
In this paper, we propose a prey-predator model with age structure which is described by the mature period. The aim of this paper is to study how mature period affect the dynamics of interaction between prey and predator. The sufficient condition of the existence of non-negative steady state is derived. By using integrated semigroup theory, we obtain the characteristic equation, by which we find that the non-negative steady state will lose its stability via Hopf bifurcation induced by mature period, and the corresponding periodic solutions emerge. Additionally, some numerical simulations are provided to illustrate the results predicted by linear analysis. Especially, the numerical results indicate that both mature period and age can affect the amplitude and period of periodic solutions.
Citation: Lijun Wang, Chuanjun Dai, Min Zhao. Hopf bifurcation in an age-structured prey-predator model with Holling Ⅲ response function[J]. Mathematical Biosciences and Engineering, 2021, 18(4): 3144-3159. doi: 10.3934/mbe.2021156
[1] | Shuangte Wang, Hengguo Yu . Stability and bifurcation analysis of the Bazykin's predator-prey ecosystem with Holling type Ⅱ functional response. Mathematical Biosciences and Engineering, 2021, 18(6): 7877-7918. doi: 10.3934/mbe.2021391 |
[2] | Yongli Cai, Malay Banerjee, Yun Kang, Weiming Wang . Spatiotemporal complexity in a predator--prey model with weak Allee effects. Mathematical Biosciences and Engineering, 2014, 11(6): 1247-1274. doi: 10.3934/mbe.2014.11.1247 |
[3] | Yuhong Huo, Gourav Mandal, Lakshmi Narayan Guin, Santabrata Chakravarty, Renji Han . Allee effect-driven complexity in a spatiotemporal predator-prey system with fear factor. Mathematical Biosciences and Engineering, 2023, 20(10): 18820-18860. doi: 10.3934/mbe.2023834 |
[4] | Lazarus Kalvein Beay, Agus Suryanto, Isnani Darti, Trisilowati . Hopf bifurcation and stability analysis of the Rosenzweig-MacArthur predator-prey model with stage-structure in prey. Mathematical Biosciences and Engineering, 2020, 17(4): 4080-4097. doi: 10.3934/mbe.2020226 |
[5] | Eric M. Takyi, Charles Ohanian, Margaret Cathcart, Nihal Kumar . Dynamical analysis of a predator-prey system with prey vigilance and hunting cooperation in predators. Mathematical Biosciences and Engineering, 2024, 21(2): 2768-2786. doi: 10.3934/mbe.2024123 |
[6] | Kawkab Al Amri, Qamar J. A Khan, David Greenhalgh . Combined impact of fear and Allee effect in predator-prey interaction models on their growth. Mathematical Biosciences and Engineering, 2024, 21(10): 7211-7252. doi: 10.3934/mbe.2024319 |
[7] | Kunlun Huang, Xintian Jia, Cuiping Li . Analysis of modified Holling-Tanner model with strong Allee effect. Mathematical Biosciences and Engineering, 2023, 20(8): 15524-15543. doi: 10.3934/mbe.2023693 |
[8] | Eduardo González-Olivares, Claudio Arancibia-Ibarra, Alejandro Rojas-Palma, Betsabé González-Yañez . Bifurcations and multistability on the May-Holling-Tanner predation model considering alternative food for the predators. Mathematical Biosciences and Engineering, 2019, 16(5): 4274-4298. doi: 10.3934/mbe.2019213 |
[9] | Peter A. Braza . A dominant predator, a predator, and a prey. Mathematical Biosciences and Engineering, 2008, 5(1): 61-73. doi: 10.3934/mbe.2008.5.61 |
[10] | Xiaoyuan Chang, Junjie Wei . Stability and Hopf bifurcation in a diffusivepredator-prey system incorporating a prey refuge. Mathematical Biosciences and Engineering, 2013, 10(4): 979-996. doi: 10.3934/mbe.2013.10.979 |
In this paper, we propose a prey-predator model with age structure which is described by the mature period. The aim of this paper is to study how mature period affect the dynamics of interaction between prey and predator. The sufficient condition of the existence of non-negative steady state is derived. By using integrated semigroup theory, we obtain the characteristic equation, by which we find that the non-negative steady state will lose its stability via Hopf bifurcation induced by mature period, and the corresponding periodic solutions emerge. Additionally, some numerical simulations are provided to illustrate the results predicted by linear analysis. Especially, the numerical results indicate that both mature period and age can affect the amplitude and period of periodic solutions.
Ecosystem consists of varieties of living organisms and nonliving substances where the living organisms are not independent and have continuous interactions, such as mutualism, competition and predation [1,2,3]. It is obvious that the prey-predator interaction is of the greatest attention among other interactions in an ecosystem because of its universal existence and immense importance [4,5]. Moreover, due to the complexity of ecosystems, mathematical models have been considered as an alternative method of improving our knowledge of the ecological processes [6]. Such an approach goes back to the work of Lotka [7] and Volterra [8], who first proposed a prey-predator model in their pioneer work, which stimulated considerable theoretical researches on mathematical models of population dynamics.
Additionally, a growing evidence suggests that the functional response plays a crucial role in the interaction between the predator and prey. In recent years, various functional response functions, which describe the number of preys consumed by predator per unit time, have been used considerably in modelling population dynamics [9,10], for example, Holling [11] introduced several prey-predator models with three functional responses (named type Ⅰ, type Ⅱ and type Ⅲ) depending on different predation pressure. Specifically, the following mathematical model with Holling type Ⅲ functional response was analyzed by [12]:
{dV(t)dt=rV(t)(1−V(t)k)−mV2(t)U(t)b+V2(t),dU(t)dt=U(t)(−d+cmV2(t)b+V2(t)), | (1) |
where V(t) and U(t) denote the biomass of prey and predator at time t, respectively. Here, r, c, k, d are positive constants that denote prey logistic growth rate, coefficient for the conversion that predator intake to per capital prey, carrying capacity of prey and mortality rate of the predator, respectively. The term mV2(t)b+V2(t) is Holling type Ⅲ functional response.
Recently, Holling type Ⅲ prey-predator models have been widely explored by incorporating some factors. Some studies have shown that a large refuge of prey can stabilize the positive equilibrium and protect prey from its extinction [13,14]. Moreover, a number of works have argued that diffusion-driven spatiotemporal patterns appear in diffusive prey-predator models with Holling type Ⅲ functional response [15,16,17]. In addition, many researchers are also interested in delay-induced dynamics in Holling type Ⅲ prey-predator models, and these studies suggested that delay can destabilize the positive equilibrium, and then the periodic oscillation occurs [18,19,20]. Apart from above researches, Holling type Ⅲ prey-predator models have been devoted by researchers to investigating the dynamics induced by other factors such as Allee effects [21,22] and competition [23,24].
Although all of above studies provide a better understanding of the dynamics of Holling type Ⅲ prey-predator models, they generally consider the fertility rate of a species as a constant for most eco-models. However, in natural ecosystem, a general mechanism for species is that a new reproducing generation needs time to mature to gain fertility, which is so called "age-dependent fertility". Actually, there exist many instances in real ecosystems which support that the fertility is age dependent, such as human, temora longicornis, hoverfly in nature [35]. The age-dependent fertility has been widely explored in prey-predator models, where age-dependent fertility is described by age-structure and there is a growing result suggested that age structure can influence the population dynamics [25,26,27]. In order to investigate the effects on model stability of several specific biological, delay causing mechanisms related to fertility, Cushing et al. [28,29,30] presented several prey-predator models and hypothesized age structure of predator can significantly influence its reproductive ability. These studies focused on the effects of mature period, of gestation periods and of age-related differentials [30]. In our investigation, the biological motivation is to understand further how mature period affect the model stability.
The study of the dynamics induced by age structure in prey-predator models is mainly focused on the boundedness of solutions, the existence and stability of steady states [31,32,33]. However, due to some factors such as noise and environmental factors, a constant population density may not exist in nature [34,35,36]. In recent years, Liu et al. investigated several prey-predator models with age structure [37,38,39,40], and found that the age-dependent equilibrium may loses its stability via Hopf bifurcation and a non-trivial periodic oscillation phenomenon appears, which implied that the oscillation behaviors exists in age-structure models. Here, we are interested in investigating how age structure affect the changes of the population density. Motivated by the idea developed in Liu et al. [39], we present the following prey-predator model with age structure:
{dV(t)dt=rV(t)(1−V(t)k)−mV2(t)∫+∞0u(t,a)dab+V2(t),∂u(t,a)∂t+∂u(t,a)∂a=−du(t,a),u(t,0)=cmV2(t)∫+∞0β(a)u(t,a)dab+V2(t),u(0,⋅)=u0∈L1((0,+∞),R). | (2) |
Here, u(t,a) is distribution function of the predators over age a at time t, and ∫+∞0u(t,a)da denotes the predator population density at time t. The function β(a) is age-specific fertility function.
In natural ecosystems, predator individuals cannot reproduce in the immature stage, that is, a∈(0,τ), so the corresponding fecundity is zero. Obviously, when predator individuals are mature, the reproduction may occur. Thus, we have the following assumption:
Assumption 1.1. Assume that
β(a)={β∗,a≥τ,0,a∈(0,τ), |
where τ>0, β∗>0.
Let R=∫+∞0β(a)e−dada, where e−da denote the survival probability. According to the results of ref. [25], the predator population is stable if R≥1, but the predator population tends to be extinct if R<1. And the results of refs. [37,38] show that the results with the condition R=1 are similar to ones with R=M, where M>1 is a constant. Thus, we assume R=1 in this paper, which implies β∗=dedτ.
In this paper, we propose a prey-predator model and investigate how the mature period effects the interaction dynamics between preys and predators. The rest of paper is organized as follows: In section 2, we express the model as a non-densely defined abstract Cauchy problem and obtain the positive equilibrium. The characteristic equation is obtained based on integrated semigroup theory, by which we derive the existence of Hopf bifurcation in Section 3. Some numerical simulations are given in Section 4. Finally, the paper ends with the conclusions in Section 5.
In this subsection, by the time-scaling ˉt=tτ and the age-scaling ˉa=aτ, the change of variables ˉV(ˉt,ˉa)=V(τˉt,τˉa) and ˉu(ˉt,ˉa)=u(τˉt,τˉa), we can normalize τ in model (2). Then we have the following model:
{dV(t)dt=τ(rV(t)(1−V(t)k)−mV2(t)∫+∞0u(t,a)dab+V2(t)),∂u(t,a)∂t+∂u(t,a)∂a=−dτu(t,a),u(t,0)=τcmV2(t)∫+∞0β(a)u(t,a)dab+V2(t),u(0,⋅)=u0∈L1((0,+∞),R). | (3) |
and we denote the new function β(a) as
β(a)={β∗,a≥1,0,otherwise. |
Denote V(t)=∫+∞0v(t,a)da in model (2), we can rewrite the first equation of model (3) as:
{∂v(t,a)∂t+∂v(t,a)∂a=−v(t,a)μτ,v(t,0)=N(u,v)τ,v(0,⋅)=v0∈L1((0,+∞),R), |
where
N(u,v)=θ∫+∞0v(t,a)da−r(∫+∞0v(t,a)da)2k−m(∫+∞0v(t,a)da)2∫+∞0u(t,a)dab+(∫+∞0v(t,a)da)2, |
and r=θ−μ represents the intrinsic growth rate of the prey, where θ is the birth rate of the prey.
Set χ(t,a)=(v(t,a)u(t,a)), model (2) can be rewritten as follows:
{∂χ(t,a)∂t+∂χ(t,a)∂a=−τQχ(t,a),χ(t,0)=τF(χ(t,⋅)),χ(0,⋅)=χ0∈L1((0,+∞),R2), | (4) |
where Q=(μ00d), and F(χ(t,⋅))=(N(u,v)cmV2(t)∫+∞0β(a)u(t,a)dab+V2(t)).
Considering the following Banach space:
X=R2×L1((0,+∞),R2), |
where ‖(αψ)‖=∥α∥R2+∥ψ∥L1((0,+∞),R2).
And we define the linear operator Aτ:D(Aτ)→X as
Aτ(0R2ψ)=(−ψ(0)ψ′(a)−τQψ), |
with D(Aτ)=({0R2})×W1,1((0,+∞),R2)⊂X. The operator H:¯D(Aτ)→X is defined by
H(0R2ψ)=(F(ψ)0L1). |
Due to the fact that X0:=¯D(Aτ)={0R2}×L1((0,+∞),R2), the linear operator Aτ is non-densely defined.
Let
x(t)=(0R2χ(t,⋅)), |
model (4) can be rewritten as the following non-densely defined abstract Cauchy problem:
{dx(t)dt=Aτ(x(t))+τH(x(t)),x(0)=(0R2χ0)∈¯D(Aτ), | (5) |
model (5) has been studied including global existence and uniqueness of solution by [41,42].
Define ¯x(t)=(0R2¯χ(a))∈X0 is a steady state of model (5), then we have
{Aτ(ˉx(t))+τH(x(t))=0,ˉx(0)∈D(Aτ). |
It follows that
{−¯χ(0)+τF(¯χ(a))=0,−¯χ′(a)−τQ¯χ(a)=0. | (6) |
From the first equation of model (6), we get
∫+∞0β(a)¯u(a)da=cm¯V2b+¯V2∫+∞0β(a)¯u(a)da, |
and
∫+∞0¯u(a)da=∫+∞0β(a)¯u(a)dad. |
According to the second equation of model (6), we have
r¯V(1−¯Vk)−m¯V2∫+∞0¯udab+¯V2=0. |
Obviously, model (5) always has a boundary equilibrium:
¯x0(a)=(0R2¯χ0(a))with¯χ0(a)=(0τμke−τμa), |
and the unique positive equilibrium:
¯xτ(a)=(0R2¯χτ(a)), |
with
¯χτ(a)=(¯V¯uτ(a))=(√bcm−1τdcre−τdak√bcm−1(k−√bcm−1)), |
if and only if cm>1, k−√bcm−1>0.
In the remainder of our paper we assume that cm>1, k−√bcm−1>0.
Here we will obtain the linearized equation of model (5) around the positive equilibrium ˉxτ. Taking the change of variable x(t)=y(t)+¯xτ into model (5), we get:
{dy(t)dt=Aτ(y(t)+ˉxτ)−τH(¯xτ),y(0)=(0R2χ0−¯χτ)=y0∈¯D(Aτ), | (7) |
which is equivalent to
dy(t)dt=Aτ(y(t))+τDH(¯xτ)y(t),fort≥0, | (8) |
where
τDH(¯xτ)(0R2ψ)=(τDB(¯xτ)(ψ)0L1),(0R2ψ)∈D(Aτ), |
with
DB(¯χτ)(ψ)=(θ−2r¯Vk−2bm¯V(b+¯V2)2∫+∞0¯u(a)da2cmb¯V(b+¯V2)2∫+∞0β(a)¯u(a)da−m¯V2b+¯V20)×∫+∞0ψda+(000cm¯V2b+¯V2)×∫+∞0β(a)ψda. | (9) |
Next we define
Fτ:=Aτ+τDH(¯xτ),M(y(t))=τH(y(t)+¯xτ)−τH(¯xτ)−τDH(¯xτ)y(t), |
satisfying M(0)=0, and DM(0)=0.
Therefore the model (7) can be rewritten as:
dy(t)dt=Fτ(y(t))+M(y(t)). | (10) |
In this subsection, we will get the characteristic equation of model (5) around the positive equilibrium ˉxτ. Denote
s:=min{d,μ}>0,Ω:={λ∈C:Re(λ)>−sτ}. |
From the results of [43], we derive the following lemma.
Lemma 3.1. For λ∈Ω∩ρ(Aτ), we can obtain:
(λI−Aτ)−1(ηφ)=(0R2ψ)⇔ψ(a)=ηe−∫a0(λI+τQ)dl+∫a0φ(s)e−∫as(λI+τQ)dlds. | (11) |
with (ηφ)∈X, and (0R2ψ)∈D(Aτ).
Furthermore, Aτ is a Hille-Yosida operator and satisfies
‖(λI−Aτ)−n‖≤1(Re(λ)+sτ)n,for∀λ∈Ω. |
Denote A0 is the part of Aτ in ¯DA(τ), for (0R2ψ)∈D(A0), we obtain:
A0(0R2ψ)=(0R2¯A0(ψ)), |
where ¯A0(ψ)=−ψ′(a)−τQψ,D(¯A0(ψ))={ψ∈W1,1((0,+∞),R2):ψ(0)=0}.
Note that τDH(¯xτ):D(Aτ)⊂X→X, is a compact bounded linear operator, we get
∥TA0(t)∥≤e−sτt,fort≥0, |
then we obtain
ω0.ess(A0)≤ω0(A0)≤−sτ, |
and from [42], we have
ω0.ess((Fτ)0)=ω0.ess((Aτ+τDH(¯xτ))0)≤−sτ<0. |
Consequently, we derive the following proposition.
Lemma 3.2. The linear operator Fτ is a Hille-Yosida operator, and its part (Fτ)0 in ¯D(Fτ) satisfies ω0.ess((Fτ)0)<0.
Set λ∈Ω, (λI−Aτ) and (λI−Fτ) are invertible and
(λI−Fτ)−1=(λI−(Aτ+τDH(¯xτ)))−1=(λI−Aτ)−1(1−τDH(¯xτ)(λI−Aτ)−1)−1, |
then, we have (I−τDH(ˉx))(λI−A−1τ) is also invertible.
Let
(I−τDH(¯xτ)(λI−Aτ)−1)(ηψ)=(χφ), |
which is equivalent to
(ηψ)−τDH(¯xτ)(λI−Aτ)−1(ηψ)=(χφ). |
We obtain
{η−τDB(¯χτ)(ηe−∫a0(λI+τQ)dl+∫a0φ(s)e−∫as(λI+τQ)dlds)=χ,ψ=φ, |
that is
{η−τDB(¯χτ)ηe−∫a0(λI+τQ)dl=χ+τDB(¯χτ)∫a0φ(s)e−∫as(λI+τQ)dlds,ψ=φ. |
Taking the formula of DB(¯χτ) into consideration, we get
{Δ(λ)η=χ+K(λ,φ),ψ=φ, |
where
Δ(λ)=I−τDB(¯χτ)(e−∫a0(λI+τQ)dl), | (12) |
and
K(λ,φ)=τDB(¯χτ)∫a0φ(s)e−∫as(λI+τQ)dlds. |
From the proof of Lemma(3.5) in [44], we derive the following lemma.
Lemma 3.3. Under Assumption 1.1, the following results hold
● σ(Fτ)∩Ω=σp(Fτ)∩Ω={λ∈Ω:det(Δ(λ))=0},
● if λ∈δ(Fτ)∩Ω, we can obtain the formula for resolvent
(λI−Fτ)−1(ηψ)=(0R2φ), |
where φ(a)=(Δ(λ))−1(χ+K(λ,ψ))e−∫a0(λI+τQ)dl+∫a0ψ(s)e−∫as(λI+τQ)dlds. On the other hand, we get
∫+∞0e−∫a0(λI+τQ)dlda=(1λ+μτ0R0R1λ+dτ),∫+∞0β(a)e−∫a0(λI+τQ)dlda=(β∗e−(λ+μτ)λ+μτ0R0Rβ∗e−(λ+dτ)λ+dτ). | (13) |
From (9) and (13), the characteristic at the positive equilibrium is:
det(Δ(λ))=λ2+τp1λ+τ2p0+(τq1λ+τ2q0)e−λ(λ+dτ)(λ+μτ)≜f(λ)h(λ), | (14) |
where
p1=d+r−2rcm−2brcmk√cm−1b+2rk√bcm−1,p0=3dr−4drcm−4dbrcmk√cm−1b+2drk√bcm−1,q1=−d,q0=−3dr+2crdc2m−2bdrcmk√cm−1b−2drk√bcm−1. |
Let λ=τξ, we get
f(λ)=τ2g(ξ)=τ2[ξ2+p1ξ+p0+(q1ξ+q0)e−τξ], |
and
{λ∈Ω:det(Δ(λ))=0}⇔{λ=τξ∈Ω:g(ξ)=0}. |
In this section, we consider the parameter τ as a bifurcation parameter and investigate the existence of Hopf bifurcation by applying the Hopf bifurcation theory in [43]. We get
g(ξ)=ξ2+p1ξ+p0+(q1ξ+q0)e−τξ. | (15) |
Obviously ξ=0 is not an eigenvalue of Eq 15 when p0+q0≠0, and the real parts of eigenvalues of Eq 15 are negative for τ=0 when p1+q1>0.
Suppose that Eq 15 has a pure imaginary root ξ=iω(ω>0), thus we get
−ω2+ip1ω+p0+(iq1ω+q0)e−iτω=0. |
Separating real and imaginary parts of the above equation, we have
{ω2−p0=q1ωsin(ωτ)+q0cos(ωτ),−p1ω=q1ωcos(ωτ)−q0sin(ωτ). | (16) |
It follows that
ω4+(p21−2p0−q21)ω2+p20−q20=0. | (17) |
Set θ=ω2, then Eq 17 turns into
θ2+(p21−2p0−q21)θ+p20−q20=0. |
Consequently, we can clearly see that Eq 17 has only one positive root ω0=√θ0 when p20−q20<0.
From Eq 16, one can get that g(ξ)=0 at τ=τk0,k=0,1,2⋯ has a pair of purely imaginary roots ±iω0, where
ω0=(−(p21−2p0−q21)+√(p21−2p0−q21)2−4(p20−q20)2)12, | (18) |
and
τk0={1ω0(arccos(q0−p1q1)ω20−p0q0q21ω20+q20+2kπ),ifω0(ω20+p1q0−p0q1)q21ω20+q20≥0,1ω0(2π−arccos(q0−p1q1)ω20−p0q0q21ω20+q20+2kπ),ifω0(ω20+p1q0−p0q1)q21ω20+q20<0. | (19) |
Assumption 3.1. Assume that cm>1, k−√bcm−1>0, and p1+q1>0, p20−q20<0.
Lemma 3.4. Under Assumptions 1.1 and 3.1, we have that Eq 17 has a simple root ω=ω0, which given by
dg(ξ)dξ|ξ=iω≠0, |
The proof is similar to [39], and hence we omit it.
Lemma 3.5. Under Assumptions 1.1 and 3.1, Eq 17 has a simple root ξ(τ)=α(τ)+iω(τ) satisfying α(τ0)=0 and ω(τ0)>0, from [39] we have
sign(dRe(ξ)dτ|τ=τk0)=sign(Re(d(τ)dξ|ξ=iω0))=sign(2ω2i+p21−2p0−q21q21ω2i+q20)>0 |
Thus we conclude the following results.
Theorem 3.1. Under Assumptions 1.1 and 3.1, a Hopf bifurcation occurs around ˉxτ for τ near τ=τ0.
In this section, we give two examples to illustrate the existence of Hopf bifurcation. The following parameter set are used
{θ=1.2,μ=0.2,r=θ−μ=1,c=0.8,d=0.3,K=20,m=3,b=3}. |
The age-specific fertility function becomes
β(a)={0.5e0.5τ,a≥τ,0,a∈(0,τ). |
By a simple computation, we can get that model (1) exists a positive equilibrium E0(1.36,0.89) with the given parameter set. Based on software Python 2.7.15, we carry out some numerical simulations for model (1) using of the finite difference methods from [45], and the corresponding time series and phase diagram are shown in Figure 1, which indicates that the positive equilibrium of model (1) is stable. Then taking age structure into model (1), that is, the model (3), we can readily get cm−1=1.4, k−√bcm−1≈18.54, such that there exists a positive steady state E∗(0.54τe−0.5τa,1.46). Additionally, we obtain the conditions p1+q1≈0.23, p20−q20≈−1.01 that guarantee the existence of critical value τ0≈3.4 from Eq 19.
Example 4.1. We give out an example corresponding to stability of the age-dependent equilibrium by choosing τ=3 in Figure 2. Obviously, the age-dependent equilibrium is locally asymptotically stable when τ=3 is below τ0, which is consistent with the theoretical results.
Example 4.2. In constrast, Figure 3 shows the oscillatory behavior induced by mature period. The results imply that when τ=8 is beyond τ0, the age-dependent equilibrium loses its stability via Hopf bifurcation and the corresponding periodic solution emerges. Under the same parameter set, Figure 1 shows that the positive equilibrium is stable in the absence of age structure. Obviously, the results shown in Figure 3 are induced by age structure.
In addition, we can clearly see that the existence of positive steady state depends on a and τ from its expression. Figure 4 (a) shows that, under a fixed mature period value, the amplitude and period of periodic solutions have a decreasing tendency with the increase of age. In contrast, Figure 4 (b) shows that the mature period can enhance the prey-predator oscillation when the age is identical. Obviously, the mature period and age of species have a significant impact on the dynamic behavior of the prey-predator models.
In this paper, we proposed a prey-predator model with age structure, and studied the effect of mature period on prey-predator dynamics. Specifically, the condition of the existence of non-negative steady state was derived, by which we found that the steady state exists whenever cm>1 and k−√bcm−1>0. According to general results of the non-densely defined abstract Cauchy problem [41,42], we obtained the global existence, positivity and uniqueness of solution of model (3). By integrated semigroup theory, the characteristic equation is derived, by which we studied the effect of the mature period τ on stability of the steady state of model (3). Particularly, the theoretical results shown that when bifurcation parameter τ passes through the critical value, the age-dependent equilibrium loses its stability via Hopf bifurcation and the corresponding periodic solutions emerge, which indicates that the oscillatory behavior exists in changes of density of prey-predator populations. Actually, in real ecosystems, population oscillations is likely to be common due to many external factors [36], and the oscillatory behavior induced by mature period indeed exist in an ecological context.
Additionally, the numerical simulations showed the variation in the amplitude of prey-predator populations corresponding to the mature period and age. The numerical results suggested that the positive equilibrium is stable for model (3) in the absence of age structure. However, when age structure is incorporated, there exists a critical value τ0 such that the positive steady state loses its stability when mature period τ exceeding τ0, which means prey-predator population produces periodic oscillatory behaviors. And the numerical analysis shown that the population density tends to zero with the increase of age, which consist with the natural law of ecosystem. Furthermore, we observed that the mature period τ can affect the characteristics of periodic solutions such as amplitude and period. To be specific, when mature period τ exceeding the critical value τ0, the intensity of prey-predator population biomass oscillations is positively correlated with the mature period, while there is a negative correlation between age and the intensity of prey-predator biomass oscillation. The numerical analysis clearly shows that the mature period is the principle factor giving raise to oscillation, and the amplitude and period of periodic solution are greatly depending on τ and a.
In fact, there is a growing evidence that a constant population density may not exist due to the existence of some factors such as noise, physical factors and so on. In addition, the experimental results and field observation [34,35] have confirmed the existence of oscillations in reality. In present paper, we found that prey-predator population exhibits fluctuation under some conditions and mature period is considered as the factor causing population oscillation, which may be more reasonable compared with prey-predator model without age structure. Our results not only will help in further investigating the dynamics of population models but also provide an in-depth understanding of the oscillating coexistence of populations.
This work was supported by the National Key Research and Development Program of China (Grant No. 2018YFE0103700), the National Natural Science Foundation of China (Grants No. 61901303 and No. 61871293), Wenzhou basic argicultural science and technology project (N0. S20180006).
The author declares that there are no conflicts of interest.
[1] |
M. L. Quammen, Predation by Shorebirds, Fish, and Crabs on Invertebrates in Intertidal Mudflats: An Experimental Test, Ecology, 65 (1984), 529–537. doi: 10.2307/1941415
![]() |
[2] |
H. D. Rundle, S. M. Vamosi, Experimental test of predation's effect on divergent selection during character displacement in sticklebacks, Proc. Natl. Acad. Sci. U.S.A., 100 (2003), 14943–14948. doi: 10.1073/pnas.2036360100
![]() |
[3] | K. G. Magnˊusson, O. K. Pˊalsson, Predator–prey interactions of cod and capelin in Icelandic waters, Proc. Natl. Acad. Sci. U.S.A., 193 (1991), 153–170. |
[4] |
J. H. Steele, E. W. Henderson, The role of predation in plankton models, J. Plankton Res., 14 (1992), 157–172. doi: 10.1093/plankt/14.1.157
![]() |
[5] |
R. Mohn, W. D. Bowen, Grey seal predation on the eastern Scotian Shelf: Modeling the impact on Atlantic cod, Can. J. Fish. Aquat. Sci., 53 (1996), 2722–2738. doi: 10.1139/f96-239
![]() |
[6] |
L. J. White, R. J. Maude, W. Pongtavornpinyo, S. Saralamba, R. Aguas, T. V. Effelterre, et al., The role of simple mathematical models in malaria elimination strategy design, Malar. J., 8 (2009), 212. doi: 10.1186/1475-2875-8-212
![]() |
[7] | V. Volterra, The role of simple mathematical models in malaria elimination strategy design, Roma:Academia Nazionale dei Lincei., 2 (1926), 31–113. |
[8] | A. J. Lotka, Elements of physical biology, Science Progress in the Twentieth Century (1919–1933), 21 (1926), 341–343. |
[9] |
W. Y. Wang, L. J. Pei, Stability and Hopf bifurcation of a delayed ratio-dependent predator-prey system, Acta Mech. Sin., 27 (2011), 285–296. doi: 10.1007/s10409-011-0424-1
![]() |
[10] |
J. H. Bao, J. H. Shao, Permanence and Extinction of Regime-Switching Predator-Prey Models, SIAM J. Math. Anal., 48 (2016), 725–739. doi: 10.1137/15M1024512
![]() |
[11] | C. S. Holling, The functional response of predators to prey density and its role in mimicry and population regulation, Mem. Ent. Soc. Can., 97 (1965), 5–60. |
[12] |
J. P. Chen, Z. H. De, The qualitative analysis of two species predator-prey model with Holling's type Ⅲ functional response, Appl. Math. Mech., 7 (1986), 77–86. doi: 10.1007/BF01896254
![]() |
[13] | Y. J. Huang, F. D. Chen, L. Zhong, Stability analysis of a prey–predator model with Holling type Ⅲ response function incorporating a prey refuge, Appl. Math. Comput., 182 (206), 672–683. |
[14] |
J. H. Wang, L. Q. Pan, Qualitative analysis of a harvested predator-prey system with Holling-type Ⅲ functional response incorporating a prey refuge, Adv. Differ. Equ., 2012 (2012), 96. doi: 10.1186/1687-1847-2012-96
![]() |
[15] |
R. Z. Yang, J. J. Wei, Stability and bifurcation analysis of a diffusive prey–predator system in Holling type Ⅲ with a prey refuge, Nonlinear Dyn., 79 (2015), 631–646. doi: 10.1007/s11071-014-1691-8
![]() |
[16] |
J. F. Wang, Spatiotemporal patterns of a homogeneous diffusive predator–prey system with Holling type Ⅲ functional response, J. Dyn. Differ. Equ., 29 (2017), 1383–1409. doi: 10.1007/s10884-016-9517-7
![]() |
[17] |
N. Apreutesei, G. Dimitriu, On a prey–predator reaction–diffusion system with Holling type Ⅲ functional response, J. Comput. Appl. Math., 235 (2010), 366–379. doi: 10.1016/j.cam.2010.05.040
![]() |
[18] |
P. J. Pal, P.K. Mandal, K. K. Lahiri, A delayed ratio-dependent predator–prey model of interacting populations with Holling type Ⅲ functional response, Nonlinear Dyn., 76 (2014), 201–220. doi: 10.1007/s11071-013-1121-3
![]() |
[19] | X. D. Wang, M. Peng, Stability and Hopf bifurcation analysis of a ratio-dependent predator–prey model with two time delays and Holling type Ⅲ functional response, Appl. Math. Comput., 268 (2015), 496–508. |
[20] | Z. W. Cai, L. H. Huang, H. B. Chen, Positive periodic solution for a multispecies competition-predator system with Holling Ⅲ functional response and time delays, Appl. Math. Comput., 217 (2011), 4866–4878. |
[21] |
E. G. Olivares, A. R. Palma, Multiple limit cycles in a Gause type predator–prey model with Holling type Ⅲ functional response and Allee effect on prey, Bull. Math. Biol., 73 (2011), 1378–1397. doi: 10.1007/s11538-010-9577-5
![]() |
[22] |
H. Liu, K. Zhang, Y. Ye, Y. M. Wei, M. Ma, Dynamic complexity and bifurcation analysis of a host–parasitoid model with Allee effect and Holling type Ⅲ functional response, Adv. Differ. Equ., 2019 (2019), 507. doi: 10.1186/s13662-019-2430-8
![]() |
[23] | S. Sengupta, P. Das, D. Mukherjee, Stochastic non-autonomous Holling type-Ⅲ prey-predator model with predator's intra-specific competition, Discrete Contin. Dyn. Syst., 23 (2018), 3275. |
[24] |
X. M. Fan, F. Q. Jiang, H. N. Zhang, Dynamics of multi-species competition–predator system with impulsive perturbations and Holling type Ⅲ functional responses, Nonlinear Anal. Theory Methods Appl., 74 (2011), 3363–3378. doi: 10.1016/j.na.2011.02.012
![]() |
[25] |
D. J. Thompson, Towards a predator-prey model incorporating age structure: the effects of predator and prey size on the predation of Daphnia magna by Ischnura elegans, J. Anim. Ecol., 44 (1975), 907–916. doi: 10.2307/3727
![]() |
[26] |
E. M. Cauley, W. G. Wilson, Dynamics of age-structured and spatially structured predator-prey interactions: individual-based models and population-level formulations, Am. Nat., 142 (1993), 412–442. doi: 10.1086/285547
![]() |
[27] |
M. E. Gurtin, D. S. Levine, On predator-prey interactions with predation dependent on age of prey, Math. Biosci., 47 (1979), 207–219. doi: 10.1016/0025-5564(79)90038-5
![]() |
[28] | J. M. Cushing, M. Saleem, A competition model with age structure, in Mathematical Ecology, Springer, Berlin (1984), 178–192. |
[29] | J. M. Cushing, M. Saleem, competition with age structure, in Population Biology, Springer, Berlin (1983), 225–230. |
[30] |
J. M. Cushing, M. Saleem, A predator prey model with age structure, J. Math. Biol., 14 (1982), 231–250. doi: 10.1007/BF01832847
![]() |
[31] |
S. X. Zhang, X. X. Xu, A mathematical model for hepatitis B with infection-age structure, Discrete Contin. Dyn. Syst., 21 (2016), 1329–1346. doi: 10.3934/dcdsb.2016.21.1329
![]() |
[32] |
M. Saleem, Egg-eating age-structured predators in interaction with age-structured prey, Math. Biosci., 70 (1984), 91–104. doi: 10.1016/0025-5564(84)90048-8
![]() |
[33] | J. M. Cushing, An introduction to structured population dynamics, Society for industrial and applied mathematics, 1998. |
[34] |
E. Benincˊa, J. Huisman, R. Heerkloss, K. D. J¨ohnk, P. Branco, E. H. V. Nes, et al., Chaos in a long-term experiment with a plankton community, Nature, 451 (2008), 822–825. doi: 10.1038/nature06512
![]() |
[35] |
E. Benincˊa, B. Ballantine, S. P. Ellner, J. Huisman, Species fluctuations sustained by a cyclic succession at the edge of chaos, Proc. Natl. Acad. Sci. U.S.A., 112 (2015), 6389–6394. doi: 10.1073/pnas.1421968112
![]() |
[36] |
J. A. Sherratt, M. J. Smith, Periodic travelling waves in cyclic populations: field studies and reaction–diffusion models, J. R. Soc. Interface, 5 (2008), 483–505. doi: 10.1098/rsif.2007.1327
![]() |
[37] |
Z. H. Liu, N. W. Li, Stability and bifurcation in a predator–prey model with age structure and delays, J. Nonlinear Sci., 25 (2015), 937–957. doi: 10.1007/s00332-015-9245-x
![]() |
[38] | X. M. Zhang, Z. H. Liu, Hopf bifurcation analysis in a predator-prey model with predator-age structure and predator-prey reaction time delay, Appl. Math. Model., 91 (2020), 530–548. |
[39] |
X. M. Zhang, Z. H. Liu, Periodic oscillations in age-structured ratio-dependent predator–prey model with Michaelis–Menten type functional response, Physica D, 389 (2019), 51–63. doi: 10.1016/j.physd.2018.10.002
![]() |
[40] |
H. Tang, Z. H. Liu, Hopf bifurcation for a predator–prey model with age structure, Appl. Math. Model., 40 (2016), 726–737. doi: 10.1016/j.apm.2015.09.015
![]() |
[41] | P. Magal, S. G. Ruan, On semilinear Cauchy problems with non-dense domain, Adv. Differ. Equ., 14 (2009), 1041–1084. |
[42] | P. Magal, Compact attractors for time-periodic age-structured population models, Electron. J. Differ. Equ., 2001 (2001), 1–35. |
[43] |
Z. H. Liu, P. Magal, S. G. Ruan, Hopf bifurcation for non-densely defined Cauchy problems, Z. Angew. Math. Phys., 62 (2011), 191–222. doi: 10.1007/s00033-010-0088-x
![]() |
[44] |
W. Zhen, Z. H. Liu, Hopf bifurcation of an age-structured compartmental pest-pathogen model, J. Math. Anal. Appl., 385 (2012), 1134–1150. doi: 10.1016/j.jmaa.2011.07.038
![]() |
[45] |
L. M. Abia, O. Angulo, J. C. Lˊopez-Marcos, Age-structured population models and their numerical solution, Ecol. Modell., 188 (2005), 112–136. doi: 10.1016/j.ecolmodel.2005.05.007
![]() |
1. | Dongxue Yan, Yuan Yuan, Xianlong Fu, Asymptotic analysis of an age-structured predator-prey model with ratio-dependent Holling Ⅲ functional response and delays, 2023, 12, 2163-2480, 391, 10.3934/eect.2022034 | |
2. | Juan Ye, Yi Wang, Zhan Jin, Chuanjun Dai, Min Zhao, Dynamics of a predator-prey model with strong Allee effect and nonconstant mortality rate, 2022, 19, 1551-0018, 3402, 10.3934/mbe.2022157 | |
3. | Yuan Yuan, Xianlong Fu, Asymptotic behavior of an age-structured prey-predator system with distributed delay, 2022, 317, 00220396, 121, 10.1016/j.jde.2022.01.062 | |
4. | Hongyu Chen, Chunrui Zhang, BIFURCATIONS AND HYDRA EFFECTS IN A REACTION-DIFFUSION PREDATOR-PREY MODEL WITH HOLLING Ⅱ FUNCTIONAL RESPONSE, 2023, 13, 2156-907X, 424, 10.11948/20220221 | |
5. | Dongxue Yan, Yu Cao, Yuan Yuan, Stability and Hopf bifurcation analysis of a delayed predator–prey model with age-structure and Holling III functional response, 2023, 74, 0044-2275, 10.1007/s00033-023-02036-3 | |
6. | Yanhe Qiao, Hui Cao, Guoming Xu, A double time-delay Holling Ⅱ predation model with weak Allee effect and age-structure, 2024, 32, 2688-1594, 1749, 10.3934/era.2024080 | |
7. | WENJUN LIU, LI ZHANG, YANNING AN, ASYMPTOTIC STABILITY FOR A FREE BOUNDARY MODEL OF AN ATHEROSCLEROTIC PLAQUE FORMATION IN THE PRESENCE OF REVERSE CHOLESTEROL TRANSPORT WITH HOLLING TYPE-III FUNCTIONAL RESPONSE, 2024, 32, 0218-3390, 683, 10.1142/S0218339024500244 | |
8. | Swagata Dutta, Gourav Mandal, Lakshmi Narayan Guin, Santabrata Chakravarty, Identifying numerical bifurcation structures of codimensions 1 and 2 in interacting species system, 2024, 47, 0170-4214, 14639, 10.1002/mma.10295 |