This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

11institutetext: Qianqian Zhang22institutetext: School of Mathematics, Harbin Institute of Technology, Harbin 150001, China 33institutetext: Yuan Yuan 44institutetext: Department of Mathematics and Statistics, Memorial University of Newfoundland, St. John’s, NL, A1C 5S7, Canada 55institutetext: Yunfei Lv, Shengqiang Liu 66institutetext: School of Mathematical Sciences, Tiangong University, Tianjin 300387, China

Global Dynamics of a Predator-Prey Model with State-Dependent Maturation-Delay

Qianqian Zhang    Yuan Yuan    Yunfei Lv    Shengqiang Liu111Corresponding author. [email protected]
Abstract

In this paper, a stage structured predator-prey model with general nonlinear type of functional response is established and analyzed. The state-dependent time delay (hereafter SDTD) is the time taken from predator’s birth to its maturity, formatted as a monotonical (ly) increasing, continuous(ly) differentiable and bounded function on the number of mature predator. The model is quite different from many previous models with SDTD, in the sense that the derivative of delay on the time is involved in the model. First, we have shown that for a large class of commonly used types of functional responses, including Holling types I, II and III, Beddington-DeAngelis-type (hereafter BD-type), etc, the predator coexists with the prey permanently if and only if the predator’s net reproduction number is larger than one unit; Secondly, we have discussed the local stability of the equilibria of the model; Finally, for the special case of BD-type functional response, we claim that if the system is permanent, that is, the derivative of SDTD on the state is small enough and the predator interference is large enough, then the coexistence equilibrium is globally asymptotically stable.

Keywords:
State-dependent mature delay Predator-prey model Permanence Extinction Global stability

1 Introduction

In the modeling of natural ecosystems, after Hutchison proposed a Logistic population model with delay, the time delay differential equation models have received widely attention 1 ; 2 . In addition, since the growth process of mammalian has gone through immature and mature stages there is mature time delay, and their behaviors are different at different stages. Therefore, it is necessary to consider the stage structure in the population model. Gurney et al. 3 ; 4 established a stage structure model of the green-headed fly with mature time delay, and numerical simulation based on the experiment data of the green-headed fly experiment of Nicholson verified the rationality of the model in the biological sense.

In 1992, motivated by the significant influences of the population densities on maturation length of juvenile seals and whales found in 5 , Aiello et al 10 argued that the maturity delay of the population should be a function of the total number of populations, which shows that due to the complexity of the ecological environment, the time lag may be adjusted continuously as the state changes, i.e., the constant time delay 6 ; 7 ; 8 ; 9 can not describe the growth of the population as well as SDTD. In 10 , the following SDTD single-population model is established and analyzed:

xi(t)\displaystyle x_{i}^{\prime}(t) =\displaystyle= αxm(t)γxi(t)αeγτ(z(t))xm(tτ(z(t))),\displaystyle\alpha x_{m}(t)-\gamma x_{i}(t)-\alpha e^{-\gamma\tau(z(t))}x_{m}(t-\tau(z(t))),
xm(t)\displaystyle x_{m}^{\prime}(t) =\displaystyle= αeγτ(z(t))xm(tτ(z(t)))βxm2(t).\displaystyle\alpha e^{-\gamma\tau(z(t))}x_{m}(t-\tau(z(t)))-\beta x_{m}^{2}(t).

where xi(t)x_{i}(t) and xm(t)x_{m}(t) are the number of immature and mature populations at time tt, α\alpha is the birth rate, β\beta is the internal competition coefficient of mature populations, and γ\gamma is the mortality rate of immature populations, the SDTD τ(z(t))\tau(z(t)) is taken to be an increasing differentiable bounded function of the total population z(t)=xi(t)+xm(t)z(t)=x_{i}(t)+x_{m}(t). An attracting region is determined for solutions, which collapses to the unique positive equilibrium in the state-independent case.

In 2005, based on the model (1), Al-Omari et al. 13 studied the following stage-dependent population model with a SDTD:

xi(t)\displaystyle x_{i}^{\prime}(t) =\displaystyle= R(xm(t))γxi(t)R(xm(tτ(z(t))))eγτ(z(t)),\displaystyle R(x_{m}(t))-\gamma x_{i}(t)-R(x_{m}(t-\tau(z(t))))e^{-\gamma\tau(z(t))},
xm(t)\displaystyle x_{m}^{\prime}(t) =\displaystyle= R(xm(tτ(z(t))))eγτ(z(t))βxm(t).\displaystyle R(x_{m}(t-\tau(z(t))))e^{-\gamma\tau(z(t))}-\beta x_{m}(t).

where the immature birth rate R(y(t))R(y(t)) is taken as a general function of the present mature population and the death rate for the mature one is a constant. They provided the sufficient conditions for the global stability of the extinction equilibrium and existence of periodic solutions. Late, Magpantay et al MagpantayWu gave an age-structured single-species population model that accounts for complex life cycles and competition for resources limiting the transition to maturity, shown an interesting numerical scheme and simulations to integrate the equations with significant applications. Late, by modeling the state-dependent delay as maturity period in the juvenile zooplankton population, Kloosterman et al. studied a closed nutrient-phytoplankton-zooplankton model that includes size structure in the juvenile zooplankton. In 2017, Lv et al. 16 studied an SDTD competitive model, where the SDTD is taken to be an increasing differentiable bounded function of the number of its own population, and completely analyzed the global stability of the equilibria by using the comparison principle and iterative method. For other non-predator-prey biological models with SDTD, we refer to Rezounenko2012 ; Rezounenko2017 ; Hu2014 ; Hu2016 ; LiGuo2017 and the references therein.

On the other hand, as one of the central goal for ecologists, delayed predator-prey interaction has attracted many attentions. Following pioneering works in age-structured predator-prey models by May May1975 , Hastings Hastings1983 ; Hastings1984 , Murdoch et al., Murdoch1987 and the stage structured predator-prey models with constant delay in 6 ; 7 ; 8 ; 9 , there are some recent works relate to the SDTD predator-prey model. For example, in 2015, based on the 7 constant delay model, Al-Omari 14 established and analyzed the following predator-prey model with state-dependent delay:

x(t)\displaystyle x^{\prime}(t) =\displaystyle= rx(t)(1x(t)K)ax(t)z(t)h1x(t),\displaystyle rx(t)(1-\frac{x(t)}{K})-ax(t)z(t)-h_{1}x(t),
y(t)\displaystyle y^{\prime}(t) =\displaystyle= bx(t)z(t)bx(tτ(u(t)))z(tτ(u(t)))e(γ+h2)τ(u(t))γy(t)h2y(t),\displaystyle bx(t)z(t)-bx\big{(}t-\tau(u(t))\big{)}z\big{(}t-\tau(u(t))\big{)}e^{-(\gamma+h_{2})\tau(u(t))}-\gamma y(t)-h_{2}y(t),
z(t)\displaystyle z^{\prime}(t) =\displaystyle= bx(tτ(u(t)))z(tτ(u(t)))e(γ+h2)τ(u(t))dz(t)h3z(t).\displaystyle bx\big{(}t-\tau(u(t))\big{)}z\big{(}t-\tau(u(t))\big{)}e^{-(\gamma+h_{2})\tau(u(t))}-dz(t)-h_{3}z(t).

where x(t),y(t),z(t)x(t),y(t),z(t) represent the number of prey, juvenile predator and adult predator, respectively, b,d,ab,d,a are the adult predator’s birth rate, mortality and capture rate, respectively, h1, h2, h3 are the prey, juvenile predator and adult predator capture rate, γ\gamma is the juvenile predator mortality rate, u(t)=y(t)+z(t)u(t)=y(t)+z(t), the state dependent delay τ(u(t))\tau(u(t)) is a function of the total number of populations, investigated the global stability of trivial and the boundary equilibria by using Liapunov functional and LaSalle invariant principle.

In 2018, Lv et al. 15 proposed and studied a predator-prey model with SDTD where the prey population is assumed to have an age structure:

xi(t)=αxm(t)γxi(t)αeγτ(t)xm(tτ(t))(1τ(t)),xm(t)=αeγτ(t)xm(tτ(t))βxm2(t)μxm(t)y(t)xm(t)+hy(t),y(t)=μ1xm(t)y(t)xm(t)+hy(t)δy(t),1τ(t)=k(z(t))k(z(tτ(t))).\begin{split}x_{i}^{\prime}(t)=&\alpha x_{m}(t)-\gamma x_{i}(t)-\alpha e^{-\gamma\tau(t)}x_{m}(t-\tau(t))(1-\tau^{\prime}(t)),\\ x_{m}^{\prime}(t)=&\alpha e^{-\gamma\tau(t)}x_{m}(t-\tau(t))-\beta x^{2}_{m}(t)-\frac{\mu x_{m}(t)y(t)}{x_{m}(t)+hy(t)},\\ y^{\prime}(t)=&\frac{\mu_{1}x_{m}(t)y(t)}{x_{m}(t)+hy(t)}-\delta y(t),\\ 1-\tau^{\prime}(t)=&\frac{k(z(t))}{k\big{(}z(t-\tau(t))\big{)}}.\end{split}

where, xi(t),xm(t)x_{i}(t),x_{m}(t) and y(t)y(t) are the number of juvenile prey, adult feast and predator at time tt, α\alpha and γ\gamma are the birth rate and mortality of juvenile and adult bait respectively, β\beta is the intraspecific competition coefficient of adult prey, δ\delta is predator mortality rate, z(t)=xi(t)+xm(t)z(t)=x_{i}(t)+x_{m}(t) is the total number of prey. For the global dynamics of the system, they discuss an attracting region which is determined by solutions, and the region collapses to the interior equilibrium in the constant delay case.

In 2018, Wang et al. 11 established and analyzed a state-dependent time-delay model with the delayed time-derivative term:

xi(t)\displaystyle x_{i}^{\prime}(t) =\displaystyle= αxm(t)γxi(t)α(1τ(z(t))z(t))eγτ(z(t))xm(tτ(z(t))),\displaystyle\alpha x_{m}(t)-\gamma x_{i}(t)-\alpha(1-\tau^{\prime}(z(t))z^{\prime}(t))e^{-\gamma\tau(z(t))}x_{m}(t-\tau(z(t))),
xm(t)\displaystyle x_{m}^{\prime}(t) =\displaystyle= α(1τ(z(t))z(t))eγτ(z(t))xm(tτ(z(t)))βxm2(t).\displaystyle\alpha(1-\tau^{\prime}(z(t))z^{\prime}(t))e^{-\gamma\tau(z(t))}x_{m}(t-\tau(z(t)))-\beta x_{m}^{2}(t).

This model is clearly different from the previous models in the sense that it includes the correction term 1τ(z(t))z(t)1-\tau^{\prime}(z(t))z^{\prime}(t) in the maturity rate. Permanence of the system was analyzed, and explicit bounds for the eventual behaviors of the immature and mature populations are established in 11 , while the global stability was not discussed.

In this paper, following the research method of Wang et al. 11 , based on the age structure model, we consider a SDTD predator-prey model where the predators was divided into immature and mature, mature delay is a monotonically increasing continuous differentiable bounded function that depends on the number of adult predators. In other words, if the number of adult predators is large, the maturity period is longer, and the population size is reduced as slow growth of adult predators. Our aim is to conduct a qualitative analysis of the model to study the persistence, extinction and stability of the population.

Our paper is organized as follows. In Sect. 2, we establish a state-dependent time-delay predator-prey model with the delay time-derivative term. In Sect. 3, we discuss the positivity and boundedness of solutions. In Sect. 4, we give the necessary and sufficient conditions for the permanence and necessity of the population. In Sect. 5, we discuss the linearized stability of all equilibria of the model. In Sect. 6, we prove the global attractiveness of positive equilibrium of the model. The summary and discussion are presented in Sect. 7.

2 Model Derivation

Motivated by the ideas in 7 ; 9 ; MagpantayWu ; 11 , we consider the growth of predator through immature and mature two stages. In order to distinguish immature individuals, yj(t)y_{j}(t), from mature ones, y(t)y(t), we introduce a threshold age τ(y(t))\tau(y(t)), which is the maturation time for an immature individual that matures at time tt depending on the number of mature predator, y(t)y(t). Let ρ(t,a)\rho(t,a) be the density of predator of age aa at time tt. Then the number of immature, yj(t)y_{j}(t), and mature y(t)y(t) is given by

yj(t)=0τ(y(t))ρ(t,a)𝑑aandy(t)=τ(y(t))ρ(t,a)𝑑a,y_{j}(t)=\int_{0}^{\tau(y(t))}\rho(t,a)da~{}~{}~{}~{}\mbox{and}~{}~{}~{}~{}y(t)=\int_{\tau(y(t))}^{\infty}\rho(t,a)da,

respectively. The dynamics of predator with age structure can be represented (see 12 ; 18 ) by the following partial differential equations:

ρ(t,a)t+ρ(t,a)a=djρ(t,a),ifaτ(y(t)),ρ(t,a)t+ρ(t,a)a=dρ(t,a),ifa>τ(y(t)).\begin{split}\frac{\partial\rho(t,a)}{\partial t}+\frac{\partial\rho(t,a)}{\partial a}&=-d_{j}\rho(t,a),~{}~{}~{}if~{}a\leq\tau(y(t)),\\ \frac{\partial\rho(t,a)}{\partial t}+\frac{\partial\rho(t,a)}{\partial a}&=-d\rho(t,a),~{}~{}~{}if~{}a>\tau(y(t)).\end{split} (1)

where each individual from yj(t)y_{j}(t) dies at a constant rate djd_{j} and that from y(t)y(t) at a constant rate dd.

Taking the derivatives of yj(t)y_{j}(t) and y(t)y(t), respectively, and combining with (1), we get

dyj(t)dt\displaystyle\frac{dy_{j}(t)}{dt} =\displaystyle= ρ(t,0)djyj(t)[1τ(y(t))y(t)]ρ(t,τ(y(t))),\displaystyle\rho(t,0)-d_{j}y_{j}(t)-[1-\tau^{\prime}(y(t))y^{\prime}(t)]\rho(t,\tau(y(t))),
dy(t)dt\displaystyle\frac{dy(t)}{dt} =\displaystyle= [1τ(y(t))y(t)]ρ(t,τ(y(t)))ρ(t,)dy(t).\displaystyle[1-\tau^{\prime}(y(t))y^{\prime}(t)]\rho(t,\tau(y(t)))-\rho(t,\infty)-dy(t).

It is necessary to note that a prime refers to differentiation with respect to yy, and a dot indicates differentiation with respect to time tt, namely, dτ(y(t))/dt=τ(y)y(t).d\tau(y(t))/dt=\tau^{\prime}(y)\cdot y^{\prime}(t).

Taking ρ(t,)\rho(t,\infty) as zero since no one can live forever. We assume that the birth rate of predator is nn and its functional response function is f(x(t),y(t))f(x(t),y(t)), where x(t)x(t) is the total number of prey, so the term ρ(t,0)=nf(x(t),y(t))y(t)\rho(t,0)=nf(x(t),y(t))y(t). Therefore, for tτM=max{τ(y(t))}t\geq{\tau_{M}}=max\{\tau(y(t))\} we obtain

ρ(t,τ(y(t)))=ρ(tτ(y(t)),0)edjτ(y)=nf(x(tτ(y),y(tτ(y)))y(tτ(y))edjτ(y).\rho(t,\tau(y(t)))=\rho(t-\tau(y(t)),0)e^{-d_{j}\tau(y)}=nf(x(t-\tau(y),y(t-\tau(y)))y(t-\tau(y))e^{-d_{j}\tau(y)}.

Further we assume that the growth of prey obays logistic growth, then we have the following predator-prey model with a state-dependent maturation delay for the predator:

x(t)=rx(t)(1x(t)K)f(x(t),y(t))y(t),yj(t)=n(1τ(y)y(t))edjτ(y)f(x(tτ(y),y(tτ(y)))y(tτ(y))+nf(x(t),y(t))y(t)djyj(t),y(t)=n(1τ(y)y(t))edjτ(y)f(x(tτ(y),y(tτ(y)))y(tτ(y))dy(t),\begin{split}x^{\prime}(t)=&rx(t)(1-\frac{x(t)}{K})-f(x(t),y(t))y(t),\\ y_{j}^{\prime}(t)=&-n(1-\tau^{\prime}(y)y^{\prime}(t))e^{-d_{j}\tau(y)}f(x(t-\tau(y),y(t-\tau(y)))y(t-\tau(y))\\ &+nf(x(t),y(t))y(t)-d_{j}y_{j}(t),\\ y^{\prime}(t)=&n(1-\tau^{\prime}(y)y^{\prime}(t))e^{-d_{j}\tau(y)}f(x(t-\tau(y),y(t-\tau(y)))y(t-\tau(y))-dy(t),\end{split} (2)

where rr and KK represent the specific growth rate of the prey and environmental carrying capacity, respectively.

From a biological point of view, it is natural that tτ(y(t))t-\tau(y(t)) is a strictly increasing function of tt, i.e., the possibility of mature individuals becoming immature only by birth. Assume that q(s)q(s) is the developmental proportion at time ss, when an immature individual moves to the mature state from tτ(y(t))t-\tau(y(t)) to tt, the cumulative rate of development qq should be equal to one, namely,

tτ(y(t))tq(s)𝑑s=1.\int_{t-\tau(y(t))}^{t}q(s)ds=1.

Taking the derivative with respect to tt, we have

1τ(y(t))y(t)=q(t)q(tτ(y(t)))>0,1-\tau^{\prime}(y(t))y^{\prime}(t)=\frac{q(t)}{q\big{(}t-\tau(y(t))\big{)}}>0,

implying that tτ(y(t))t-\tau(y(t)) is a strictly increasing function of tt and the variation of maturity time delay τ(y(t))\tau(y(t)) is bounded by one.

Furthermore, from the biological point of view, we give the following hypotheses for the model (2):

  • (A1)(A_{1})

    Parameters r,K,n,dj,dr,K,n,d_{j},d are all positive constants;

  • (A2)(A_{2})

    The state-dependent maturity time delay τ(y)\tau(y) is an increasing continuously differentiable bounded function of the mature predator population y(t)y(t), i.e., τ(y)0\tau^{\prime}(y)\geq 0, and 0τmτ(y)τM<+0\leq\tau_{m}\leq\tau(y)\leq\tau_{M}<+\infty with τ(+)=τM,τ(0)=τm.\tau(+\infty)=\tau_{M},\tau(0)=\tau_{m}.

  • (A3)(A_{3})

    The functional response function f(x,y)f(x,y) satisfies the following conditions:

  • (i)

    f(x,y)f(x,y) is continuously differentiable; f(x,y)>0f(x,y)>0 for x,y(0,)\forall\ x,y\in(0,\infty), and f(x,y)=0f(x,y)=0 if and only if x=0x=0; |fx(0,0)|<+|f_{x}^{\prime}(0,0)|<+\infty, where fx(0,0)f_{x}^{\prime}(0,0) denotes fx(x,y)|x=0,y=0\frac{\partial f}{\partial x}(x,y)\big{|}_{x=0,y=0}.

  • (ii)

    f(x,y)f(x,y) is increasing of xx and decreasing of yy.

In fact, in the literature, most of the commonly used functional response functions f(x,y)f(x,y) satisfies the condition (A3)(A_{3}). We can summarize the forms in the following:

  • (1)

    Linear type 11 : f(x,y)=bxf(x,y)=bx;

  • (2)

    Nonlinear type 29 : f(x,y)=bxk,k>1f(x,y)=bx^{k},~{}~{}k>1;

  • (3)

    Holling type I (Blackman type): 27 f(x,y)={axxbabx>b.f(x,y)=\left\{\begin{array}[]{rcl}ax&&{x\leq b}\\ ab&&{x>b.}\\ \end{array}\right.

  • (4)

    Holling type II (Michaelis-Menten type) 27 : f(x,y)=bx1+bhxf(x,y)=\frac{bx}{1+bhx};

  • (5)

    Holling type III 27 : f(x,y)=bx21+bhx2f(x,y)=\frac{bx^{2}}{1+bhx^{2}};

  • (6)

    Saturation type 28 f(x,y)=bxk1+bhxk,k>1f(x,y)=\frac{bx^{k}}{1+bhx^{k}},\ \ k>1;

  • (7)

    Ivlev type 24 : f(x,y)=b(1ecx)f(x,y)=b(1-e^{-cx});

  • (8)

    Beddington-DeAngelis type 26 ; DA : f(x,y)=bx1+k1x+k2yf(x,y)=\frac{bx}{1+k_{1}x+k_{2}y};

  • (9)

    Crowley-Martin type 25 : f(x,y)=bx1+k1x+k2y+k1k2xyf(x,y)=\frac{bx}{1+k_{1}x+k_{2}y+k_{1}k_{2}xy}.

3 Positivity and Boundedness

In this section, we shall address the positivity and boundedness of the solution of system (2). From the standpoint of biology, positivity means that the species persists, i.e., the populations can not be extincted. Boundedness may be viewed as a natural restriction to growth as a result of limited resources in an closed environment.

Denote C:=C([τM,0),R3)C:=C([-\tau_{M},0),R^{3}). For ϕ=(ϕ1,ϕ2,ϕ3)C\phi=(\phi_{1},\phi_{2},\phi_{3})\in C, define ϕ=i=13ϕi\Arrowvert\phi\Arrowvert=\sum_{i=1}^{3}\Arrowvert\phi_{i}\Arrowvert_{\infty}, where

ϕi=maxθ[τM,0]|ϕi(θ)|.\Arrowvert\phi_{i}\Arrowvert_{\infty}=max_{\theta\in[-\tau_{M},0]}|\phi_{i}(\theta)|.

Then CC is a Banach space and C+={ϕC:ϕi(θ)0,i{1,2,3},θ[τM,0]}C_{+}=\{\phi\in C:\phi_{i}(\theta)\geq 0,i\in\{1,2,3\},\theta\in[-\tau_{M},0]\} is a normal cone of CC with nonempty interior in CC. The initial conditions for the system (2) are

x(t)=ϕ1(t)0,yj(t)=ϕ2(t)0,y(t)=ϕ3(t)0for allτMt0,x(t)=\phi_{1}(t)\geq 0,\ \ y_{j}(t)=\phi_{2}(t)\geq 0,\ \ y(t)=\phi_{3}(t)\geq 0~{}~{}~{}\mbox{for~{}all}-\tau_{M}\leq t\leq 0, (3)

with positive ϕ1(0),ϕ2(0),ϕ3(0)\phi_{1}(0),\phi_{2}(0),\phi_{3}(0) and

ϕ2(0)=τ(y(0))0f(x(s),y(s))y(s)edjs𝑑s\phi_{2}(0)=\int_{-\tau(y(0))}^{0}f\big{(}x(s),y(s)\big{)}y(s)e^{d_{j}s}ds

presenting the size of the immature population surviving to time t=0t=0, where τ(y(0))\tau(y(0)) is the maturation time at t=0t=0.

The following theorem demonstrates that the solutions of (2) are positive and bounded.

Theorem 3.1

Assume that the initial condition (3) holds,then every solution
(x(t),y(t),yj(t))(x(t),y(t),y_{j}(t)) of system (2) is positive for all t>0t>0.

Proof

Clearly, x(t)>0x(t)>0 for all t>0t>0 (why? seems it’s not obvious, better to give some reason).

Now to prove the positivity of y(t)y(t) as a solution of (2). Assume that,there exists t>0t^{*}>0, such that y(t)=0y(t^{*})=0 and y(t)>0y(t)>0 for every t(0,t)t\in(0,t^{*}). Then by the initial conditions (3) we have y(t)dyt[τM,t].y^{\prime}(t)\geq-dy~{}~{}~{}\forall t\in[-\tau_{M},t^{*}]. Thus we get y(t)y0edt>0y(t)\geq y_{0}e^{-dt}>0 for t[τM,t]\forall t\in[-\tau_{M},t^{*}]. Let t=tt=t^{*}, then we have 0=y(t)y0edt>00=y(t^{*})\geq y_{0}e^{-dt^{*}}>0, which is a contradiction. So no such tt^{*} exists, and we obtain the results.

Next we prove the positivity of yj(t)y_{j}(t) as a solution of (2).

Integrating the third equation of the system (2), we have

yj(t)=edjt[ϕ3(0)+0tedjsnf(x(s),y(s))y(s)𝑑sτ(y(0))tτ(y(t))nf(x(s),y(s))y(s)edjs𝑑s]=tτ(y)tnf(x(s),y(s))y(s)edj(ts)𝑑s.\begin{split}y_{j}(t)=&e^{-d_{j}t}\bigg{[}\phi_{3}(0)+\int_{0}^{t}e^{d_{j}s}nf(x(s),y(s))y(s)ds-\int_{-\tau(y(0))}^{t-\tau(y(t))}nf(x(s),y(s))y(s)e^{d_{j}s}ds\bigg{]}\\ =&\int_{t-\tau(y)}^{t}nf(x(s),y(s))y(s)e^{d_{j}(t-s)}ds.\end{split} (4)

By the positivity of x(t),y(t),τ(t)x(t),y(t),\tau(t), we have yj(t)>0,t>0y_{j}(t)>0,\ \forall t>0. The proof is complete.

In proving our main results, a comparison principle will be used. As we know, the comparison principles do not always hold for SDTD equations, which depends very much on how the delay term appears in the equations. The comparison principle of SDTD equations without delay time-derivative term is discussed in 16 . Now we extend the result to the system with delay time-derivative term.

Lemma 1

Let y1(t)y_{1}(t) be the solution of

y1(t)=[1τ(y1)y1(t)]edjτ(y1)f(tτ(y1))dy1(t),t>0y_{1}^{\prime}(t)=\big{[}1-\tau^{\prime}(y_{1})y_{1}^{\prime}(t)\big{]}e^{-d_{j}\tau(y_{1})}f(t-\tau(y_{1}))-dy_{1}(t),~{}~{}t>0

and y2(t)y_{2}(t) be a function satisfying

y2(t)[1τ(y2)y2(t)]edjτ(y2)f(tτ(y2))dy2(t),t>0y_{2}^{\prime}(t)\leq\big{[}1-\tau^{\prime}(y_{2})y_{2}^{\prime}(t)\big{]}e^{-d_{j}\tau(y_{2})}f(t-\tau(y_{2}))-dy_{2}(t),~{}~{}t>0 (5)

and y2(s)y1(s)y_{2}(s)\leq y_{1}(s) for all s[τM,0]s\in[-\tau_{M},0], where f()f(\cdot) is a continuous function. Then y2(t)y1(t)y_{2}(t)\leq y_{1}(t) holds true for all t>0t>0.

Proof

I think the lemma has problem, and the proof

With the assumption y2(s)<y1(s)y_{2}(s)<y_{1}(s) for all s[τM,0]s\in[-\tau_{M},0], first we claim that y2(t)<y1(t)y_{2}(t)<y_{1}(t) for all t>0t>0. If this is false, there must exist some t0>0t_{0}>0 such that y2(t)<y1(t)y_{2}(t)<y_{1}(t), t[τM,t0)t\in[-\tau_{M},t_{0}) and y2(t0)=y1(t0)y_{2}(t_{0})=y_{1}(t_{0}). It follows that y2(t0)y1(t0)y_{2}^{\prime}(t_{0})\geq y_{1}^{\prime}(t_{0}). But

y1(t0)=[1τ(y1(t0))y1(t0)]edjτ(y1(t0))f(tτ(y1(t0)))dy1(t0)[1τ(y2(t0))y2(t0)]edjτ(y2(t0))f(tτ(y2(t0)))dy2(t0)y2(t0)\begin{split}y_{1}^{\prime}(t_{0})&=\big{[}1-\tau^{\prime}(y_{1}(t_{0}))y_{1}^{\prime}(t_{0})\big{]}e^{-d_{j}\tau(y_{1}(t_{0}))}f(t-\tau(y_{1}(t_{0})))-dy_{1}(t_{0})\\ &\geq\big{[}1-\tau^{\prime}(y_{2}(t_{0}))y_{2}^{\prime}(t_{0})\big{]}e^{-d_{j}\tau(y_{2}(t_{0}))}f(t-\tau(y_{2}(t_{0})))-dy_{2}(t_{0})\\ &\geq y_{2}^{\prime}(t_{0})\end{split}

For the general case, let ε>0\varepsilon>0 and yε(t)y_{\varepsilon}(t) be the solution of

y1(t)=[1τ(y1)y1(t)]edjτ(y1)f(tτ(y1))dy1(t)+εy_{1}^{\prime}(t)=\big{[}1-\tau^{\prime}(y_{1})y_{1}^{\prime}(t)\big{]}e^{-d_{j}\tau(y_{1})}f(t-\tau(y_{1}))-dy_{1}(t)+\varepsilon

corresponding to initial data yε(t)=y1(t)+εy_{\varepsilon}(t)=y_{1}(t)+\varepsilon, t[τM,0)t\in[-\tau_{M},0). Following the previous result, we can conclude that yε(t)>y2(t)y_{\varepsilon}(t)>y_{2}(t) for all t>0t>0 for which yε(t)y_{\varepsilon}(t) is defined. It can be shown that for sufficiently small ε>0\varepsilon>0, the solution yε(t)y1(t)y_{\varepsilon}(t)\to y_{1}(t) as ε0\varepsilon\to 0 for all tτMt\geq-\tau_{M}. Consequently, y1(t)=limε0yε(t)y2(t).y_{1}(t)=\lim_{\varepsilon\to 0}y_{\varepsilon}(t)\geq y_{2}(t).

For the comparison principles, the other reversed inequality follows analogously, and will be used later. Not sure what you want to claim for this part: Furthermore, an differential inequality of the form (5), which holds only for t above some value, say t1t_{1}, and not for all t>0t>0, will be often used in applications of these comparison results. That is the initial time is simply thought of as t1t_{1} rather than 0, and y2(t)y1(t)y_{2}(t)\leq y_{1}(t) is arranged to hold for tt1t\leq t_{1} by appropriate definition of y1(t)y_{1}(t) for values of tt1t\leq t_{1}. In the interests of clarity, this latter case in detail will not be always explained.

In the next theorem we give boundedness results.

Theorem 3.2

Assume that the initial condition (3) holds, then every solution (x(t),y(t),yj(t))(x(t),y(t),y_{j}(t)) of system (2)(2) is eventually uniformly bounded.

Proof

Define a function as follows:

V(t)=nx(t)+y(t)+yj(t).V(t)=nx(t)+y(t)+y_{j}(t).

Calculating the time derivative of V(t)V(t) along the solutions of system (2), we obtain

V(t)=nrx(t)(1x(t)K)dydjyj=min{dj,d}V+n(min{dj,d}+r)xnrKx2+(min{dj,d}max{dj,d})ymin{dj,d}V+M.\begin{split}V^{\prime}(t)&=nrx(t)(1-\frac{x(t)}{K})-dy-d_{j}y_{j}\\ &=-\min\{d_{j},d\}V+n(\min\{d_{j},d\}+r)x-\frac{nr}{K}x^{2}+(\min\{d_{j},d\}-\max\{d_{j},d\})y\\ &\leq-\min\{d_{j},d\}V+M.\end{split}

where M>0M>0 is the maximum of the quadratic function n(min{dj,d}+r)xnrKx2n(\min\{d_{j},d\}+r)x-\frac{nr}{K}x^{2}. Therefore, lim suptV(t)M/min{dj,d}\limsup_{t\rightarrow\infty}V(t)\leq M/min\{d_{j},d\} and the solution of system (2) is eventually uniformly bounded.

In particular, according to the first equation of system (2), we have

x(t)rx(t)(1x(t)K)x^{\prime}(t)\leq rx(t)\big{(}1-\frac{x(t)}{K}\big{)}

By comparison principleCP , we can get

lim suptx(t)K\limsup_{t\to\infty}x(t)\leq K

The proof is complete.

Remark 1

Due to the existence of the delay time-derivative term 1τ(y(t))y(t)1-\tau^{\prime}(y(t))y^{\prime}(t), the method of proving boundedness in 6 is not applicable. Here, the method of constructing the VV function is used to subtly prove the eventually boundedness of x(t),y(t),yj(t)x(t),y(t),y_{j}(t).

4 Permanence and Extinction

It is clear that system (2) has a trivial equilibrium E0(0,0,0)E_{0}(0,0,0) and a predator extinction equilibrium E1(K,0,0)E_{1}(K,0,0). Mathematically, permanence is equivalent to the existence of positive equilibriumE(x,y,yj)E^{*}(x^{*},y^{*},y_{j}^{*}). The following results give the necessary and sufficient conditions for the permanence/extinction of the system (2).

Theorem 4.1

System (2) is permanent if and only the nedjτ(0)f(K,0)>dne^{-d_{j}\tau(0)}f(K,0)>d holds.

To prove this theorem, we engage the persistence theory by Hale and Waltmann 19 for infinite dimensional systems (see 20 as well), in the following:

Consider a metric space X with metric d. T is a continuous semiflow on X,i.e., a continuous mapping T:[0,)×XXT:[0,\infty)\times X\to X with the following properties:

TtTs=Tt+s,t,s0;T0(x)=x,xX.T_{t}\circ T_{s}=T_{t+s},~{}~{}~{}~{}t,s\geq 0;~{}~{}~{}~{}~{}T_{0}(x)=x,~{}~{}x\in X.

Here TtT_{t} denotes the mapping from X to X given by Tt(x)=T(t,X)T_{t}(x)=T(t,X). The distance d(x,Y)d(x,Y) of a point xXx\in X from a subset Y of X is defined by

d(x,Y)=infyYd(x,y).d(x,Y)=\inf_{y\in Y}d(x,y).

Recall that the positive orbit γ+(x)\gamma^{+}(x) through xx is defined as γ+(x)=t0{T(t)x}\gamma^{+}(x)=\cup_{t\geq 0}\{T(t)x\}, and its ω\omega-limit set ω(x)=τ0CLtτ{T(t)x}\omega(x)=\cap_{\tau\geq 0}CL\cup_{t\geq\tau}\{T(t)x\}, where CL means closure. Define Ws(A)W^{s}(A) the stable set of an compact invariant set A as

Ws(A)={x:xX,ω(x)ϕ,ω(x)A};W^{s}(A)=\{x:x\in X,\omega(x)\neq\phi,\omega(x)\subset A\};

define A~\widetilde{A_{\partial}} the particular invariant set of interest as

A~=xAω(x).\widetilde{A_{\partial}}=\bigcup_{x\in A}\omega(x).
Lemma 2

(19 ) Suppose T(t) satisfies (H1)(H_{1}):

(H1)(H_{1}). Assume X is the closure of open set X0X^{0}; X0\partial X^{0} is nonempty and is the boundary of X0X^{0}; and the C0C^{0}-semigroup T(t)T(t) on X satisfies

T(t):X0×X0,T(t):X0X0.T(t):X^{0}\times X^{0},~{}~{}~{}~{}~{}T(t):\partial X^{0}\to\partial X^{0}.

and

  • (i)

    there is a t00t_{0}\geq 0 such that T(t)T(t) is compact for t>t0t>t_{0};

  • (ii)

    T(t) is point dissipative in X;

  • (iii)

    A~\widetilde{A_{\partial}} is isolated and has an acyclic covering M.

Then T(t)T(t) is uniformly persistent iff for each MiMM_{i}\in M, Ws(Mi)X0=W^{s}(M_{i})\cap X^{0}=\emptyset.

To prove the theorem (4.1), first we have the following claims.

Claim

A: If nedjτ(0)f(K,0)>dne^{-d_{j}\tau(0)}f(K,0)>d, then the system (2) is permanent.

Proof

We begin by showing the claim holds true for the system (6),

x(t)=rx(t)(1x(t)K)f(x(t),y(t))y(t),y(t)=n(1τ(y)y(t))edjτ(y)f[x(tτ(y),y(tτ(y))]y(tτ(y))dy(t),\begin{split}x^{\prime}(t)&=rx(t)(1-\frac{x(t)}{K})-f(x(t),y(t))y(t),\\ y^{\prime}(t)&=n(1-\tau^{\prime}(y)y^{\prime}(t))e^{-d_{j}\tau(y)}f[x(t-\tau(y),y(t-\tau(y))]y(t-\tau(y))-dy(t),\end{split} (6)

which is a subsystem of system (2). As the first step, we verify that the boundary plan R+2={(x,y):x0,y0}R_{+}^{2}=\{(x,y):x\geq 0,y\geq 0\} repel the positive solutions to systems (6) uniformly.

Let C+([τM,0],R+2)C^{+}([-\tau_{M},0],R_{+}^{2}) denote the space of continuous functions mapping [τM,0][-\tau_{M},0] into R+2R_{+}^{2}. We choose

C1={(ϕ1,ϕ2)C+([τM,0],R+2):ϕ1(θ)0,ϕ2(θ)0,θ[τM,0]},C2={(ϕ1,ϕ2)C+([τM,0],R+2):ϕ1(θ)>0,ϕ2(θ)0,θ[τM,0]}.\begin{split}C_{1}=\{(\phi_{1},\phi_{2})\in C^{+}([-\tau_{M},0],R_{+}^{2}):\phi_{1}(\theta)\equiv 0,\phi_{2}(\theta)\geq 0,\theta\in[-\tau_{M},0]\},\\ C_{2}=\{(\phi_{1},\phi_{2})\in C^{+}([-\tau_{M},0],R_{+}^{2}):\phi_{1}(\theta)>0,\phi_{2}(\theta)\equiv 0,\theta\in[-\tau_{M},0]\}.\end{split}

Denote C=C1C2C=C_{1}\cup C_{2}, X=C+([τM,0],R+2)X=C^{+}([-\tau_{M},0],R_{+}^{2}), and X0=IntC+([τM,0],R+2)X_{0}=IntC^{+}([-\tau_{M},0],R_{+}^{2}); then C=X0C=\partial X^{0}. It is easy to see that the system (2) possesses two constant solutions in C=X0C=\partial X^{0}: E0~C1,E1~C2\widetilde{E_{0}}\in C_{1},\widetilde{E_{1}}\in C_{2} with

E0~={(ϕ1,ϕ2)C+([τM,0],R+2):ϕ1(θ)=ϕ2(θ)0,θ[τM,0]},E1~={(ϕ1,ϕ2)C+([τM,0],R+2):ϕ1(θ)K,ϕ2(θ)0,θ[τM,0]}.\begin{split}\widetilde{E_{0}}=&\{(\phi_{1},\phi_{2})\in C^{+}([-\tau_{M},0],R_{+}^{2}):\phi_{1}(\theta)=\phi_{2}(\theta)\equiv 0,\theta\in[-\tau_{M},0]\},\\ \widetilde{E_{1}}=&\{(\phi_{1},\phi_{2})\in C^{+}([-\tau_{M},0],R_{+}^{2}):\phi_{1}(\theta)\equiv K,\phi_{2}(\theta)\equiv 0,\theta\in[-\tau_{M},0]\}.\end{split}

We verify below that the conditions of Lemma 22 are satisfied. By the definition of X0X^{0} and X0\partial X^{0}, it is easy to see that condition (i) and (ii) of Lemma 22 are satisfied for the system (6)and that X0X^{0} and X0\partial X^{0} are invariant. Hence (H1)(H_{1}) is satisfied.

Consider condition (iii) of Lemma 22. We have

x(t)|(ϕ1,ϕ2)C10,x^{\prime}(t)|_{(\phi_{1},\phi_{2})\in C_{1}}\equiv 0,

thus x(t)|(ϕ1,ϕ2)C10x(t)|_{(\phi_{1},\phi_{2})\in C_{1}}\equiv 0 for all t0t\geq 0. Consequently we have y(t)|(ϕ1,ϕ2)C1=dy(t)0,y^{\prime}(t)|_{(\phi_{1},\phi_{2})\in C_{1}}=-dy(t)\leq 0, implying all the points in C1C_{1} approach E0~,\widetilde{E_{0}}, i.e., C1=Ws(E0~).C_{1}=W^{s}(\widetilde{E_{0}}). Similarly we have C2=Ws(E1~).C_{2}=W^{s}(\widetilde{E_{1}}). Hence A~=E0~E1~\widetilde{A_{\partial}}=\widetilde{E_{0}}\cup\widetilde{E_{1}} and clearly it is isolated. Noting that C1C2=C_{1}\cap C_{2}=\emptyset, it follows from these structural features that the flow in A~\widetilde{A_{\partial}} is acyclic, satisfying condition (iii) in Lemma 22.

Now we show that Ws(Ei~)X0=,(i=0,1).W^{s}(\widetilde{E_{i}})\cap X^{0}=\emptyset,(i=0,1). By Theorem (3.1), we have x(t),y(t)>0x(t),y(t)>0 for all t>0t>0. Assume Ws(E0~)X0W^{s}(\widetilde{E_{0}})\cap X^{0}\neq\emptyset, i.e., there exists a positive solution (x(t),y(t))(x(t),y(t)) with limt(x(t),y(t))=(0,0)\lim_{t\rightarrow\infty}(x(t),y(t))=(0,0), by (A3)(A3)-(i), we have limx0,y0f(x(t),y(t))x(t)=fx(0,0)<+\lim_{x\to 0,y\to 0}\frac{f(x(t),y(t))}{x(t)}=f^{\prime}_{x}(0,0)<+\infty. Thus limx0,y0f(x(t),y(t))y(t)x(t)=0\lim_{x\to 0,y\to 0}\frac{f(x(t),y(t))y(t)}{x(t)}=0. Then from the first equation of (6), we have

d(lnx(t))dt=r(1x(t)K)f(x(t),y(t))y(t)x(t)>r2\frac{d(lnx(t))}{dt}=r(1-\frac{x(t)}{K})-\frac{f(x(t),y(t))y(t)}{x(t)}>\frac{r}{2}

for all sufficiently large tt. Hence we have limtx(t)=+,\lim_{t\rightarrow\infty}x(t)=+\infty, contradicting limtx(t)=0\lim_{t\rightarrow\infty}x(t)=0; this proves Ws(E0~)X0=W^{s}(\widetilde{E_{0}})\cap X^{0}=\emptyset.

Now we verify Ws(E1~)X0=W^{s}(\widetilde{E_{1}})\cap X^{0}=\emptyset. If it is not the case, i.e., Ws(E1~)X0W^{s}(\widetilde{E_{1}})\cap X^{0}\neq\emptyset. Then there exists a positive solution (x(t),y(t))(x(t),y(t)) to system (6) with limt(x(t),y(t))=(K,0)\lim_{t\rightarrow\infty}(x(t),y(t))=(K,0), and for sufficiently small positive constant ε\varepsilon with nedjτ(ε)f(Kε,ε)>dne^{-d_{j}\tau(\varepsilon)}f(K-\varepsilon,\varepsilon)>d, there exists a positive constant T=T(ε)T=T(\varepsilon) such that

x(t)>Kε>0,y(t)<εfor  alltT.x(t)>K-\varepsilon>0,y(t)<\varepsilon~{}~{}~{}\mbox{for~{}~{}all}~{}~{}t\geq T.

Consider the function

V=y(t)+dtτ(y)ty(s)𝑑s.V=y(t)+d\int_{t-\tau(y)}^{t}y(s)ds.

Based on hypotheses A3A_{3}, we have

dVdt|(2)=nedjτ(y)(1τ(y)y(t))f[x(tτ(y)),y(tτ(y))]y(tτ(y))dy(t)+d[y(t)y(tτ(y))(1τ(y)y(t))]=y(tτ(y))(1τ(y)y(t))[nedjτ(y)f[x(tτ(y)),y(tτ(y))]d]>y(tτ(y))(1τ(y)y(t))[nedjτ(ε)f(Kε,ε)d]>0.\begin{split}\frac{dV}{dt}|_{(2)}=&ne^{-d_{j}\tau(y)}(1-\tau^{\prime}(y)y^{\prime}(t))f[x(t-\tau(y)),y(t-\tau(y))]y(t-\tau(y))-dy(t)\\ &+d\big{[}y(t)-y(t-\tau(y))(1-\tau^{\prime}(y)y^{\prime}(t))\big{]}\\ =&y(t-\tau(y))(1-\tau^{\prime}(y)y^{\prime}(t))\big{[}ne^{-d_{j}\tau(y)}f[x(t-\tau(y)),y(t-\tau(y))]-d\big{]}\\ >&y(t-\tau(y))(1-\tau^{\prime}(y)y^{\prime}(t))\big{[}ne^{-d_{j}\tau(\varepsilon)}f(K-\varepsilon,\varepsilon)-d\big{]}\\ >&0.\end{split}

Hence we have limty(t)=+,\lim_{t\rightarrow\infty}y(t)=+\infty, contradicting with limty(t)=0\lim_{t\rightarrow\infty}y(t)=0, this proves Ws(E1~X0)={W^{s}(\widetilde{E_{1}}}\cap X^{0})=\emptyset.

Therefore we know that the system (6) satisfies all the conditions in Lemma 1, thus (x(t),y(t))(x(t),y(t)) is uniformly persistent, i.e., there exists positive constants ε\varepsilon and T=T(ε)T=T(\varepsilon) such that x(t),y(t)εx(t),y(t)\geq\varepsilon for all tTt\geq T; In addition, from Theorem 22,we have that (x(t),y(t))(x(t),y(t)) is eventually bounded, implying the permanence of the system (6). Obviously yj(t)y_{j}(t) is permanent from (4), so the permanence of system (2) is straightforward.

The following claim is need for the proof of necessity.

Claim

B: limt(x(t),y(t),yj(t))=(K,0,0)\lim_{t\rightarrow\infty}(x(t),y(t),y_{j}(t))=(K,0,0) holds true iff nedjτ(0)f(K,0)dne^{-d_{j}\tau(0)}f(K,0)\leq d.

Proof

By the first equation of system (2), x(t)x(t) is always decreasing when x(t)>Kx(t)>K. We can show that if there exists some t0>0t_{0}>0 such that x(t0)<Kx(t_{0})<K, then x(t)<Kx(t)<K for all t>t0t>t_{0}. Otherwise there must exist some t1>t0t_{1}>t_{0} such that x(t1)=Kx(t_{1})=K and x(t1)0x^{\prime}(t_{1})\geq 0. This is impossible. Hence, there are two possible cases, either

(1) x(t)>Kx(t)>K and x(t)Kx(t){\rightarrow K} as tt\rightarrow\infty, or

(2) there exists some t0>0t_{0}>0 such that x(t0)<Kx(t_{0})<K.

For the first of these cases, we only need to prove that limty(t)=0\lim_{t\rightarrow\infty}y(t)=0, since this implies limtyj(t)=0\lim_{t\rightarrow\infty}y_{j}(t)=0. Integrating the equation for x(t)x(t) in (2), we have

x(t)x(0)=0trx(s)(1x(s)K)𝑑s0tf(x(s),y(s))y(s)𝑑s<0trx(s)(1x(s)K)x(s)K𝑑s0tf(K,y(s))y(s)𝑑s\begin{split}x(t)-x(0)&=\int_{0}^{t}rx(s)(1-\frac{x(s)}{K})ds-\int_{0}^{t}\ f(x(s),y(s))y(s)ds\\ &<\int_{0}^{t}\underbrace{rx(s)(1-\frac{x(s)}{K})}_{x(s)\geq K}ds-\int_{0}^{t}\ f(K,y(s))y(s)ds\end{split}

for all t0t\geq 0, and then

0tf(K,y(s))y(s)𝑑s<x(0)x(t)+0trx(s)(1x(s)K)0𝑑s<x(0).\int_{0}^{t}\ f(K,y(s))y(s)ds<x(0)-x(t)+\int_{0}^{t}\underbrace{rx(s)(1-\frac{x(s)}{K})}_{\leq 0}ds<x(0).

By the boundedness of y(t)y(t),then 0ty(s)𝑑s\int_{0}^{t}y(s)ds is bounded for all tt0t\geq t_{0},and this implies limty(t)=0\lim_{t\rightarrow\infty}y(t)=0.

For the second of these cases, consider the function

V=y(t)+dtτ(y)ty(s)𝑑sV=y(t)+d\int_{t-\tau(y)}^{t}y(s)ds

Then based on hypotheses A3iiA3-ii, for all tt0+τMt\geq t_{0}+\tau_{M}, we have

dVdt|(2)=nedjτ(y)(1τ(y)y(t))f(x(tτ(y)),y(tτ(y)))y(tτ(y))dy(t)+d[y(t)y(tτ(y))(1τ(y)y(t))]=y(tτ(y))(1τ(y)y(t))[nedjτ(y)f(x(tτ(y)),y(tτ(y)))d]<y(tτ(y))(1τ(y)y(t))[nedjτ(0)f(K,0)d].\begin{split}\frac{dV}{dt}|_{(2)}=&ne^{-d_{j}\tau(y)}(1-\tau^{\prime}(y)y^{\prime}(t))f(x(t-\tau(y)),y(t-\tau(y)))y(t-\tau(y))-dy(t)\\ &+d\big{[}y(t)-y(t-\tau(y))(1-\tau^{\prime}(y)y^{\prime}(t))\big{]}\\ =&y(t-\tau(y))(1-\tau^{\prime}(y)y^{\prime}(t))\big{[}ne^{-d_{j}\tau(y)}f(x(t-\tau(y)),y(t-\tau(y)))-d\big{]}\\ <&y(t-\tau(y))(1-\tau^{\prime}(y)y^{\prime}(t))\big{[}ne^{-d_{j}\tau(0)}f(K,0)-d\big{]}.\end{split}

When nedjτ(0)f(K,0)dne^{-d_{j}\tau(0)}f(K,0)\leq d, dVdt|(2)<0\frac{dV}{dt}|_{(2)}<0 yields limt(x(t),y(t),yj(t))=(K,0,0)\lim_{t\rightarrow\infty}(x(t),y(t),y_{j}(t))=(K,0,0)i from the positivity of y(t)y(t).

The prove the necessary condition for limt(x(t),y(t),yj(t))=(K,0,0)\lim_{t\rightarrow\infty}(x(t),y(t),y_{j}(t))=(K,0,0), assume in contrast, i.e., nedjτ(0)f(K,0)>dne^{-d_{j}\tau(0)}f(K,0)>d, there exists a positive equilibrium (x,y,yj)(x^{*},y^{*},y_{j}^{*}) in the system (2), contradicting with limt(x(t),y(t),yj(t))=(K,0,0)\lim_{t\rightarrow\infty}(x(t),y(t),y_{j}(t))=(K,0,0) for all solution (x(t),y(t),yj(t))(x(t),y(t),y_{j}(t)). Hence there must be nedjτ(0)f(K,0)dne^{-d_{j}\tau(0)}f(K,0)\leq d which is the sufficient condition in Theorem 33.

To show the necessity of Theorem 33, we assume, by contrast, i.e.,
nedjτ(0)f(K,0)dne^{-d_{j}\tau(0)}f(K,0)\leq d; then by Claim B, x(t)K,y(t)0x(t)\to K,y(t)\to 0 as tt\to\infty, which contradict the permanence of (2). This ends of the prove in Theorem 33.

Remark 2

Due to the existence of delay time-derivative term 1τ(y)y(t)1-\tau^{\prime}(y)y^{\prime}(t) in our model, the comparison principle cannot be used directly. We overcome the difficulty by constructing some Lyapunov-like function. Here Theorems 3 is an extension of (6, , Theorem 3.2) and The claim B extends (6, , Theorem 3.1).

5 Linearized Stability

In this section,we study the linearized stability of the equilibria E0E_{0}, E1E_{1} and EE^{*} in the system (6). Different from the linearization for constant delay system, here the delay is a function depending on the state variable yy, thus linearizing the system with SDTD is not completely straightforward. Adopting the “freezing the delay ” idea introduced in 21 , we linearize the system (6) first.

By equation (4), we can get the same conclusions for system (2) and System (6). Let E(x,y)E^{*}(x^{*},y^{*}) be an arbitrary equilibrium. we can get the linearized system of (6) as :

x(t)=Ax(t)By(t),y(t)=Cx(tτ(y))+(ηd)y(t)+Dy(tτ(y)).\begin{split}x^{\prime}(t)&=Ax(t)-By(t),\\ y^{\prime}(t)&=Cx(t-\tau(y^{*}))+(\eta-d)y(t)+Dy(t-\tau(y^{*})).\end{split} (7)

with

A\displaystyle A =r2rKxfxy,\displaystyle=r-\frac{2r}{K}x^{*}-f^{*}_{x}y^{*}, B\displaystyle B =f+fyy,\displaystyle=f^{*}+f^{*}_{y}y^{*},
C\displaystyle C =nedjτ(y)fxy,\displaystyle=ne^{-d_{j}\tau(y^{*})}f^{*}_{x}y^{*}, D\displaystyle D =nedjτ(y)(f+fyy),\displaystyle=ne^{-d_{j}\tau(y^{*})}(f^{*}+f^{*}_{y}y^{*}),
η\displaystyle\eta =nedjτ(y)τ(y)fy(ddj).\displaystyle=ne^{-d_{j}\tau(y^{*})}\tau^{\prime}(y^{*})f^{*}y^{*}(d-d_{j}).

where f=f(x,y)|x=x,y=yf^{*}=f(x,y)\Big{|}_{x=x^{*},y=y^{*}}, fx=fx(x,y)|x=x,y=yf_{x}^{*}=\frac{\partial f}{\partial x}(x,y)\Big{|}_{x=x^{*},y=y^{*}}, fy=fy(x,y)|x=x,y=yf_{y}^{*}=\frac{\partial f}{\partial y}(x,y)\Big{|}_{x=x^{*},y=y^{*}}.
This leads to the following characteristic equation:

(λA)(λ+dηDedjτ(y))+BCeλτ(y)=0.(\lambda-A)\big{(}\lambda+d-\eta-De^{-d_{j}\tau(y^{*})}\big{)}+BCe^{-\lambda\tau(y^{*})}=0.
Theorem 5.1

The trivial equilibrium E0=(0,0)E_{0}=(0,0) is unstable.

Proof

For the trivial equilibrium E0=(0,0)E_{0}=(0,0), we have

A=r,B=C=D=η=0.A=r,\ B=C=D=\eta=0.

The characteristic equation

(λr)(λ+d)=0(\lambda-r)(\lambda+d)=0 (8)

Obviously, there is a positive real root λ=r\lambda=r in (8). Hence E0=(0,0)E_{0}=(0,0) is unstable.

Theorem 5.2

The predator extinction equilibrium E1(K,0)E_{1}(K,0) is

  • (i)

    unstable if nedjτ(0)f(K,0)>dne^{-d_{j}\tau(0)}f(K,0)>d;

  • (ii)

    linearly neutrally stable if nedjτ(0)f(K,0)=dne^{-d_{j}\tau(0)}f(K,0)=d;

  • (iii)

    locally asymptotically stable if nedjτ(0)f(K,0)<dne^{-d_{j}\tau(0)}f(K,0)<d.

Proof

For the extinction equilibrium E1=(K,0)E_{1}=(K,0), we have

A=r,B=f(K,0),D=nedjτ(0)f(K,0),C=η=0.A=-r,\ B=f(K,0),\ D=ne^{-d_{j}\tau(0)}f(K,0),\ C=\eta=0.

The characteristic equation

(λ+r)[λ+dnedjτ(0)f(K,0)eλτ(0)]=0(\lambda+r)\big{[}\lambda+d-ne^{-d_{j}\tau(0)}f(K,0)e^{-\lambda\tau(0)}\big{]}=0 (9)

Obviously, the stability of E1E_{1} is determined by the roots in

G(λ)=λ+dnedjτ(0)f(K,0)eλτ(0)=0G(\lambda)=\lambda+d-ne^{-d_{j}\tau(0)}f(K,0)e^{-\lambda\tau(0)}=0 (10)
  • (i)

    When nedjτ(0)f(K,0)>dne^{-d_{j}\tau(0)}f(K,0)>d, we have G(0)=dnedjτ(0)f(K,0)<0G(0)=d-ne^{-d_{j}\tau(0)}f(K,0)<0, and G(+)=+G(+\infty)=+\infty. Hence G(λ)G(\lambda) has at least one positive root and E1E_{1} is unstable.

  • (ii)

    As nedjτ(0)f(K,0)=dne^{-d_{j}\tau(0)}f(K,0)=d, G(λ)=λ+ddeλτ(0)G(\lambda)=\lambda+d-de^{-\lambda\tau(0)}, so λ=0\lambda=0 is a root of G(λ)=0G(\lambda)=0. Furthermore, since G(λ)=1+τ(0)deλτ(0)G^{\prime}(\lambda)=1+\tau(0)de^{-\lambda\tau(0)}, we have G(0)>0G^{\prime}(0)>0. Then, the root λ=0\lambda=0 is simple.

    Denote all the other roots as λ=u+iv\lambda=u+iv, then u,vu,v must satisfy

    (u+d)2+v2=d2e2uτ(0).(u+d)^{2}+v^{2}=d^{2}e^{-2u\tau(0)}.

    Thus u0u\leq 0, i.e., all the other roots have real nonpositive parts. Therefore E1E_{1} is linearly neutrally stable.

  • (iii)

    If nedjτ(0)f(K,0)<dne^{-d_{j}\tau(0)}f(K,0)<d, i.e.

    dnf(K,0)edjτ(0)eλτ(0)>0d-nf(K,0)e^{-d_{j}\tau(0)}e^{-\lambda\tau(0)}>0

    Then G(λ)=0G(\lambda)=0 implies that

    λ+d=nf(K,0)edjτ(0)eλτ(0).\lambda+d=nf(K,0)e^{-d_{j}\tau(0)}e^{-\lambda\tau(0)}.

    Assume Re(λ)0Re(\lambda)\geq 0, then

    |λ+d|>d>nf(K,0)edjτ(0)eλτ(0)|\lambda+d|>d>nf(K,0)e^{-d_{j}\tau(0)}e^{-\lambda\tau(0)}

    yields contradiction. This shows that all roots of G(λ)=0G(\lambda)=0 must have negative real parts, and therefore E1E_{1} is locally asymptotically stable.

Combining the result in Theorem 4, we have:

Corollary 1

The equilibrium E1(K,0)E_{1}(K,0) of system (6) is globally asymptotically stable iff nedjτ(0)f(K,0)<dne^{-d_{j}\tau(0)}f(K,0)<d holds true.

To discuss the stability of the positive equilibria E2E_{2}, For simplicity, in the rest of the manuscript, we chose

f(x,y)=bx1+k1x+k2y,f(x,y)=\frac{bx}{1+k_{1}x+k_{2}y},

for simplicity, in the rest of the manuscript. Then the system (6) becomes

x(t)=rx(t)(1x(t)K)bx(t)y(t)1+k1x(t)+k2y(t),y(t)=n(1τ(y)y(t))edjτ(y)bx(tτ(y))y(tτ(y))1+k1x(tτ(y))+k2y(tτ(y))dy(t).\begin{split}x^{\prime}(t)&=rx(t)(1-\frac{x(t)}{K})-\frac{bx(t)y(t)}{1+k_{1}x(t)+k_{2}y(t)},\\ y^{\prime}(t)&=n(1-\tau^{\prime}(y)y^{\prime}(t))e^{-d_{j}\tau(y)}\frac{bx(t-\tau(y))y(t-\tau(y))}{1+k_{1}x(t-\tau(y))+k_{2}y(t-\tau(y))}-dy(t).\end{split} (11)

and

A\displaystyle A =r2rKx(1+k2y)by(1+k1x+k2y)2,\displaystyle=r-\frac{2r}{K}x^{*}-\frac{(1+k_{2}y^{*})by^{*}}{(1+k_{1}x^{*}+k_{2}y^{*})^{2}}, B\displaystyle B =(1+k1x)bx(1+k1x+k2y)2,\displaystyle=\frac{(1+k_{1}x^{*})bx^{*}}{(1+k_{1}x^{*}+k_{2}y^{*})^{2}},
C\displaystyle C =nbyedjτ(y)(1+k2y)(1+k1x+k2y)2,\displaystyle=\frac{nby^{*}e^{-d_{j}\tau(y^{*})}(1+k_{2}y^{*})}{(1+k_{1}x^{*}+k_{2}y^{*})^{2}}, D\displaystyle D =nbxedjτ(y)(1+k1x)(1+k1x+k2y)2,\displaystyle=\frac{nbx^{*}e^{-d_{j}\tau(y^{*})}(1+k_{1}x^{*})}{(1+k_{1}x^{*}+k_{2}y^{*})^{2}},
η\displaystyle\eta =nbxyedjτ(y)τ(y)(ddj)1+k1x+k2y.\displaystyle=\frac{nbx^{*}y^{*}e^{-d_{j}\tau(y^{*})}\tau^{\prime}(y^{*})(d-d_{j})}{1+k_{1}x^{*}+k_{2}y^{*}}.

From 22 , we know the distribution of the roods in an 4th4^{th}-order polynomial

v4+Q1v3+Q2v2+Q3v+Q4=0,v^{4}+Q_{1}v^{3}+Q_{2}v^{2}+Q_{3}v+Q_{4}=0, (12)

which is:

Lemma 3

Considering the positive real root of equation (12), there are the following conclusions:

  • (i)

    If Q4<0Q_{4}<0, then equation (12) has at least one positive real root;

  • (ii)

    If Q40Q_{4}\geq 0, and 0\bigtriangleup\geq 0, then equation (12) have positive real roots iff v1>0v_{1}>0, and h(v1)<0h(v_{1})<0;

  • (iii)

    If Q40Q_{4}\geq 0, and <0\bigtriangleup<0, then equation (12) have positive real roots iff exists v{v1,v2,v3}v^{*}\in\{v_{1},v_{2},v_{3}\} such that v>0v^{*}>0, and h(v)0h(v^{*})\leq 0 holds true.

where

h(v)\displaystyle h(v) =v4+Q1v3+Q2v2+Q3v+Q4,\displaystyle=v^{4}+Q_{1}v^{3}+Q_{2}v^{2}+Q_{3}v+Q_{4},
M\displaystyle M =12Q2316Q12,\displaystyle=\frac{1}{2}Q_{2}-\frac{3}{16}Q_{1}^{2}, N\displaystyle N =132Q1318Q1Q2+Q3,\displaystyle=\frac{1}{32}Q_{1}^{3}-\frac{1}{8}Q_{1}Q_{2}+Q_{3},
\displaystyle\bigtriangleup =(N2)2+(M2)3,\displaystyle=(\frac{N}{2})^{2}+(\frac{M}{2})^{3}, σ\displaystyle\sigma =1+3i2,\displaystyle=\frac{-1+\sqrt{3}i}{2},
Y1\displaystyle Y_{1} =N2+3+N23,\displaystyle=\sqrt[3]{-\frac{N}{2}+\sqrt{\bigtriangleup}}+\sqrt[3]{-\frac{N}{2}-\sqrt{\bigtriangleup}}, Y2\displaystyle Y_{2} =σN2+3+σ2N23\displaystyle=\sigma\sqrt[3]{-\frac{N}{2}+\sqrt{\bigtriangleup}}+\sigma^{2}\sqrt[3]{-\frac{N}{2}-\sqrt{\bigtriangleup}}
Y3\displaystyle Y_{3} =σ2N2+3+σN23,\displaystyle=\sigma^{2}\sqrt[3]{-\frac{N}{2}+\sqrt{\bigtriangleup}}+\sigma\sqrt[3]{-\frac{N}{2}-\sqrt{\bigtriangleup}}, vi\displaystyle v_{i} =Yi3Q14,i=1,2,3.\displaystyle=Y_{i}-\frac{3Q_{1}}{4},~{}~{}~{}i=1,2,3.

Consequently, we have the following:

Theorem 5.3

The positive equilibrium (x,y)(x^{*},y^{*}) in system (11) is locally asymptotically stable provided that system (6) is permanent and

k2>2nbedjτ(0)dk1nredjτ(0),ddjk_{2}>2\cdot\frac{nbe^{-d_{j}\tau(0)}-dk_{1}}{nre^{-d_{j}\tau(0)}},\ \ \ \ d\leq d_{j} (13)

holds true.

Proof

For the positive equilibrium E(x,y)E^{*}(x^{*},y^{*}), we have

r(1xK)by1+k1x+k2y=0,nbxedjτ(y)1+k1x+k2yd=0.r(1-\frac{x^{*}}{K})-\frac{by^{*}}{1+k_{1}x^{*}+k_{2}y^{*}}=0,~{}~{}~{}~{}~{}~{}~{}\frac{nbx^{*}e^{-d_{j}\tau(y^{*})}}{1+k_{1}x^{*}+k_{2}y^{*}}-d=0. (14)

Thus

x=12(α+α2+4β).x^{*}=\frac{1}{2}(-\alpha+\sqrt{\alpha^{2}+4\beta}).

where

α=Kr(nbedjτ(y)dk1nk2edjτ(y)r),β=Kdnrk2edjτ(y).\alpha=\frac{K}{r}(\frac{nbe^{-d_{j}\tau(y^{*})}-dk_{1}}{nk_{2}e^{-d_{j}\tau(y^{*})}-r}),~{}~{}~{}~{}~{}~{}\beta=\frac{Kd}{nrk_{2}e^{-d_{j}\tau(y^{*})}}.

nothing that x<Kx^{*}<K, then we have

x>12(α+|α|)=α>K(1nbedjτ(0)dk1nrk2edjτ(0))>K/2>0.x^{*}>\frac{1}{2}(-\alpha+|\alpha|)=-\alpha>K(1-\frac{nbe^{-d_{j}\tau(0)}-dk_{1}}{nrk_{2}e^{-d_{j}\tau(0)}})>K/2>0.

Since ddjd\leq d_{j}, we have

A<0,η0,B>0,C>0,Dd,r2rKx<0.A<0,\eta\leq 0,B>0,C>0,D\leq d,r-\frac{2r}{K}x^{*}<0.

The characteristic equation

(λA)(λ+dηDedjτ(y))+BCeλτ(y)=0.(\lambda-A)(\lambda+d-\eta-De^{-d_{j}\tau(y^{*})})+BCe^{-\lambda\tau(y^{*})}=0. (15)

Thus the roots are given by the following equation:

G(λ)=λ2+H1λ+H2+(N1λ+N2)eλτ(y)=0G(\lambda)=\lambda^{2}+H_{1}\lambda+H_{2}+(N_{1}\lambda+N_{2})e^{-\lambda\tau(y^{*})}=0 (16)

where

H1\displaystyle H_{1} =dηA\displaystyle=d-\eta-A H2\displaystyle H_{2} =AηAd\displaystyle=A\eta-Ad
N1\displaystyle N_{1} =D\displaystyle=-D N2\displaystyle N_{2} =AD+BC\displaystyle=AD+BC

Since

H2+N2\displaystyle H_{2}+N_{2} =AηAd+AD+EC\displaystyle=A\eta-Ad+AD+EC
=A(Dd+η)+BC\displaystyle=A(D-d+\eta)+BC
>0\displaystyle>0

Hence, zero is not the root of equation (16).

Now, let us prove that equation (16) has no purely imaginary roots. Assume that equation (16) has a purely imaginary root λ=iv\lambda=iv, where v>0v>0. Substituting it into equation (16) and separating the real and the imaginary parts, we obtain

v2H2\displaystyle v^{2}-H_{2} =N1vsin(τ(yv)+N2cos(τ(yv),\displaystyle=N_{1}vsin(\tau(y^{*}v)+N_{2}cos(\tau(y^{*}v),
H1v\displaystyle-H_{1}v =N1vcos(τ(yv)N2sin(τ(yv).\displaystyle=N_{1}vcos(\tau(y^{*}v)-N_{2}sin(\tau(y^{*}v).

Add the square of the above equation, we have

v4+B1v2+B2=0v^{4}+B_{1}v^{2}+B_{2}=0 (17)

where B1=H122H2N12,B2=H22N22.B_{1}=H_{1}^{2}-2H_{2}-N_{1}^{2},B_{2}=H_{2}^{2}-N_{2}^{2}.

Since the model assumes τ(y)0\tau^{\prime}(y^{*})\geq 0, we will discuss the following two cases. Case 1. τ(y)=0\tau^{\prime}(y^{*})=0, we have η=0\eta=0, then

B1=(dA)2+2AdD2=A2+(d+D)(dD)>0,B2=(H2+N2)(H2N2)=(H2+N2)(AdADBC)=(H2+N2)[A(nbxedjτ(y)1+k1x+k2y+nbxedjτ(y)1+k1x+k2y1+k1x1+K1x+k2y)nb2xyedjτ(y)(1+k1x)(1+k2y)(1+K1x+k2y)4]=(H2+N2)[nbxedjτ(y)1+k1x+k2y(A+1+k1x1+k1x+k2y(r2rKx))]>0.\begin{split}B_{1}=&(d-A)^{2}+2Ad-D^{2}=A^{2}+(d+D)(d-D)>0,\\ B_{2}=&(H_{2}+N_{2})(H_{2}-N_{2})\\ =&(H_{2}+N_{2})(-Ad-AD-BC)\\ =&(H_{2}+N_{2})\Big{[}-A(\frac{nbx^{*}e^{-d_{j}\tau(y^{*})}}{1+k_{1}x^{*}+k_{2}y^{*}}+\frac{nbx^{*}e^{-d_{j}\tau(y^{*})}}{1+k_{1}x^{*}+k_{2}y^{*}}\frac{1+k_{1}x^{*}}{1+K_{1}x^{*}+k_{2}y^{*}})\\ &-\frac{nb^{2}x^{*}y^{*}e^{-d_{j}\tau(y^{*})}(1+k_{1}x^{*})(1+k_{2}y^{*})}{(1+K_{1}x^{*}+k_{2}y^{*})^{4}}\Big{]}\\ =&(H_{2}+N_{2})\Big{[}-\frac{nbx^{*}e^{-d_{j}\tau(y^{*})}}{1+k_{1}x^{*}+k_{2}y^{*}}(A+\frac{1+k_{1}x^{*}}{1+k_{1}x^{*}+k_{2}y^{*}}(r-\frac{2r}{K}x^{*}))\Big{]}\\ >&0.\end{split}

Then the equation (17) has no positive roots, i.e. equation (16) has no purely imaginary roots.

Case 2. τ(y)>0\tau^{\prime}(y^{*})>0, we have η0\eta\leq 0, then by Lemma 2, for the equation (12),

Q1\displaystyle Q_{1} =0,Q2=B1,Q3=0,\displaystyle=0,Q_{2}=B_{1},Q_{3}=0,
Q4\displaystyle Q_{4} =B2=(H2+N2)((H2N2)\displaystyle=B_{2}=(H_{2}+N_{2})((H_{2}-N_{2})
=(H2+N2)(AηAdADBC)\displaystyle=(H_{2}+N_{2})(A\eta-Ad-AD-BC)
>0.\displaystyle>0.

Therefore, equation (17) has positive real roots can only be the case (ii) and (iii) in the Lemma 33, however, vi=Yi3Q1/4=Yi=0,i=1,2,3,v_{i}=Y_{i}-3Q_{1}/4=Y_{i}=0,~{}i=1,2,3, the case (ii) and (iii) in the lemma 33 cannot be established. Then the equation (17) has no positive roots, i.e., equation (16) has no purely imaginary roots, and each root of the characteristic equation has a negative real part. The proof is complete.

6 Global attractiveness

In this section, we consider the global stability of (x,y)(x^{*},y^{*}) in system (11). The next two lemmas are elementary and useful in the following discussion, which can be found in Gopalsamy GK and Hirsh et al. HW .

Lemma 4

(Barba˘latBarb\breve{a}lat Lemma). Let aa be a finite number and f:[a,)Rf:[a,\infty)\to R be a differentiable function. If limtf(t)lim_{t\to\infty}f(t) exists (finite) and ff^{\prime} is uniformly continuous on [a,)[a,\infty), then limtf(t)=0\lim_{t\to\infty}f^{\prime}(t)=0.

Lemma 5

(Fluctuation Lemma). Let aa be a finite number and f:[a,)Rf:[a,\infty)\to R be a differentiable function. If lim inftf(t)<lim suptf(t)\liminf_{t\to\infty}f(t)<\limsup_{t\to\infty}f(t), then there exist sequences {tn}\{t_{n}\}\uparrow\infty and {sn}\{s_{n}\}\uparrow\infty such that limnf(tn)=lim suptf(t)\lim_{n\to\infty}f(t_{n})=\limsup_{t\to\infty}f(t), f(tn)=0f^{\prime}(t_{n})=0 and limnf(sn)=lim inftf(t)\lim_{n\to\infty}f(s_{n})=\liminf_{t\to\infty}f(t), f(sn)=0f^{\prime}(s_{n})=0.

Now, we are mainly interested in the global asymptotic stability of (x,y)(x^{*},y^{*}). Before proceeding, we will need the following lemma.

Lemma 6

Let v(t)v(t) be the solution of

v(t)=[1τ(v)v(t)]a1edjτ(v)v(tτ(v))1+a2v(tτ(v))a3v(t)v^{\prime}(t)=\big{[}1-\tau^{\prime}(v)v^{\prime}(t)\big{]}\frac{a_{1}e^{-d_{j}\tau(v)}v(t-\tau(v))}{1+a_{2}v(t-\tau(v))}-a_{3}v(t) (18)

where the initial date v(t)=ϕ(t)0v(t)=\phi(t)\geq 0, v(0)>0v(0)>0, for t[τM,0]t\in[-\tau_{M},0] and ai>0,fori=1,2,3a_{i}>0,\ \mbox{for}\ i=1,2,3. Then limtv(t)=v~\lim_{t\to\infty}v(t)=\tilde{v} providing

a1edjτ(v~)>a3,wherev~=a1edjτ(v~)a3a2a3.a_{1}e^{-d_{j}\tau(\tilde{v})}>a_{3},\ \mbox{where}\ \ \tilde{v}=\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}-a_{3}}{a_{2}a_{3}}.
Proof

First we prove the positiveness and eventually uniformly boundedness of v(t)v(t). Assume that, there exists t>0t^{*}>0, such that v(t)=0v(t^{*})=0 and y(t)>0y(t)>0 for every t(0,t)t\in(0,t^{*}). Then we have

v(t)a3vt[τM,t]v^{\prime}(t)\geq-a_{3}v~{}~{}~{}\forall t\in[-\tau_{M},t^{*}]

Therefore

v(t)v0ea3t>0v(t)\geq v_{0}e^{-a_{3}t}>0

let t=tt=t^{*}, we have

0=v(t)v0ea3t>00=v(t^{*})\geq v_{0}e^{-a_{3}t^{*}}>0

which is a contradiction. So no such tt^{*} exists, and we obtain the positiveness results.

Proof of boundedness is divided into two steps.

  • (i)

    Suppose that v(t)0v^{\prime}(t)\geq 0 for all t>Tt>T for some T0T\geq 0. Then for t>T+τMt>T+\tau_{M},

    0v(t)=[1τ(v)v(t)]a1edjτ(v)v(tτ(v))1+a2v(tτ(v))a3v(t)a1edjτ(v)v(t)1+a2v(t)a3v(t)=v(t)[a1edjτ(v)1+a2v(t)a3]\begin{split}0\leq v^{\prime}(t)&=\big{[}1-\tau^{\prime}(v)v^{\prime}(t)\big{]}\frac{a_{1}e^{-d_{j}\tau(v)}v(t-\tau(v))}{1+a_{2}v(t-\tau(v))}-a_{3}v(t)\\ &\leq\frac{a_{1}e^{-d_{j}\tau(v)}v(t)}{1+a_{2}v(t)}-a_{3}v(t)\\ &=v(t)\Big{[}\frac{a_{1}e^{-d_{j}\tau(v)}}{1+a_{2}v(t)}-a_{3}\Big{]}\end{split}

    since v(tτ(v))v(t)v(t-\tau(v))\leq v(t). This means that

    v(t)a1edjτ(v)a3a2a3a1edjτ(0)a3a2a3,fort>T+τM.v(t)\leq\frac{a_{1}e^{-d_{j}\tau(v)}-a_{3}}{a_{2}a_{3}}\leq\frac{a_{1}e^{-d_{j}\tau(0)}-a_{3}}{a_{2}a_{3}},\ \ \mbox{for}\ \ t>T+\tau_{M}.
  • (ii)

    Assume that v(t)v(t) is not eventually monotonic. There is a sequence {tn}n=1\{t_{n}\}_{n=1}^{\infty} such that v(tn)=0v^{\prime}(t_{n})=0, and v(tn)v(t_{n}) is a local maximum. We can further choose the subsequence (still denote {tn}n=1)\{t_{n}\}_{n=1}^{\infty}) such that v(t)v(tn)v(t)\leq v(t_{n}) for all 0<t<tn0<t<t_{n} and nn. Then by a similar analysis at t=tnt=t_{n}, it follows that the solutions v(t)v(t) is bounded above by a bound.

Let us next deal with the case when v(t)v(t) is eventually monotonic. For this case, there exists 0v<0\leq v<\infty such that limtv(t)=v~\lim_{t\to\infty}v(t)=\tilde{v} and limtv(t)=0\lim_{t\to\infty}v^{\prime}(t)=0. Hence from system (18), taking the limit as tt\to\infty, we get that

0=limtv(t)=v~(a1edjτ(v~)1+a2v~a3)0=\lim_{t\to\infty}v^{\prime}(t)=\tilde{v}\Big{(}\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}}{1+a_{2}\tilde{v}}-a_{3}\big{)}

Thus v~=0\tilde{v}=0 or v~=a1edjτ(v~)a3a2a3\tilde{v}=\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}-a_{3}}{a_{2}a_{3}} if a1edjτ(v~)>a3a_{1}e^{-d_{j}\tau(\tilde{v})}>a_{3}. So, this limit must be an equilibrium of (18) and is therefore either zero or the value stated. Zero is ruled out since a standard linearized analysis yields that the zero solution of (18) is linearly unstable under the stated condition on a1edjτ(v~)>a3a_{1}e^{-d_{j}\tau(\tilde{v})}>a_{3}. Therefore, v~=a1edjτ(v~)a3a2a3\tilde{v}=\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}-a_{3}}{a_{2}a_{3}}.

The rest of the case to discuss is that in which v(t)v(t) is neither eventually monotonically increasing nor decreasing. Now, we assume that v(t)v(t) is oscillatory. Then v(t)v(t) has an infinite sequence of local maxima and define the sequence {tj}\{t_{j}\} as those times for which v(tj)=0v^{\prime}(t_{j})=0 and v′′(tj)<0v^{\prime\prime}(t_{j})<0. Here, we will only discuss in detail the case of the local maximum v(tj)>v~v(t_{j})>\tilde{v} for all j=1,2,3,j=1,2,3,\cdots, and other cases can be dealt with analogously.

Now, we prove that suptt1v(t)=v(tk)\sup_{t\geq t_{1}}v(t)=v(t_{k}) for some integer kk. Otherwise, after every local maximum v(tj)v(t_{j}) there is another that is higher, and therefore a subsequence of {tj}\{t_{j}\} (still relabelled {tj}\{t_{j}\}) can be chosen with the property that v(t)<v(tj)v(t)<v(t_{j}) for all t1t<tjt_{1}\leq t<t_{j} and each jj. The subsequence is selected by including each local maximum which is smaller than every one before it. By assumption τ(v)>0\tau(v)>0 and tjτ(v(tj))<tjt_{j}-\tau(v(t_{j}))<t_{j}, for each jj

0=v(tj)=[1τ(v(tj))v(tj)]a1edjτ(v(tj))v(tjτ(v(tj)))1+a2v(tjτ(v(tj)))a3v(tj)<v(tj)[a1edjτ(v~)1+a2v(tj)a3]<v(tj)[a1edjτ(v~)1+a2v~a3]=0\begin{split}0=v^{\prime}(t_{j})&=\big{[}1-\tau^{\prime}(v(t_{j}))v^{\prime}(t_{j})\big{]}\frac{a_{1}e^{-d_{j}\tau(v(t_{j}))}v\big{(}t_{j}-\tau(v(t_{j}))\big{)}}{1+a_{2}v(t_{j}-\tau(v(t_{j})))}-a_{3}v(t_{j})\\ &<v(t_{j})\Big{[}\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}}{1+a_{2}v(t_{j})}-a_{3}\Big{]}\\ &<v(t_{j})\Big{[}\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}}{1+a_{2}\tilde{v}}-a_{3}\Big{]}=0\end{split}

this is a contradiction. So, suptt1v(t)=v(tk)\sup_{t\geq t_{1}}v(t)=v(t_{k}) for some integer kk and we let s1=tks_{1}=t_{k}. Now, by applying the same analysis to the interval ttk+1t\geq t_{k+1}, the existence of a tl(l>k)t_{l}(l>k) with supttk+1v(t)=v(tl)\sup_{t\geq t_{k+1}}v(t)=v(t_{l}) can be obtained, and we set s2=tls_{2}=t_{l}. Continuing this process, we obtain an infinite sequence {sj}\{s_{j}\} of times such that sj+1>sjs_{j+1}>s_{j} as sjs_{j}\to\infty, and v(t)v(sj)v(t)\geq v(s_{j}) for all t>sjt>s_{j}, and v(sj)=0v^{\prime}(s_{j})=0.

Let y(t)=v(t)v~y(t)=v(t)-\tilde{v}. Next, we will prove that y(t)0y(t)\to 0 as tt\to\infty. We have got a sequence y(sj)>y(sj+l)>0y(s_{j})>y(s_{j+l})>0 (since v(sj)v(sj+l)v(s_{j})\geq v(s_{j+l}) and v(sj)>v~v(s_{j})>\tilde{v}), and it is now enough to show that y(sj)0y(s_{j})\to 0 as jj\to\infty. In terms of yy, equation (18) becomes, at t=sjt=s_{j},

0=y(sj)=[1τ(v(sj))v(sj)]a1edjτ(v(sj))v[sjτ(v(sj))]1+a2v[sjτ(v(sj))]a3v(sj)=[a1[y(sjτ(y(sj)+v~)]+a1v~]edjτ(y(sj)+v~)1a2y[sjτ(y(sj)+v~)]+a2v~a3(y(sj)+v~).<[a1[y(sjτ(y(sj)+v~)]+a1v~]edjτ(v~)1a2y[sjτ(y(sj)+v~)]+a2v~a3(y(sj)+v~).\begin{split}0=y^{\prime}(s_{j})&=\big{[}1-\tau^{\prime}(v(s_{j}))v^{\prime}(s_{j})\big{]}\frac{a_{1}e^{-d_{j}\tau(v(s_{j}))}v[s_{j}-\tau(v(s_{j}))]}{1+a_{2}v[s_{j}-\tau(v(s_{j}))]}-a_{3}v(s_{j})\\ &=\frac{\Big{[}a_{1}[y(s_{j}-\tau(y(s_{j})+\tilde{v})]+a_{1}\tilde{v}\Big{]}e^{-d_{j}\tau(y(s_{j})+\tilde{v})}}{1-a_{2}y[s_{j}-\tau(y(s_{j})+\tilde{v})]+a_{2}\tilde{v}}-a_{3}(y(s_{j})+\tilde{v}).\\ &<\frac{\Big{[}a_{1}[y(s_{j}-\tau(y(s_{j})+\tilde{v})]+a_{1}\tilde{v}\Big{]}e^{-d_{j}\tau(\tilde{v})}}{1-a_{2}y[s_{j}-\tau(y(s_{j})+\tilde{v})]+a_{2}\tilde{v}}-a_{3}(y(s_{j})+\tilde{v}).\end{split}

since τ(v~)<τ(y(sj)+v~).\tau(\tilde{v})<\tau\big{(}y(s_{j})+\tilde{v}\big{)}. By the sequence {sj}\{s_{j}\}, we choose a final subsequence, once again denoted {sj}\{s_{j}\}, so that sjτMsj1s_{j}-\tau_{M}\geq s_{j-1}. Then y(sjs)<y(sj1)y(s_{j}-s)<y(s_{j-1}) for all s[0,τM]s\in[0,\tau_{M}] and therefore

a3(y(sj)+v~)<[a1y[sjτ(y(sj)+v~)]+a1v~]edjτ(v~)1+a2y[sjτ(y(sj)+v~)]+a2v~<[a1y(sj1)+a1v~]edjτ(v~)1+a2y(sj1)+a2v~<[a1y(sj1)+a1v~]edjτ(v~)1+a2v~\begin{split}a_{3}(y(s_{j})+\tilde{v})&<\frac{\Big{[}a_{1}y[s_{j}-\tau(y(s_{j})+\tilde{v})]+a_{1}\tilde{v}\Big{]}e^{-d_{j}\tau(\tilde{v})}}{1+a_{2}y[s_{j}-\tau(y(s_{j})+\tilde{v})]+a_{2}\tilde{v}}\\ &<\frac{[a_{1}y(s_{j-1})+a_{1}\tilde{v}]e^{-d_{j}\tau(\tilde{v})}}{1+a_{2}y(s_{j-1})+a_{2}\tilde{v}}\\ &<\frac{[a_{1}y(s_{j-1})+a_{1}\tilde{v}]e^{-d_{j}\tau(\tilde{v})}}{1+a_{2}\tilde{v}}\end{split}

so that

y(sj)<a1edjτ(v~)a3(1+a2v~)y(sj1)y(s_{j})<\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}}{a_{3}(1+a_{2}\tilde{v})}y(s_{j-1})

noting that v~=a1edjτ(v~)a3a2a3\tilde{v}=\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}-a_{3}}{a_{2}a_{3}}, we have a1edjτ(v~)a3(1+a2v~)=1\frac{a_{1}e^{-d_{j}\tau(\tilde{v})}}{a_{3}(1+a_{2}\tilde{v})}=1 and it is independent of jj. Therefore, y(sj)0y(s_{j})\to 0 as jj\to\infty. We summarize that limtv(t)=v~\lim_{t\to\infty}v(t)=\tilde{v} and complete the proof of this theorem.

Remark 3

The global stability of a single-population model without delay time-derivative term is discussed in 16 . Now we use the same method to prove that for a single-population model with delay time-derivative term, the conclusion is still true.

Next, we prove the global stability of (x,y)(x^{*},y^{*}) in system (11).

Theorem 6.1

The positive equilibrium (x,y)(x^{*},y^{*}) in system (11) is globally attractive provided that nbedjτ(y)K1+k1K>d\frac{nbe^{-d_{j}\tau(y^{*})}K}{1+k_{1}K}>d and

k2>max{bK(nbedjτ(y)dk1)r[(nbedjτ(y)dk1)Kd],bK(nbedjτ(y)dk1)rd,br}k_{2}>\max\Big{\{}\frac{bK(nbe^{-d_{j}\tau(y^{*})}-dk_{1})}{r\big{[}(nbe^{-d_{j}\tau(y^{*})}-dk_{1})K-d\big{]}},\frac{bK(nbe^{-d_{j}\tau(y^{*})}-dk_{1})}{rd},\frac{b}{r}\Big{\}} (19)

holds true.

Proof

By the first equation of (11) and the arguments to Theorem 22, for sufficiently small ε>0\varepsilon>0, there is a T1>0T_{1}>0 such that x(t)<K+ε=x¯1x(t)<K+\varepsilon=\overline{x}_{1} for tT1t\geq T_{1}. It is easy to see x<x¯1x^{*}<\overline{x}_{1} for tT1t\geq T_{1}. Replacing this inequality into the second equation of (11), since 1τ(y)y(t)>01-\tau^{\prime}(y)y^{\prime}(t)>0, we have

y(t)<[1τ(y)y(t)]nbedjτ(y)x¯1y(tτ(y))1+k1x¯1+k2y(tτ(y))dy(t),tT1+τMy^{\prime}(t)<\big{[}1-\tau^{\prime}(y)y^{\prime}(t)\big{]}\frac{nbe^{-d_{j}\tau(y)}\overline{x}_{1}y(t-\tau(y))}{1+k_{1}\overline{x}_{1}+k_{2}y(t-\tau(y))}-dy(t),~{}~{}t\geq T_{1}+\tau_{M}

Consider the system

{v(t)=[1τ(y)y(t)]nbedjτ(y)x¯1v(tτ(v))1+k1x¯1+k2v(tτ(v))dv(t),tT1+τMv(t)=y(t),t[t1,T1+τM]\left\{\begin{split}v^{\prime}(t)=&\big{[}1-\tau^{\prime}(y)y^{\prime}(t)\big{]}\frac{nbe^{-d_{j}\tau(y)}\overline{x}_{1}v(t-\tau(v))}{1+k_{1}\overline{x}_{1}+k_{2}v(t-\tau(v))}-dv(t),~{}~{}t\geq T_{1}+\tau_{M}\\ v(t)=&y(t),~{}~{}t\in[t_{1},T_{1}+\tau_{M}]\end{split}\right.

Noting nbedjτ(y)x¯1d(1+k1x¯1)>nbedjτ(y)Kd(1+k1K)>0nbe^{-d_{j}\tau(y^{*})}\overline{x}_{1}-d(1+k_{1}\overline{x}_{1})>nbe^{-d_{j}\tau(y^{*})}K-d(1+k_{1}K)>0. Thus by Lemma 66, we have

limtv(t)=nbedjτ(y)x¯1d(1+k1x¯1)k2d>0.\lim_{t\to\infty}v(t)=\frac{nbe^{-d_{j}\tau(y^{*})}\overline{x}_{1}-d(1+k_{1}\overline{x}_{1})}{k_{2}d}>0.

By the Lemma 11, we have y(t)v(t)y(t)\leq v(t), tT1+τMt\geq T_{1}+\tau_{M}. Then for the sufficiently small ε>0\varepsilon>0, there exists T2>T1+τMT_{2}>T_{1}+\tau_{M} such that

y(t)<nbedjτ(y)x¯1d(1+k1x¯1)k2d+ε=y¯1,tT2.y(t)<\frac{nbe^{-d_{j}\tau(y^{*})}\overline{x}_{1}-d(1+k_{1}\overline{x}_{1})}{k_{2}d}+\varepsilon=\overline{y}_{1},~{}~{}t\geq T_{2}. (20)

By the first equation of (14), we have

y<nbedjτ(y)x¯1d(1+k1x¯1)k2d<y¯1,tT2.y^{*}<\frac{nbe^{-d_{j}\tau(y^{*})}\overline{x}_{1}-d(1+k_{1}\overline{x}_{1})}{k_{2}d}<\overline{y}_{1},~{}~{}t\geq T_{2}.

Replacing (20) into the first equation of (11), we have

x(t)>rx(t)(1x(t)K)bx(t)y¯11+k2y¯1,tT2.x^{\prime}(t)>rx(t)(1-\frac{x(t)}{K})-\frac{bx(t)\overline{y}_{1}}{1+k_{2}\overline{y}_{1}},~{}~{}t\geq T_{2}.

By (19), r>bk2>by¯11+k2y¯1r>\frac{b}{k_{2}}>\frac{b\overline{y}_{1}}{1+k_{2}\overline{y}_{1}}. Using the comparison theorem, for sufficiently small ε>0\varepsilon>0, there is a T3>T2T_{3}>T_{2} such that

x(t)>zε=x¯1>0,tT3.x(t)>z^{*}-\varepsilon=\underline{x}_{1}>0,~{}~{}t\geq T_{3}. (21)

where z=K[1by¯1r(1+k2y¯1)]>0z^{*}=K\big{[}1-\frac{b\overline{y}_{1}}{r(1+k_{2}\overline{y}_{1})}\big{]}>0 is the positive root for the equation

rx(t)(1x(t)K)bx(t)y¯11+k2y¯1=0.rx(t)(1-\frac{x(t)}{K})-\frac{bx(t)\overline{y}_{1}}{1+k_{2}\overline{y}_{1}}=0.

By the second equation of (14), we have

r(1xK)<by¯11+k2y¯1=r(1zK),tT3.r(1-\frac{x^{*}}{K})<\frac{b\overline{y}_{1}}{1+k_{2}\overline{y}_{1}}=r(1-\frac{z^{*}}{K}),~{}~{}t\geq T_{3}.

i.e.,for sufficiently small ε>0\varepsilon>0,

x>zε=x¯1,tT3.x^{*}>z^{*}-\varepsilon=\underline{x}_{1},~{}~{}t\geq T_{3}.

Replacing (21) into the second equation of (11), we have

y(t)>[1τ(y)y(t)]nbedjτ(y)x¯1y(tτ(y))1+k1x¯1+k2y(tτ(y))dy(t),tT3+τMy^{\prime}(t)>\big{[}1-\tau^{\prime}(y)y^{\prime}(t)\big{]}\frac{nbe^{-d_{j}\tau(y)}\underline{x}_{1}y(t-\tau(y))}{1+k_{1}\underline{x}_{1}+k_{2}y(t-\tau(y))}-dy(t),~{}~{}t\geq T_{3}+\tau_{M}

By (21), we have

nbedjτ(y)x¯1d(1+k1x¯1)=(nbedjτ(y)dk1){K[1by¯1r(1+k2y¯1)]ε}d>(nbedjτ(y)dk1){K[1brk2]ε}d=(nbedjτ(y)dk1)(Kε)dk2{k2bK(nbedjτ(y)dk1)r[(nbedjτ(y)dk1)(Kε)d]}.\begin{split}nbe^{-d_{j}\tau(y^{*})}\underline{x}_{1}-d(1+k_{1}\underline{x}_{1})=&(nbe^{-d_{j}\tau(y^{*})}-dk_{1})\big{\{}K\big{[}1-\frac{b\overline{y}-1}{r(1+k_{2}\overline{y}_{1})}\big{]}-\varepsilon\big{\}}-d\\ >&(nbe^{-d_{j}\tau(y^{*})}-dk_{1})\big{\{}K\big{[}1-\frac{b}{rk_{2}}\big{]}-\varepsilon\big{\}}-d\\ =&\frac{(nbe^{-d_{j}\tau(y^{*})}-dk_{1})(K-\varepsilon)-d}{k_{2}}\\ &\cdot\big{\{}k_{2}-\frac{bK(nbe^{-d_{j}\tau(y^{*})}-dk_{1})}{r[(nbe^{-d_{j}\tau(y^{*})}-dk_{1})(K-\varepsilon)-d]}\big{\}}.\end{split}

Using (19), we can get

nbedjτ(y)x¯1d(1+k1x¯1)>0for sufficiently smallε.nbe^{-d_{j}\tau(y^{*})}\underline{x}_{1}-d(1+k_{1}\underline{x}_{1})>0\ \ \ \mbox{for sufficiently small}\ \ \varepsilon. (22)

By Lemma 66 and the similar arguments to y¯1\overline{y}_{1}, for the above selected ε>0\varepsilon>0, there exists T4>T3+τT_{4}>T_{3}+\tau such that

y(t)>nbedjτ(y)x¯1d(1+k1x¯1)k2dε=y¯1,tT4.y(t)>\frac{nbe^{-d_{j}\tau(y^{*})}\underline{x}_{1}-d(1+k_{1}\underline{x}_{1})}{k_{2}d}-\varepsilon=\underline{y}_{1},~{}~{}t\geq T_{4}. (23)

and

y>y¯1,tT4.y^{*}>\underline{y}_{1},~{}~{}t\geq T_{4}.

Therefore we have that

x¯1<x(t)<x¯1,y¯1<y(t)<y¯1,tT4,\underline{x}_{1}<x(t)<\overline{x}_{1},\ \ \underline{y}_{1}<y(t)<\overline{y}_{1},\ \ t\geq T_{4},
x¯1<x<x¯1,y¯1<y<y¯1,tT4,\underline{x}_{1}<x^{*}<\overline{x}_{1},\ \ \underline{y}_{1}<y^{*}<\overline{y}_{1},\ \ t\geq T_{4},

hold for system (11).

Replacing (23) into the first equation of (11), we have

x(t)<rx(t)(1x(t)K)bx(t)y¯11+k1x¯1+k2y¯1,tT4.x^{\prime}(t)<rx(t)(1-\frac{x(t)}{K})-\frac{bx(t)\underline{y}_{1}}{1+k_{1}\overline{x}_{1}+k_{2}\underline{y}_{1}},~{}~{}t\geq T_{4}.

Since rby¯11+k1x¯1+k2y¯1>rby¯11+k2y¯1>0r-\frac{b\underline{y}_{1}}{1+k_{1}\overline{x}_{1}+k_{2}\underline{y}_{1}}>r-\frac{b\overline{y}_{1}}{1+k_{2}\overline{y}_{1}}>0, by the comparison theorem, for sufficiently small ε>0\varepsilon>0, there is a T5>T4T_{5}>T_{4} such that

x(t)<z1+ε=x¯2>0,tT5,x(t)<z_{1}^{*}+\varepsilon=\overline{x}_{2}>0,~{}~{}t\geq T_{5}, (24)

with z1=K[1by¯1r(1+k2y¯1)]>0z_{1}^{*}=K\big{[}1-\frac{b\underline{y}_{1}}{r(1+k_{2}\underline{y}_{1})}\big{]}>0. By the similar arguments to x¯1\underline{x}_{1}, we have

x<x¯2,tT5.x^{*}<\overline{x}_{2},t\geq T_{5}.

From the definition of x¯2\overline{x}_{2} we get

x¯2<K<x¯1.\overline{x}_{2}<K<\overline{x}_{1}.

Replacing (24) into the second equation of (11), we have

y(t)<[1τ(y)y(t)]nbedjτ(y)x¯2y(tτ(y))1+k1x¯2+k2y(tτ(y))dy(t),tT5+τMy^{\prime}(t)<\big{[}1-\tau^{\prime}(y)y^{\prime}(t)\big{]}\frac{nbe^{-d_{j}\tau(y)}\overline{x}_{2}y(t-\tau(y))}{1+k_{1}\overline{x}_{2}+k_{2}y(t-\tau(y))}-dy(t),~{}~{}t\geq T_{5}+\tau_{M}

Since x¯2>x¯1\overline{x}_{2}>\underline{x}_{1} and noting (22), we have nbedjτ(y)x¯2d(1+k1x¯2)>nbedjτ(y)x¯1d(1+k1x¯1)>0nbe^{-d_{j}\tau(y^{*})}\overline{x}_{2}-d(1+k_{1}\overline{x}_{2})>nbe^{-d_{j}\tau(y^{*})}\underline{x}_{1}-d(1+k_{1}\underline{x}_{1})>0. Thus using arguments similar to above, for the sufficiently small ε>0\varepsilon>0, there is a T6>T5+τMT_{6}>T_{5}+\tau_{M} such that

y(t)<nbedjτ(y)x¯2d(1+k1x¯2)k2d+ε=y¯2,tT6.y(t)<\frac{nbe^{-d_{j}\tau(y^{*})}\overline{x}_{2}-d(1+k_{1}\overline{x}_{2})}{k_{2}d}+\varepsilon=\overline{y}_{2},~{}~{}t\geq T_{6}. (25)

by (20), (25) we get y<y¯2<y¯1.y^{*}<\overline{y}_{2}<\overline{y}_{1}.

Replacing (25) into the first equation of (11), we have

x(t)>rx(t)(1x(t)K)bx(t)y¯21+k2y¯2,tT6.x^{\prime}(t)>rx(t)(1-\frac{x(t)}{K})-\frac{bx(t)\overline{y}_{2}}{1+k_{2}\overline{y}_{2}},~{}~{}t\geq T_{6}.

From (19), r>bk2>by¯11+k2y¯1>by¯21+k2y¯2r>\frac{b}{k_{2}}>\frac{b\overline{y}_{1}}{1+k_{2}\overline{y}_{1}}>\frac{b\overline{y}_{2}}{1+k_{2}\overline{y}_{2}}. Then by the comparison theorem, for sufficiently small ε>0\varepsilon>0, there is a T7>T6T_{7}>T_{6} such that

x(t)>z2ε=x¯2>0,tT7,x(t)>z_{2}^{*}-\varepsilon=\underline{x}_{2}>0,~{}~{}t\geq T_{7}, (26)

where z2=K[1by¯2r(1+k2y¯2)]>0z_{2}^{*}=K\big{[}1-\frac{b\overline{y}_{2}}{r(1+k_{2}\overline{y}_{2})}\big{]}>0. By the definition of x¯2\underline{x}_{2}, we have x¯2>x¯1>x\underline{x}_{2}>\underline{x}_{1}>x^{*}.

Replacing (26) into the second equation of (11), then by arguments similar to those for y¯2\overline{y}_{2}, we get that there exists a T8>T7+τMT_{8}>T_{7}+\tau_{M} such that

y(t)>nbedjτ(y)x¯2d(1+k1x¯2)k2dε=y¯2,tT8.y(t)>\frac{nbe^{-d_{j}\tau(y)}\underline{x}_{2}-d(1+k_{1}\underline{x}_{2})}{k_{2}d}-\varepsilon=\underline{y}_{2},~{}~{}t\geq T_{8}. (27)

and we get y>y¯2>y¯1y^{*}>\underline{y}_{2}>\underline{y}_{1}.

Therefore we have that

0<x¯1<x¯2<x(t)<x¯2<x¯1, 0<y¯1<y¯2<y(t)<y¯2<y¯1,tT8.0<\underline{x}_{1}<\underline{x}_{2}<x(t)<\overline{x}_{2}<\overline{x}_{1},\ \ 0<\underline{y}_{1}<\underline{y}_{2}<y(t)<\overline{y}_{2}<\overline{y}_{1},\ \ t\geq T_{8}.
0<x¯1<x¯2<x<x¯2<x¯1, 0<y¯1<y¯2<y<y¯2<y¯1,tT8.0<\underline{x}_{1}<\underline{x}_{2}<x^{*}<\overline{x}_{2}<\overline{x}_{1},\ \ 0<\underline{y}_{1}<\underline{y}_{2}<y^{*}<\overline{y}_{2}<\overline{y}_{1},\ \ t\geq T_{8}.

Repeating the above arguments, we get the four sequences {x¯n}n=1\{\overline{x}_{n}\}_{n=1}^{\infty}, {x¯n}n=1\{\underline{x}_{n}\}_{n=1}^{\infty}, {y¯n}n=1\{\overline{y}_{n}\}_{n=1}^{\infty}, {y¯n}n=1\{\underline{y}_{n}\}_{n=1}^{\infty} with

0<x¯1<x¯2<<x¯n<x(t)<x¯n<<x¯2<x¯1,0<y¯1<y¯2<<y¯n<y(t)<y¯n<<y¯2<y¯1,tT4n.\begin{split}0<&\underline{x}_{1}<\underline{x}_{2}<\cdots<\underline{x}_{n}<x(t)<\overline{x}_{n}<\cdots<\overline{x}_{2}<\overline{x}_{1},\\ 0<&\underline{y}_{1}<\underline{y}_{2}<\cdots<\underline{y}_{n}<y(t)<\overline{y}_{n}<\cdots<\overline{y}_{2}<\overline{y}_{1},\ \ t\geq T_{4n}.\end{split} (28)
0<x¯1<x¯2<<x¯n<x<x¯n<<x¯2<x¯1,0<y¯1<y¯2<<y¯n<y<y¯n<<y¯2<y¯1,tT4n.\begin{split}0<&\underline{x}_{1}<\underline{x}_{2}<\cdots<\underline{x}_{n}<x^{*}<\overline{x}_{n}<\cdots<\overline{x}_{2}<\overline{x}_{1},\\ 0<&\underline{y}_{1}<\underline{y}_{2}<\cdots<\underline{y}_{n}<y^{*}<\overline{y}_{n}<\cdots<\overline{y}_{2}<\overline{y}_{1},\ \ t\geq T_{4n}.\end{split} (29)

From (28) follows that the limit of each sequence in {x¯n}n=1\{\overline{x}_{n}\}_{n=1}^{\infty}, {x¯n}n=1\{\underline{x}_{n}\}_{n=1}^{\infty}, {y¯n}n=1\{\overline{y}_{n}\}_{n=1}^{\infty}, {y¯n}n=1\{\underline{y}_{n}\}_{n=1}^{\infty} exist. Denote

x¯=limnx¯n,y¯=limny¯n,x¯=limnx¯n,y¯=limny¯n.\overline{x}=\lim_{n\rightarrow\infty}\overline{x}_{n},\ \ \overline{y}=\lim_{n\rightarrow\infty}\overline{y}_{n},\ \ \underline{x}=\lim_{n\rightarrow\infty}\underline{x}_{n},\ \ \underline{y}=\lim_{n\rightarrow\infty}\underline{y}_{n}.

thus we get x¯x¯,y¯y¯.\overline{x}\geq\underline{x},\ \overline{y}\geq\underline{y}. To complete the proof, it suffices to prove x¯=x¯=x,y¯=y¯=y.\overline{x}=\underline{x}=x^{*},\ \overline{y}=\underline{y}=y^{*}.

By the definition of y¯n,y¯m,\overline{y}_{n},\ \underline{y}_{m}, we have

y¯n=nbedjτ(0)x¯nd(1+k1x¯n)k2d+ε,y¯m=nbedjτ(0)x¯md(1+k1x¯m)k2dε.\overline{y}_{n}=\frac{nbe^{-d_{j}\tau(0)}\overline{x}_{n}-d(1+k_{1}\overline{x}_{n})}{k_{2}d}+\varepsilon,\ \ \underline{y}_{m}=\frac{nbe^{-d_{j}\tau(0)}\underline{x}_{m}-d(1+k_{1}\underline{x}_{m})}{k_{2}d}-\varepsilon.

then we get

y¯ny¯m=nbedjτ(0)dk1k2d(x¯nx¯m)+2ε.\overline{y}_{n}-\underline{y}_{m}=\frac{nbe^{-d_{j}\tau(0)}-dk_{1}}{k_{2}d}(\overline{x}_{n}-\underline{x}_{m})+2\varepsilon. (30)

By the definition of x¯n,x¯n\overline{x}_{n},\ \underline{x}_{n} and using (30), we have

x¯nx¯n=K[1by¯n1r(1+k2y¯n1)]K[1by¯nr(1+k2y¯n)]+2ε=bKr[y¯ny¯n1(1+k2y¯n1)(1+k2y¯n)]+2ε=bKr[nbedjτ(0)dk1]/k2d(x¯nx¯n1)+2ε(1+k2y¯n1)(1+k2y¯n)+2ε<bKk2dr[nbedjτ(0)dk1](x¯nx¯n1)+2ε(1+bKr).\begin{split}\overline{x}_{n}-\underline{x}_{n}&=K\Big{[}1-\frac{b\underline{y}_{n-1}}{r(1+k_{2}\underline{y}_{n-1})}\Big{]}-K\Big{[}1-\frac{b\underline{y}_{n}}{r(1+k_{2}\underline{y}_{n})}\Big{]}+2\varepsilon\\ &=\frac{bK}{r}\Big{[}\frac{\overline{y}_{n}-\underline{y}_{n-1}}{(1+k_{2}\underline{y}_{n-1})(1+k_{2}\underline{y}_{n})}\Big{]}+2\varepsilon\\ &=\frac{bK}{r}\frac{[nbe^{-d_{j}\tau(0)}-dk_{1}]/k_{2}d(\overline{x}_{n}-\underline{x}_{n-1})+2\varepsilon}{(1+k_{2}\underline{y}_{n-1})(1+k_{2}\underline{y}_{n})}+2\varepsilon\\ &<\frac{bK}{k_{2}dr}[nbe^{-d_{j}\tau(0)}-dk_{1}](\overline{x}_{n}-\underline{x}_{n-1})+2\varepsilon\big{(}1+\frac{bK}{r}\big{)}.\end{split}

Let nn\to\infty, then we have

x¯x¯bKk2dr[nbedjτ(0)dk1](x¯x¯)+2ε(1+bKr),\overline{x}-\underline{x}\leq\frac{bK}{k_{2}dr}[nbe^{-d_{j}\tau(0)}-dk_{1}](\overline{x}-\underline{x})+2\varepsilon\big{(}1+\frac{bK}{r}\big{)},

thus

{1bKk2dr[nbedjτ(0)dk1]}(x¯x¯)2ε(1+bKr).\Big{\{}1-\frac{bK}{k_{2}dr}[nbe^{-d_{j}\tau(0)}-dk_{1}]\Big{\}}(\overline{x}-\underline{x})\leq 2\varepsilon\big{(}1+\frac{bK}{r}\big{)}.

From (19), we have 1bKk2dr[nbedjτ(0)dk1]>01-\frac{bK}{k_{2}dr}[nbe^{-d_{j}\tau(0)}-dk_{1}]>0, and noting that ε>0\varepsilon>0 can be arbitrarily small, then we have x¯=x¯.\overline{x}=\underline{x}. By (30) and let n,m,n,m\rightarrow\infty, we get y¯=y¯.\overline{y}=\underline{y}. From (29), x¯=x¯=x,y¯=y¯=y.\overline{x}=\underline{x}=x^{*},\ \overline{y}=\underline{y}=y^{*}. This proves Theorem 88.

Corollary 2

The positive equilibrium (x,y)(x^{*},y^{*}) in system (11) is globally asymptotically stable provided that nbedjτ(y)K1+k1K>d\frac{nbe^{-d_{j}\tau(y^{*})}K}{1+k_{1}K}>d and that the conditions (13), (19) hold true.

Remark 4

From Theorem and Corollary 2, it is shown that if the system (11) is permanent, that the derivative of SDTD on the state yy is small enough such that |τ(y)τ(0)||\tau(y^{*})-\tau(0)| is small enough, and that the predator interference k2k_{2} is large enough, then the coexistence equilibrium (x,y)(x^{*},y^{*}) in system (11) is globally asymptotically stable. Theorem 8 and Corollary 2 directly extend (6, , Theorem 4.1)

7 Conclusions and discussions

In this paper, based on the biological observations that during World War II the maturation time of seals and whales was not a fixed value, but depended on the mature population, starting with an age-structured model (1), we formulated and analyzed a prey-predator stage-structured model with SDTD.

Compared with the previous SDTD models 10 ; 13 ; 14 , model (2) is not directly changing the constant delay τ\tau into a SDTD τ(y(t))\tau(y(t)) but was obtained by reducing the age-structured population model, which has the delay time-derivative term τ(y(t))y(t)1\tau^{\prime}(y(t))\cdot y^{\prime}(t)\leq 1. Biologically speaking, model (2) is appropriate in terms of population modeling. On the one hand, with the SDTD, the changes in the number of mature individuals depend on reproduction and death and the changing definition of maturity, which is in line with the delay time-derivative term τ(y(t))y(t)1\tau^{\prime}(y(t))\cdot y^{\prime}(t)\leq 1. On the other hand, we can represent y(t)y(t) and yj(t)y_{j}(t) in an integral form by some biological inductions, namely

y(t)=edt0tedsn(1τ(y(s))y(s))edjτ(y(s))f(x(sτ(y(s)),y(sτ(y(s)))y(sτ(y(s)))dsyj(t)=tτ(y)tnf(x(s),y(s))y(s)edj(ts)ds\begin{split}y(t)&=e^{-dt}\int_{0}^{t}e^{ds}n(1-\tau^{\prime}(y(s))y^{\prime}(s))e^{-d_{j}\tau(y(s))}f(x(s-\tau(y(s)),y(s-\tau(y(s)))\\ &y(s-\tau(y(s)))ds\\ y_{j}(t)&=\int_{t-\tau(y)}^{t}nf(x(s),y(s))y(s)e^{d_{j}(t-s)}ds\end{split}

Taking the derivatives of y(t)y(t) and yj(t)y_{j}(t), we obtain the second and third equation of model (2).

From a biological point of view, we show that tτ(y(t))t-\tau(y(t)) should be a strictly increasing function of tt without any conditions and the derivative with respect to time of the SDTD τ(y(t))\tau(y(t)) is strictly less than one. In addition, it is biologically reasonable for the assumption of the delay τ(y(t))\tau(y(t)). The biological phenomenon mentioned above, a non-decreasing delay, implies that a more mature predator leads to a longer developmental duration, and it makes clear the stabilizing effect 5 ; 23 .

Mathematically compared with 13 ; 14 , first of all, the positivity and boundedness of solutions are discussed, which do not need the stringent condition on τ(y(t))\tau(y(t)) to ensure the positivity of yy and yjy_{j}. Then we give the conditions which are both necessary and sufficient for the permanence and extinction of system (2), and local asymptotic stability of trivial and the boundary equilibria is investigated. Finally, taking the BD-type functional response function as an example, the local asymptotic stability and global attractiveness of positive equilibrium of the model is discussed. It shows that as long as k2k_{2} is big enough, the positive equilibrium E(x,y,yj)E(x^{*},y^{*},y_{j}^{*}) will reach a steady state.

Further research in this direction may consider more realistic complex models, for example,

x(t)=rx(t)(1x(t)K)f(x(t),y(t))y(t),y(t)=n(1τ(x)x(t))edjτ(x)f(x(tτ(x),y(tτ(x)))y(tτ(x))dy(t),yj(t)=n(1τ(x)x(t))edjτ(x)f(x(tτ(x),y(tτ(x)))y(tτ(x))+nf(x(t),y(t))y(t)djyj(t).\begin{split}x^{\prime}(t)=&rx(t)(1-\frac{x(t)}{K})-f(x(t),y(t))y(t),\\ y^{\prime}(t)=&n(1-\tau^{\prime}(x)x^{\prime}(t))e^{-d_{j}\tau(x)}f(x(t-\tau(x),y(t-\tau(x)))y(t-\tau(x))-dy(t),\\ y_{j}^{\prime}(t)=&-n(1-\tau^{\prime}(x)x^{\prime}(t))e^{-d_{j}\tau(x)}f(x(t-\tau(x),y(t-\tau(x)))y(t-\tau(x))\\ &+nf(x(t),y(t))y(t)-d_{j}y_{j}(t).\end{split}

where τ(x)\tau(x) is a decreasing continuously differentiable bounded function of the prey population x(t)x(t). And,

x(t)=rx(t)(1x(t)K)f(x(t),y(t))y(t),y(t)=n(1τ(z)z(t))edjτ(z)f(x(tτ(z),y(tτ(z)))y(tτ(z))dy(t),yj(t)=n(1τ(z)z(t))edjτ(z)f(x(tτ(z),y(tτ(z)))y(tτ(z))+nf(x(t),y(t))y(t)djyj(t).\begin{split}x^{\prime}(t)=&rx(t)(1-\frac{x(t)}{K})-f(x(t),y(t))y(t),\\ y^{\prime}(t)=&n(1-\tau^{\prime}(z)z^{\prime}(t))e^{-d_{j}\tau(z)}f(x(t-\tau(z),y(t-\tau(z)))y(t-\tau(z))-dy(t),\\ y_{j}^{\prime}(t)=&-n(1-\tau^{\prime}(z)z^{\prime}(t))e^{-d_{j}\tau(z)}f(x(t-\tau(z),y(t-\tau(z)))y(t-\tau(z))\\ &+nf(x(t),y(t))y(t)-d_{j}y_{j}(t).\end{split}

where z(t)z(t) is the total number of mature and immature predators y(t)+yj(t)y(t)+y_{j}(t), and τ(z)\tau(z) is an increasing continuously differentiable bounded function of the predator population y(t)+yj(t)y(t)+y_{j}(t), which will be left as our future work.

Acknowledgments

The authors would like to thank Dr Xianning Liu for his valuable discussions. Q. Zhang and S. Liu are supported by the National Natural Science Foundation of China (No.11871179 and No.11771374). Y. Yuan is supported by …… Y. Lv is supported by the National Natural Science Foundation of China (No.11871371).

References

  • (1) P. J. Wangersky, W. J. Cunningham, Time lag in prey-predator population models, Ecology, 38, 136-139 (1957)
  • (2) H. D. Landahl, B. D. Hansen, A three stage population model with cannibasism, Bulletin of Mathematical Biology, 37, 11-17 (1975)
  • (3) W. S. C. Gurney, S. P. Blythe, Nicholson’s blowflies revisited, Nature, 287, 17-21 (1980)
  • (4) A. J. Nicholson, Compensatory reactions of populations to stresses, and their evolutionary significance, Australian Journal of Zoology, 2, 1-8 (1954)
  • (5) S. Liu, E. Beretta, A stage-structured predator-prey model of beddington-deangelis type, SIAM J. Appl. Math, 66, 1101-1129 (2006)
  • (6) S. A. Gourley, Y. Kuang, A stage structured predator-prey model and its dependence on maturation delay and death rate, J. Math. Biol, 49, 188-200 (2004)
  • (7) J. Fang, S. A. Gourley, Y. Lou, Stage-structured models of intra- and inter-specific competition within age classes, J. Differential Equations, 260, 1918-1953 (2016)
  • (8) I. A. Darabsah, Y. Yuan, A Stage-structured mathematical model for fish stock with harvesting, SIAM J. Appl. Math, 78, 145-170 (2018)
  • (9) R. Gambell, Birds and mammals-Antarctic whales, Pergamon Press, New York (1985)
  • (10) W. G. Aiello, H. I. Freedman, J. Wu, Analysis of a model representing stage-structured population growth with state-dependent time delay. SIAM J. Appl. Math, 52(3), 855-869 (1992)
  • (11) J. F. M. Al-Omari, S. A. Gourley, Dynamics of a stage-structured population model incorporating a state-dependent maturation delay. Nonlinear Anal. Real World Appl, 6(1), 13-33 (2005)
  • (12) M. Kloosterman, S. A. Campbell, F. J. Poulin, An NPZ model with state-dependent delay due to size-structure in juvenile zooplankton, SIAM J. Appl. Math, 76, 551-577 (2016)
  • (13) F. Magpantay, N. Kosovalic, J. Wu, An age-structured population model with state-dependent delay: Derivation and numerical integration, SIAM J. Numer. Anal., 52(2), 735-756 (2014)
  • (14) Y. Lv, Y. Pei, R. Yuan, Global stability of a competitive model with state-dependent delay, J. Dyn. Diff. Equat., 29, 501-521 (2017)
  • (15) A. Rezounenko, A condition on delay for differential equations with discrete state-dependent delay, J. Math. Anal. Appl., 385(1), 506-516 (2012)
  • (16) A. Rezounenko, Stability of a viral infection model with state-dependent delay, CTL and antibody immune responses, Discrete Cont. Dyn-B, 22(4), 506-516 (2017)
  • (17) Q. Hu, Global Hopf bifurcation of differential equations with threshold type state-dependent delay, J. Diff. Equat., 257(7), 2622-2670 (2014)
  • (18) Q. Hu, A Model of Cold Metal Rolling Processes with State-Dependent Delay, SIAM J. Appl. Math., 76(3), 1076-1100 (2016)
  • (19) S. Li, S. Guo, Dynamics of a two-species stage-structured model incorporating state-dependent maturation delays, Discrete Cont. Dyn-B, 22(4), 1393-1423 (2017)
  • (20) R. May, Stability and Complexity in Model Ecosystems, 2nd edn. Princeton University Press, Princeton, NJ. (1974)
  • (21) A. Hastings, Age-dependent predation is not a simple process. I. Continuous time models. Theoretical Population Biology 23, 347-362 (1983)
  • (22) A. Hastings, Age-dependent predation is not a simple process. II. Wolves, ungulates and a discretetime model for predation on juveniles with a stabilizing tail. Theoretical Population Biology, 26, 271-282 (1984)
  • (23) W. Murdoch, R. Nisbet, S. Blythe, W. Gurney, and J. Reeve, An invulnerable age class and stability in delay-differential host-parasitoid models. American Naturalist, 134 288-310 (1987)
  • (24) J. F. M. Al-Omari, The effect of state dependent delay and harvesting on a stage-structured predator-prey model, Applied Mathematics and Computation, 271, 142-153 (2015)
  • (25) Y. Lv, Y. Pei, R. Yuan, Modeling and analysis of a predator-prey model with state-dependent delay, Int. J. Biomath, 11(2), 1850026 (2018)
  • (26) Y. Wang, X. Liu, Y. Wei, Dynamics of a stage-structured single population model with state-dependent delay, Advances in Difference Equations, 364, 1-5 (2018)
  • (27) J. M. Cushing, An introduction to structured population dynamics, SIAM, Philadelphia, (1998)
  • (28) G. F. Webb, Theory of Nonlinear Age-Dependent Population Dynamics, Dekker, NewYork (1985)
  • (29) S. Liu, X. Wang, L. Wang, H. Song, Competitive exclusion in delayed chemostat models with differential removal rates, SIAM J. Appl. Math., 74, 634-648 (2014)
  • (30) C. S. Holling, Some characteristics of simple types of predation and parasitism, Canad. Entomologist, 91, 385-395 (1959)
  • (31) G. S. K. Wolkowicz, H. Xia, Global asymptotic behavior of a chemostat model with discrete delays, SIAM J. Appl. Math., 57 1019-1043 (1997)
  • (32) M. L. Rosenzweig, Paradox of Enrichment: Destabilization of Exploitation Ecosystems in Ecological Time, Science, New Series, 171, 385-387 (1971)
  • (33) J. R. Beddington, Mutual interference between parasites or predators and its effect on searching efficiency, J. Animal Ecol., 44 331-340 (1975)
  • (34) D. L. DeAngelis, R. A. Goldstein, and R. Neill, A model for trophic interaction, Ecology, 56, 881-892 (1975).
  • (35) P. H. Crowley, E. K. Martin, Functional responses and interference within and between year classes of a dragonfly population, J. North American Benthological Soc., 8, 211-221 (1989)
  • (36) R. Martin,H. Smith, Abstract functional differential equations and reaction-diffusion systems. Trans. Am. Math. Soc. 321, 1-44 (1990)
  • (37) J. K. Hale and P. Waltman, Persistence in infinite-dimensional systems, SIAM J. Math. Anal., 20, 388-395 (1989)
  • (38) H. R. Thieme, Persistence under relaxed point-dissipativity (with application to an endemic model), SIAM J. Math. Anal., 24, 407-435 (1993)
  • (39) K. L. Cooke, W. Z. Huang, On the problem of linearization for state dependent delay differential equations, Proc. Am. Math. Soc. 124(5), 1417-1426 (1996)
  • (40) X. Li, J. Wei, On the zeros of a fourth degree exponential polynomial with applications to a neural network model with delays, Chaos, Solitons and Fractals 26, 519-526 (2005)
  • (41) K. Gopalsamy, Stability and Oscillations in Delay Differential Equations of Population Dynamics. Kluwer, Dordrecht (1992)
  • (42) W. Hirsch, H. Hanisch, J. Gabriel, Differential equations models of some parasitic infections, methods for the study of asymptotic behavior. Comm. Pure. Appl. Math. 38, 733-753 (1985)
  • (43) E. A. Trippel, Age at maturity as a stress indicator in fisheries, Bioscience 45(11), 759-771 (1995)
  • (44) W. G. Aiello, H. I. Freedman, A time-delay model of single-species growth with stage structure, Math, Biosc., 101, 139-153 (1990)