
A two-patch model with additive Allee effect is proposed and studied in this paper. Our objective is to investigate how dispersal and additive Allee effect have an impact on the above model's dynamical behaviours. We discuss the local and global asymptotic stability of equilibria and the existence of the saddle-node bifurcation. Complete qualitative analysis on the model demonstrates that dispersal and Allee effect may lead to persistence or extinction in both patches. Also, combining mathematical analysis with numerical simulation, we verify that the total population abundance will increase when the Allee effect constant a increases or m decreases. And the total population density increases when the dispersal rate D1 increases or the dispersal rate D2 decreases.
Citation: Lijuan Chen, Tingting Liu, Fengde Chen. Stability and bifurcation in a two-patch model with additive Allee effect[J]. AIMS Mathematics, 2022, 7(1): 536-551. doi: 10.3934/math.2022034
[1] | Yudan Ma, Ming Zhao, Yunfei Du . Impact of the strong Allee effect in a predator-prey model. AIMS Mathematics, 2022, 7(9): 16296-16314. doi: 10.3934/math.2022890 |
[2] | Na Min, Hongyang Zhang, Xiaobin Gao, Pengyu Zeng . Impacts of hunting cooperation and prey harvesting in a Leslie-Gower prey-predator system with strong Allee effect. AIMS Mathematics, 2024, 9(12): 34618-34646. doi: 10.3934/math.20241649 |
[3] | Fatao Wang, Ruizhi Yang, Yining Xie, Jing Zhao . Hopf bifurcation in a delayed reaction diffusion predator-prey model with weak Allee effect on prey and fear effect on predator. AIMS Mathematics, 2023, 8(8): 17719-17743. doi: 10.3934/math.2023905 |
[4] | Weili Kong, Yuanfu Shao . Bifurcations of a Leslie-Gower predator-prey model with fear, strong Allee effect and hunting cooperation. AIMS Mathematics, 2024, 9(11): 31607-31635. doi: 10.3934/math.20241520 |
[5] | Khairul Saleh . Basins of attraction in a modified ratio-dependent predator-prey model with prey refugee. AIMS Mathematics, 2022, 7(8): 14875-14894. doi: 10.3934/math.2022816 |
[6] | Ali Al Khabyah, Rizwan Ahmed, Muhammad Saeed Akram, Shehraz Akhtar . Stability, bifurcation, and chaos control in a discrete predator-prey model with strong Allee effect. AIMS Mathematics, 2023, 8(4): 8060-8081. doi: 10.3934/math.2023408 |
[7] | Reshma K P, Ankit Kumar . Stability and bifurcation in a predator-prey system with effect of fear and additional food. AIMS Mathematics, 2024, 9(2): 4211-4240. doi: 10.3934/math.2024208 |
[8] | Vinoth Seralan, R. Vadivel, Dimplekumar Chalishajar, Nallappan Gunasekaran . Dynamical complexities and chaos control in a Ricker type predator-prey model with additive Allee effect. AIMS Mathematics, 2023, 8(10): 22896-22923. doi: 10.3934/math.20231165 |
[9] | Xiaoming Su, Jiahui Wang, Adiya Bao . Stability analysis and chaos control in a discrete predator-prey system with Allee effect, fear effect, and refuge. AIMS Mathematics, 2024, 9(5): 13462-13491. doi: 10.3934/math.2024656 |
[10] | Chengchong Lu, Xinxin Liu, Zhicheng Li . The dynamics and harvesting strategies of a predator-prey system with Allee effect on prey. AIMS Mathematics, 2023, 8(12): 28897-28925. doi: 10.3934/math.20231481 |
A two-patch model with additive Allee effect is proposed and studied in this paper. Our objective is to investigate how dispersal and additive Allee effect have an impact on the above model's dynamical behaviours. We discuss the local and global asymptotic stability of equilibria and the existence of the saddle-node bifurcation. Complete qualitative analysis on the model demonstrates that dispersal and Allee effect may lead to persistence or extinction in both patches. Also, combining mathematical analysis with numerical simulation, we verify that the total population abundance will increase when the Allee effect constant a increases or m decreases. And the total population density increases when the dispersal rate D1 increases or the dispersal rate D2 decreases.
With the continuous development of the country and on-going progress in society, there are more and more empirical results show that the destruction of nature by human has affected the habitat of the population [1]. The habitat of the population is first fragmented, then splits into patches, and finally becomes extinct, which will seriously threaten the survival and diversity of biological populations. However, if ecological corridor can be established to allow the species to migrate between different patches, it may become one of the meaningful measures to protect the species from extinction. Therefore, there have been a large body of literatures in modeling population migration and analyzing the protection and development of ecosystem, which provide biological insight for ecological management.
The dynamics of plaque population can be traced back to the year of 1951 (see [2]). Subsequently, in 1977, Freedman and Waltman [3] consider a two-patch model with a single species in logistic population growth as follows.
dN1dt=r1N1(1−N1K1)+D(N2−N1),dN2dt=r2N2(1−N2K2)+D(N1−N2), | (1.1) |
where Ni(i=1,2) represents the population density in patch i, parameter ri is the intrinsic growth rate, Ki is carrying capacity and D is the dispersal rate. Freedman and Waltman show that under certain conditions, the total population abundance of a spatially distributed population with dispersal can be larger than the total carrying capacities K1+K2. After that, a more in-depth study of this model is derived by DeAngelis et al. [4], Holt [5], DeAngelis and Zhang [6] and Roger Arditi et al. [7].
However, in nature, the dispersal between patches is often asymmetric. It is necessary to consider asymmetric dispersal with general dispersal rates. Arditi et al. [8] extend the model (1.1) by considering asymmetric dispersals. Moreover, recently, Wu et al. [9] characterize the following two-patch model:
dN1dt=r1N1(1−N1K1)+D(N2−sN1),dN2dt=−ˉr2N2+D(sN1−N2). | (1.2) |
Parameter D represents the dispersal intensity and s reflects the dispersal asymmetry. The authors show that the population persists in both patches but reaches a density which is less (higher) than that without dispersal at intermediate (small) dispersal rates, while extremely large dispersal could result in the population extinction. And asymmetric dispersal is more favorable than symmetric dispersal under certain conditions. For more relevant works, one can see [10,11]. In recent years, there has been growing interest in understanding the roles that Allee effect [12] plays in the interaction between population. Due to its strong possible impact on population dynamics, Allee effect has been observed extensively in ecology, biology and epidemiology and so on [13,14,15]. For example, Kang et al. [16] have considered two-patch model with Allee effect and dispersal
dxdt=rx(x−θ)(1−x)+μ(y−x),dydt=ry(y−θ)(1−y)+μ(x−y) | (1.3) |
where x(t) and y(t) denote the population density at time t in different patches. The parameters μ∈[0,1] and θ represent the dispersal intensity and Allee threshold, respectively. It was shown that the dispersal parameter μ and the Allee threshold θ will affect the global dynamics. Subsequently, about two-patch models with dispersal, studies of Allee effect on interacting population have been done recently in [17,18].
Particularly, the equation
dxdt=r[(1−xK)−mx+a]x | (1.4) |
is suffered from the additive Allee effect which was firstly introduced in [19]. The positive parameters m and a are the Allee effect constants. The additive Allee effect consists of two cases, i.e., weak and strong Allee effects. That is, if 0<m<a, it is the weak Allee effect; if m>a, it is the strong Allee effect. Consequently, there are a lot of excellent works in mathematical, biology and ecology models with additive Allee effect [20,21,22,23]. For example, in [20], the authors propose a single species logistic model with feedback control and additive Allee effect in the growth of species. The result shows that Allee effect can make the above system's dynamic behaviors more complex. In [21], the authors consider an SI epidemic model incorporating additive Allee effect and time delay. They obtain that both additive Allee effect and time delay have vital effect on the prevalence of the disease. It brings to our attention that, the literature which has investigated the two-patch model with additive Allee effect is rather few. To the best of the authors' knowledge, in [24], analytical conditions for the global stability of the patchy model in the case of weak Allee effect. However, in the case of strong Allee effect, only numerical simulations are provided. In other words, the author [24] has not provided a complete qualitative analysis since it involves an algebraic equation of nine-degree. Moreover, the occurrence of bifurcation phenomenon due to the Allee effect still remain unknown. Motivated by the above, in this paper, we will consider the following one species model with additive Allee effect and dispersal:
dx1dt=−x1+D2x2−D1x1,dx2dt=x2(1−x2−mx2+a)+D1x1−D2x2 | (1.5) |
where x1 and x2 represent the densities of the species in patch 1 and 2, respectively. Similar to that in [9], in this paper, we assume that patch 1 is the sink (i.e., the coefficient of x1 is negative). And the species in patch 2 is affected by the additive Allee effect (patch 2 is the source, i.e., the coefficient of x2 is positive). From the view point of biology, we only need to focus on the nonnegative solution of system (1.5). Thus it is assumed that the initial conditions of (1.5) are x1(0)≥0, x2(0)≥0.
Different from the strong Allee effect in [16,17,18], this paper will pay attention to investigate the patchy model with additive Allee effect, i.e., the system (1.5). We will try our best to present the possible qualitative behavior and bifurcation phenomena. Firstly, we discuss the existence and stability of all nonnegative equilibria. Throughout this paper, we will denote local asymptotically stable by LAS. Then, by the Bendixson-Dulac discriminant, if certain condition holds, the unique positive equilibrium in system (1.5) is globally asymptotically stable which indicates that the species will persist in both patches. Secondly, we obtain that system (1.5) undergoes the saddle-node bifurcation. Finally, we investigate the effect of Allee effect and dispersal on total population abundance. We notice that comparing with [9,24], system (1.5) is experiencing more complex dynamics such as saddle-node bifurcation when incorporating the additive Allee effect. The main result in this paper is a good extension and supplement that of [9,24].
The paper is organized as follows. Sections 2 represents the existence and stability of the equilibria. In Section 3, we will discuss the saddle-node bifurcation of the system. In Section 4, we show the effect of Allee effect and dispersal on total population abundance. Brief discussions are presented finally in Section 5.
In this section, we shall discuss the existence and stability of all nonnegative equilibria for system (1.5). It is easy to check that the system (1.5) always has a trivial equilibrium E0(0,0). For other non-trivial equilibria, we shall investigate them as follows.
The equilibrium of system (1.5) is determined by the following equation:
{−x1+D2x2−D1x1=0,x2(1−x2−mx2+a)+D1x1−D2x2=0. | (2.1) |
For the possible equilibria, we consider the positive solution of the following equation:
x22−(1−D21+D1−a)x2+[m−a(1−D21+D1)]=0. | (2.2) |
Let △(m) be the discriminant of Eq (2.2), i.e.,
△(m)=(1−D21+D1+a)2−4m. | (2.3) |
From the above, we have the unique root of △(m)=0 and denote it by m∗ as follows.
m∗=(1−D21+D1+a)24. | (2.4) |
It follows that △(m)>0, △(m)=0 and △(m)<0 when m<m∗, m=m∗ and m>m∗, respectively.
The existence of the positive equilibria is obtained as follows.
Theorem 2.1. Suppose that a<1−D21+D1.
(i) If m≤a(1−D21+D1), system (1.5) has a unique positive equilibrium E1(x11,x21).
(ii) If a(1−D21+D1)<m<m∗, system (1.5) has two distinct positive equilibria E1(x11,x21) and E2(x12,x22).
(iii) If m=m∗, system (1.5) has a unique positive equilibrium E3(x13,x23).
(iv) If m>m∗, system (1.5) has no positive equilibria.
Theorem 2.2. Suppose that a≥1−D21+D1.
(i) If m<a(1−D21+D1), then system (1.5) has a unique positive equilibrium E1(x11,x21).
(ii) If m≥a(1−D21+D1), then system (1.5) has no positive equilibrium.
In the above, x21=12(1−D21+D1−a+√△(m)),x22=12(1−D21+D1−a−√△(m)),x23=12(1−D21+D1−a) and x1i=D21+D1x2i, i=1,2,3.
Moreover, the Jacobian matrix at E(x1, x2) of system (1.5) is
JE=[−1−D1D2D11−2x2−ma(x2+a)2−D2]. | (2.5) |
Thus, the determinant and the trace of the Jacobian matrix at E0(0,0) is
Det[JE0]=(ma−1)(1+D1)+D2, |
Tr[JE0]=−ma−D1−D2<0. |
So if m<a(1−D21+D1), we have Det[JE0]<0, then E0(0,0) is a saddle. If m>a(1−D21+D1), we have Det[JE0]>0, then E0(0,0) is stable. If m=a(1−D21+D1), we have Det[JE0]=0, i.e., JE0 has a unique zero eigenvalue. We will analyze that E0 is an attracting saddle-node as follows. For system (1.5), we apply the Taylor expansion of mx2+a, it can be rewritten as
{˙x1=(−1−D1)x1+D2x2,˙x2=−D1D21+D1x2+D1x1+(−1+ma2)x22−ma3x32+Q(x2) | (2.6) |
where Q(x2) denotes the power sery with term xj2 satisfying j≥4.
Moreover, we make the following transformation:
(x1x2)=(1−1+D1D11+D1D21)(ˉx1ˉx2), | (2.7) |
then system (1.5) becomes
{˙ˉx1=p20ˉx21+p11ˉx1ˉx2+p02ˉx22+p30ˉx31+p12ˉx1ˉx22+p21ˉx21ˉx2+p03ˉx32+P1(ˉx1,ˉx2),˙ˉx2=q01ˉx2+q02ˉx22+q11ˉx1ˉx2+q20ˉx21+q21ˉx21ˉx2+q12ˉx1ˉx22+q03ˉx32+q30ˉx31+P2(ˉx1,ˉx2) | (2.8) |
where P1(ˉx1,ˉx2), P2(ˉx1,ˉx2) denote the power series with term ˉxi1ˉxj2 satisfying i+j≥4 and
p20=(1+D1)3(m−a2)a2D2[D1D2+(1+D1)2], p11=2(1+D1)2(m−a2)a2[D1D2+(1+D1)2], p02=D2(1+D1)(m−a2)a2[D1D2+(1+D1)2],
p30=−m(1+D1)4a3D22[D1D2+(1+D1)2], p12=−3m(1+D1)2a3[D1D2+(1+D1)2], p21=−3m(1+D1)3a3D2[D1D2+(1+D1)2],
p03=−m(1+D1)D2a3[D1D2+(1+D1)2], q01=−D1D2+(1+D1)21+D1, q02=D1D2(m−a2)a2[D1D2+(1+D1)2],
q11=2(D1+D21)(m−a2)a2[D1D2+(1+D1)2], q20=(1+D1)2D1(m−a2)a2D2[D1D2+(1+D1)2], q21=−3m(1+D1)2D1a3D2[D1D2+(1+D1)2],
q12=−3m(1+D1)D1a3[D1D2+(1+D1)2], q03=−mD1D2a3[D1D2+(1+D1)2], q30=−m(1+D1)3D1a3D22[D1D2+(1+D1)2].
Introducing a new time variable τ=−D1D2+(1+D1)21+D1t, for the sake of simplicity, we still retain t to τ and obtain
{˙ˉx1=v20ˉx21+v11ˉx1ˉx2+v02ˉx22+v30ˉx31+v12ˉx1ˉx22+v21ˉx21ˉx2+v03ˉx32+Q1(ˉx1,ˉx2),˙ˉx2=ˉx2+w02ˉx22+w11ˉx1ˉx2+w20ˉx21+w21ˉx21ˉx2+w12ˉx1ˉx22+w03ˉx32+w30ˉx31+Q2(ˉx1,ˉx2) | (2.9) |
where vij=−pij(1+D1)D1D2+(1+D1)2, wij=−qij(1+D1)D1D2+(1+D1)2, pij, qij(i+j=2,3) are the same as those in (2.8). When m=a(1−D21+D1), the coefficient of ˉx21 can be simplified as v20=−(1+D1)4(1−D21+D1−a)aD2[D1D2+(1+D1)2]2<0. Considering the new time variable τ and using Theorem 7.1 in [25], we can conclude that the equilibrium E0 is an attracting saddle-node.
Following, we concentrate on the stability of coexistence equilibria Ei(x1i,x2i), i=1,2. The corresponding Jacobian matrix at Ei(x1i, x2i) is given by
JEi=[−1−D1D2D11−2x2i−ma(x2i+a)2−D2]. | (2.10) |
Combining with (2.1), the determinant and trace of the Jacobian matrix can be simplified as
Det[JEi]=x2i(x2i+a)2((x2i+a)2−m)(1+D1), |
Tr[JEi]=−D1−2x2i−ma(x2i+a)2−D2<0. |
Furthermore, we can get
Det[JE1]=x21√Δ(m)(√Δ(m)+√Δ(m)+4m)(1+D1)2(x21+a)2>0, |
Det[JE2]=x22√Δ(m)(√Δ(m)−√Δ(m)+4m)(1+D1)2(x22+a)2<0. |
Thus E1 is LAS and E2 is a saddle.
We present the above discussion in Theorem 2.3 and Theorem 2.4.
Theorem 2.3. Suppose that a<1−D21+D1.
(i) If m<a(1−D21+D1), then E0(0,0) is a saddle and E1(x11,x21) is LAS;
(ii) If m=a(1−D21+D1), then E0(0,0) is an attracting saddle-node and E1(x11,x21) is LAS;
(iii) If a(1−D21+D1)<m<m∗=(1−D21+D1+a)24, then E0(0,0) and E1(x11,x21) are LAS and E2(x12,x22) is a saddle;
(iv) If m=m∗, then E0(0,0) is LAS and E3(x13,x23) is an attracting saddle-node;
(v) If m>m∗, then E0(0,0) is LAS and there exists no positive equilibrium.
Proof. We will only need to prove that E3 is an attracting saddle-node when m=m∗. Let X=x1−x13, ˉX=x2−x23, system (1.5) can be transformed to the following system:
{˙X=−X+D2ˉX−D1X,˙ˉX=(ˉX+x23)[1−ˉX−x23−mˉX+x23+a]+D1(X+x13)−D2(ˉX+x23). | (2.11) |
Applying the Taylor expansion of mˉX+x23+a, system (2.11) can be rewritten as
{˙X=−X+D2ˉX−D1X,˙ˉX=−D1D21+D1ˉX+D1X+(−1+ma(x23+a)3)ˉX2−m(x23+a)3ˉX3+Q(ˉX) | (2.12) |
where Q(ˉX) denotes the power sery with term ˉXj satisfying j≥4. The Jacobian matrix of system (2.12) evaluated at the origin can be calculated as
JE3=[−1−D1D2D1−D1D21+D1], | (2.13) |
then the eigenvalues of JE3 are λ1=0 and λ2=−1−D1−D1D21+D1. We make the following transformation:
(XˉX)=(1−1+D1D11+D1D21)(x1x2), | (2.14) |
then system (2.12) becomes
{˙x1=c20x21+c11x1x2+c02x22+c30x31+c12x1x22+c21x21x2+c03x32+P1(x1,x2),˙x2=d01x2+d02x22+d11x1x2+d20x21+d21x21x2+d12x1x22+d03x32+d30x31+P2(x1,x2) | (2.15) |
where P1(x1,x2), P2(x1,x2) denote the power series with term xi1xj2 satisfying i+j≥4 and
c20=(1+D1)3[−(x23+a)3+ma](x23+a)3D2[D1D2+(1+D1)2], c11=2(1+D1)2[−(x23+a)3+ma](x23+a)3[D1D2+(1+D1)2], c02=D2(1+D1)[−(x23+a)3+ma](x23+a)3[D1D2+(1+D1)2],
c30=−m(1+D1)4(x23+a)3D22[D1D2+(1+D1)2], c12=−3m(1+D1)2(x23+a)3[D1D2+(1+D1)2], c21=−3m(1+D1)3(x23+a)3D2[D1D2+(1+D1)2],
c03=−m(1+D1)D2(x23+a)3[D1D2+(1+D1)2], d01=−1+2D1+D1D2+D211+D1, d02=D1D2[−(x23+a)3+ma](x23+a)3[D1D2+(1+D1)2],
d11=2(D1+D21)[−(x23+a)3+ma](x23+a)3[D1D2+(1+D1)2], d20=(1+D1)2D1[−(x23+a)3+ma](x23+a)3D2[D1D2+(1+D1)2], d21=−3m(1+D1)2D1(x23+a)3D2[D1D2+(1+D1)2],
d12=−3m(1+D1)D1(x23+a)3[D1D2+(1+D1)2], d03=−mD1D2(x23+a)3[D1D2+(1+D1)2], d30=−m(1+D1)3D1(x23+a)3D22[D1D2+(1+D1)2].
Introducing a new time variable τ=−D1D2+(1+D1)21+D1t, for the sake of simplicity, we still retain t to denote τ and obtain
{˙x1=e20x21+e11x1x2+e02x22+e30x31+e12x1x22+e21x21x2+e03x32+Q1(x1,x2),˙x2=x2+f02x22+f11x1x2+f20x21+f21x21x2+f12x1x22+f03x32+f30x31+Q2(x1,x2) | (2.16) |
where eij=−cij(1+D1)D1D2+(1+D1)2, fij=−dij(1+D1)D1D2+(1+D1)2, cij, dij(i+j=2,3) are the same as those in (2.15). When m=m∗, the coefficient of x2 can be simplified as e20=−(1+D1)4[−(x23+a)3+m∗a](x23+a)3D2[D1D2+(1+D1)2]2>0. Here we use −(x23+a)3+m∗a=18(a−1+D21+D1)(a+1−D21+D1)2<0. Considering the new time variable τ and using Theorem 7.1 in [25], we can conclude that the equilibrium E3 is an attracting saddle-node. This completes the proof.
Similarly, we can obtain the result as follows.
Theorem 2.4. Suppose that a≥1−D21+D1.
(i) If m<a(1−D21+D1), then E0(0,0) is a saddle and E1(x11,x21) is LAS;
(ii) If m=a(1−D21+D1), then E0(0,0) is an attracting saddle-node and E1(x11,x21) is LAS;
(iii) if m>a(1−D21+D1), then E0(0,0) is LAS and there exists no positive equilibrium.
Corollary 2.1. When m≤a(1−D21+D1), for system (1.5), the unique positive equilibrium E1(x11,x21) is globally asymptotically stable.
Proof. As was shown in Theorems 2.3 and 2.4, E1 is LAS when m≤a(1−D21+D1). To prove that E1 is globally asymptotically stable, we consider the Dulac function u(x1,x2)=1. We can get
∂(uP)∂x1+∂(uQ)∂x2=−2−D1−2x2−ma(x2+a)2−D2<0 |
where
P(x1,x2)=−x1+D2x2−D1x1, |
Q(x1,x2)=x2(1−x2−mx2+a)+D1x1−D2x2. |
According to the Bendixson-Dulac discriminant, system (1.5) has no limit cycle in the first quadrant. And when m≤a(1−D21+D1), E0 is unstable in the first quadrant. Thus one can deduce that E1(x11,x21) is global asymptotically stable when m≤a(1−D21+D1). The proof of Corollary 2.1 is finished.
Corollary 2.2. When a<1−D21+D1 and m>m∗, for system (1.5), the trivial equilibrium E0(0,0) is globally asymptotically stable.
Proof. As was mentioned in Theorem 2.3, E0 is LAS when a<1−D21+D1 and m>m∗. Also, for system (1.5), we can declare that under the initial conditions x1(0)≥0 and x2(0)≥0, the solution (x1(t), x2(t)) satisfies xi(t)≥0(i=1, 2) when t≥0. Actually, we can prove x1(t)≥0(t≥0) as follows. Moreover, from the first equation of system (1.5), we can obtain that ˙x1(t0)=D2x2(t0)≥0. The above leads to a contradiction with ˙x1(t0)<0. So it follows that x1(t)≥0(t≥0). Similarly, we have x2(t)≥0(t≥0). Moreover, we have d(x1+x2)dt=−x1+x2(1−x2−mx2+a)≤−x1+x2(1−x2)=−(x1+x2)+x2(2−x2)≤1−(x1+x2) which shows that limt→+∞supxi(t)≤1, i=1, 2. In other words, the solution of system (1.5) is ultimately bounded and then the solution can't extend to the infinity when t is large sufficiently. Also, there exists no limit cycle since system (1.5) has no other equilibrium. Thus E0(0,0) is global asymptotically stable. The proof of Corollary 2.2 is finished.
Similarly, we can obtain the result as follows.
Corollary 2.3. When a≥1−D21+D1 and m>a(1−D21+D1), for system (1.5), the trivial equilibrium E0(0,0) is globally asymptotically stable.\
From Corollaries 2.1–2.3, we can conclude that the additive Allee effect m and a is closely related with the population survival. In detail, when m≤a(1−D21+D1), the species in both patches can survive. However, when m>a(1−D21+D1), the species may still be persistent or become extinguished. Actually, when a≥1−D21+D1, m>a(1−D21+D1) or a<1−D21+D1, m>m∗, both species will be extinct. In other words, the larger the Allee effect constant m, the more likely the species will become extinct. The phase ports of system (1.5) is given in Figure 1.
From Section 2, we know that if a<m<m∗, system (1.5) has two positive equilibria, i.e., E1(x11,x21) and E2(x12,x22); if m=m∗, the system has a unique positive equilibrium E3(x13,x23); if m>m∗, the system has no positive equilibria. The above indicates that there will be the occurrence of saddle-node bifurcation. Next we will prove that system (1.5) undergoes the saddle-node bifurcation when m=mSN≜(1−D21+D1+a)24. And the numerical simulation of the saddle-node bifurcation is shown in Figure 2.
Theorem 3.1. System (1.5) undergoes the saddle-node bifurcation when m=mSN.
Proof. We use Sotomayor's theorem in [26] to verify the transversality condition for the occurrence of saddle-node bifurcation at m=mSN. Notice that the determinant and the trace of the corresponding Jacobian matrix can be given by Det[JE3]=0 and Tr[JE3]=−1−D1−D1D21+D1<0. Then JE3 has a unique zero eigenvalue, named by λ1.
For the matrices JE3 and JTE3, let V and W represent two eigenvectors corresponding to the eigenvalue λ1, respectively. Then we can get
V=(V1V2)=(11+D1D2),W=(W1W2)=(11+D1D1). | (3.1) |
Furthermore, we obtain
Fm(E3;mSN)=(0(1−D21+D1−a)[2(−1+D21+D1−a)+(1−D21+D1+a)2]2(1−D21+D1+a)2), | (3.2) |
D2F(E3;mSN)(V,V)=(∂2F1∂x21V21+2∂2F1∂x1∂x2V1V2+∂2F1∂x22V22∂2F2∂x21V21+2∂2F2∂x1∂x2V1V2+∂2F2∂x22V22)(E3;mSN)=(02(−1+D21+D1+a)(1+D1)2(1−D21+D1+a)D22). | (3.3) |
Here F1=−x1+D2x2−D1x1 and F2=x2(1−x2−mx2+a)+D1x1−D2x2. According to (3.2) and (3.3), it follows that
WTFm(E3;mSN)=(1+D1)(1−D21+D1−a)[2(−1+D21+D1−a)+(1−D21+D1+a)2]2D1(1−D21+D1+a)2≠0,WT[D2F(E3;mSN)(V,V)]=2(−1+D21+D1+a)(1+D1)3(1−D21+D1+a)D1D22≠0 |
which means that the saddle-node bifurcation occurs at E3 when m=mSN. Thus we complete the proof.
Through the previous analysis, we can see that when the parameters satisfy certain conditions, the unique positive equilibrium E1(x11,x21) is globally asymptotically stable. At this time, the total population abundance is
T1=x11+x21=(1+D1+D2)(1−D21+D1−a+√△(m))2(1+D1). |
Following, we will explore the effect of Allee effect and dispersal on the total population abundance, respectively.
Firstly, by simple calculation, we have
dT1da=1+D1+D22(1+D1)(−1+1−D21+D1+a√Δ(m))>0, |
dT1dm=−1+D1+D2(1+D1)√Δ(m)<0. |
Thus we can obtain that the total population abundance will increase when the Allee effect constant a increases or m decreases. Let D1=1, D2=0.5. In Figure 3(a), (b), we have used the software of matlab to plot the bifurcation diagram when a=0.2 or m=0.2, respectively. The bifurcation diagram shows the maximum and minimum density of species 1, i.e., x11 and x21 as Allee effect m or a is increasing. Here x11 and x12 are regarded as functions of m, a which are plotted in Figure 3(a), (b), respectively. It follows that Allee effect has a vital influence on the species. In detail, the density of the species decreases when the Allee effect coefficients m increases or a decreases.
Secondly, we will explore the possible impact of dispersal on the total population abundance through numerical simulation. We regard the total population abundance T1 as a function of D1 or D2, respectively. In Figure 4(a), let m=0.2, a=0.5, D2=1, it follows that T1 increases when D1 increases. In Figure 4(b), let m=0.02, a=0.5, D1=10, it follows T1 decreases when D2 increases. Especially, when m=m∗, the system (1.5) has a unique positive equilibrium E3. Then the corresponding total population abundance T1=x13+x23=(1+D1+D2)(1−D21+D1−a)2(1+D1). A direct computation shows dT1dD1=D22(1+D1)2[a+2D21+D1]>0, dT1dD2=−12(1+D1)(a+2D21+D1)<0. This observation is consistent with the numerical simulation in Figure 4. Actually, it is not difficult to understand the above result biologically. Notice that D1(D2) is the dispersal rate of the species from patch 1(2) to patch 2(1) and the intrinsic growth rate of the species is negative (positive) in patch 1(2) when there is no dispersal. As a result, once D1 increases, the species easily migrates from patch 1 to patch 2, that is, migrates to a patch with a larger growth rate. The migration will result in the increase of the total population abundance. On the contrary, when D2 increases, the species easily migrates from patch 2 to patch 1, which makes the species more likely to be extinct and then reduce the total population abundance.
In this paper, we have proposed and studied a two-patch model with additive Allee effect and dispersal. Our goal of this investigation is to discuss how the above model's dynamical behaviour such as the persistence and extinction is influenced by the additive Allee effect and dispersal. The model always has a trivial equilibrium E0(0,0) which is a saddle or a stable node when m<a(1−D21−D1) or m>a(1−D21−D1), respectively. For the existence and stability of the positive equilibrium, we have considered two cases, i.e., a<1−D21−D1 and a≥1−D21−D1 in Theorems 2.3 and 2.4. In detail, when a<1−D21−D1, there may exist a stable positive equilibrium E1, two different positive equilibria E1, E2(E1 is stable and E2 is a saddle), an attracting saddle-node E3 or no positive equilibrium under the case m≤a(1−D21−D1), a(1−D21−D1)<m<m∗, m=m∗, m>m∗, respectively. On the other hand, when a<1−D21−D1, there may exist a stable positive equilibrium E1 or no positive equilibrium under the case m≤a(1−D21−D1), m>a(1−D21−D1), respectively. Furthermore, in Theorem 3.1, we have revealed that the system can undergo the saddle-node bifurcation around the unique positive equilibrium E3.
In all, the above results indicate that the additive Allee effect and dispersal have great impact on the species permanence, extinction and stability. Especially, according to Corollary 2.1, E1 is globally asymptotically stable which shows that the species in both patches coexist when m≤a(1−D21−D1). From Corollaries 2.2 and 2.3, one can find that both species go extinct when a<1−D21−D1, m>m∗ or a≥1−D21−D1, m>a(1−D21−D1). Furthermore, in order to study the effect of the Allee effect and dispersal on the population more concisely, we also investigate the total population abundance and have found that it will increase when the Allee effect constant a increases or m decreases. Also, we find that the total population density increases when the dispersal rate D1 increases or D2 decreases. Additive Allee effect can reduce complex dynamics and bifurcation. Actually, many literatures have investigated system (1.5) when m=0. For example, in [9], system (1.5) with m=0 has at most a positive equilibrium. However, when m>0, there may exist one or two positive equilibria and even the emergence of the bifurcation. Also, for the strong Allee effect case in (1.5), we provide a rigorous qualitative analysis other than numerical simulation which has been done in [24]. The main results in this paper can be seen as a good supplement to the work in [9] and [24].
Here we have to point out that this paper investigates a patchy model in which the species in patch 1 is assumed to die exponentially and the system only admits saddle-node bifurcation. As we know, it is also meaningful to consider the logistic growth rate for the species in one of the patches and the species in the other patch is suffered from the additive Allee effect. In this case, the above system may exhibit at most six positive equilibria and even undergo complex bifurcation such as Hopf bifurcation and Bogdanov-Takens bifurcation. It is rather difficult for us to present a rigorous qualitative analysis at present. We hope to consider this aspect for future work by use of "bifurcation control" and "parametric normal forms" in [27,28,29,30,31,32,33].
This work was supported the Natural Science Foundation of Fujian Province (2021J01614, 2021J01613).
The authors declare that there is no conflict of interest regarding the publication of this paper.
[1] |
H. Andren, Effects of habitat fragmentation on birds and mammals in landscapes with different proportions of suitable habitat: A review, Oikos, 71 (1994), 355–366. doi: 10.2307/3545823. doi: 10.2307/3545823
![]() |
[2] |
J. G. Skellam, Random dispersal in theoretical populations, Biometrika, 38 (1951), 196–218. doi: 10.2307/2332328. doi: 10.2307/2332328
![]() |
[3] |
H. I. Freedman, P. Waltman, Mathematical models of population interactions with dispersal. I: Stability of two habitats with and without a predator, SIAM J. App. Math., 32 (1977), 631–648. doi: 10.1137/0132052. doi: 10.1137/0132052
![]() |
[4] |
D. L. DeAngelis, C. C. Travis, W. M. Post, Persistence and stability of seed-dispersed species in a patchy environment, Theor. Popul. Biol., 16 (1979), 107–125. doi: 10.1016/0040-5809(79)90008-X. doi: 10.1016/0040-5809(79)90008-X
![]() |
[5] |
R. D. Holt, Population dynamics in two-patch environments: Some anomalous consequences of an optimal habitat distribution, Theor. Popul. Biol., 28 (1985), 181–208. doi: 10.1016/0040-5809(85)90027-9. doi: 10.1016/0040-5809(85)90027-9
![]() |
[6] |
D. L. DeAngelis, B. Zhang, Effects of dispersal in a non-uniform environment on population dynamics and competition: A patch model approach, Discrete Cont. Dyn-A., 19 (2014), 3087–3104. doi: 10.3934/dcdsb.2014.19.3087. doi: 10.3934/dcdsb.2014.19.3087
![]() |
[7] |
A. Roger, L. Claude, S. Tewfik, Is dispersal always beneficial to carrying capacity? New insights from the multi-patch logistic equation, Theor. Popul. Biol., 106 (2015), 45–59. doi: 10.1016/j.tpb.2015.10.001. doi: 10.1016/j.tpb.2015.10.001
![]() |
[8] |
R. Arditi, C. Lobry, T. Sari, Asymmetric dispersal in the multi-patch logistic equation, Theor. Popul. Biol., 120 (2018), 11–15. doi: 10.1016/j.tpb.2017.12.006. doi: 10.1016/j.tpb.2017.12.006
![]() |
[9] |
H. Wu, Y. S. Wang, Y. F. Li, D. L. DeAngelis, Dispersal asymmetry in a two-patch system with source-sink populations, Theor. Popul. Biol., 131 (2020), 54–65. doi: 10.1016/j.tpb.2019.11.004. doi: 10.1016/j.tpb.2019.11.004
![]() |
[10] |
R. Huang, Y. S. Wang, H. Wu, Population abundance in predator-prey systems with predator's dispersal between two patches, Theor. Popul. Biol., 135 (2020), 1–8. doi: 10.1016/j.tpb.2020.06.002. doi: 10.1016/j.tpb.2020.06.002
![]() |
[11] |
M. Fang, Y. Wang, M. Chen, L. DeAngelis, Asymptotic population abundance of a two-patch system with asymmetric diffusion, Discrete Cont. Dyn-A., 40 (2020), 3411–3425. doi: 10.3934/dcds.2020031. doi: 10.3934/dcds.2020031
![]() |
[12] |
P. A. Stephens, W. J. Sutherland, R. P. Freckleton, What is the Allee effect, Oikos, 87 (1999), 185–190. doi: 10.2307/3547011. doi: 10.2307/3547011
![]() |
[13] |
P. M. Hao, X. C. Wang, J. J. Wei, Hopf bifurcation analysis of a diffusive single species model with stage structure and strong Allee effect, Math. Comput. Simulat., 153 (2018), 1–14. doi: 10.1016/j.matcom.2018.05.004. doi: 10.1016/j.matcom.2018.05.004
![]() |
[14] |
L. M. Zhang, C. F. Zhang, Z. R. He, Codimension-one and codimension-two bifurcations of a discrete predator-prey system with strong Allee effect, Math. Comput. Simulat., 162 (2019), 155–178. doi: 10.1016/j.matcom.2019.01.006. doi: 10.1016/j.matcom.2019.01.006
![]() |
[15] |
C. Arancibia-Ibarra, J. Flores, Dynamics of a Leslie-Gower predator-prey model with Holling type II functional response, Allee effect and a generalist predator, Math. Comput. Simulat., 188 (2021), 1–12. doi: 10.1016/j.matcom.2021.03.035. doi: 10.1016/j.matcom.2021.03.035
![]() |
[16] |
Y. Kang, N. Lanchier, Expansion or extinction: Deterministic and stochastic two-patch models with Allee effects, J. Math. Biol., 62 (2011), 925–973. doi: 10.1007/s00285-010-0359-3. doi: 10.1007/s00285-010-0359-3
![]() |
[17] |
D. Pal, G. P. Samanta, Effects of dispersal speed and strong Allee effect on stability of a two-patch predator-prey model, Int. J. Dyn. Control, 6 (2018), 1484–1495. doi: 10.1007/s40435-018-0407-1. doi: 10.1007/s40435-018-0407-1
![]() |
[18] |
S. Saha, G. P. Samanta, Influence of dispersal and strong Allee effect on a two-patch predator-prey model, Int. J. Dyn. Control, 7 (2019), 1321–1349. doi: 10.1007/s40435-018-0490-3. doi: 10.1007/s40435-018-0490-3
![]() |
[19] |
B. Dennis, Allee effects: Population growth, critical density and the chance of extinction, Nat. Resour. Model, 3 (1989), 481–538. doi: 10.1111/j.1939-7445.1989.tb00119.x. doi: 10.1111/j.1939-7445.1989.tb00119.x
![]() |
[20] |
Y. Y. Lv, L. J. Chen, F. D. Chen, Stability and bifurcation in a single species logistic model with additive Allee effect and feedback control, Adv. Differ. Equ., 8 (2020), 2686–2697. doi: 10.1186/s13662-020-02586-0. doi: 10.1186/s13662-020-02586-0
![]() |
[21] |
Y. Y. Lv, L. J. Chen, F. D. Chen, Z. Li, Stability and bifurcation in an SI epidemic model with additive Allee effect and time delay, Int. J. Bifurcat. Chaos, 31 (2021), 2150060. doi: 10.1142/S0218127421500607. doi: 10.1142/S0218127421500607
![]() |
[22] |
J. Y. Xu, T. H. Zhang, M. A. Han, A regime switching model for species subject to environmental noises and additive Allee effect, Physica A, 527 (2019), 121300. doi: 10.1016/j.physa.2019.121300. doi: 10.1016/j.physa.2019.121300
![]() |
[23] |
Y. L. Cai, C. D. Zhao, W. M. Wang, J. F. Wang, Dynamics of a Leslie-Gower predator-prey model with additive Allee effect, App. Math. Model., 39 (2015), 2092–2106. doi: 10.1016/j.apm.2014.09.038. doi: 10.1016/j.apm.2014.09.038
![]() |
[24] |
W. D. Wang, Population dispersal and Allee effect, Ric. Mat., 65 (2016), 535–548. doi: 10.1007/s11587-016-0273-0. doi: 10.1007/s11587-016-0273-0
![]() |
[25] | Z. F. Zhang, T. R. Ding, W. Z. Huang, Z. X. Dong, Qualitative Theory of Differential Equation, Bei Jing: Science Press, 1992. doi: 10.1007/978-1-4757-4969-4_4. |
[26] | L. Perko, Differential Equations and Dynamical Systems, 3 Eds., New York: Springer-Verlag, 2001. doi: 10.1007/978-1-4613-0003-8. |
[27] |
M. Gazor, M. Kazemi, Singularity: A maple library for local zero bifurcation control of scalar smooth maps, ASME J. Comput. Nonlinear Dyn., 15 (2020), 011001. doi: 10.1115/1.4045286. doi: 10.1115/1.4045286
![]() |
[28] |
M. Gazor, M. Kazemi, Normal Form Analysis of Z2-Equivariant Singularities, Int. J. Bifurcat. Chaos, 29 (2019), 1950015. doi: 10.1142/S0218127419500159. doi: 10.1142/S0218127419500159
![]() |
[29] |
B. W. Qin, K. W. Chung, A. Algaba, A. J. Rodríguez-Luis, High-order approximation of heteroclinic bifurcations in truncated 2D-normal forms for the generic cases of Hopf-zero and nonresonant double Hopf singularities, SIAM J. Appl. Dyn. Syst., 20 (2021), 403–437. doi: 10.1137/20M1325101. doi: 10.1137/20M1325101
![]() |
[30] |
M. Gazor, N. Sadri, Bifurcation controller designs for the generalized cusp plants of Bogdanov–Takens singularity with an application to ship control, SIAM J. Control Optim., 57 (2019), 2122–2151. doi: 10.1137/18M1210769. doi: 10.1137/18M1210769
![]() |
[31] |
M. Gazor, N. Sadri, Bifurcation control and universal unfolding for Hopf-zero singularities with Leading solenoidal terms, SIAM J. Appl. Dyn. Syst., 15 (2016), 870–903. doi: 10.1137/141000403. doi: 10.1137/141000403
![]() |
[32] |
A. J. Krener, W. Kang, D. E. Chang, Control bifurcations, IEEE T. Automat. Contr., 49 (2004), 1231–1246. doi: 10.1109/TAC.2004.832199. doi: 10.1109/TAC.2004.832199
![]() |
[33] |
B. Hamzi, W. Kang, J. P. Barbot, Analysis and control of Hopf bifurcations, SIAM J. Control Optim., 42 (2004), 2200–2220. doi: 10.1137/S0363012900372714. doi: 10.1137/S0363012900372714
![]() |
1. | Shangming Chen, Yanbo Chong, Fengde Chen, Periodic solution of a discrete commensal symbiosis model with Hassell-Varley type functional response, 2022, 9, 2353-0626, 170, 10.1515/msds-2022-0154 | |
2. | A.Q. Khan, S.M. Qureshi, A.M. Alotaibi, Bifurcation analysis of a three species discrete-time predator-prey model, 2022, 61, 11100168, 7853, 10.1016/j.aej.2021.12.068 | |
3. | Qimei Zhou, Sijia Lin, Fengde Chen, Runxin Wu, Positive Periodic Solution of a Discrete Lotka-volterra Commensal Symbiosis Model with Michaelis-menten Type Harvesting, 2022, 21, 2224-2880, 515, 10.37394/23206.2022.21.57 | |
4. | Lili Xu, Yalong Xue, Qifa Lin, Chaoquan Lei, Global Attractivity of Symbiotic Model of Commensalism in Four Populations with Michaelis–Menten Type Harvesting in the First Commensal Populations, 2022, 11, 2075-1680, 337, 10.3390/axioms11070337 | |
5. | Fengde , Chen, Yanbo , Chong, Shangming Chen, Global Stability of a Commensal Symbiosis Model With Holling Ii Functional Response and Feedback Controls, 2022, 17, 2224-2856, 279, 10.37394/23203.2022.17.32 | |
6. | Fengde Chen, Zhong Li, Lijuan Chen, Dynamic Behaviors of a Stage Structure Commensalism System with Holling type II Commensalistic Benefits, 2022, 21, 2224-2880, 810, 10.37394/23206.2022.21.93 | |
7. | Xiaowan Liu, Qin Yue, Stability property of the boundary equilibria of a symbiotic model of commensalism and parasitism with harvesting in commensal populations, 2022, 7, 2473-6988, 18793, 10.3934/math.20221034 | |
8. | Xiaoran Li, Qin Yue, Fengde Chen, Global Stability of the Positive Equilibrium of a Leslie-Gower Predator-Prey Model Incorporating Predator Cannibalism, 2022, 21, 2224-2678, 400, 10.37394/23202.2022.21.44 | |
9. | Weijie Lu, Yonghui Xia, Multiple Periodicity in a Predator–Prey Model with Prey Refuge, 2022, 10, 2227-7390, 421, 10.3390/math10030421 | |
10. | Lili Xu, Yalong Xue, Xiangdong Xie, Qifa Lin, Dynamic Behaviors of an Obligate Commensal Symbiosis Model with Crowley–Martin Functional Responses, 2022, 11, 2075-1680, 298, 10.3390/axioms11060298 | |
11. | Fengde Chen, Qimei Zhou, Sijia Lin, Global Stability of Symbiotic Model of Commensalism and Parasitism with Harvesting in Commensal Populations, 2022, 21, 2224-2880, 424, 10.37394/23206.2022.21.50 | |
12. | Bilel Elbetch , Effect of dispersal in two-patch environment with Richards growth on population dynamics, 2022, 2, 2773-4196, 41, 10.58205/jiamcs.v2i3.47 | |
13. | Yue Xia, Lijuan Chen, Vaibhava Srivastava, Rana D. Parshad, Stability and bifurcation analysis of a two-patch model with the Allee effect and dispersal, 2023, 20, 1551-0018, 19781, 10.3934/mbe.2023876 | |
14. | S. KALABUŠIĆ, E. PILAV, THE BEHAVIOR OF A CLASS HOST-PARASITOID MODELS WITH HOST REFUGE AND STRONG ALLEE EFFECT UPON THE HOST POPULATION, 2023, 31, 0218-3390, 755, 10.1142/S0218339023500274 | |
15. | Yalong Xue, Analysis of a prey-predator system incorporating the additive Allee effect and intraspecific cooperation, 2024, 9, 2473-6988, 1273, 10.3934/math.2024063 | |
16. | Minjuan Gao, Lijuan Chen, Fengde Chen, Dynamical analysis of a discrete two-patch model with the Allee effect and nonlinear dispersal, 2024, 21, 1551-0018, 5499, 10.3934/mbe.2024242 | |
17. | Carolin Grumbach, Frank M. Hilker, Allee pits in metapopulations: critical dispersal rates for connectivity to be beneficial, 2025, 18, 1874-1738, 10.1007/s12080-024-00597-1 |