
In this paper, we study a system of nonlinear partial differential equations that models the population dynamics of two competitive species both under Allee effects. The consideration of the model includes Logistic growth with Allee effects, Lotka-Volterra competition, diffusion, initial density and boundary conditions on the habitat. In the reaction-diffusion system, we employ the method of upper and lower solutions to address questions on self-elimination or persistence, as well as permanence or competitive exclusion. Specific conditions on biological parameters are explicitly given for extinction, coexistence and competitive exclusion of the species under various boundary conditions. Numerical simulations for the model are demonstrated to illustrate our results from mathematical analysis.
Citation: Yaw Chang, Wei Feng, Michael Freeze, Xin Lu, Charles Smith. Elimination, permanence, and exclusion in a competition model under Allee effects[J]. AIMS Mathematics, 2023, 8(4): 7787-7805. doi: 10.3934/math.2023391
[1] | Shanshan Yu, Jiang Liu, Xiaojie Lin . Multiple positive periodic solutions of a Gause-type predator-prey model with Allee effect and functional responses. AIMS Mathematics, 2020, 5(6): 6135-6148. doi: 10.3934/math.2020394 |
[2] | Lingling Li, Xuechen Li . The spatiotemporal dynamics of a diffusive predator-prey model with double Allee effect. AIMS Mathematics, 2024, 9(10): 26902-26915. doi: 10.3934/math.20241309 |
[3] | Ziran Cheng, Sze-Bi Hsu . Dynamics of drug on-drug off models with mutations in morbidostat — Dedicated to the seventieth birthday of Professor Gail Wolkowicz. AIMS Mathematics, 2023, 8(9): 20815-20840. doi: 10.3934/math.20231061 |
[4] | 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 |
[5] | 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 |
[6] | Yalong Xue . Analysis of a prey-predator system incorporating the additive Allee effect and intraspecific cooperation. AIMS Mathematics, 2024, 9(1): 1273-1290. doi: 10.3934/math.2024063 |
[7] | Liye Wang, Wenlong Wang, Ruizhi Yang . Stability switch and Hopf bifurcations for a diffusive plankton system with nonlocal competition and toxic effect. AIMS Mathematics, 2023, 8(4): 9716-9739. doi: 10.3934/math.2023490 |
[8] | Chaohong Pan, Xiaowen Xu, Yong Liang . Speed determinacy of traveling waves for a lattice stream-population model with Allee effect. AIMS Mathematics, 2024, 9(7): 18763-18776. doi: 10.3934/math.2024913 |
[9] | Marcello Delitala, Mario Ferraro . Is the Allee effect relevant in cancer evolution and therapy?. AIMS Mathematics, 2020, 5(6): 7649-7660. doi: 10.3934/math.2020489 |
[10] | Eric M. Takyi, Charles Ohanian, Margaret Cathcart, Nihal Kumar . Sex-biased predation and predator intraspecific competition effects in a prey mating system. AIMS Mathematics, 2024, 9(1): 2435-2453. doi: 10.3934/math.2024120 |
In this paper, we study a system of nonlinear partial differential equations that models the population dynamics of two competitive species both under Allee effects. The consideration of the model includes Logistic growth with Allee effects, Lotka-Volterra competition, diffusion, initial density and boundary conditions on the habitat. In the reaction-diffusion system, we employ the method of upper and lower solutions to address questions on self-elimination or persistence, as well as permanence or competitive exclusion. Specific conditions on biological parameters are explicitly given for extinction, coexistence and competitive exclusion of the species under various boundary conditions. Numerical simulations for the model are demonstrated to illustrate our results from mathematical analysis.
In traditional methods of estimating population growth rates, people consider that most species reproduce and die proportionally to the current population size. A simple representation of this concept can be expressed with the exponential growth of the populations. There are at least two major reasons to further modify the system: 1) Infinite growth is impossible due to limitation of habitat and resources; 2) Competition between various species should be addressed. We can now introduce the Lotka-Volterra system to address inter-species competition, where u(t) and v(t) represent the population size of two biological species, bu and bv are intrinsic birth rates, du and dv are death rates. Throughout this paper we will assume that for all species considered the intrinsic birth rate is larger than death rate. This type of competition, known as Lotka-Volterra competition [2,5], implies that species v diminishes resources available to u by a factor of α, and u diminishes resources available to v by the factor β.
dudt=u[bu(1−(u+αv))−du], | (1.1) |
dvdt=v[bv(1−(v+βu))−dv]. | (1.2) |
Due to the possibility that small population size can make a species vulnerable to extinction, further modifications can be made to the above equations. The Allee effect (given as early as 1932, by Allee [1]) models this possibility and was introduced to the multi-species models in recent years [9,14,17,18,20,21].
dudt=u[bu(1−u−αv)(uu+ku)−du], | (1.3) |
dvdt=v[bv(1−v−βu)(vv+kv)−dv], | (1.4) |
where ku and kv are called half-saturation constants. We see that, for large population sizes (u and v much larger than ku and kv respectively) the Allee term, (uu+ku) or (vv+kv), approaches 1 and has little effect on the modeled change in population. As population size u or v approaches 0, the Allee effect term gets near 0 and the death rate (du or dv) dominates the differential equation, which drives a small population toward extinction.
We now have the above system of ordinary differential equations that addresses total population sizes, boundedness, and the vulnerability of species for sizes getting too small, as well as the competitive nature of the two species. Our final extension of the system (1.4) is to include the diffusion of both populations, local density and migration, as well as boundary conditions on the habitat. This leads to the following reaction-diffusion system for density functions u(t,x) and v(t,x) with respective diffusion rates Du, Dv>0 for both species:
∂u∂t−Du∇2u=u[bu(1−u−αv)(uu+ku)−du]in (0,∞)×Ω,∂v∂t−Dv∇2v=v[bv(1−v−βu)(vv+kv)−dv] in (0,∞)×Ω,Bu[u]=0 and Bv[v]=0 on (0,∞)×∂Ω,u(0,x)=u0(x) and v(0,x)=v0(x)on¯Ω. | (1.5) |
Here the boundary conditions are given as
Bu[u]=uorBu[u]=∂u∂ν+γu(x)u,andBv[v]=vorBv[v]=∂v∂ν+γv(x)v, | (1.6) |
with ν as the normal vector on ∂Ω, γu and γv∈C1+α(∂Ω), and γu(x), γv(x)≥0 on ∂Ω. This way, we include three commonly used types of boundary conditions: Dirichlet, Neumann, and Robin types.
The above reaction-diffusion system (with one-side Allee effect) on infinite spatial domain was recently studied in [9] where the asymptotic stability of the equilibria are given and the existence of traveling wave solutions are proven. In research studies on reaction-diffusion systems modeling multi-species population dynamics in bounded spatial domain (competition, predator-prey, food chain, etc.), much attention has been given to extinction, permanence, and competition or predation caused exclusion [3,6,7,10,11,12,13,15,17,19]. These discussions are also extended to many variations of the Lotka-Volterra models under different boundary conditions [4,8,9]. In this paper, we apply the method of upper-lower solutions to the reaction-diffusion system 1.5 and find conditions for the competing species to self-eliminate (because of Allee effect), coexist (with balanced biological parameters and initial density functions), or competitively exclude (through resource competition and Allee effect). It is seen that the ultimate outcomes in the biological system also depend on the size of the habitat and boundary conditions for both species. According to the theoretical results on permanence and competitive exclusion, we will demonstrate numerical simulations under parameters satisfying conditions obtained.
Examining the reaction functions in system (1.5) by taking the partial derivative of
f(u,v)=u[bu(1−u−αv)(uu+ku)−du] |
with respect to v, and the partial derivative of
g(u,v)=v[bv(1−v−βu)(vv+kv)−dv] |
with respect to u, we find that (for u and v≥0):
∂f∂v=−αbuu2(u+ku)≤0and∂g∂u=−βbvv2(v+kv)≤0. | (1.7) |
As defined in ([15], page 383), the system (1.5) is quasi-monotone non-increasing. The upper-lower solutions (˜u,˜v) and (ˆu,ˆv) defined as following ensure the existence-comparison result given in the below lemma.
Definition 1.1 Upper and lower solutions [16].
A pair of smooth functions in ˜U=(˜u,˜v), and ˆU=(ˆu,ˆv) in C((0,∞)×ˉΩ)∩C1,2((0,∞)×Ω) are ordered upper and lower solutions of system (1.5) if they satisfy the relation ˜U≥ˆU and if:
˜ut−Du∇2˜u≥˜u[bu(1−˜u−αˆv)(˜u˜u+ku)−du]in(0,∞)×Ω,˜vt−Dv∇2˜v≥˜v[bv(1−˜v−βˆu)(˜v˜v+kv)−dv]in(0,∞)×Ω,ˆut−Du∇2ˆu≤ˆu[bu(1−ˆu−α˜v)(ˆuˆu+ku)−du]in(0,∞)×Ω,ˆvt−Dv∇2ˆv≤ˆv[bv(1−ˆv−β˜u)(ˆvˆv+kv)−dv]in(0,∞)×Ω,Bu[˜u]≥0≥Bu[ˆu]andBv[˜v]≥0≥Bv[ˆv]on(0,∞)×∂Ω,˜u(0,x)≥u0(x)≥ˆu(0,x)and˜v(0,x)≥v0(x)≥ˆv(0,x)on¯Ω. | (1.8) |
Lemma 1.2. Existence and comparison [16].
If there are a pair of smooth functions ˜U=(˜u,˜v) and ˆU=(ˆu,ˆv) as ordered upper and lower solutions of (1.5) (defined in Definition 1.1), then the reaction-diffusion system (1.5) has a unique solution U=(u,v) with (˜u,˜v)≥(u,v)≥(ˆu,ˆv) on (0,∞)×ˉΩ.
It can be easily verified that for any constants M1 and M2 satisfying the relations
M1=max{‖u0‖∞,bu−dubu},M2=max{‖v0‖∞,bv−dvbv}, | (1.9) |
the constant functions (M1,M2) and (0,0) are a pair of ordered upper and lower solutions of (1.5) on (0,∞)×¯Ω under any combination of the boundary conditions given in (1.6).
Theorem 1.3. Global existence and boundedness.
For any smooth function u0(x), v0(x) on ¯Ω the reaction-diffusion system (1.5) has a unique solution U=(u,v) with (M1,M2)≥(u,v)≥(0,0) on (0,∞)×ˉΩ, with constants M1 and M2 given in (1.9).
Throughout this paper, for each type of the boundary conditions in (1.6), say B[⋅], We let λ0 and ϕ0(x) (with ‖ϕ0‖∞=1) be the principal eigenvalue and associated eigenfunction of the eigenvalue problem
∇2ϕ+λϕ=0inΩ,B[ϕ]=0on∂Ω. | (2.1) |
It is well-known that for Neumann Boundary condition, ϕ0(x)=1 on ˉΩ with λ0=0. Also, for Dirichlet or Robin boundary condition we have ϕ0(x)>0 in Ω with λ0>0. We demonstrate how the Allee effect, birth and death rates, as well as diffusion rates affect the long-term survival of each species with relatively small initial population size.
Theorem 2.1. Self-elimination by Allee effect (Neumann boundary condition).
Under the Neumann boundary conditions for the u-species
∂u(t,x)∂ν=0on(0,∞)×∂Ω, |
if 0≤u0(x)<kudubu−du then limt→∞u(t,x)=0 uniformly on ˉΩ.
Proof. For some σ>0, let ˜u=Me−σt and ˆu=0, ˜v=(bv−dv)/bv and ˆv=0. Since all the defined upper and lower solutions are independent of x, they satisfy the boundary condition inequalities in (1.8) with nonnegative function values and normal derivatives as 0. Also, ˜u≥ˆu and ˜v≥ˆv on [0,∞)×ˉΩ. It is obvious that the differential inequalities for ˆu, ˜v and ˆv are satisfied in (1.8).
To verify that ˜u=Me−σt satisfies the corresponding differential inequality in (1.8), we find that the following relation must be satisfied:
−σMe−σt≥Me−σt[bu(1−Me−σt)(Me−σtMe−σt+ku)−du]. |
After simplifying, the following must hold for all t∈(0,∞):
−σ≥bu(1−Me−σt)(Me−σtMe−σt+ku)−du. | (2.2) |
It suffices to have
−σ≥bu(MM+ku)−du. | (2.3) |
Allowing for a small and positive σ, for ˜u to function as an upper solution defined in (1.8) we just need:
M<kudubu−du. | (2.4) |
By Definition 1.1 and Theorem 1.2, for any initial density function u0(x) with 0≤u0(x)<kudubu−du on ¯Ω, we have
0≤u(t,x)<Me−σton[0,∞)×ˉΩ, |
with M=‖u0‖∞<kudubu−du. Therefore, limt→∞u(t,x)=0 uniformly on ˉΩ.
We now analyze the self-elimination by Allee effect of one species (say u) under Dirichlet or Robin boundary condition. For this case, the principal eigenvalue for Bu[⋅] is λ0>0 and the corresponding eigenfunction ϕ0(x)>0 in Ω (with ||ϕ0||∞=1).
Theorem 2.2. Self-elimination by Allee effect (Dirichlet or Robin boundary condition).
Under the Dirichlet or Robin (with non-trivial γu(x)) boundary condition for the u-species
u(t,x)=0or∂u(t,x)∂ν+γu(x)u(t,x)=0on∂Ω, |
a) If Duλ0≥bu−du, then for any positive M>0 with initial density u0(x)≤MΦ0(x), we have limt→∞u(t,x)=0 uniformly on ˉΩ.
b) If Duλ0<bu−du and 0≤u0(x)≤MΦ0(x) for some M<ku(λ0Du+du)bu−du−λ0Du, then limt→∞u(t,x)=0 uniformly on ˉΩ.
Proof. For some positive σ, Let ˜u=Mϕ0e−σt, and ˆu=0, ˜v=(bv−dv)/bv and ˆv=0, so (˜u,˜v)≥(ˆu,ˆv) on [0,∞)×ˉΩ. Since ϕ0(x)=0 on ∂Ω, the boundary inequality in (1.8) for ˜u is satisfied. As seen in the proof of the previous theorem, the differential and boundary inequalities in (1.8) for ˆu, ˜v and ˆv are also satisfied.
For part a), assume that Duλ0>bu−du. To satisfy the differential inequality in (1.8) for ˜u we need:
−σMϕ0e−σt+Duλ0Mϕ0e−σt≥Mϕ0e−σt[bu(1−Mϕ0e−σt)(Mϕ0e−σtMϕ0e−σt+ku)−du]. | (2.5) |
Simplifying above and allowing for a small positive σ, it is suffice to have the following hold for any t>0:
Duλ0>bu(1−Mϕ0e−σt)(Mϕ0e−σtMϕ0e−σt+ku)−du. | (2.6) |
We recognize that 1−Mϕ0e−σt<1 and 0<Mϕ0e−σtMϕ0e−σt+ku<1, so we can ensure that:
Duλ0>bu−du>bu(1−Mϕ0e−σt)(Mϕ0e−σtMϕ0e−σt+ku)−du. | (2.7) |
Hence the differential inequality for the upper solution is satisfied by any ˜u=Mϕ0e−σt with any M>0. This proves part a).
For part b), assume that Duλ0<bu−du. As in part a), the differential inequality for ˜u in (1.8 needs to be satisfied for all t>0:
Duλ0−σ≥[bu(1−Mϕ0e−σt)(Mϕ0e−σtMϕ0e−σt+ku)−du], | (2.8) |
which can be ensured by
Duλ0−σ≥bu(MM+ku)−du. | (2.9) |
The above relation holds for a small and positive σ, and constant M:
M<ku(λ0Du+du)bu−du−λ0Du. | (2.10) |
By Definition 1.1 and Theorem 1.2, for any initial density function u0(x) with 0≤u0(x)≤Mϕ0(x) on ˉΩ, we have 0≤u(t,x)≤Mϕ0e−σt and limt→∞u(t,x)=0 uniformly on ˉΩ.
Theorems 2.1 and 2.2 shows that larger Allee effect coefficient, higher death rate, and faster diffusion will drive a population to extinction as long as the initial density is within the given ranges for self-elimination.
In the next theorem, we give a result of global extinction under the significance of the Allee effect, for one of the competing species (say u) with any initial density size and boundary condition.
Theorem 2.3. Global extinction of u-species under Allee effect (any boundary condition).
For the boundary condition Bu[⋅]=0, let λ0≥0 be the principal eigenvalue with the corresponding eigenfunction ϕ0(x)>0 in Ω. If
du+Duλ0>bu(1+ku−√ku(1+ku))(√ku(1+ku)−ku)√ku(ku+1), | (2.11) |
then limt→∞u(t,x)=0 uniformly on ˉΩ with all initial density 0≤u0(x)≤Nϕ0(x) for any N>0.
Proof. Again, for any N>0 and some small σ>0, Let ˜u=Nϕ0e−σt, and ˆu=0, ˜v=(bv−dv)/bv and ˆv=0, so (˜u,˜v)≥(ˆu,ˆv) on [0,∞)×ˉΩ. We can show see that the differential and boundary inequalities are satisfied are all satisfied for ˆu, ˜v, and ˆv under any boundary conditions for u and v, as in the proofs for previous theorems.
Also, as seen in (2.6), for the upper solution ˜u to satisfy the required differential inequality given in (1.8), we need to ensure that for all t>0,
bu(1−Nϕ0e−σt)Nϕ0e−σtNϕ0e−σt+ku<du+Duλ0. |
We now examine the positive maximum of G(X)=bu(1−X)XX+ku for X∈[0,∞). Note that G(0)=G(1)=0, and G(X)<0 for X>1.
G′(X)=bu(−X2−2kuX+ku)(X+ku)2, |
so G′(X)=0 when X=−ku±√ku(ku+1). We then allocate the only positive maximum point of G(X), at X=√ku(ku+1)−ku∈(0,1). This implies that for all t>0,
bu(1−Nϕ0e−σt)Nϕ0e−σtNϕ0e−σt+ku≤G(√ku(ku+1)−ku)=bu(1+ku−√ku(1+ku))(√ku(1+ku)−ku)√ku(ku+1). | (2.12) |
We can now conclude that if
du+Duλ0>bu(1+ku−√ku(1+ku))(√ku(1+ku)−ku)√ku(ku+1), |
then for any positive N and some small positive σ, 0≤u(t,x)≤Nϕ0e−σt on [0,∞)×ˉΩ as long as 0≤u0(x)≤Nϕ0(x) on ˉΩ. The global extinction of u-species given in the theorem then follows.
In this section, we explore on conditions for permanence (long-term survival of both species) in the competition model (1.5). Our approach is to find the possibility of a pair of upper-lower solutions given in Definition 1.1 with nontrivial lower solutions for both u and v.
Theorem 3.1. Permanence of both species (Neumann boundary condition).
Let u and v both satisfy the no-flux boundary condition ∂u∂ν=∂v∂ν=0 on ∂Ω. If
0<α<bv(bu−du)bu(bv−dv)and0<β<bu(bv−dv)bv(bu−du),ku≤bu4du(bu−dubu−αbv−dvbv)2andkv≤bv4dv(bv−dvbv−βbu−dubu)2, | (3.1) |
then the competition model (1.5) is permanent as long as the initial density functions satisfy
12(bu−dubu−αbv−dvbv)<u0(x)<bu−dubuon ˉΩ, | (3.2) |
and
12(bv−dvbv−βbu−dubu)<v0(x)<bv−dvbvon ˉΩ. | (3.3) |
Proof. Let the conditions in (3.1) and (3.2) hold. We will show that (˜u,˜v)=(bu−dubu,bv−dvbv) and (ˆu,ˆv)=(ϵu,ϵv), where bu−dubu>ϵu>0 and bv−dvbv>ϵv>0, are a pair of ordered upper and lower solutions given in Definition 1.1. It is clear that (˜u,˜v)≥(ˆu,ˆv). Since ˜u, ˆu, ˜v, and ˆv are all independent of x, so they satisfy the boundary condition inequalities in (1.8).
In order to satisfy the differential inequalities in (1.8), we need the following relations to hold:
0≥bu−dubu[bu(1−bu−dubu−αϵv)(bu−dububu−dubu+ku)−du],0≥bv−dvbv[bv(1−bv−dvbv−βϵu)(bv−dvbvbv−dvbv+kv)−dv],0≤ϵu[bu(1−ϵu−αbv−dvbv)(ϵuϵu+ku)−du],0≤ϵv[bv(1−ϵv−βbu−dubu)(ϵvϵv+kv)−dv]. |
Consider the following factors in the first two inequalities for upper solutions:
bu(1−bu−dubu−αϵv)bu−du(1+ku)bu−du<du−αϵvbu<du, |
and
bv(1−bv−dvbv−αϵu)bv−dv(1+kv)bv−dv<dv−αϵubv<dv. |
It's clear to see that the right-hand sides of the first two inequalities are strictly negative, so the differential inequalities for the upper solutions are satisfied.
According to the required differential inequalities for the lower solutions, we need the following to hold:
bu(1−ϵu−αbv−dvbv)(ϵuϵu+ku)−du≥0,bv(1−ϵv−βbu−dubu)(ϵvϵv+kv)−dv≥0. | (3.4) |
Multiplying the inequalities in (3.4) by −ϵu+kubu and −ϵv+kvbv respectively, we now need to analyze the following functions to find ϵu>0 and ϵv>0 to ensure
F1(ϵu)=ϵ2u+(αbv−dvbv+dubu−1)ϵu+kudubu≤0,F2(ϵv)=ϵ2v+(βbu−dubu+dvbv−1)ϵv+kvdvbv≤0. | (3.5) |
The functions F1 and F2 (as parabolas opening upward and with F1(0), F2(0)>0) have their respective vertices at:
ϵ∗u=12(bu−dubu−αbv−dvbv)<bu−dubu,ϵ∗v=12(bv−dvbv−βbu−dubu)<bv−dvbv. | (3.6) |
We then see that the following conditions on competition coefficients α and β ensure that ϵ∗u and ϵ∗v>0 in (3.6):
0<α<bv(bu−du)bu(bv−dv),and0<β<bu(bv−dv)bv(bu−du). | (3.7) |
Finally, we place a constraint on F1 and F2 for the function values at the vertices to be negative. For
F1(ϵ∗u)=14(bu−dubu−αbv−dvbv)2−12(bu−dubu−αbv−dvbv)2+kudubu<0, |
and doing the same computation for F2(ϵ∗v)<0, we can now find the permanence conditions with respect to the Allee effect coefficients ku and kv:
ku<bu4du(bu−dubu−αbv−dvbv)2andkv<bv4dv(bv−dvbv−βbu−dubu)2. | (3.8) |
Given constraints (3.7) and (3.8), by Definition 1.1 and Lemma 1.2, for any initial density functions u0(x) and v0(x) with ϵ∗u≤u0(x)≤bu−dubu and ϵ∗v≤v0(x)≤bv−dvbv, the unique solution of (1.5) exists in ⟨(ˆu,ˆv),(˜u,˜v)⟩. This implies permanence in the ecological system and proves the theorem.
In model (1.5), to secure long-term survival of both species we have obtained constraints (3.1) on α, β (the Lotka-Volterra competition coefficients from one species to another) and ku, kv (the Allee effect coefficients for each species) in relation with the birth and death rates. Also, the initial density functions need to balanced in ranges given in (3.3). To demonstrate a numerical example (in Figure 1), we make the following choices on the biological parameters which satisfy all conditions for permanence in (3.1):
bu=0.8,bv=0.6,du=0.4,dv=0.2,ku=0.045,kv=0.12,α=0.3,β=0.5, |
D1=0.08,D2=0.03. |
Also, we set the initial density functions as follows to satisfy the permanence conditions (3.3) on ¯Ω=[0,1] and the no-flux boundary condition on ∂Ω:
u0(x)=0.2−0.05cos(2πx),v0(x)=0.26−0.05cos(2πx). |
In previous sections, we have studied criteria for self-elimination of each species and long-term survival of both species. Now in this section, we examine conditions for competitive exclusion, that is, one species survives and forces the other species into extinction in (1.5). Various competitive exclusion results can be derived from different combinations of boundary conditions for the two species.
Theorem 4.1. Competitive exclusion of one species (u) (Neumann boundary condition).
Let u and v both satisfy the no-flux boundary condition ∂u∂ν=∂v∂ν=0 on ∂Ω.Assume that the following inequalities are satisfied
0<β<bu(bv−dv)bv(bu−du)andkv≤bv4dv(bv−dvbv−βbu−dubu)2. | (4.1) |
If
v0(x)≥ϵ∗v=12(bv−dvbv−βbu−dubu)andˉΩ, |
then
limt→∞u(t,x)=0and u0(x)<min{bu−dubu,kudubu−du−αbuϵ∗v}onˉΩ, |
then
limt→∞u(t,x)=0andlim inft→∞v(t,x)≥12(bv−dvbv−βbu−dubu) |
uniformly on ˉΩ.
Proof. Let inequalities in (4.1) be satisfied. We have
M1<min{bu−dubu,kudubu−du−αbuϵ∗v}, |
where M1 is the maximum value of u0(x) on ˉΩ. We are going to verify that for some σ1>0 (to be determined), (˜u,˜v)=(M1e−σ1t,bv−dvbv), and (ˆu,ˆv)=(0,ϵ∗v) are a pair of ordered upper-lower solutions. It is obvious that (˜u,˜v)≥(ˆu,ˆv) in [0,∞)×ˉΩ. ˜u, ˆu, ˜v, and ˆv are all independent of x, so their normal derivatives are equal to 0 on ∂Ω and the boundary inequalities in (1.8) are satisfied.
In order to satisfy the differential inequalities in (1.8), first notice that the differential inequality for ˆu=0 trivially holds. For the rest, we need the following relations to hold:
−σ1M1e−σ1t≥M1e−σ1t[bu(1−M1e−σ1t−αϵ∗v)(M1e−σ1tM1e−σ1t+ku)−du],0≥bv−dvbv[bv(1−bv−dvbv)(bv−dvbvbv−dvbv+kv)−dv],0≤ϵ∗v[bv(1−ϵ∗v−βM1e−σ1t)(ϵ∗vϵ∗v+kv)−dv]. | (4.2) |
We now analyze the differential inequalities in (4.2). The second of these inequalities is clearly satisfied by the fact that
bv(1−bv−dvbv)(bv−dvbvbv−dvbv+kv)−dv=dv(bv−dvbvbv−dvbv+kv)−dv≤0. |
Also, by the fact that M1e−σ1t≤bu−dubu and the argument related to ϵ∗v in the proof of Theorem 3, we can conclude that the condition kv≤bv4dv(bv−dvbv−βbu−dubu)2 ensures the third inequality in (4.2).
We further analyze the first inequality in (4.2). By the fact that 0≤e−σ1t≤1, it suffices to have the following holds and allow a small σ1>0:
bu(1−αϵ∗v)(M1M1+ku)<du. |
Solving for M from above, we conclude that for M1<kudubu−du−αbuϵ∗v and a positive number σ1 which is small enough, the first inequality in (4.2) will also be satisfied. Hence we have verified all required inequalities for the coupled upper-lower solutions.
By Definition 1.1 and Lemma 1.2, for any smooth initial density function u0(x) and v0(x) with u0(x)<M1 and v0(x)≥ϵ∗v, there is a unique solution of (1.5) with
(0,ϵ∗v)≤(u(t,x),v(t,x))≤(M1e−σ1t,bv−dvbv). |
This implies that: limt→∞u(t,x)=0 uniformly on ˉΩ and v(t,x)≥12(bv−dvbv−βbu−dubu) in [0,∞)×ˉΩ.
To demonstrate a numerical example (in Figure 2) for competitive exclusion of u-species given above, we make the following choices on the biological parameters which satisfy all conditions given in (4.1):
bu=0.5,bv=0.8,du=0.1,dv=0.2,ku=0.7,kv=0.25,α=0.5,β=0.3, |
D1=0.06,D2=0.04. |
Also, we set the initial density functions as follows to satisfy the conditions given in Theorem 4.1 on ¯Ω=[0,1] and the no-flux boundary condition on ∂Ω:
u0(x)=0.178−0.03cos(2πx),v0(x)=0.27−0.015cos(2πx). |
As seen in the theorem above, from the Allee coefficient ku and death rate bu for u-species and the persistence strength ϵ∗v for v-species, we can determine the magnitude M1 of ‖u0(⋅)‖∞ which will drive the u population to extinction. Moreover, with the presence of a competitor, large competition coefficient (α) and small difference between the birth and death rates (ratio dubu being close to 1) also will drive the population (u) with larger initial size to extinction. The following theorem shows those detrimental effects to the survival of u-species under reasonably large initial population size. We now examine conditions for competitive exclusion of u-species globally for 0≤u0(x)≤bu−dubu.
Theorem 4.2. Global exclusion of one species (u) under stronger competitor (v) (Neumann boundary condition).
Let u and v both satisfy the no-flux boundary condition ∂u∂ν=∂v∂ν=0 on ∂Ω. Assume that the inequalities in (4.1) hold. If
du>bu(1−αϵ∗v−√k2u+ku−αϵ∗vku+ku)(√k2u+ku−αϵ∗vku−ku)√k2u+ku−αϵ∗vku, | (4.3) |
with ϵ∗v=12(bv−dvbv−βbu−dubu), and the initial population functions satisfy v0(x)≥ϵ∗v and u0(x)<bu−dubu, then limt→∞u(t,x)=0 and lim inft→∞v(t,x)≥12(bv−dvbv−βbu−dubu) uniformly on ˉΩ.
Proof. As in the proof of Theorem 4.1, we will verify that for some small σ1>0 and the maximum of u0(x) M1<bu−dubu, (˜u,˜v)=(M1e−σ1t,bv−dvbv), and (ˆu,ˆv)=(0,ϵ∗v) are a pair of ordered upper-lower solutions. It is already known that (˜u,˜v)≥(ˆu,ˆv) in [0,∞)×ˉΩ, and all the boundary inequalities in (1.8) are satisfied. Among the four differential inequalities for upper-lower solution given in (4.2), we also know from the previous proof that the second, third, and fourth inequalities already hold.
We now focus on analyzing the first differential inequality in (4.2). Let M1e−σ1t=X, we need to find a condition for
du−σ1≥bu(1−X−αϵ∗v)(XX+ku) |
with all X>0 and some σ1>0. Define the function
H(X)=bu[(1−αϵ∗v)−X]XX+ku, |
and note that H(0)=H(1−αϵ∗v)=0, and H(X)<0 for X>1−αϵ∗v. Since H(x)>0 in the interval (0,1−αϵ∗v), we look for the maximum value of H(X) there. Setting
H′(X)=[bu[(1−αϵ∗v)−2X](X+ku)−bu(1−αϵ∗v−X)X(X+ku)2=0, |
we find the local extremum at √k2u+ku−αϵ∗vku−ku in the interval [0,1−αϵ∗v].
Since
H(√k2u+ku−αϵ∗vku−ku)=bu(1−αϵ∗v−√k2u+ku−αϵ∗vku+ku)(√k2u+ku−αϵ∗vku−ku)√k2u+ku−αϵ∗vku, | (4.4) |
allowing for a small σ1>0, the first differential inequality (4.2) is satisfied if
bu(1−αϵ∗v−√k2u+ku−αϵ∗vku+ku)(√k2u+ku−αϵ∗vku−ku)√k2u+ku−αϵ∗v<du. |
By Definition 1.1 and Lemma 1.2, for any smooth initial density function u0(x) and v0(x) with u0(x)<bu−dubu and v0(x)≥ϵ∗v, there is a unique solution of (1.5) with (0,ϵ∗v)≤(u(t,x),v(t,x))≤(M1e−σ1t,bv−dvbv). This implies that: limt→∞u(t,x)=0 uniformly on ˉΩ and v(t,x)≥12(bv−dvbv−βbu−dubu) on [0,∞)×ˉΩ.
We now examine an numerical example (in Figure 3) for the above global exclusion result, which shows competitive exclusion of u-species with any initial density u0(x)<(bu−du)/bu. We choose the following biological parameters that satisfy all the conditions in (4.1) and (4.3). Also, we use initial density functions with relatively larger u0(x) which satisfy the conditions given in Theorem 4.2 on ¯Ω=[0,1] and the no-flux boundary condition on ∂Ω.
bu=0.8,bv=0.6,du=0.18,dv=0.2,ku=0.5,kv=0.12,α=0.5,β=0.2, |
D1=0.03,D2=0.08.u0(x)=0.5−0.27cos(2πx),v0(x)=0.3−0.044cos(2πx). |
At last, we consider the case that the population density function u(t,x) is subject to the Dirichlet or Robin boundary condition (with λ0>0) and the population density function v(t,x) is subject to the Neumann boundary condition. This way, the u-species will be under greater strain at the boundary, so we study the conditions that will drive u-species to extinction under Allee effect and competition from v-species.
Theorem 4.3. Competitive exclusion of one species (mixed boundary conditions).
Let u and v satisfy the boundary condition
∂v(t,x)∂ν=0,u(t,x)=0or∂u(t,x)∂ν+γu(x)u(t,x)=0on∂Ω, |
with nontrivial γu(x) on ∂Ω.
Assume that the inequalities in (4.1) hold and v0(x)≥ϵ∗v=12(bv−dvbv−βbu−dubu), u0(x)≤Mϕ0(x) on ˉΩ for some M>0, with eigenfunction ϕ0(x) and eigenvalue λ0 for Bu[⋅] given in (2.1). If one of the following conditions holds:
a)
bu(1−αϵ∗v)≤Duλ0+du |
and M≤bu−dubu, or
b)
bu(1−αϵ∗v)>Duλ0+du |
and M<ku(Duλ0+dubu(1−αϵ∗v)−Duλ0−du, or
c)
Duλ0+du>kubu(1−αϵ∗v−√k2u+ku−αϵ∗vku+ku)(√k2u+ku−αϵ∗vku−ku)√k2u+ku−αϵ∗v |
and M≤bu−dubu, then limt→∞u(t,x)=0 uniformly on ˉΩ and v(t,x)>12(bv−dvbv−βbu−dubu) in [0,∞)×ˉΩ.
Proof. Once again, for M and σ>0, we set up the ordered upper and lower solutions (˜u,˜v)=(Mϕ0e−σt,bv−dvbv), and (ˆu,ˆv)=(0,ϵ∗v). Noting that ˜v and ˆv are independent of x, with respective normal derivatives =0 on ∂Ω. ˆu=0 on ˉΩ and Bu[ϕ0]=0 on ∂Ω. Therefore all boundary inequalities in (1.8) are satisfied.
As seen in the proofs of previous theorems, the differential inequalities in (1.8) for ˜v, ˆu and ˆv also hold as long as M≤bu−dubu.
We now analyze the differential inequality (1.8) for ˜u=Mϕ0e−σt:
−σMe−σt+Duλ0Mϕ0e−σt≥Mϕ0e−σt[bu(1−Mϕ0e−σt−αϵ∗v))(Mϕ0e−σtMϕ0e−σt+ku)−du]. | (4.5) |
To satisfy (4.5), it suffices to have the following relation hold which allows a small σ>0:
Duλ0+du>bu(1−αϵ∗v)(MM+ku). | (4.6) |
If bu(1−αϵ∗v)≤Duλ0+du, the first differential inequality (4.5) is satisfied for all M>0. Therefore the condition a) will ensure that (˜u,˜v)=(Mϕ0e−σt,bv−dvbv) and (ˆu,ˆv)=(0,ϵ∗v) are a pair of ordered upper-lower solutions under the mixed boundary conditions.
On the other hand, if bu(1−αϵ∗v)>Duλ0+du, solving for M from the inequality (4.6) leads to M<ku(Duλ0+dubu(1−αϵ∗v)−Duλ0−du. Hence the condition b) also ensures the ordered upper-lower solutions as designed.
Furthermore, the first differential inequality (4.5) is satisfied if the following holds for any X>0:
Duλ0+du>bu(1−X−αϵ∗v)(XX+ku). | (4.7) |
Maximizing the right side of the inequality as in the proof of Theorem 4.2, we can conclude that condition c) and a small σ>0 ensures the first differential inequality in (4.5) and verifies the ordered upper-lower solutions as designed.
By Definition 1.1 and Lemma 1.2, under the assumptions given and with one condition from a), b), or c), there exists a unique solution of (1.5) with (0,ϵ∗v)≤(u(t,x),v(t,x))≤(Mϕ0e−σt,bv−dvbv). This concludes that: limt→∞u(t,x)=0 uniformly on ˉΩ and v(t,x)≥12(bv−dvbv−βbu−dubu) in [0,∞)×ˉΩ.
We will end this section with two numerical examples for Theorem 4.3 on competitive exclusion of one species (u) which is under Dirichlet or Robin boundary condition. Setting Ω=(0,1), we let u(t,x) be under the Dirichlet boundary condition u(t,0)=u(t,1)=0. This way, the principal eigenvalue λ0=π2 with corresponding eigenfunction ϕ0(x)=sin(πx). For condition b) to be satisfied, we choose the following biological parameters and initial functions to demonstrate our numerical simulation (in Figure 4):
bu=0.8,bv=0.6,du=0.3,dv=0.2,ku=0.3,kv=0.1,α=0.5,β=0.3, |
D1=0.04,D2=0.1.u0(x)=0.9sin(πx),v0(x)=0.3−0.06cos(2πx). |
For condition c) to be satisfied, we choose the following biological parameters and initial functions to demonstrate our numerical simulation (in Figure 5):
bu=0.5,bv=0.8,du=0.2,dv=0.2,ku=0.5,kv=0.2,α=0.3,β=0.5, |
D1=0.3,D2=0.1.u0(x)=0.6sin(πx),v0(x)=0.28−0.04cos(2πx). |
The results shown in this article demonstrate how the all the biological parameters in the model, especially the coefficients representing Allee effects, ultimately determine the long-term survival of each competing species. We also pay special attention to the impacts of the dimensions of the habitat Ω and boundary conditions for both species on ∂Ω. As we know, Dirichlet boundary condition (u=0 on ∂Ω) indicates the biological species diminishes on the boundary of the habitat, while Neumann boundary condition (∂u∂ν=0 on ∂Ω) indicates no flux on the boundary of the habitat as the advection and diffusion fluxes of the population are exactly balanced. And, the Robin boundary condition in (1.6) is a linear combination of those two.
In Section 2, we give several results on self-elimination by Allee effect of one species despite the behavior of other. In Theorem 2.1 (Neumann boundary condition) and Theorem 2.2 (Dirichlet or Robin boundary condition), we find ranges for initial density function u0
0≤u0(x)<kudubu−duor0≤u0(x)<ku(λ0Du+du)bu−du−λ0DuΦ0(x) |
which ensure extinction of species u. And, in Theorem 2.3, the following self-elimination condition on all the biological parameters is obtained under all boundary conditions and any feasible initial density (global elimination):
du+Duλ0>bu(1+ku−√ku(1+ku))(√ku(1+ku)−ku)√ku(ku+1). |
We note that
limku→∞bu(1−(√ku(ku+1)−ku))(√ku(ku+1)−ku)√ku(ku+1)=0. |
Therefore, for sufficiently large Allee effect coefficient ku, relatively large death rate and limited birth rate, as well as faster diffusion that makes the global elimination condition hold, a species with any nonnegative initial density will be driven to extinction under all three boundary conditions.
In section 3, we explore the possibility of permanence for both species under Neumann (no flux) boundary conditions. This requires balanced constraints on birth, death, and competition rates, as well as the Allee effect coefficients:
0<α<bv(bu−du)bu(bv−dv)and0<β<bu(bv−dv)bv(bu−du),ku≤bu4du(bu−dubu−αbv−dvbv)2andkv≤bv4dv(bv−dvbv−βbu−dubu)2. |
Also, the initial density functions are required to be above some minimum thresholds:
12(bu−dubu−αbv−dvbv)<u0(x)<bu−dubu on ˉΩ, |
and
12(bv−dvbv−βbu−dubu)<v0(x)<bv−dvbv on ˉΩ. |
In Section 4, conditions for competitive exclusion are investigated under various boundary conditions. Under Neumann boundary conditions for both species and with the persistence constraints for v-species hold, we give the range for initial density u0 ensuring extinction of u-species and survival of v species in Theorem 4.1:
u0(x)<min{bu−dubu,kudubu−du−αbuϵ∗v}. |
In Theorems 4.2 and 4.3, under Neumann boundary condition for v-species and any boundary condition for u-species, global exclusion of u-species are ensured by the following condition for all feasible initial density u0:
Duλ0+du>kubu(1−αϵ∗v−√k2u+ku−αϵ∗vku+ku)(√k2u+ku−αϵ∗vku−ku)√k2u+ku−αϵ∗v. |
Here for the u-species, λ0=0 under Neumann boundary condition and λ0>0 under Dirichlet or Robin boundary condition. Specifically, with Dirichlet or Robin boundary condition, limited habitat size can also be detrimental to the persistence of the u-species. Examining a special case of the eigenvalue problem with Dirichlet boundary condition on a 1-dimensional habitat Ω=(0,L):
∇2ϕ0+λ0ϕ0=0in(0,L),ϕ0(0)=ϕ0(L)=0, |
we see that
λ0=π2L2andϕ0(x)=sin(πx/L). |
This shows that smaller habitat size (L) makes much bigger λ0 which validates the above competitive exclusion condition while all other biological parameters remaining unchanged. In addition to the self-elimination and competitive exclusion constraints on all biological parameters mentioned in Sections 2 and 4, this example illustrates that limited habitat size can also drive a biological population into extinction under Dirichlet or Robin boundary condition.
The authors would like to thank the anonymous referees for their valuable suggestions to strengthen the results in this article.
The authors declare no conflicts of interest.
[1] |
W. C. Allee, E. S. Bowen, Studies in animal aggregations: mass protection against aolloidal silver among goldfishes, J. Exp. Biol., 61 (1932), 185–207. https://doi.org/10.1002/jez.1400610202 doi: 10.1002/jez.1400610202
![]() |
[2] |
J. Blat, K. J. Brown, Bifurcation of steady-state solutions in predator-prey and competition systems, Proceedings of the Royal Society of Edinburgh, 97 (1984), 21–34. https://doi.org/10.1017/S0308210500031802 doi: 10.1017/S0308210500031802
![]() |
[3] |
R. S. Cantrell, C. Cosner, V. Hutson, Permanence in ecological systems with spatial heterogeneity, Proceedings of the Royal Society of Edinburgh, 123 (1993), 533–559. https://doi.org/10.1017/S0308210500025877 doi: 10.1017/S0308210500025877
![]() |
[4] |
Y. Chang, W. Feng, M. Freeze, X. Lu, Permanence and coexistence in a diffusive complex ratio-dependent food chain, Int. J. Dyn. Control, 3 (2015), 262–274. https://doi.org/10.1007/s40435-014-0131-4 doi: 10.1007/s40435-014-0131-4
![]() |
[5] | L. Edelstein-Keshet, Mathematical models in biology, Philadelphia: Society for Industrial and Applied Mathematics, 2005. https://doi.org/10.1137/1.9780898719147 |
[6] |
W. Feng, Coexistence, stability, and limiting behavior in a one-predator-two-prey model, J. Math. Anal. Appl., 179 (1993), 592–609. https://doi.org/10.1006/jmaa.1993.1371 doi: 10.1006/jmaa.1993.1371
![]() |
[7] |
W. Feng, Permanence effect in a three-species food chain model, Appl. Anal., 54 (1994), 195–209. https://doi.org/10.1080/00036819408840277 doi: 10.1080/00036819408840277
![]() |
[8] |
W. Feng, Competitive exclusion and persistence in models of resource and sexual competition, J. Math. Biol., 35 (1997), 683–694. https://doi.org/10.1007/s002850050071 doi: 10.1007/s002850050071
![]() |
[9] |
W. Feng, M. Freeze, X. Lu, On competition models under Allee effect: asymptotic behavior and traveling waves, Commun. Pure Appl. Anal., 19 (2020), 5609–5626. http://dx.doi.org/10.3934/cpaa.2020256 doi: 10.3934/cpaa.2020256
![]() |
[10] |
W. Feng, X. Lu, Some coexistence and extinction results for a 3-species ecological system, Differ. Integral Equ., 8 (1995), 617–626. http://dx.doi.org/10.57262/die/1369316510 doi: 10.57262/die/1369316510
![]() |
[11] |
W. Feng, X. Lu, Harmless delays for permanence in a class of population models with diffusion effects, J. Math. Anal. Appl., 206 (1997), 547–566. https://doi.org/10.1006/jmaa.1997.5265 doi: 10.1006/jmaa.1997.5265
![]() |
[12] |
W. Feng, W. Ruan, Coexistence, permanence, and stability in a three species competition model, Acta Math. Appl. Sinica (English Ser.), 12 (1996), 443–446. https://doi.org/10.1007/BF02029074 doi: 10.1007/BF02029074
![]() |
[13] | A. Leung, Nonlinear systems of partial differential equations: applications to life and physical sciences, Singapore: World Scientific, 2009. https://doi.org/10.1142/7353 |
[14] |
H. Liu, Y. Ye, Y. Wei, W. Ma, M. Ma, K. Zhang, Pattern formation in a reaction-diffusion predator-prey model with weak Allee effect and delay, Complexity, 2019 (2019), 6282958. https://doi.org/10.1155/2019/6282958 doi: 10.1155/2019/6282958
![]() |
[15] |
X. Lu, W. Feng, Dynamics and numerical simulations of food-chain populations, Appl. Math. Comput., 65 (1994), 335–344. https://doi.org/10.1016/0096-3003(94)90186-4 doi: 10.1016/0096-3003(94)90186-4
![]() |
[16] | C. V. Pao, Nonlinear parabolic and elliptic equations, New York: Plenum Press, 1992. https://doi.org/10.1007/978-1-4615-3034-3 |
[17] |
J. Shi, R. Shivaji, Persistence in reaction diffusion models with weak Allee effect, J. Math. Biol., 52 (2006), 807–829. http://dx.doi.org/10.1007/s00285-006-0373-7 doi: 10.1007/s00285-006-0373-7
![]() |
[18] |
G. Wang, X. Liang, F. Wang, The competitive dynamics of populations subject to an Allee effect, Ecol. Modell., 124 (1999), 183–192. https://doi.org/10.1016/S0304-3800(99)00160-X doi: 10.1016/S0304-3800(99)00160-X
![]() |
[19] |
L. Zhou, C. V. Pao, Asymptotic behavior of a competition-diffusion system in population dynamics, Nonlinear Anal.: Theory Method Appl., 6 (1982), 1163–1184. https://doi.org/10.1016/0362-546X(82)90028-1 doi: 10.1016/0362-546X(82)90028-1
![]() |
[20] |
S. Zhou, C. Liu, G. Wang, The competitive dynamics of metapopulationss subject to the Allee-like effect, Theor. Popul. Biol., 65 (2004), 29–37. https://doi.org/10.1016/j.tpb.2003.08.002 doi: 10.1016/j.tpb.2003.08.002
![]() |
[21] |
S. Zhou, Y. Liu, G. Wang, The stability of predator-prey systems subject to the Allee effects, Theor. Popul. Biol., 67 (2005), 23–31. https://doi.org/10.1016/j.tpb.2004.06.007 doi: 10.1016/j.tpb.2004.06.007
![]() |
1. | Xiaoming Su, Jiahui Wang, Adiya Bao, Stability analysis and chaos control in a discrete predator-prey system with Allee effect, fear effect, and refuge, 2024, 9, 2473-6988, 13462, 10.3934/math.2024656 |