,I(t)<η0,t∈(t_,¯t). |
If ¯t−t_≤τ, then the system (3.1) implies that
I(t)≥I(t_)e−∫tt_μ(s)ds∏k:tk∈[t_,t)(1−θ)≥η0e−∫t+τt_μ(s)ds∏k:tk∈[t_,t_+τ)(1−θ)t∈[t_,¯t]. |
Because μ(t) is ω-periodic and tn+q=tn+ω, it is obvious that e−∫t+τt_μ(s)ds and ∏k:tk∈[t_,t_+τ), for any given τ, have a minimum value which is independent of ˉt and t_, denoted by c. Then I(t)≥η0c=q, if ¯t−t_≤τ.
If ¯t−t_>τ, then for t∈[t_,t_+τ],
I(t)≥η0c=q, |
For t∈(t_+τ,¯t], we consider the following comparison equation
{I′(t)≥p(t)(∂f(t−τ,S∗(t−τ),0)∂I−δ)I(t−τ)−μ(t)I(t),a.e.t>0,t≠tn,I(t+)=(1−θ)I(t),t=tn. |
Note that v(t) is ω-periodic, let m, M be the minimum and maximum of v(t), respectively, then choose a sufficiently small real number k>0 such that
I(t)≥q=keμδ(t_+τ)M≥keμδtvδ(t)≥keμδtm≥keμδt_m=qeμτM=q′,t∈[t_,t_+τ], |
where q′ is a constant independent of t_ and k. Then combined with the principle of comparison, we conclude that
I(t)>q′,t∈(t_+τ,¯t], |
due to I(t)>keμδtvδ(t)>q′. Consequently, we get I(t)>η for t∈[t_,¯t], where η=min{q,q′}. Since this kind of interval [t_,¯t] is chosen arbitrarily, we get lim inft→∞I(t)≥η.
From the above theorem, we find that R0 is the threshold parameter for the extinction of the disease. Now we verify the theoretical results through numerical simulations. According to [27], we set δ(t)=3+0.2×sin(πt/30),μ(t)=0.2×(1+0.2×sin(πt/30)),f=βS(t)I(t)1+0.5×I(t),τ=3,tn=4n,n∈N,θ=0.8, When β=0.055, then R0=0.9665<1. Figure 1 shows that the disease-free periodic solution is globally attractive. When β=0.2, then R0=3.5148>1. Figure 2 shows that the non-zero solution is uniformly persistent.
In this paper, we develop the basic reproduction ratio R0 of more general periodic and time-delayed impulsive model. Note that the period of model coefficients is not same as that of fixed impulsive moments. We extend some known results of Bai et al.[7]. We show that R0 is the threshold parameter for the stability of the zero solution of the associated linear system.Furthermore, We use the developed theory to a swine parasitic disease model with pulse therapy and obtain threshold results on its global dynamics in terms of R0. We hope that our method can also be applied to more general population models or epidemic model with impulse.
In the early stage of the disease outbreak, the corresponding measures have not been implemented. If the time of the outbreak of the infectious disease can be known, preparing in advance is of great significance for controlling the spread of the infectious disease. Turkyilmazoglu [28] proposed a explicit formulae for the peak time of SIR model. Kröger et al. [29] compared different forecasting methods and this is a very meaningful but challenging task for non-autonomous impulsive systems.
We are grateful to the anonymous referees and the editors for their valuable comments and suggestions which improved the quality of the paper. This work is supported by the NNSF of China (11861044), the NSF of Gansu of China(21JR7RA212 and 21JR7RA535) and the HongLiu first-class disciplines Development Program of Lanzhou University of Technology.
The authors declare there is no conflict of interest.
[1] |
O. Diekmann, J. A. P. Heesterbeek, J. A. J. Metz, On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations, J. Math. Biol., 28 (1990), 365–382. doi: 10.1007/bf00178324. doi: 10.1007/bf00178324
![]() |
[2] |
P. Van den Driessche, J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002), 29–48. doi: 10.1016/s0025-5564(02)00108-6. doi: 10.1016/s0025-5564(02)00108-6
![]() |
[3] |
N. Bacaër, S. Guernaoui, The epidemic threshold of vector-borne diseases with seasonality, J. Math. Biol., 53 (2006), 421–436. doi: 10.1007/s00285-006-0015-0. doi: 10.1007/s00285-006-0015-0
![]() |
[4] |
W. Wang, X. Q. Zhao, Threshold dynamics for compartmental epidemic models in periodic environments, J. Dyn. Diff. Equat., 20 (2008), 699–717. doi: 10.1007/s10884-008-9111-8. doi: 10.1007/s10884-008-9111-8
![]() |
[5] |
X. Liang, L. Zhang, X. Q. Zhao, Basic reproduction ratios for periodic abstract functional differential equations (with application to a spatial model for Lyme disease, J. Dyn. Diff. Equat., 31 (2019), 1247–1278. doi: 10.1007/s10884-017-9601-7. doi: 10.1007/s10884-017-9601-7
![]() |
[6] |
X.Q. Zhao, Basic reproduction ratios for periodic compartmental models with time delay, J. Dyn. Diff. Equat., 29 (2017), 67–82. doi: 10.1007/s10884-015-9425-2. doi: 10.1007/s10884-015-9425-2
![]() |
[7] |
Z. Bai, X. Q. Zhao, Basic reproduction ratios for periodic and time-delayed compartmental models with impulses, J. Math. Biol., 80 (2020), 1095–1117. doi: 10.1007/s00285-019-01452-2. doi: 10.1007/s00285-019-01452-2
![]() |
[8] |
H. R. Thieme, Spectral bound and reproduction number for infinite-dimensional population structure and time heterogeneity, SIAM J. Appl. Math., 70 (2009), 188–211. doi: 10.1137/080732870. doi: 10.1137/080732870
![]() |
[9] |
N. Bacaër, On the biological interpretation of a definition for the parameter R0 in periodic population models, J. Math. Biol., 65 (2012), 601–621. doi: 10.1007/s00285-011-0479-4. doi: 10.1007/s00285-011-0479-4
![]() |
[10] |
H. Inaba, On a new perspective of the basic reproduction number in heterogeneous environments. J. Math. Biol., 65 (2012), 309–348. doi: 10.1007/s00285-011-0463-z. doi: 10.1007/s00285-011-0463-z
![]() |
[11] |
W. T. Li, H. F. Huo, Global attractivity of positive periodic solutions for an impulsive delay periodic model of respiratory dynamics, J. Comput. Appl. Math., 174 (2005), 227–238. doi: 10.1016/j.cam.2004.04.010. doi: 10.1016/j.cam.2004.04.010
![]() |
[12] |
S. A. Gourley, R. Liu, J. Wu, Eradicating vector-borne diseases via age-structured culling, J. Math. Biol., 54 (2007), 309–335. doi: 10.1007/s00285-006-0050-x. doi: 10.1007/s00285-006-0050-x
![]() |
[13] |
J. Shen, J. Li, Existence and global attractivity of positive periodic solutions for impulsive predator–prey model with dispersion and time delays, Nonlinear Anal. Real. World. Appl., 10 (2009), 227–243. doi: 10.1016/j.nonrwa.2007.08.026. doi: 10.1016/j.nonrwa.2007.08.026
![]() |
[14] |
G. Ballinger, X. Liu, Existence, uniqueness and boundedness results for impulsive delay differential equations, Appl. Anal., 74 (2000), 71–93. doi: 10.1080/00036810008840804. doi: 10.1080/00036810008840804
![]() |
[15] | D. Bainov, P. Simeonov, Impulsive differential equations: periodic solutions and applications, CRC Press, 1993. |
[16] | Y. Du, Order structure and topological methods in nonlinear partial differential equations: Vol. 1: Maximum principles and applications, World Scientific, 2006. doi: 10.1142/5999. |
[17] |
L. Burlando, Monotonicity of spectral radius for positive operators on ordered Banach spaces, Arch. Math., 56 (1991), 49–57. doi: 10.1007/bf01190081. doi: 10.1007/bf01190081
![]() |
[18] |
G. Degla, An overview of semi-continuity results on the spectral radius and positivity, J. Math. Anal. Appl., 338 (2008), 101–110. doi: 10.1016/j.jmaa.2007.05.011. doi: 10.1016/j.jmaa.2007.05.011
![]() |
[19] |
N. Sharma, A. K. Gupta, Impact of time delay on the dynamics of SEIR epidemic model using cellular automata, Phys. A., 471 (2017), 114–125. doi: 10.1016/j.physa.2016.12.010. doi: 10.1016/j.physa.2016.12.010
![]() |
[20] |
N. Sharma, A. K. Verma, A. K. Gupta, Spatial network based model forecasting transmission and control of COVID-19, Phys. A., 581 (2021), 126–223. doi: 10.1016/j.physa.2021.126223. doi: 10.1016/j.physa.2021.126223
![]() |
[21] | K. E. M. Church, X. Liu, Delayed SIR model with pulse vaccination and temporary immunity, in Bifurcation Theory of Impulsive Dynamical Systems (eds. K. E. M. Church, X. Liu), Springer, Cham, (2021) 325–341. doi: 10.1007/978-3-030-64533-5. |
[22] |
Y. Yang, Y. Xiao, Threshold dynamics for compartmental epidemic models with impulses, Nonlinear Anal. Real. World. Appl., 13 (2012), 224–234. doi: 10.1016/j.nonrwa.2011.07.028. doi: 10.1016/j.nonrwa.2011.07.028
![]() |
[23] |
J. Boes, A. L. Willingham, S. Fuhui, H. Xuguang, Prevalence and distribution of pig helminths in the Dongting Lake Region (Hunan Province) of the People's Republic of China, J. Helminthol., 74 (2000), 45–52. doi: 10.1017/s0022149x00000068. doi: 10.1017/s0022149x00000068
![]() |
[24] | H. R. Thieme, Global asymptotic stability in epidemic models, in Proceesings Equadiff 82 (eds. H. W. Knobloch, K.Schmitt.), Springer, Berlin (1983), 608–615. doi: 10.1007/bfb0103284. |
[25] |
S. Busenberg, K. L. Cooke, The effect of integral conditions in certain equations modelling epidemics and population growth, J. Math. Biol., 10 (1980), 13–32. doi: 10.1007/bf00276393. doi: 10.1007/bf00276393
![]() |
[26] |
X. Liang, X. Q. Zhao, Asymptotic speeds of spread and traveling waves for monotone semiflows with applications, Commun. Pur. Appl. Math., 60 (2007), 1–40. doi: 10.1002/cpa.20154. doi: 10.1002/cpa.20154
![]() |
[27] |
Y. B. Weng, Y. J. Hu, Y. Li, B. S. Li, R. Q. Lin, D. H. Xie, Survey of intestinal parasites in pigs from intensive farms in Guangdong Province, People's Republic of China, Vet. Parasitol., 127 (2005), 333–336. doi: 10.1016/j.vetpar.2004.09.030. doi: 10.1016/j.vetpar.2004.09.030
![]() |
[28] |
M. Turkyilmazoglu, Explicit formulae for the peak time of an epidemic from the SIR model, Phys. D., 422 (2021), 132902. doi: 10.1016/j.physd.2021.132902. doi: 10.1016/j.physd.2021.132902
![]() |
[29] |
M. Kröger, M. Turkyilmazoglu, R. Schlickeiser, Explicit formulae for the peak time of an epidemic from the SIR model. Which approximant to use?, Phys. D., 425 (2021), 132–981. doi:10.1016/j.physd.2021.132981. doi: 10.1016/j.physd.2021.132981
![]() |
1. | Wenjie Li, Ying Zhang, Lihong Huang, Dynamics analysis of a predator–prey model with nonmonotonic functional response and impulsive control, 2023, 204, 03784754, 529, 10.1016/j.matcom.2022.09.002 |