Abstract

We consider a periodic time-dependent predator-prey system with stage structure and impulsive harvesting, in which the prey has a life history that takes them through two stages, immature and mature. A set of sufficient and necessary conditions which guarantee the permanence of the system are obtained. Finally, we give a brief discussion of our results.

1. Introduction

In the natural world, there are many species whose individual members have a life history that takes them through two stages, immature and mature. In particular, we have mammalian populations and some amphibious animals in mind, which exhibit these two stages. From the view point of mathematics, the description of the stage structure of the population in the life history is also an interesting problem in population dynamics. The permanence and extinction of species are significant concepts for those stage-structured population dynamical systems. Recently, stage structure models have been studied by many authors [13]. This is not only because they are much more simple than the models governed by partial differential equations but also because they can exhibit phenomena similar to those of partial differential models [4], and many important physiological parameters can be incorporated. Much research has been devoted to the models concerning single-species population growth with the stage structure of immature and mature [5, 6]. Two species models with stage structure were investigated by Wang and Chen (1997), Magnusson (1999), Xiao and Chen (2003), as well as Cui and Song (2004). Also, Zhang, Chen, and Neumann (2000) proposed the following autonomous stage structure predator-prey system:𝑥1=𝛼𝑥2𝑟1𝑥1𝛽𝑥1𝜂𝑥21𝛽1𝑥1𝑥3,𝑥2=𝛽𝑥1𝑟2𝑥2,𝑥3=𝑥3𝑟+𝑘𝛽1𝑥1𝜂1𝑥3,(1.1)where 𝛼, 𝛽, 𝛽1, 𝜂, 𝜂1, 𝑟, 𝑟1, 𝑟2, and 𝑘 are all positive constants, 𝑘 is a digesting constant.

On the other hand, since biological and environmental parameters are naturally subject to fluctuation in time, the effects of a periodically varying environment are considered as important selective forces on systems in a fluctuating environment. So more realistic and interesting models should take into account the seasonality of the changing environment [7, 8]. This motivated Cui and Y. Takeuchi (2006) to consider the following periodically nonautonomous predator-prey model with stage structure for prey:̇𝑥1=𝑎(𝑡)𝑥2𝑏(𝑡)𝑥1𝑑(𝑡)𝑥21𝑝(𝑡)𝜙𝑡,𝑥1𝑥1𝑦,̇𝑥2=𝑐(𝑡)𝑥1𝑓(𝑡)𝑥22,̇𝑦=𝑦𝑔(𝑡)+(𝑡)𝜙𝑡,𝑥1𝑥1,𝑞(𝑡)𝑦(1.2)in which 𝑎(𝑡), 𝑏(𝑡), 𝑐(𝑡), 𝑑(𝑡), 𝑓(𝑡), 𝑔(𝑡), (𝑡), 𝑝(𝑡), 𝑞(𝑡), and 𝜙(𝑡,𝑥1) are all continuous 𝜔-periodic functions; 𝑎(𝑡), 𝑏(𝑡), 𝑐(𝑡), 𝑑(𝑡), 𝑓(𝑡), (𝑡), 𝑝(𝑡), 𝑞(𝑡) are all positive; 𝑔(𝑡), 𝜙(𝑡,𝑥1) are nonnegative; 𝑥1 and 𝑥2 denote the density of immature and mature population (prey), respectively; and 𝑦 is the density of the predator that only prey on 𝑥1 (immature prey). They provided a set of sufficient and necessary conditions to guarantee the permanence of the above system.

Systems with impulsive effects describing evolution processes are characterized by the fact that at certain moments of time, they experience a change of state abruptly. Processes of such type are studied in almost every domain of applied science. Impulsive equations [9, 10] have been recently used in population dynamics in relation to impulsive vaccination [11, 12], population ecology [13, 14], the chemotherapeutic treatment of disease [15], birth pulses [16], as well as the theory of the chemostat [17].

Let us assume that the predator population is affected by harvesting (e.g., fishing or hunting). Further, as we all know that the harvesting does not occur continuously, that is, the harvesting occurs in regular pulses, then let us assume that at some fixed moments, the predator population in system (1.2) is subject to a perturbation which incorporates the proportional decrease. After a perturbation at step 𝜏𝑘>0(𝑘𝐍), the size of the population 𝑦(𝜏+𝑘) becomes𝑦𝜏+𝑘=1𝑢𝑘𝑦𝜏𝑘,(1.3)where 𝑦(𝜏𝑘) is the size of the predator population at step 𝜏𝑘 before the impulsive perturbation, and 0<𝑢𝑘<1 represents the rate at which the predator is harvested.

In this article, we extend the model (1.2) to the case when the predator population is omnivorous and affected by impulsive effects, which is governed by the following system:̇𝑥1=𝑎(𝑡)𝑥2𝑏(𝑡)𝑥1𝑑(𝑡)𝑥21𝑝1(𝑡)𝜙1𝑡,𝑥1𝑥21𝑦,𝑡𝜏𝑘,̇𝑥2=𝑐(𝑡)𝑥1𝑓(𝑡)𝑥22𝑝2(𝑡)𝜙2𝑡,𝑥2𝑥22𝑦,𝑡𝜏𝑘,̇𝑦=𝑦𝑔(𝑡)+2𝑖=1𝑖(𝑡)𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖𝑞(𝑡)𝑦(𝑡),𝑡𝜏𝑘,Δ𝑦(𝑡)=𝑢𝑘𝑦(𝑡),𝑡=𝜏𝑘,𝑘𝐍,(1.4)with the initial value conditions𝑥1(0)=𝑥100,𝑥2(0)=𝑥200,𝑦(0)=𝑦00,(1.5)in which 𝑥1 and 𝑥2 are the densities of immature and mature prey, respectively; and 𝑦 is the density of the predator that can prey on 𝑥1and𝑥2. When its favorite food is severely scarce, population 𝑦 can eat other resources: Δ𝑦(𝑡)=𝑦(𝑡+)𝑦(𝑡). Also, there exists a positive integer 𝑞 such that𝑢𝑘+𝑞=𝑢𝑘,𝜏𝑘+𝑞=𝜏𝑘+𝜔,0=𝜏0<𝜏1<𝜏2<<𝜏𝑘<𝜏𝑘+1<,𝑘𝐍.(1.6) Here, 𝑥2𝑖𝜙𝑖(𝑡,𝑥𝑖), the number of the prey 𝑥𝑖 consumed per predator in unit time, is called the predator functional response. We assume that there exists a positive constant 𝐿 such that0<𝜙𝑖𝑡,𝑥𝑖𝜕<𝐿;𝜕𝑥𝑖𝑥2𝑖𝜙𝑖𝑡,𝑥𝑖0for𝑥𝑖>0,𝑖=1,2.(1.7)The last condition in (1.7) implies that as the prey population increases, the consumption rate of prey consumed per predator increases. The birth rate of the immature prey population is proportional to the existing mature prey population with a proportionality function 𝑎(𝑡). For the immature prey population, the death rate is proportional to the existing immature prey population with a proportionality function 𝑏(𝑡). The variable parameter 𝑑(𝑡) represents that the immature prey population is density restriction. The transition rate from immature individuals to the mature individuals is assumed to be proportional to the existing immature population with proportionality coefficient 𝑐(𝑡). The death rate of the mature population is of a logistic nature with proportionality coefficient 𝑓(𝑡). Also, 𝑝𝑖(𝑡) and 𝑖(𝑡)(𝑖=1,2) are the coefficients that relate to conversion rates of the immature and, respectively, mature prey biomass into predator biomass. The coefficients in (1.4) are all continuous 𝜔-periodic for 𝑡0. In fact, 𝑎(𝑡), 𝑏(𝑡), 𝑐(𝑡), 𝑑(𝑡), 𝑓(𝑡), 𝑝𝑖(𝑡), 𝑖(𝑡), and 𝑞(𝑡) are all strictly positive, and 𝜙𝑖(𝑡,𝑥𝑖) is nonnegative (𝑖=1,2).

The organization of this paper is as follows. In Section 2, we provide some preliminary results which will be useful. In Section 3, we investigate the permanence and extinction of system (1.4) by using analysis technique. In the last section, we give a biological example and a brief discussion of our result.

2. Preliminary Results

Before stating and proving our main results, we give the following definitions, notations, and lemmas which will be useful in the following section.

Let 𝐑+=[0,), 𝐑𝑛+={(𝑥1,,𝑥𝑛)𝑥𝑖>0,𝑖=1,,𝑛} and 𝛼(𝑡) be a continuous 𝜔-periodic function defined on [0,+), then we set𝐀𝜔(𝛼)=𝜔1𝜔0𝛼(𝑡)𝑑𝑡,𝛼𝑈=max𝑡[0,𝜔]𝛼(𝑡),𝛼𝐿=min𝑡[0,𝜔]𝛼(𝑡).(2.1)Lemma 2.1. System (1.4) is dissipative.Proof. Define a function 𝑉(𝑡,𝑥1,𝑥2,𝑦) such that𝑉𝑡,𝑥1,𝑥2𝑥,𝑦=1𝑡+𝑥2(𝑡)+𝑝𝑦(𝑡),(2.2)in which =max𝑡[0,𝜔]{1(𝑡),2(𝑡)}, 𝑝=min𝑡[0,𝜔]{𝑝1(𝑡),𝑝2(𝑡)}. After a simple computations, we have𝐷+𝑉||(4)+𝑤𝑉<𝑎𝑈+𝑤𝑥2𝑓𝐿𝑥22+𝑐𝑈+𝑤𝑏𝐿𝑥1𝑑𝐿𝑥21+𝑤+𝑝𝑔𝑈𝑦𝑝𝑞𝐿𝑦2,(2.3)in which 𝑤 is a positive constant. Obviously, the right-hand side of the above inequality is bounded above for all (𝑥1(𝑡),𝑥2(𝑡),𝑦(𝑡))𝑅3+. Hence,𝐷+𝑉||(4)+𝑤𝑉<𝜆,(2.4)where 𝜆 is a positive constant. When 𝑡=𝜏𝑘, we get𝑉𝜏+𝑘𝜏𝑉𝑘.(2.5)According to Lemma 2.2 in [9, page 23 ], we derive that𝑉(𝑡)=𝑉(0)𝑒𝑤𝑡+𝑡0𝜆𝑒𝑤(𝑡𝑠)𝑑𝑠<2𝜆𝑤,as𝑡,(2.6)which implies that system (1.4) is dissipative. This completes the proof.Next, we consider the following two subsystems of system (1.4):̇𝑥1=𝑎(𝑡)𝑥2𝑏(𝑡)𝑥1𝑑(𝑡)𝑥21,̇𝑥2=𝑐(𝑡)𝑥1𝑓(𝑡)𝑥22,(2.7)̇𝑦=𝑦𝑔(𝑡)𝑞(𝑡)𝑦(𝑡),𝑡𝜏𝑘,Δ𝑦(𝑡)=𝑢𝑘𝑦(𝑡),𝑡=𝜏𝑘,𝑘𝐍.(2.8)Lemma 2.2 ([18]). The system (2.7) has a positive 𝜔-periodic solution (𝑥1(𝑡),𝑥2(𝑡)) which is globally asymptotically stable with respect to 𝐑2+.Lemma 2.3. If the following conditions 𝐀𝜔(𝑔)>0,𝑞𝑘=11𝑢𝑘>exp𝜔𝐀𝜔(𝑔)(2.9)hold, then system (2.8) has a unique 𝜔-periodic solution: 𝑦(𝑡)=0<𝜏𝑘<𝑡1/1𝑢𝑘exp𝑡0𝑔(𝜉)𝑑𝜉𝜔0𝑠𝜏𝑘<𝜔1/1𝑢𝑘exp𝜔𝑠𝑔(𝜉)𝑑𝜉𝑞(𝑠)𝑑𝑠1𝑞𝑘=11/1𝑢𝑘exp𝜔0+𝑔(𝜉)𝑑𝜉𝑡0𝑠𝜏𝑘<𝑡11𝑢𝑘exp𝑡𝑠𝑔(𝜉)𝑑𝜉𝑞(𝑠)𝑑𝑠1(2.10)and for every solution 𝑦(𝑡) of system (2.8), |||𝑦(𝑡)𝑦|||(𝑡)0as𝑡.(2.11)Proof. Let 𝑦(𝑡)=1/𝑧(𝑡) and obtain the linear nonhomogeneous impulsive equatioṅ𝑧(𝑡)=𝑔(𝑡)𝑧(𝑡)+𝑞(𝑡),𝑡𝜏𝑘,𝑧𝑡+=11𝑢𝑘𝑧(𝑡),𝑡=𝜏𝑘.(2.12)Let 𝑊(𝑡,𝑠)=𝑠𝜏𝑘<𝑡(1/(1𝑢𝑘))exp(𝑡𝑠𝑔(𝜉)𝑑𝜉) be the Cauchy matrix for the relevant homogeneous equation, then the solution of (2.12) has the form𝑧(𝑡)=𝑊(𝑡,0)𝑧(0)+𝑡0𝑊(𝑡,𝑠)𝑞(𝑠)𝑑𝑠.(2.13)The solution 𝑧(𝑡) will be 𝜔-periodic if 𝑧(𝜔)=𝑧(0), or if(1𝑊(𝜔,0))𝑧(0)=𝜔0𝑊(𝜔,𝑠)𝑞(𝑠)𝑑𝑠.(2.14)In view of conditions (2.9), (2.14) has a unique solution𝑧(0)=𝜔0𝑊(𝜔,𝑠)𝑞(𝑠)𝑑𝑠1𝑊(𝜔,0).(2.15)Then, (2.13) has a unique 𝜔-periodic solution𝑧(𝑡)=𝑊(𝑡,0)𝜔0𝑊(𝜔,𝑠)𝑞(𝑠)𝑑𝑠+1𝑊(𝜔,0)𝑡0𝑊(𝑡,𝑠)𝑞(𝑠)𝑑𝑠.(2.16)Hence, (2.8) has a unique 𝜔-periodic solution𝑦(𝑡)=𝑊(𝑡,0)𝜔0𝑊(𝜔,𝑠)𝑞(𝑠)𝑑𝑠+1𝑊(𝜔,0)𝑡0𝑊(𝑡,𝑠)𝑞(𝑠)𝑑𝑠1=0𝜏𝑘<𝑡1/1𝑢𝑘exp𝑡0𝑔(𝜉)𝑑𝜉𝜔0𝑠𝜏𝑘<𝜔1/1𝑢𝑘exp𝜔𝑠𝑔(𝜉)𝑑𝜉𝑞(𝑠)𝑑𝑠1𝑞𝑘=11/1𝑢𝑘exp𝜔0+𝑔(𝜉)𝑑𝜉)𝑡0𝑠𝜏𝑘<𝑡11𝑢𝑘exp𝑡𝑠𝑔(𝜉)𝑑𝜉𝑞(𝑠)𝑑𝑠1.(2.17)From (2.12), we derive that𝑧(𝑡)𝑧(𝑡)=𝑔(𝑡)𝑧(𝑡)𝑧(𝑡),𝜏𝑘<𝑡𝜏𝑘+1.(2.18)Let𝜅𝑞𝑘=111𝑢𝑘exp𝜔0𝑔(𝑡)𝑑𝑡.(2.19)Since conditions (2.9) hold, we note that 0<𝜅<1. Then, from Lemma 2.1, the solution of (2.8) is governed by|||𝑦(𝑡)𝑦|||(𝑡)2𝜆𝑤𝑦𝑈|||𝑧0+)𝑧0+|||0<𝜏𝑘<𝑡11𝑢𝑘exp𝑡0,𝑔(𝑠)𝑑𝑠2𝜆𝑤𝑦𝑈|||𝑧0+𝑧0+|||𝜅[𝑡/𝜔]+10as𝑡,(2.20)in which 𝜆 and 𝑤 are defined in Lemma 2.1. This completes the proof.

3. Permanence and Extinction of System

Theorem 3.1. System (1.4) is permanent if and only if 𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖>0,𝑞𝑘=11𝑢𝑘>exp𝜔𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖,(3.1)in which (𝑥1(𝑡),𝑥2(𝑡)) is the positive 𝜔-periodic solution of system (2.7).We need the following lemmas to prove Theorem 3.1.Lemma 3.2. There exist positive constants 𝑀𝑥 and 𝑀𝑦 such that 𝑙𝑖𝑚𝑠𝑢𝑝𝑡𝑥𝑖(𝑡)𝑀𝑥(𝑖=1,2),𝑙𝑖𝑚𝑠𝑢𝑝𝑡𝑦(𝑡)𝑀𝑦.(3.2)Proof. In fact, from Lemmas 2.1 and 2.3, the proof of the lemma is obvious. This completes the proof.Lemma 3.3. There exists a positive constant 𝜌𝑥(𝜌𝑥<𝑀𝑥) such that 𝑙𝑖𝑚𝑖𝑛𝑓𝑡𝑥𝑖(𝑡)𝜌𝑥(𝑖=1,2).(3.3)Proof. From Lemma 3.2, one notes that there exists a positive constant 𝑇0 such that0<𝑦(𝑡)𝑀𝑦for𝑡𝑇0,(3.4)we then havė𝑥1𝑎(𝑡)𝑥2𝑏(𝑡)𝑥1𝑑(𝑡)+𝑝1(𝑡)𝐿𝑀𝑦𝑥21,̇𝑥2𝑐(𝑡)𝑥1𝑓(𝑡)+𝑝2(𝑡)𝐿𝑀𝑦𝑥22(3.5)for 𝑡𝑇0. According to Lemma 2.2, the following auxiliary equations:̇𝑢1=𝑎(𝑡)𝑢2𝑏(𝑡)𝑢1𝑑(𝑡)+𝑝1(𝑡)𝐿𝑀𝑦𝑢21,̇𝑢2=𝑐(𝑡)𝑢1𝑓(𝑡)+𝑝2(𝑡)𝐿𝑀𝑦𝑢22(3.6)has a globally asymptotically stable 𝜔-periodic solution (𝑢1(𝑡),𝑢2(𝑡)). Let (𝑢1(𝑡),𝑢2(𝑡)) be the solution of (3.6) with (𝑢1(0),𝑢2(0))=(𝑥1(0),𝑥2(0)). By the vector comparison theorem [19], we obtain𝑥𝑖(𝑡)𝑢𝑖(𝑡),𝑖=1,2𝑡0.(3.7)According to the global asymptotic stability of (𝑢1(𝑡),𝑢2(𝑡)), for any positive constant 𝜀(min𝑡[0,𝜔]{𝑢𝑖(𝑡)/3,𝑖=1,2}), there exists a 𝑇1(>𝑇0) such that for all 𝑡𝑇1,|||𝑢𝑖(𝑡)𝑢𝑖|||(𝑡)<𝜀,𝑖=1,2.(3.8)Hence, for all 𝑡𝑇1, we derive that𝑢𝑖(𝑡)𝑢𝑖(𝑡)𝜀,𝑖=1,2.(3.9)Let𝜌𝑥=max𝑡[0,𝜔]𝑢𝑖(𝑡)2,𝑖=1,2,(3.10)then𝑥𝑖(𝑡)𝜌𝑥,𝑖=1,2.(3.11)Consequently,𝑙𝑖𝑚𝑖𝑛𝑓𝑡𝑥𝑖(𝑡)𝜌𝑥,𝑖=1,2.(3.12)This completes the proof.Lemma 3.4. Suppose that (3.1) holds, then there exists a positive constant 𝜚𝑦(𝜚𝑦<𝑀𝑦) such that 𝑙𝑖𝑚𝑠𝑢𝑝𝑡𝑦(𝑡)𝜚𝑦.(3.13)Proof. In view of (3.1), we can choose a positive constant 𝜀0 such that𝐀𝜔𝜓𝜀0(𝑡)>0,𝑞𝑘=11𝑢𝑘>exp(𝜔𝐀𝜔𝜓𝜀0,(𝑡))(3.14)in which𝜓𝜀0(𝑡)=𝑔(𝑡)+2𝑖=1𝑖(𝑡)𝜙𝑖𝑥𝑡,𝑖(𝑡)𝜀0𝑥𝑖(𝑡)𝜀02𝑞(𝑡)𝜀0.(3.15)Consider the following system with a positive parameter 𝜇: ̇𝑥1=𝑎(𝑡)𝑥2𝑏(𝑡)𝑥1𝑑(𝑡)+𝐿𝜇𝑝1𝑥(𝑡)21,̇𝑥2=𝑐(𝑡)𝑥1𝑓(𝑡)+𝐿𝜇𝑝2𝑥(𝑡)22.(3.16)By Lemma 2.2, system (3.16) has a positive 𝜔-periodic solution (𝑥1𝜇(𝑡),𝑥2𝜇(𝑡)), which is globally asymptotically stable. Let (𝑥1𝜇(𝑡),𝑥2𝜇(𝑡)) be the solution of (3.16) with initial condition 𝑥𝑖𝜇(0)=𝑥𝑖(0), 𝑖=1,2, where (𝑥1(𝑡),𝑥2(𝑡)) is the positive periodic solution of (2.7). Hence, for the above 𝜀0, there exists 𝑇2>𝑇1 such that|||𝑥𝑖𝜇(𝑡)𝑥𝑖𝜇|||<𝜀(𝑡)04(3.17)for 𝑡𝑇2, 𝑖=1,2. According to the continuity of the solution in the parameter 𝜇, we have 𝑥𝑖𝜇(𝑡)𝑥𝑖(𝑡)(𝑖=1,2) uniformly in [𝑇2,𝑇2+𝜔] as 𝜇0. Hence, for 𝜀0>0, there exists 𝜇0=𝜇0(𝜀0)(0<𝜇0<𝜀0) such that|||𝑥𝑖𝜇(𝑡)𝑥𝑖|||<𝜀(𝑡)04,0𝜇𝜇0,(3.18)𝑡[𝑇2,𝑇2+𝜔], 𝑖=1,2. Thus, from (3.17) and (3.18), we get|||𝑥𝑖𝜇(𝑡)𝑥𝑖|||<𝜀(𝑡)02,0𝜇𝜇0,(3.19)𝑡[𝑇2,𝑇2+𝜔], 𝑖=1,2. Since 𝑥𝑖𝜇(𝑡) and 𝑥𝑖(𝑡) are all 𝜔-periodic, we have|||𝑥𝑖𝜇(𝑡)𝑥𝑖|||<𝜀(𝑡)02,0𝜇𝜇0,(3.20)𝑡0,𝑖=1,2.
Choose a constant 𝜇1(0<𝜇1<𝜇0,𝜇1<𝜀0), from (3.20), we derive𝑥𝑖𝜇1(𝑡)𝑥𝑖𝜀(𝑡)02,𝑡0,𝑖=1,2.(3.21)Suppose that (3.13) is not true, then for the above 𝜀0, there exists a 𝜈𝐑3+ such that𝑙𝑖𝑚𝑠𝑢𝑝𝑡𝑦(𝑡)<𝜇1,(3.22)where (𝑥1(𝑡),𝑥2(𝑡),𝑦(𝑡)) is the solution of (1.4) with the initial condition (𝑥10,𝑥20,𝑦0)=𝜈. So there exists a constant 𝑇3(>𝑇2) such that𝑦(𝑡)<𝜇1,𝑡𝑇3,(3.23)then we derive thaṫ𝑥1𝑎(𝑡)𝑥2𝑏(𝑡)𝑥1𝑑(𝑡)+𝐿𝜇1𝑝1𝑥(𝑡)21,̇𝑥2𝑐(𝑡)𝑥1𝑓(𝑡)+𝐿𝜇1𝑝2𝑥(𝑡)22(3.24)for 𝑡𝑇3. Let (𝑥1𝜇1,𝑥2𝜇1) be the solution of (3.16) with 𝜇=𝜇1 and (𝑥1𝜇1(𝑇3),𝑥2𝜇1(𝑇3))=(𝑥1(𝑇3),𝑥2(𝑇3)), then by the vector comparison theorem, we obtain𝑥𝑖(𝑡)𝑥𝑖𝜇1(𝑡),𝑖=1,2,(3.25)𝑡𝑇3. By the global asymptotic stability of (𝑥1𝜇1(𝑡),𝑥2𝜇1(𝑡)), for the given 𝜀0>0, there exists 𝑇4>𝑇3 such that𝑥𝑖𝜇1(𝑡)>𝑥𝑖𝜇1𝜀(𝑡)02,𝑡𝑇4,𝑖=1,2,(3.26) and hence, by (3.21), we get𝑥𝑖(𝑡)>𝑥𝑖(𝑡)𝜀0,𝑡𝑇4,𝑖=1,2.(3.27)Since 0<𝑦(𝑡)<𝜇1<𝜀0 together with (1.4) and (1.7), we havė𝑦𝑦𝑔(𝑡)+2𝑖=1𝑖(𝑡)𝜙𝑖𝑥𝑡,𝑖(𝑡)𝜀0𝑥𝑖(𝑡)𝜀02𝑞(𝑡)𝜀0,𝑡𝜏𝑘,𝑦𝑡+=1𝑢𝑘𝑦(𝑡),𝑡=𝜏𝑘,𝑘𝐍.(3.28)Hence, it follows from Lemma 2.2 in [9, page 23] that𝑦(𝑡)𝑦00<𝜏𝑘<𝑡1𝑢𝑘exp𝑡0𝜓𝜀0(𝑠)𝑑𝑠,(3.29)that is,𝑦(𝑡)𝑦00<𝜏𝑘<𝜔1𝑢𝑘exp(𝜔𝐀𝜔𝜓𝜀0[𝑡/𝜔].(3.30)By (3.14), we know that 𝑦(𝑡) as 𝑡, which leads to a contradiction. This completes the proof.
Lemma 3.5. Assume that (3.1) holds, then there exists a positive constant 𝛿𝑦(𝛿𝑦<𝑀𝑦) such that any solution (𝑥1,𝑥2,𝑦) of system (1.4) with initial conditions satisfies 𝑙𝑖𝑚𝑖𝑛𝑓𝑡𝑦(𝑡)𝛿𝑦.(3.31)Proof. Suppose that (3.31) is not true, there must exists a time sequence {𝑡(̂𝑘)𝑡𝑘}𝑘=1𝐑+, ̂𝑘𝐙 such that𝑙𝑖𝑚𝑖𝑛𝑓𝑡(̂𝑘)𝑘𝑦𝑡̂𝑘𝑘<𝜚𝑦(̂𝑘+1)2,(3.32)and by Lemma 3.4, we have 𝑙𝑖𝑚𝑠𝑢𝑝𝑡(̂𝑘)𝑘𝑦(𝑡(̂𝑘)𝑘)𝜚𝑦, ̂𝑘=1,2,. Hence, for each ̂𝑘, we choose two time sequences {𝑠(̂𝑘)𝑞} and {𝑡(̂𝑘)𝑞}, satisfying 0<𝑠(̂𝑘)1<𝑡(̂𝑘)1<𝑠(̂𝑘)2<𝑡(̂𝑘)2<<𝑠(̂𝑘)𝑞<𝑡(̂𝑘)𝑞< and 𝑠(̂𝑘)𝑞 as 𝑞, as well as𝑦𝑠(̂𝑘)𝑞=𝜚𝑦̂𝑡𝑘+1,𝑦(̂𝑘)𝑞=0<𝜏𝑘<𝑡(̂𝑘)𝑞1𝑢𝑘𝜚𝑦(̂𝑘+1)2,(3.33)0<𝜏𝑘<𝑡(̂𝑘)𝑞1𝑢𝑘𝜚𝑦(̂𝑘+1)2𝜚<𝑦(𝑡)<𝑦̂𝑠𝑘+1,𝑡(̂𝑘)𝑞,𝑡(̂𝑘)𝑞.(3.34)̂𝑘By Lemma 3.2, for a given positive integer 𝑇(̂𝑘)>0, there exists 𝑥𝑖(𝑡)𝑀𝑥(𝑖=1,2) such that 𝑦(𝑡)𝑀𝑦 and 𝑇𝑡(̂𝑘) for all 𝑠(̂𝑘)𝑞. In view of 𝑞 as 𝐾(̂𝑘)1, there exists a positive integer 𝑠(̂𝑘)𝑞>𝑇(̂𝑘) such that 𝑞𝐾(̂𝑘)1 as 𝑇𝑡(̂𝑘). Hence, for any ||||||̇𝑦𝑦𝑔(𝑡)𝑀𝑦𝑞(𝑡),𝑡𝜏𝑘,𝑦𝑡+=1𝑢𝑘𝑦(𝑡),𝑡=𝜏𝑘,𝑘𝐍.(3.35), we have𝑠(̂𝑘)𝑞Integrating the above inequality from 𝑡(̂𝑘)𝑞 to 𝑞𝐾(̂𝑘)1, for any 𝑦(𝑡(̂𝑘)𝑞)𝑦(𝑠(̂𝑘)𝑞)𝑠(̂𝑘)𝑞<𝜏𝑘<𝑡(̂𝑘)𝑞(1𝑢𝑘)exp𝑡(̂𝑘)𝑞𝑠(̂𝑘)𝑞||||||𝑔(𝑡)𝑀𝑦𝑞(𝑡)𝑑𝑡.(3.36), then we have𝑡(̂𝑘)𝑞𝑠(̂𝑘)𝑞||||||𝑔(𝑡)+𝑀𝑦̂𝑞(𝑡)𝑑𝑡>ln(𝑘+1)for𝑞𝐾(̂𝑘)1.(3.37)Obviously, it follows from (3.33) that𝑔(𝑡)Hence, in view of the periodicity of 𝑞(𝑡) and 𝑡(̂𝑘)𝑞𝑠(̂𝑘)𝑞̂as𝑘,𝑞𝐾(̂𝑘)1.(3.38), we get𝑇By (3.14), (3.33), and (3.38), there are positive constants 𝑁0 and 𝑦𝑠(̂𝑘)𝑞=0<𝜏𝑘𝑠(̂𝑘)𝑞1𝑢𝑘𝜚𝑦̂𝑘+1<𝜀0𝑡,(3.39)(̂𝑘)𝑞𝑠(̂𝑘)𝑞>2𝑇,(3.40)0<𝜏𝑘<𝜅1𝑢𝑘exp𝜅0𝜓𝜀0(𝑡)𝑑𝑡>1(3.41) such that̂𝑘𝑁0,𝑞𝐾(̂𝑘)1𝜅>𝑇for 𝑦(𝑡)<𝜀0𝑠,𝑡(̂𝑘)𝑞,𝑡(̂𝑘)𝑞(3.42) for ̂𝑘𝑁0,𝑞𝐾(̂𝑘)1., and for all ̇𝑥1𝑎(𝑡)𝑥2𝑏(𝑡)𝑥1𝑑(𝑡)+𝐿𝜀0𝑝1𝑥(𝑡)21,̇𝑥2𝑐(𝑡)𝑥1𝑓(𝑡)+𝐿𝜀0𝑝2𝑥(𝑡)22.(3.43). However, (3.39) implies that(𝑥1𝜀0,𝑥2𝜀0)for 𝜇=𝜀0 Then, we get(𝑥1𝜀0(𝑇3),𝑥2𝜀0(𝑇3))=(𝑥1(𝑇3),𝑥2(𝑇3))Let 𝑥𝑖(𝑡)𝑥𝑖𝜀0𝑠(𝑡),𝑖=1,2,𝑡(̂𝑘)𝑞,𝑡(̂𝑘)𝑞.(3.44) be the solution of (3.16) with lim𝑞𝑠(̂𝑘)𝑞= and ̂𝑘, then by the vector comparison theorem, we obtain𝐾(̂𝑘)2>𝐾(̂𝑘)1From 𝑞𝐾(̂𝑘)2 and Lemmas 3.2 and 3.3, we obtain that for any 𝜌𝑥𝑥𝑖𝑠(̂𝑘)𝑞𝑀𝑥,𝑖=1,2.(3.45), there is a 𝜇=𝜀0 such that for any 𝜔,(𝑥1𝜀0(𝑡),𝑥2𝜀0(𝑡))
For (𝑥1𝜀0(𝑡),𝑥2𝜀0(𝑡)), (3.16) has a globally asymptotically stable positive 𝜀0>0-periodic solution 𝑇5>𝑇4. By the global asymptotic stability of 𝑇5, for the given ̂𝑘, there exists 𝑞, and 𝑥𝑖𝜀0(𝑡)>𝑥𝑖𝜀0𝜀(𝑡)02,𝑡𝑇5+𝑠(̂𝑘)𝑞,𝑖=1,2,𝑞𝐾(̂𝑘)2,(3.46) is dependent of any 𝑥𝑖𝜀0(𝑡)>𝑥𝑖(𝑡)𝜀0,𝑡𝑇5+𝑠(̂𝑘)𝑞,𝑖=1,2,𝑞𝐾(̂𝑘)2.(3.47) and 𝑁1𝑁0 such that𝑡(̂𝑘)𝑞𝑠(̂𝑘)𝑞2𝑇and hence, by (3.21), we get̂𝑘𝑁1By (3.44), there is an 𝑞𝐾(̂𝑘)2 such that 𝑇𝑇5 for all 𝑥𝑖(𝑡)𝑥𝑖(𝑡)𝜀0,𝑖=1,2(3.48) and 𝑡[𝑇+𝑠(̂𝑘)𝑞,𝑡(̂𝑘)𝑞], where ̂𝑘𝑁1. Hence, from (3.44) and (3.47), we obtain𝑞𝐾(̂𝑘)2for all 𝑡[𝑇+𝑠(̂𝑘)𝑞,𝑡(̂𝑘)𝑞], 𝑘𝑁1, and 𝑞𝐾(̂𝑘)2. Since, for any ̇𝑦𝑦𝑔(𝑡)+2𝑖=1𝑖(𝑡)𝜙𝑖𝑥𝑡,𝑖(𝑡)𝜀0𝑥𝑖(𝑡)𝜀02𝑞(𝑡)𝜀0,𝑡𝜏𝑘̂𝑘,𝑡𝑇+𝑠𝑞,𝑡(̂𝑘)𝑞,𝑦𝑡+=1𝑢𝑘𝑦(𝑡),𝑡=𝜏𝑘,𝑘𝐍.(3.49), 𝑇+𝑠(̂𝑘)𝑞 and 𝑡(̂𝑘)𝑞, by (1.4) and (3.42), we have 𝑘𝑁1Integrating from 𝑞𝐾(̂𝑘)2 to 𝑦𝑡(̂𝑘𝑞𝑠𝑦(̂𝑘)𝑞+𝑇𝑠(̂𝑘)𝑞+𝑇<𝜏𝑘<𝑡(̂𝑘)𝑞1𝑢𝑘exp𝑡(̂𝑘)𝑞𝑠(̂𝑘)𝑞+𝑇𝜓𝜀0(𝑡)𝑑𝑡.(3.50) for any 0<𝜏𝑘<𝑡(̂𝑘)𝑞1𝑢𝑘𝜚𝑦(̂𝑘+1)20<𝜏𝑘𝑡(̂𝑘)𝑞1𝑢𝑘𝜚𝑦(̂𝑘+1)2𝑠(̂𝑘)𝑞+𝑇<𝜏𝑘<𝑡(̂𝑘)𝑞1𝑢𝑘exp𝑡(̂𝑘)𝑞𝑠(̂𝑘)𝑞+𝑇𝜓𝜀0>(𝑡)𝑑𝑡0<𝜏𝑘<𝑡(̂𝑘)𝑞1𝑢𝑘𝜚𝑦(̂𝑘+1)2.(3.51) and 𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖0,(3.52), we obtain𝑦(𝑡)Hence, by (3.34), (3.40), and (3.41), we finally derive that𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖>0,𝑞𝑘=11𝑢𝑘exp𝜔𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖.(3.53)This leads to a contradiction. This completes the proof.
According to Lemmas 3.23.5, we can directly prove the sufficient part of the theorem. Next, we are ready to show the necessity of this theorem.

Consider the case oflim𝑡𝑦(𝑡)=0.(3.54)it is then easy to derive that the predator population 0<𝜖<1 must be extinct because of sterilization and impulsive harvesting. Suppose that𝜖1>0We will show that𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑥𝑡,𝑖+𝜖1𝑥𝑖+𝜖12𝑞𝜖>0,𝑞𝑘=11𝑢𝑘<exp𝜔𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑥𝑡,𝑖+𝜖1𝑥𝑖+𝜖12.𝑞𝜖(3.55)In fact, from (1.7) and (3.53), we know that for any given 𝑞(𝑡)>0, there exists 𝑡0 such thaṫ𝑥1𝑎(𝑡)𝑥2𝑏(𝑡)𝑥1𝑑(𝑡)𝑥21,̇𝑥2𝑐(𝑡)𝑥1𝑓(𝑡)𝑥22,(3.56)Note that 𝜖1>0 for 𝑇(1)>0. Since𝑥𝑖(𝑡)𝑥𝑖(𝑡)+𝜖1for𝑡>𝑇(1),𝑖=1,2,(3.57)for the given 𝑞𝑘=11𝑢𝑘<exp𝜔𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥𝑖2𝑞𝜖.(3.58), it is easy to show that there exists 𝑇(2)(>𝑇(1)) such that𝑦(𝑇(2))<𝜖we then have𝑇𝜖𝑦(𝑡)𝑦(1)𝑇(1)<𝜏𝑘<𝑡(1𝑢𝑘)exp𝑡𝑇(1)𝑔(𝑠)+2𝑖=1𝑖(𝑠)𝜙𝑖𝑠,𝑥𝑖𝑥(𝑠)2𝑖(𝑠)𝑞(𝑠)𝜖𝑑𝑠0as𝑡.(3.59)We now show that there must exist 𝜖0 such that 𝑀(𝜖)=max𝑡[0,𝜔]{𝑔(𝑡)+2𝑖=1𝑖(𝑡)𝜙𝑖(𝑠,𝑥𝑖(𝑡))𝑥2𝑖(𝑡)+𝑞(𝑡)𝜖}. Otherwise, by the last two equations in system (1.4), we have𝑀(𝜖)This implies that 𝑀𝜖𝜔𝑦(𝑡)𝜖expfor𝑡𝑇(2).(3.60), which is a contradiction.

Let 𝑇(3)>𝑇(2). Note that 𝑇(3)(𝑇(2)+𝑃1𝜔,𝑇(2)+(𝑃1+1)𝜔], is bounded. We then show that𝑃1𝐙+={0,1,2,}Otherwise, there exists 𝑦(𝑇(3))>𝜖exp(𝑀(𝜖)𝜔). such that 𝑇𝜖exp𝑀(𝜀)𝜔<𝑦(3)𝑇=𝑦(2)𝑇(2)<𝜏𝑘<𝑇(3)1𝑢𝑘exp𝑇(3)𝑇(2)𝑔(𝑠)+2𝑖=1𝑖(𝑠)𝜙𝑖𝑠,𝑥𝑖𝑥(𝑠)2𝑖𝑇(𝑠)𝑞(𝑠)𝜖𝑑𝑠𝑦(2)𝑞𝑘=11𝑢𝑘exp𝜔0𝑔(𝑠)+2𝑖=1𝑖(𝑠)𝜙𝑖𝑠,𝑥𝑖𝑥(𝑠)2𝑖(𝑠)𝑞(𝑠)𝜖𝑑𝑠𝑃1×𝑇(2)+𝑃1𝜔<𝜏𝑘<𝑇(3)1𝑢𝑘exp𝜔0𝑔(𝑠)+2𝑖=1𝑖(𝑠)𝜙𝑖𝑠,𝑥𝑖𝑥(𝑠)2𝑖,(𝑠)𝑞(𝑠)𝜖𝑑𝑠<𝜖exp𝑀(𝜀)𝜔(3.61) in which 𝜖, and 𝑦(𝑡)0 By (3.58), we have𝑡which leads to a contradiction. This implies that (3.60) holds. In view of the arbitrariness of 𝑦, we get that 𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖0,(3.62) as 𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖>0,𝑞𝑘=11𝑢𝑘exp𝜔𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖.(3.63). This completes the proof.

Remark 3.6. From the proof of Theorem 3.1 above, we note that the predator population 𝑥1(𝑡) in system (1.4) will be extinct provided that𝑥2(𝑡)or excessive harvesting, that is,𝑡

4. Discussion

Our result could provide a useful insight into the conservation of beneficial animals, especially rare animals. As an example, we depict the case of Oreolalax omeimontis tadpole, Oreolalax omeimontis, and red-eared slider (T. scripta elegans). Oreolalax omeimontis is a rare species of frog found near Mt. Omei in Sichuan (China). The red-eared slider is a native of the Mississippi Valley area of the United States [8, 20]. Since the 1970s, large numbers of red-eared sliders have been produced on turtle farms in the USA for the international pet trade. Red-eared Turtles are traded as pet animals and have been introduced to many countries. They are omnivorous and will eat insects, tadpoles, crayfish, shrimp, worms, snails, amphibians and small fish, as well as aquatic plants, and they hardly may be controlled by a natural enemy. In our model, the variables 𝑦(𝑡) and 𝑡 represent the density of Oreolalax omeimontis tadpole and Oreolalax omeimontis at time 𝑢𝑘(𝑘=1,2𝑞), respectively. The variable 𝑦 describes the density of the red-eared slider at time 𝑥1. It is well known that Oreolalax omeimontis is beneficial to humans because they eat so many insect pests. Ironically, although red-eared sliders have been widely introduced throughout the world, its detrimental effects have been reported by many researchers [21, 22]. The red-eared slider is one of main enemies of Oreolalax omeimontis. To protect these beneficial and rare toads, we must control the amount of red-eared sliders in the habitat of Oreolalax omeimontis. From a control point of view, our aim is to keep red-eared sliders at an acceptably low level with a minimum use of artificial control measures, not to eradicate all red-eared sliders. Hence, in the above example, the problem of nonextinction of populations becomes a description of reasonable harvesting rates 𝑥2.

According to our main result, system (1.4) is permanent if and only if the growth of the predator 𝜔 by foraging prey populations 𝑢𝑘(𝑘=1,2,𝑞) and 𝜔 plus its intrinsic rate of increase is positive on average during the period 𝑞𝑘=11𝑢𝑘>exp𝜔𝐀𝜔𝑔+2𝑖=1𝑖𝜙𝑖𝑡,𝑥𝑖𝑥2𝑖.(4.1), and harvesting rates 𝑔+2𝑖=1𝑖𝜙𝑖(𝑡,𝑥𝑖)𝑥2𝑖<0 of red-eared sliders during the period 0𝑡𝜔. are small enough to satisfy thatThese seem to be reasonable from a biological point of view; but it should be noted that condition (3.1) allows the predator to be for some time intervals among That is, under reasonable harvesting rates, the predator can survive together with prey populations even if the growth rate of the former is negative for some seasons during a period. We hope that our result can be used to help protect beneficial animals found in their habitats.

Acknowledgment

This work is supported by the National Natural Science Foundation of China (10671001).