Holling type | Definition | Generalized form |
Ⅱ | p(u)=muu+k | |
Ⅲ | p(u)=mu2u2+k | p(u)=mu2k1u2+k2u+1(k2>−2√k1) |
Ⅳ | p(u)=muu2+k | p(u)=muk1u2+k2u+1(k2>−2√k1) |
In this paper, we consider the dynamics of a slow-fast Bazykin's model with piecewise-smooth Holling type Ⅰ functional response. We show that the model has Saddle-node bifurcation and Boundary equilibrium bifurcation. Furthermore, it is also proven that the model has a homoclinic cycle, a heteroclinic cycle or two relaxation oscillation cycles for different parameters conditions. These results imply the dynamical behavior of the model is sensitive to the predator competition rate and the initial densities of prey and predators. In order to support the theoretical analysis, we present some phase portraits corresponding to different values of parameters by numerical simulation. These phase portraits include two relaxation oscillation cycles, an unstable relaxation oscillation cycle surrounded by a stable homoclinic cycle; the coexistence of a heteroclinic cycle and an unstable relaxation oscillation cycle. These results reveal far richer and much more complex dynamics compared to the model without different time scale or with smooth Holling type Ⅰ functional response.
Citation: Xiao Wu, Shuying Lu, Feng Xie. Relaxation oscillations of a piecewise-smooth slow-fast Bazykin's model with Holling type Ⅰ functional response[J]. Mathematical Biosciences and Engineering, 2023, 20(10): 17608-17624. doi: 10.3934/mbe.2023782
[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] | 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 |
[3] | 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 |
[4] | Christian Cortés García, Jasmidt Vera Cuenca . Impact of alternative food on predator diet in a Leslie-Gower model with prey refuge and Holling Ⅱ functional response. Mathematical Biosciences and Engineering, 2023, 20(8): 13681-13703. doi: 10.3934/mbe.2023610 |
[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] | 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 |
[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] | 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 |
[9] | Pei Yu, Xiangyu Wang . Analysis on recurrence behavior in oscillating networks of biologically relevant organic reactions. Mathematical Biosciences and Engineering, 2019, 16(5): 5263-5286. doi: 10.3934/mbe.2019263 |
[10] | Shengyu Huang, Hengguo Yu, Chuanjun Dai, Zengling Ma, Qi Wang, Min Zhao . Dynamics of a harvested cyanobacteria-fish model with modified Holling type Ⅳ functional response. Mathematical Biosciences and Engineering, 2023, 20(7): 12599-12624. doi: 10.3934/mbe.2023561 |
In this paper, we consider the dynamics of a slow-fast Bazykin's model with piecewise-smooth Holling type Ⅰ functional response. We show that the model has Saddle-node bifurcation and Boundary equilibrium bifurcation. Furthermore, it is also proven that the model has a homoclinic cycle, a heteroclinic cycle or two relaxation oscillation cycles for different parameters conditions. These results imply the dynamical behavior of the model is sensitive to the predator competition rate and the initial densities of prey and predators. In order to support the theoretical analysis, we present some phase portraits corresponding to different values of parameters by numerical simulation. These phase portraits include two relaxation oscillation cycles, an unstable relaxation oscillation cycle surrounded by a stable homoclinic cycle; the coexistence of a heteroclinic cycle and an unstable relaxation oscillation cycle. These results reveal far richer and much more complex dynamics compared to the model without different time scale or with smooth Holling type Ⅰ functional response.
The study of the long term symbiosis and complex dynamics among different interacting species is a hot topic and has attracted more and more attention of mathematicians and biologists over decades. Under the observations of interactions between various species, researchers purposed many appropriated mathematical models. Among these models, the Lotka-Volterra predator-prey model proposed by Lotka[1] and Volterra[2] is a widely known model due to its universal application and significance. Now, the various Lotka-Volterra models are developed to study the complex population dynamics in different ecological situations. Hence, taking into consideration prey competition and different functional responses, the original Lotka-Volterra model is extended to the generalized Gause type case[3,4]
dudt=au−bu2−vp(u),dvdt=v(−c+dp(u)), | (1.1) |
where u and v separately represent the amount of prey and predators; a and b separately represent the intrinsic growth rate and compete rate of prey; c and d separately represent the death rate and maximal growth rate of predators; p(u) is the functional response describes the variation of the amount of prey affected by the attacks of predators.
However, it is worth noticing that not only is there competition among prey, but predators also compete each other for the limited resource such as the size of the habitat to live and reproduce [5]. Hence, by introducing predator competition, system (1.1) can be rewritten as
dudt=au−bu2−vp(u),dvdt=v(−c+dp(u))−hv2, | (1.2) |
where h is the competition rate of predators. This is the well-known Bazykin's predator-prey model proposed in [6].
The functional response p(u) is usually classified into four types firstly purposed by the biologist Holling [7,8], for example the Holling Type Ⅰ functional response
p(u)={muα,u<α,m,u>α, | (1.3) |
where m is a maximal per capita consumption rate and α is the half-saturation constant-the prey's number at which the per capita consumption rate is half of its maximum m, and other three Holling Type functional responses, see Table 1. Note that these functional response are all bounded function which is suitable for actual field data and the Holling type Ⅰ functional response (1.3) is continuous but not smooth, which is also called the piecewise-smooth Holling type Ⅰ functional response.
Holling type | Definition | Generalized form |
Ⅱ | p(u)=muu+k | |
Ⅲ | p(u)=mu2u2+k | p(u)=mu2k1u2+k2u+1(k2>−2√k1) |
Ⅳ | p(u)=muu2+k | p(u)=muk1u2+k2u+1(k2>−2√k1) |
Recent researches related to the Bazykin's model with different functional responses obtain some interesting complex dynamics and bifurcations. For example, Bzaykin [9] showed that there is a threshold value c1 such that system (1.2) with the functional response p(u)=mu has the globally asymptotically stable boundary equilibrium (ab,0) if c>c1 and the unique globally asymptotically stable positive equilibrium if c<c1; many researchers [5,6,9,10,11,12,13,14,15,16,17] including Bazykin et al.[5,6,9,10], Hainzl et al.[11,12] and Lu and Huang[13] investigated system (1.2) with Holling type Ⅱ functional response by the theoretical analysis or numerical methods and got complex dynamics and rich bifurcation phenomenons such as the location and stability of equilibriums, the existence of limit cycles, the degenerate Bogdanov-Takens bifurcation, the Hopf bifurcation, etc.
In this paper, we investigate Bazykin's predator-prey model (1.2) with piecewise smooth Holling type Ⅰ functional response (1.3). Before going into details, we apply the following rescaling transformation
ˉu=uα,ˉv=mvaα,ˉt=at |
and removing the bar notation, then the system (1.2) can be rewrote as the following non-dimensional system
dudt=u(1−bu)−vp(u)=f(u,v),dvdt=ϵv(p(u)−c−hv)=ϵg(u,v) | (1.4) |
with
p(u)={u,u<1,1,u>1, |
where u and v are the non-dimensional variables; b=bda, ϵ=dma, h=haαdm2 and c=cdm are non-dimensional parameters. Furthermore, we assume the growth rate of predators is much smaller than that of prey which means that system (1.4) is a piecewise-smooth slow-fast system with the small parameter 0<ϵ≪1. Note that our assumption is reasonable because species at different trophic levels have different growth rates, which may vary several orders of magnitude, and the growth time of individuals increases gradually from the bottom of the food chain to the top [18]. Moreover, many observations of interactions between prey and predators such as hares and lynx [19], phytoplankton and zooplankton [20], insects and birds [21], etc. indicate the prey grow much faster than predators.
So far, there are few studies on the slow-fast Bazykin's model and these studies mainly focus on the dynamics of the model (1.2) with smooth Holling Ⅱ functional response [15,16] such as relaxation oscillation cycles, canard phenomenon and so on. Hence, Our main aim in this paper is to study the dynamics of the slow-fast Bazykin's model with a piecewise-smooth functional response (1.4) by using the geometrical singular perturbation theory [22,23,24,25,26,27,28,29] and piecewise-smooth dynamical system theory [30,31,32,33,34,35]. For the various values of parameters, we will show that there are the coexistence of two relaxation oscillation cycles with different stability, an unstable relaxation oscillation cycle surrounded by a stable homoclinic cycle and a heteroclinic cycle enclosing an unstable relaxation oscillation cycle. Furthermore, the system (1.4) undergoes a series of bifurcations such as Saddle-node bifurcation and Discontinuous saddle-node bifurcation of codimension 1. Moreover, we also give some conditions of the global stability of the unique positive equilibrium. Numerical simulations with the help of the "PPlane8" tool of Matlab [36] are presented to illustrate the theoretical results.
The rest of this paper is organized as follows. In Section 2, we study the critical manifold and the existence and types of equilibriums of system (1.4). In Section 3, we show that the dynamics and bifurcations of system (1.4) such as relaxation oscillation cycles, homoclinic cycle, heteroclinic cycle, etc. A brief discussion is given in the last section.
Based on the ecological viewpoint, we are keen on the dynamics of the system (1.4) in the first quadrant R2+={(u,v)∣u≥0,v≥0} and the switching boundary Σ={(u,v)∣u=1} splits the first quadrant into two regions denoted by Σ(−)={(u,v)∣0<u<1,v≥0} and Σ(+)={(u,v)∣u>1,v≥0}. The system (1.4) is smooth in Σ(∓) and determined by the following systems
dudt=u(1−bu−v)=f(−)(u,v),dudt=ϵv(u−c−hv)=ϵg(−)(u,v),(u,v)∈Σ(−) |
and
dudt=u(1−bu)−v=f(+)(u,v),dudt=ϵv(1−c−hv)=ϵg(+)(u,v),(u,v)∈Σ(+). |
Since the vector field is locally Lipschitz, the fundamental existence and uniqueness theory is true for the system (1.4) [31]. Note that the trajectories of the system (1.4) transversally pass through the switching boundary Σ. Moreover, in order to guarantee the density of preys can support the growth of predators, we assume throughout the paper that 0<b<14 and 0<c<1. These are also the necessary condition for the existence of relaxation oscillation cycles. Hence, we will study the system (1.4) in the parametric region
D={η=(b,h,c)∣0<b<14,0<c<1,h>0}. |
Before going into the detail dynamics, we firstly show some basic properties of the system (1.4).
Lemma 2.1. The system (1.4) has the invariant set
Ω={(u,v)∣0≤u≤1b,0≤v≤1−bcbh} |
and it also has no limit cycle which is entirely lied in the region Σ(−).
Proof. It is clear that u=0 and v=0 are two invariant straight lines of system (1.4) and the boundary of set Ω is
∂Ω=L1∪L2∪L3∪L4,L1={(u,v)∣u=1b,0≤v≤1−bcbh},L2={(u,v)∣0≤u≤1b,v=1−bcbh},L3={(u,v)∣u=0,0≤v≤1−eceh},L4={(u,v)∣0≤u≤1b,v=0}. |
For (u,v)∈∂Ω, we have
dudt|(u,v)∈L1=−v<0,dvdt|(u,v)∈L2={ϵ(1−c)(bu−1)bh<0,0<u<1,ϵ(1−1b)<0,0≤u≤1b,dvdt|u=0=−v(c+hv)<0,dudt|v=0=u(1−bu){≥0,0<u≤1b,<0,u>1b, |
which means that the vector field of the system (1.4) in the line ∂Ω never points outside. Hence, any trajectories of the system (1.4) are confined in the set Ω as they enter Ω in finite time.
For (u,v)∈Σ(−), we construct the Dulac function φ(u,v)=1uv and have
∂φf∂u+ϵ∂φg∂v=−bv−ϵhu |
which implies the system (1.4) has no limit cycles in the region Σ(−) because of the Dulac's criteria.
Hence, if the system (1.4) has a limit cycle, it will either entirely locate in the region Σ(+) or consist of trajectories in the regions Σ(+) and Σ(−) and transversally pass through the switching boundary Σ.
Under the time scale transformation τ=ϵt, we can get the equivalent system
ϵdudτ=u(1−bu)−vp(u),dvdτ=v(p(u)−c−hv). | (2.1) |
The systems (1.4) and (2.1) are individually known as the fast system and the slow system because of their different time scales. By setting ϵ=0 in systems (1.4) and (2.1), we get the degenerate system
0=u(1−bu)−vp(u),dvdt=v(p(u)−c−hv) | (2.2) |
and the layer system
dudt=u(1−bu)−vp(u),dvdt=0. | (2.3) |
Hence, the critical manifold is
S0={(u,v)∣u(1−bu)−vp(u)=0}=S10∪S20∪S30, |
which is composed of the singular points of layer system (2.3) and can be divided into three parts as follows, see Figure 1,
S10={(u,v)∣u=0,v>0},S20={(u,v)∣0<u<1,v=1−bu≜H1(u)},S30={(u,v)∣1<u<1b,v=u(1−bu)≜H2(u)}. |
Note that the sub-manifold S20 is a normally hyperbolic attracting sub-manifold and the non-normally hyperbolic points A(0,1) and M(12b,14b) separately split the sub-manifolds S10 and S30 into the normally hyperbolic attracting sub-manifolds
S1a0={(u,v)∣u=0,v>1},S3a0={(u,v)∣12b<u<1b,v=H2(u)} |
and normally hyperbolic repelling sub-manifolds
S1r0={(u,v)∣u=0,0<v<1},S3r0={(u,v)∣1<u<12b,v=H2(u)}. |
For 0<ϵ≪1, the Fenichel theory [22,23] indicates the normally hyperbolic attracting sub-manifolds S20, S1a0 and S3a0 can be perturbed to the attracting slow manifolds S2ϵ, S1aϵ and S3aϵ and the normally hyperbolic repelling sub-manifolds S1r0 and S3r0 can be perturbed to the repelling slow manifolds S1rϵ and S3rϵ. Hence, the trajectory starting near S1a0 is attracted to the v-axis, then it moves down with O(ϵ) speed. It is clear that the trajectory passes the non-hyperbolic point A(0,1) and leaves the O(ϵ) neighborhood of S1r0 at the point (0,pϵ(v)) satisfying limϵ→0pϵ(v)=p0(v). The function p0(v) is called the entry-exit function [28,29] and satisfies the following lemma.
Lemma 2.2. For system (1.4) and v0∈(1,+∞), there is a unique p0(v0)∈(0,1) satisfying
∫p0(v0)v0s−1s(hs+c)ds=0. | (2.4) |
Proof. Let
I(˜v)=∫˜vv0s−1s(hs+c)ds=(1h+1c)lnh˜v+chv0+c−1cln˜vv0, | (2.5) |
it is easy to verify that I(1)<0 and lim˜v→0I(˜v)=+∞. Since
I′(˜v)=˜v−1˜v(h˜v+c)<0,˜v∈(0,1), |
there is a unique ˜v∗∈(0,1) such that I(v∗)=0 which implies p0(v0)=v∗.
Note that the above proof indicates the entry-exit function p0(v0) is monotone decreasing function in (1,+∞).
In this subsection, we mainly study the existence and stability of equilibriums of system (1.4). We can get the equilibriums by a straight calculation of the following equations
f(−)(u,v)=0,g(−)(u,v)=0,0<u<1,f(+)(u,v)=0,g(+)(u,v)=0,u>1 | (2.6) |
and the dynamics of system (1.4) near these equilibriums are determined by the Jacobian matrix of system (1.4) at these equilibriums, which is given by
J(−)=(1−2bu∗−v∗−u∗ϵu∗ϵ(u∗−c−2hu∗)),(u∗,v∗)∈Σ(−) |
and
J(+)=(1−2bu∗−10ϵ(1−c−2hu∗)),(u∗,v∗)∈Σ(+) |
here (u∗,v∗) is the coordinate of these equilibriums. Now, set
h1=4b(1−c)andh2=1−c1−b, |
we give the following lemmas about the existence and stability of equilibriums.
Lemma 2.3. For system (1.4), the following conclusions hold.
1) System (1.4) always has two trivial equilibriums O(0,0) and B(1b,0) which are both saddle.
2) If h<h1 or h>h2, then system (1.4) has a stable equilibrium E1(u1,v1) which are separately located in S20 and S3a0, see Figure 1(a), (e). More precisely, the equilibrium E1(u1,v1) is a globally stable node when h>h2, see Figure 2(a).
Proof. The number and locations of equilibriums of system (1.4) can be obtained by the straight calculation of Eq (2.6). Next, we determine the types of equilibriums O(0,0), B(1b,0) and E1(u1,v1). For equilibrium O(0,0), the eigenvalues of the Jacobian matrix J(−) are λ1=1 and λ2=−ϵc<0 which implies the equilibrium O(0,0) is a saddle point. Similarly, the eigenvalues of the Jacobian matrix J(+) at B(1b,0) are λ1=−1 and λ2=ϵ(1−c)>0 which indicates the equilibrium B(1b,0) is a saddle point. For equilibrium E1(u1,v1)∈S3a0, the similar calculation gives the eigenvalues of the Jacobian matrix J(+) as λ1=1−2bu1<0 and λ2=ϵ(c−1)<0, which implies E1(u1,v1)∈S3a0 is a stable node. For equilibrium E1(u1,v1)∈S20, we calculate the determinant and trace of Jacobian matrix J(−) as
Tr(J(−)(E1))=−bu1−ϵ(u1−c)<0,Det(J(−)(E1))=ϵ(h+c)(h+c+hb−hcb2)(1+hb)2>0, |
which implies E1(u1,v1)∈S20 is stable.
Next, we will prove the global stability of equilibrium E1 when h>h2. We construct the trapping region
Ω1={(u,v)∣12b≤u≤1b,0≤v≤14b} | (2.7) |
and assert that the vector field of system (1.4) point inside at the boundary ∂Ω1. So the trajectories cannot leave after entering the region Ω1. Furthermore, we construct the same Dulac function in Lemma 2.1 as φ(u,v)=1uv and have
∂φf(+)∂u+ϵ∂φg(+)∂v=v−bu2u2v−ϵ1−cuv2<0,(u,v)∈Ω1. |
Hence, based on the Dulac's criteria, the system (1.4) has no limit cycles in the region Ω1, which indicate the equilibrium E1 is globally stable in Ω1. Moreover, the trajectories starting in the region Σ∖Ω1 enter into the region Ω in finite time because of the Fenichel theory and Lemma 2.2. Then we can assert that system (1.4) has a globally stable node E1 when h>h2.
Lemma 2.4. If h=h1 or h=h2, then system (1.4) has two positive equilibriums. More precisely,
1) If h=h1, then system (1.4) has a stable equilibrium E1(u1,v1) located in S20 and an attracting saddle-node EM(12b,14b) which is the vertex point in S30, see Figure 1(b).
2) If h=h2, then system (1.4) has a stable equilibrium E1(u1,v1) located in S20 and a boundary equilibrium EB(1,1−b) located in the switching boundary Σ, see Figures 1(d) and 2(b).
Proof. The existence and location of two positive admissible equilibriums follows from the straight calculation of Eq (2.6). By some simple calculations, we get
Tr(J(−)(E1))=−bu1−ϵ(u1−c)<0,Det(J(−)(E1))=ϵ(h+c)(h+c+hb−hcb2)(1+hb)2>0, |
and the eigenvalues of equilibrium EM(12b,14b) are λ1=0 and λ2=−ϵ(1−c)<0. Hence, the equilibrium E1 is stable and the equilibrium EM is an attracting saddle-node which can be proven based on the conditions in [37].
Lemma 2.5. If h1<h<h2, then system (1.4) has three positive equilibriums E1(u1,v1), E2(u2,v2) and E3(u3,v3). More precisely, E1(u1,v1) is a stable equilibrium in S20; E2(u2,v2) is a saddle in S3r0 and E3(u3,v3) is a stable node in S3a0, see Figure 1(c).
The proof of Lemma 2.5 is similar to that given in Lemmas 2.3 and 2.4 and so is omitted.
In this section, we are keen on various possible dynamics of the slow-fast system (1.4) including relaxation oscillation, saddle-node bifurcation, boundary equilibrium bifurcation and so on.
According to Lemmas 2.3–2.5, we know that
SN={(b,c,h)∣h=h1=4b(1−c)} |
is a saddle-node bifurcation surface. When parameters change from one side of the surface SN to another one, the number of positive equilibriums of system (1.4) in Σ(+) changes from zero to two. So there is a critical competition rate h1 such that the two species coexist in the form of a positive equilibrium for suitable choices of initial values when h=h1
In what follows, we show that system (1.4) has homoclinic orbit, heteroclinic orbit and relaxation oscillation cycles in both sides of saddle-node bifurcation surface SN.
First, we investigate the relaxation oscillation cycles of system (1.4) when h<h1. Now, we denote the function ˆI(˜v) as the function (2.5) with v0=14e and state the main theorem as follows
Theorem 3.1. When 0<b<14, 0<c<1 and 0<h<h1, system (1.4) has a positive equilibrium E1(u1,v1) which is stable. Moreover, if
ˆI(1−b)=1hln4b(1−b)h+4bch+4bc+1cln4b(1−b)h+4bc4b(1−b)h+16b2(1−b)c<0, |
then system (1.4) has a hyperbolically stable relaxation oscillation cycle Γ1ϵ and a hyperbolically unstable relaxation oscillation cycle γϵ which separately converges to
Γ10=⌢MK1∪⌢K1Q1∪⌢Q1D1∪⌢D1M |
and
γ0=⌢D2K2∪⌢K2Q2∪⌢Q2B∪⌢BD2 |
in the Hausdorff distance as ϵ→0, see Figures 3(a) and 4(a).
Proof. Our first goal is to prove the existence of the stable relaxation oscillation cycle Γ1ϵ. To begin with, it is easy to see that the vertex point M(12b,14b) is a general fold and there exists ˆv∗∈(0,1−b) such that p0(14b)=ˆv∗ based on ˆI(1−b)<0 and the properties of function ˆI(˜v). Hence, we construct the singular slow-fast cycle Γ10, see Figure 3(a), as
Γ10=⌢MK1∪⌢K1Q1∪⌢Q1D1∪⌢D1M, |
where M(12b,14b), K1(0,14b), Q1(0,ˆv∗), D1(ˆu∗,ˆv∗) and ˆu∗ is the larger root of equation H2(u)=ˆv∗.
Set
Δ0={(1,v)∣v∈(14b−δ,14b+δ)}andΔ1={(1,v)∣v∈(v∗−δ,v∗+δ)} |
which separately are vertical region of ⌢MK1 and ⌢Q1D1 and δ is a small positive parameter. We construct the Poincaré map Π as
Π=Π1∘Π0:Δ0→Δ0 |
which consists of two maps Π0:Δ0→Δ1 and Π1:Δ1→Δ0. For the trajectory starting in Δ0, it will arrive at the neighborhood of critical manifold S1a0 and move downwards until reach the neighborhood of Q1. The trajectory will move along the layers of system (2.3) and pass through the region Δ1 before reaching near the critical manifold S3a0 because system (1.4) is continuous on the switching boundary Σ. Then the trajectory will move along S3a0 and jump into Δ0 in the neighborhood of fold point M. Hence, the relaxation oscillation cycle Γ1ϵ is equivalent to the fixed point of Poincaré map Π which is a contraction map with exponential contraction rate O(e−1ϵ) by Lemma 2.2 and Theorem 2.1 in [26]. According to the Contraction Mapping Theorem, there is an unique fixed point of Π corresponding to the relaxation oscillation cycle Γ1ϵ. Furthermore, it is clear that Γ1ϵ is a hyperbolically stable limit cycle and converges to the singular slow-fast cycle Γ10 as ϵ→0.
Next we prove the existence of the unstable relaxation oscillation cycle γϵ. Since E1(u1,v1) is stable and the relaxation oscillation cycle Γ1ϵ is hyperbolically stable, we can conclude that there exists at least an unstable limit cycle inside Γϵ based on Poincaré-Bendixson Theorem. According to the Geometric Singular Perturbation Theory [23], the limit cycles of slow-fast system (1.4) are usually perturbed by singular slow-fast cycles. Clearly, system (1.4) has an unique singular slow-fast cycle γ0 inside the relaxation oscillation cycle Γ1ϵ for ϵ→0, see Figure 3(a), which are constructed as
γ0=⌢D2K2∪⌢K2Q2∪⌢Q2B∪⌢BD2, |
with Q2(0,1−b), B(1,1−b), D2(ˉu∗,ˉv∗) and K2(0,ˉv∗). Note that p0(ˉv∗)=1−b and ˉu∗ is the smaller root of equation H2(u)=ˉv∗. Hence, if the limit cycles exist, they must be in the neighborhood of γ0. By the Corollary 4.3 in [38], the sign of the following integral
∮γϵ(∂f∂u+ϵ∂g∂v)dt=∫D2A2(1−2bu)dt+O(ϵ)>0. |
determines the limit cycle is hyperbolically unstable if it exists. Thus, we can claim that there is an unique relaxation oscillation cycle γϵ near γ0 because two adjacent limit cycles don't have the same stabilities.
Combined with the above analysis and applying numerical simulations by the "PPlane8" tool in Matlab, we give the phase portrait of system (1.4) when b=0.2, c=0.5 and h=0.3, see Figure 4(a). Note that system (1.4) has a hyperbolically unstable relaxation oscillation cycle γϵ surrounded by a hyperbolically stable relaxation oscillation cycle Γ1ϵ.
Second, we investigate the existence of a homoclinic cycle of system (1.4) with h=h1 and have the following theorem.
Theorem 3.2. When 0<b<14, 0<c<1 and h=h1, system (1.4) has a stable equilibrium E1(u1,v1) and a attracting saddle-node EM(12b,14b). Moreover, if ˆI(1−b)<0, then system (1.4) has a stable homoclinic cycles Γ2ϵ and a unstable relaxation oscillation cycle γϵ which separately converge to
Γ20=⌢EMK1∪⌢K1Q1∪⌢Q1D1∪⌢D1EM |
and
γ0=⌢D2K2∪⌢K2Q2∪⌢Q2B∪⌢BD2 |
in Hausdorff distance as ϵ→0, see Figures 3(b) and 4(b).
Proof. For ϵ=0, it is clear that there is a singular orbit Γ20 in saddle-node EM, which can be constructed by the same way of Γ10 in Theorem 3.1, see Figure 3(b). Furthermore, the layer of system (2.3) at EM will be perturbed to one dimensional unstable manifold Wuϵ(EM) of system (1.4) at EM if 0<ϵ≪1. By Lemma 2.2, the unstable manifold Wuϵ(EM) will be attracted to the neighborhood of S1a0 and move down until it reach near Q1. Then Wuϵ(EM) move quickly from the neighborhood of S1r0 to the neighborhood of S3a0. Next, we show that the unstable manifold Wuϵ(EM) tends to saddle-node EM. We construct the trapping region Ω1 which is same to (2.7) in Lemma 2.4 and the orbit cannot leave if it enters the region Ω1. Since the unstable manifold Wuϵ(EM) enter this region Ω1 and the orbits is monotone increasing in v, the unstable manifold Wuϵ(EM) tends to saddle-node EM along its stable manifold Wsϵ(EM) which is perturbed by Ws(S3a0) based on the Fenichel theorem. Hence, when 0<ϵ≪1, there is a homoclinic cycle Γ2ϵ which converges to singular orbit Γ20 as ϵ→0.
Based on the Theorem 5 in [37], the stability of the homoclinic cycle Γ2ϵ is determined by the sign of the following integral
∮Γ2ϵ(∂f∂u+ϵ∂g∂v)dt=∫EMD1(1−2bu)dt<0. |
Hence, the homoclinic cycle Γ2ϵ is stable.
The proof of existence and stability of the relaxation oscillation cycle γϵ is omitted because it is similar to that in Theorem 3.1.
According to the analysis of Theorem 3.2 and using numerical simulation by the "PPlane8" tool in Matlab, the phase portrait of system (1.4) with b=0.2, c=0.5 and h=0.4 is presented in Figure 4(b). Note that system (1.4) has a stable homoclinic cycle Γ2ϵ enclosing a hyperbolically unstable relaxation oscillation cycle γϵ.
Third, we study the existence of a heterocilinic cycle of system (1.4) when h>h2. Similarly, we denote ˜I(˜u) as the function (2.5) with v0=1−ch. By applying the similar analysis methods in Theorem 3.2 and Lemma 1 in [39], we can derive the following theorem.
Theorem 3.3. When 0<b<14, 0<c<1 and h1<h<h2, system (1.4) has a saddle point E2(u2,v2) and two stable equilibirums E1(u1,v1) and E3(u3,v3). Moreover, if
˜I(1−b)=1hln[h(1−b)+c]+1cln(1−c)[h(1−b)+c](1−e)h<0, | (3.1) |
then system (1.4) has a heteroclinic cycle Γ3ϵ and a unstable relaxation oscillation cycle γϵ which separately converge to
Γ30=⌢E2K3∪⌢K3Q3∪⌢Q3D3∪⌢D3E3∪⌢E2E3 |
and
γ0=⌢D2K2∪⌢K2Q2∪⌢Q2B∪⌢BD2 |
in Hausdorff distance as ϵ→0, see Figures 3(c) and 4(c).
According to the above analysis and some numerical simulations by the "PPlane8" tool in Matlab, we give the phase portrait of system (1.4) with b=0.2, c=0.5 and h=0.45, see Figure 4(c). Note that the two unstable manifolds of saddle E2(u2,v2) tend to the stable node E3(u3,v3) and form the heteroclinic cycle. From Figure 4, we know that the number of equilibriums of system (1.4) changes from one to three when parameters pass through the saddle-node bifurcation surface. Furthermore, under some conditions, system (1.4) always has a small relaxation oscillation cycle γϵ which is separately surrounded by a big relaxation oscillation cycle, homoclinic cycle and heteroclinic cycle for different values of parameter h in the neighborhood of h1.
In this subsection, we will discuss the boundary equilibrium bifurcation of system (1.4). From Lemma 2.4, we know that the boundary equilibrium EB(1,1−b) is an admissible equilibrium for both systems in Σ(−) and Σ(+), which is a mark of boundary equilibrium bifurcation of codimension 1. Hence, incorporating Lemmas 2.3 and 2.5, we present the discontinuous saddle-node bifurcation surface as follows
DSN={(b,c,h)∣h=h2=1−c1−b}. |
When the parameters vary from one side of the surface to another, system (1.4) has two positive equilibriums E1(u1,v1) and E2(u2,v2) which are separately located in regions Σ(−) and Σ(+). Then the discontinuous saddle-node bifurcation gets two positive equilibriums.
In this paper, we consider the slow-fast Bazykin's predator-prey model with piecewise smooth Holling Ⅰ functional response. Our qualitative analysis on system (1.4) reveals that system has complex dynamics and bifurcations and the competition rate h plays a critical role which affects not only the number and type of equilibriums but also the type of bifurcations in this model. When the values of parameters vary, system (1.4) undergoes the saddle-node bifurcation and the discontinuous saddle-node bifurcation. For different values of parameters, it is clear that system (1.4) has a hyperbolically unstable relaxation oscillation cycle which is respectively surrounded by a hyperbolically stable relaxation oscillation cycle, a homoclinic cycle and a heteroclinic cycle. These complex dynamics cannot occur in the system (1.4) with single time scale and smooth Holling type Ⅰ functional response. In fact, the system (1.4) with ϵ>1 and smooth Holling type Ⅰ functional response [9] has at least one positive equilibrium which is globally stable if c<c1, here c1 is a threshold value. Hence, compared with our result in this paper, it is clear that the different time scale and piecewise-smooth functional response in system (1.4) can introduce more complex dynamical behaviour.
These complex dynamical phenomenons of system (1.4) show that the complexity of dynamical behaviors of the Bazykin's model. For the biological interpretations of these complex dynamical phenomenons, we interpret biological meaning of the big relaxation oscillation cycle in Theorem 3.1 as an example. By Figure 4(a), it is clear that the big relaxation oscillation cycle is hyperbolically stable, which indicates that that the prey and predator will coexist in system (1.4). The slow manifolds represent the density of prey and predator vary slowly under the influence of interactions or low prey density, and the fast movements implies the fast changes of prey density. The existence of the limit cycle suggests that the predator density increases slowly with sufficient food, and once the predator density exceeds the vertex point, the prey density decreases rapidly but not extinct due to large consumption. In the absence of food, the predator density also slowly declines until the prey density increases again quickly enough to support predator's reproduction, where the cycle begins again. Therefore, the relaxation oscillation cycle may show sudden outbreaks of pests in biology many years after extinction.
It is worth noticing that Kooij and Zegeling [40] extend the piecewise-smooth Holling type Ⅰ functional response, which replace the linear function p(x)=x, 0<x<1 by cubic function p(x)=x(1+(x−1)(a0+a1x)), 0<x<1. Note that the new functional response is not only a non-monotonic function but also contains different cases for different values of a1 and a2. Furthermore, there may be more complex dynamics and bifurcations for system (1.4) with this new function response. We leave these for future consideration.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The first author is supported by the Fundamental Research Funds for the Central Universities (No.2232023D-22, No.2232022G-13) and The third author was supported by the National Natural Science Foundation of China (12271088) and the Natural Science Foundation of Shanghai (21ZR1401000).
The authors declare there is no conflict of interest.
[1] | A. J. Lotka, Elements of Physical Biology, Williams and Wilkins, Baltimore, 1925. |
[2] |
V. Volterra, Fluctuations in the abundance of a species considered mathematically, Nature, (1926), 558–560. https://doi.org/10.1038/118558a0 doi: 10.1038/118558a0
![]() |
[3] | H. I. Freedman, Deterministic Mathematical Models in Population Ecology, Marcel Dekker, New York, 1980. |
[4] | G. F. Gause, The Struggle for Existence, Williams and Wilkins, Baltimore, 1935. |
[5] | A. D. Bazykin, Nonlinear Dynamics of Interacting Populations, World Scientific, Singapore, 1998. |
[6] | A. D. Bazykin, Volterra system and Michaelis-Menten equation, Problems of Mathematical Genetics, Novosibirsk State University, Novosibirsk, (1974), 103–143. |
[7] |
C. S. Holling, The components of predation as revealed by a study of small-mammal predation of the European pine swafly, Canad. Entomol., 91 (1959), 293–320. https://doi.org/10.4039/Ent91293-5 doi: 10.4039/Ent91293-5
![]() |
[8] |
C. S. Holling, The functional response of predators to prey density and its role in mimicry and population regulation, Mem. Entomol. Soc. Can., 97 (1965), 5–60. https://doi.org/10.4039/entm9745fv doi: 10.4039/entm9745fv
![]() |
[9] | A. D. Bazykin, Structural and dynamic stability of model predator-prey systems, Int. Inst. Appl. Syst. Anal., 1976. |
[10] | A. D. Bazykin, F. S. Berezovskaya, T. I. Buriev, Dynamics of predator–prey system including predator saturation and competition, in Faktory Raznoobraziya v Matematicheskoi Ekologii i Populyatsionnoi Genetike, Pushchino, (1980), 6–33. |
[11] |
J. Hainzl, Stability and Hopf bifurcation in a predator–prey system with several parameters, SIAM J. Appl. Math., 48 (1998), 170–190. https://doi.org/10.1137/0148008 doi: 10.1137/0148008
![]() |
[12] |
J. Hainzl, Multiparameter bifurcation of a predator-prey system, SIAM J. Math. Anal., 23 (1992), 150–180. https://doi.org/10.1137/0523008 doi: 10.1137/0523008
![]() |
[13] | M. Lu, J. C. Huang, Global analysis in Bazykin's model with Holling Ⅱ functional response and predator competition, J. Diff. Equation, 280 (2021), 99–138. |
[14] | Y. Kuznetsov, Elements of Applied Bifurcation Theory, 3nd edition, Springer, New York, 2004. |
[15] |
P. Chowdhury, S. Petrovskii, V. Volpert, M. Banerjee, Attractors and long transients in a spatio-temporal slow-fast Bazykin's model, Commun. Nonlinear Sci. Numer. Simul., 118 (2023), 107014. https://doi.org/10.1016/j.cnsns.2022.107014 doi: 10.1016/j.cnsns.2022.107014
![]() |
[16] |
M. Banerjee, S. Ghorai, N. Mukherjee, Approximated spiral and target patterns in Bazykin's prey-predator model: multiscale perturbation analysis, Int. J. Bifur. Chaos Appl. Sci. Engrg., 27(3) (2017), 1750038. https://doi.org/10.1142/S0218127417500389 doi: 10.1142/S0218127417500389
![]() |
[17] |
J. M. Zhang, L. J. Zhang, C. M. Khalique, Stability and Hopf bifurcation analysis on a Bazykin model with delay, Abstr. Appl. Anal., (2014), 539684. https://doi.org/10.1155/2014/539684 doi: 10.1155/2014/539684
![]() |
[18] |
S. Muratori, S. Rinaldi, Remarks on competitive coexistence, SIAM J. Appl. Math., 49(5) (1989), 1462–1472. https://doi.org/10.1137/0149088 doi: 10.1137/0149088
![]() |
[19] |
N. C. Stenseth, W. Falck, O. N. Bjornstad, C. J. Krebs, Population regulation in snowshoe hare and Canadian lynx: Asymmetric food web configurations between hare and lynx, Proc. Natl. Acad. Sci. USA, 94 (1997), 5147–5152. https://doi.org/10.1073/pnas.94.10.5147 doi: 10.1073/pnas.94.10.5147
![]() |
[20] |
M. Scheffer, S. Rinaldi, Y. A. Kuznetsov, E. H. Van Nes, Seasonal dynamics of Daphnia and algae explained as a periodically forced predator-prey system, Oikos, 80 (1997), 519–532. https://doi.org/10.2307/3546625 doi: 10.2307/3546625
![]() |
[21] |
D. Ludwig, D. D. Jones, C. S. Holling, Qualitative analysis of insect outbreak systems: the spruce budworm and forest, J. Anim. Ecol., 47 (1978), 315–332. https://doi.org/10.2307/3939 doi: 10.2307/3939
![]() |
[22] |
N. Finichel, Geometric singular perturbation theory for ordinary differential equations, J. Diff. Equation, 55 (1979), 763–783. https://doi.org/10.1016/0022-0396(79)90152-9 doi: 10.1016/0022-0396(79)90152-9
![]() |
[23] | C. Kuehn, Multiple Time Scale Dynamics, Springer, 2015. |
[24] |
W. Liu, Exchange lemmas for singular perturbation problems with certain turning points, J. Diff. Equation, 167 (2000), 134–180. https://doi.org/10.1006/jdeq.2000.3778 doi: 10.1006/jdeq.2000.3778
![]() |
[25] |
W. Liu, Geometric singular perturbations for multiple turning points: invariant manifolds and exchange lemmas, J. Dyn. Differ. Equations, 18 (2006), 667–691. https://doi.org/10.1007/s10884-006-9020-7 doi: 10.1007/s10884-006-9020-7
![]() |
[26] |
M. Krupa, P. Szmolyan, Extending geometric singular perturbation theory to nonhyperbolic points-fold and canard points in two dimensions, SIAM J. Math. Anal., 33 (2001), 286–314. https://doi.org/10.1137/S0036141099360919 doi: 10.1137/S0036141099360919
![]() |
[27] |
M. Krupa, P. Szmolyan, Relaxation oscillation and canard explosion, J. Diff. Equation, 174 (2001), 312–368. https://doi.org/10.1006/jdeq.2000.3929 doi: 10.1006/jdeq.2000.3929
![]() |
[28] |
P. De Maesschalck, S. Schecter, The entry-exit function and geometric singular perturbation theory, J. Diff. Equation, 260 (2016), 6697–6715. https://doi.org/10.1016/j.jde.2016.01.008 doi: 10.1016/j.jde.2016.01.008
![]() |
[29] |
C. Wang, X. Zhang, Stability loss delay and smoothness of the return map in slow-fast systems, SIAM J. Appl. Dyn. Syst., 17 (2018), 788–822. https://doi.org/10.1137/17M1130010 doi: 10.1137/17M1130010
![]() |
[30] | M. di Bernardo, C. J. Budd, A. R. Champneys, P. Kowalczyk, Piecewise-smooth dynamical systems: Theory and applications, Springer-Verlag London, London, 2008. |
[31] | D. J. W. Simpson, Bifurcations in Piecewise-Smooth Continuous Systems, 70nd edition, World Scientific, 2010. |
[32] |
A. Roberts, Canard explosion and relaxation oscillation in planar, piecewise-smooth, continuous systems, SIAM J. Appl. Dyn. Syst., 15(1) 2016,609–624. https://doi.org/10.1137/140998147 doi: 10.1137/140998147
![]() |
[33] |
S. M. Li, X. L. Wang, X. L. Li, K. L. Wu, Relaxation oscillations for Leslie-type predator-prey model with Holling type Ⅰ response functional function, Appl. Math. Lett., 120 (2021), 107328. https://doi.org/10.1016/j.aml.2021.107328 doi: 10.1016/j.aml.2021.107328
![]() |
[34] |
S. M. Li, C. Wang, K. L. Wu, Relaxation oscillations of a slow-fast predator-prey model with a piecewise smooth functional response, Appl. Math. Lett., 113 (2021), 106852. https://doi.org/10.1016/j.aml.2020.106852 doi: 10.1016/j.aml.2020.106852
![]() |
[35] |
T. Saha, P. J. Pal, M. Banerjee, Slow-fast analysis of a modified Lesile-Gower model with Holling type Ⅰ functional response, Nonlinear Dyn., 108 (2022), 4531–4555. https://doi.org/10.1007/s11071-022-07370-1 doi: 10.1007/s11071-022-07370-1
![]() |
[36] | D. Hanselman, Mastering Matlab, University of Maine, 2001. |
[37] | L. Perko, Differential Equations and Dynamical Systems, Springer, New York, 1991. |
[38] |
J. A. C. Medrado, J. Torregrosa, Uniqueness of limit cycles for sewing planar piecewise linear systems, J. Math. Anal. Appl., 431 (2015), 529–544. https://doi.org/10.1016/j.jmaa.2015.05.064 doi: 10.1016/j.jmaa.2015.05.064
![]() |
[39] |
X. Wu, M. K. Ni, Dynamics in diffusive Leslie-Gower prey-predator model with weak diffusion, Nonlinear Anal. Model. Control, 27 (2022), 1168–1188. https://doi.org/10.15388/namc.2022.27.29535 doi: 10.15388/namc.2022.27.29535
![]() |
[40] |
R. E. Kooij, A. Zegeling, Predator-prey models with non-analytical functional response, Chaos Solitons Fractals, 123 (2019), 163–172. https://doi.org/10.3934/dcdsb.2004.4.1065 doi: 10.3934/dcdsb.2004.4.1065
![]() |
1. | Xiao Wu, Zilai Zhou, Feng Xie, Multi-scale dynamics of a piecewise-smooth Bazykin’s prey–predator system, 2024, 0924-090X, 10.1007/s11071-024-10292-9 |
Holling type | Definition | Generalized form |
Ⅱ | p(u)=muu+k | |
Ⅲ | p(u)=mu2u2+k | p(u)=mu2k1u2+k2u+1(k2>−2√k1) |
Ⅳ | p(u)=muu2+k | p(u)=muk1u2+k2u+1(k2>−2√k1) |