Loading [MathJax]/jax/output/SVG/jax.js
Research article

Global dynamics of a Lotka-Volterra competition-diffusion-advection system for small diffusion rates in heterogenous environment

  • Received: 20 September 2020 Accepted: 23 November 2020 Published: 14 December 2020
  • We investigate the global dynamics of a Lotka-Volterra competition-diffusion-advection system for small diffusion rates in heterogenous environment. Our result suggests that the sign of L0(m1m2)ekxdx plays a significant role in understanding the global dynamics. In addition, the limiting behavior of coexistence steady state is obtained when diffusion rates of two species tend to zero meanwhile.

    Citation: Jinyu Wei, Bin Liu. Global dynamics of a Lotka-Volterra competition-diffusion-advection system for small diffusion rates in heterogenous environment[J]. Mathematical Biosciences and Engineering, 2021, 18(1): 564-582. doi: 10.3934/mbe.2021031

    Related Papers:

    [1] Jinyu Wei, Bin Liu . Coexistence in a competition-diffusion-advection system with equal amount of total resources. Mathematical Biosciences and Engineering, 2021, 18(4): 3543-3558. doi: 10.3934/mbe.2021178
    [2] Abdelrazig K. Tarboush, Jing Ge, Zhigui Lin . Coexistence of a cross-diffusive West Nile virus model in a heterogenous environment. Mathematical Biosciences and Engineering, 2018, 15(6): 1479-1494. doi: 10.3934/mbe.2018068
    [3] Azmy S. Ackleh, Keng Deng, Yixiang Wu . Competitive exclusion and coexistence in a two-strain pathogen model with diffusion. Mathematical Biosciences and Engineering, 2016, 13(1): 1-18. doi: 10.3934/mbe.2016.13.1
    [4] Pengfei Liu, Yantao Luo, Zhidong Teng . Role of media coverage in a SVEIR-I epidemic model with nonlinear incidence and spatial heterogeneous environment. Mathematical Biosciences and Engineering, 2023, 20(9): 15641-15671. doi: 10.3934/mbe.2023698
    [5] Wenzhang Huang, Maoan Han, Kaiyu Liu . Dynamics of an SIS reaction-diffusion epidemic model for disease transmission. Mathematical Biosciences and Engineering, 2010, 7(1): 51-66. doi: 10.3934/mbe.2010.7.51
    [6] Monika Joanna Piotrowska, Urszula Foryś, Marek Bodnar, Jan Poleszczuk . A simple model of carcinogenic mutations with time delay and diffusion. Mathematical Biosciences and Engineering, 2013, 10(3): 861-872. doi: 10.3934/mbe.2013.10.861
    [7] Xiaowei An, Xianfa Song . A spatial SIS model in heterogeneous environments with vary advective rate. Mathematical Biosciences and Engineering, 2021, 18(5): 5449-5477. doi: 10.3934/mbe.2021276
    [8] Yunfeng Liu, Guowei Sun, Lin Wang, Zhiming Guo . Establishing Wolbachia in the wild mosquito population: The effects of wind and critical patch size. Mathematical Biosciences and Engineering, 2019, 16(5): 4399-4414. doi: 10.3934/mbe.2019219
    [9] Yongli Cai, Yun Kang, Weiming Wang . Global stability of the steady states of an epidemic model incorporating intervention strategies. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1071-1089. doi: 10.3934/mbe.2017056
    [10] Changwook Yoon, Sewoong Kim, Hyung Ju Hwang . Global well-posedness and pattern formations of the immune system induced by chemotaxis. Mathematical Biosciences and Engineering, 2020, 17(4): 3426-3449. doi: 10.3934/mbe.2020194
  • We investigate the global dynamics of a Lotka-Volterra competition-diffusion-advection system for small diffusion rates in heterogenous environment. Our result suggests that the sign of L0(m1m2)ekxdx plays a significant role in understanding the global dynamics. In addition, the limiting behavior of coexistence steady state is obtained when diffusion rates of two species tend to zero meanwhile.


    Many ecologists and evolutionary biologists are attracted by the issue of why individuals disperse for a number of years. So far, a great deal of researches are addressed to understand the mechanism of dispersal [1,2]. Over the last few decades, researchers from both biology and mathematics have used the reaction-diffusion equations to model population dynamics in spatial ecology and evolution. Among these models, the two species Lotka-Volterra competition-diffusion system perhaps is the most salient example; refer to, e.g., the books [1,3] and some recent works in [4,5,6].

    With further development, researchers become more interested in the study of spatial population dynamics in advective environments which forces organisms to move in certain directions(biased movements) modeled by reaction-diffusion-advection equations. For instance, a part of researchers pay attention to an active research area concerning the population dynamics in which the individuals are very intelligent so that they can sense the surroundings and move upward along the gradients of a resource distribution. Belgacem and Consner [7] was the first to raise the single species model and then Cantrell et al. [2] presented the two-species model. From the direction of different research, a portion of researchers investigate population dynamics of some aquatic ecosystem modeled by reaction-diffusion-advection equations. The aquatic ecosystems are environments which are featured by a predominantly unidirectional flow, such as rivers, water columns or streams [8,9,10,11,12,13,14]. The organisms will move passively toward the downstream end due to unidirectional water flow if species lives in rivers and streams, or move upward(downward) due to buoyancy(gravity) if species is in water columns. Specially, environmental conditions shift can produce biased motion, such as, the movement of temperature isoclines, which is caused by global climate change[15]. Speirs and Gurney [16] proposed the following single compartment model with diffusion, advection and a logistic growth from river ecosystems.

    {ut=duxxαux+u[mu],0<x<L,t>0,dux(0,t)αu(0,t)=0,t>0,u(L,t)=0,t>0,u(x,0)=u0,0,0<x<L, (1.1)

    where u(x,t) denotes the population density at location x and time t, d is the diffusion rate, L is the size of the habitat, and in the sequel, we call x=0 the upstream end and x=L the downstream end, α measures the tendency of the biased movement by water flow (sometimes we call α the advection speed/rate) and we point out here that α should be positive since it is defined that x=L is the downstream end. At the upstream end x=0, the no-flux type condition is imposed, which indicates that there is no individuals movement in or out through the upstream end. However, at the downstream end x=L, the hostile condition is assumed which means that once individuals pass through the downstream end they do not return back(e.g., stream to ocean [16]). The constant m>0 accounts for intrinsic growth rate, which indicates that the environment is spatially homogeneous. However, as we know, spatial characteristics of the environment play a vital role in ecology and evolution, and the uneven distribution of resources caused by the effect of geological and environmental heterogeneity could create very interesting phenomena in population dynamics. Moreover, the phenomenon of spatial heterogeneity of resources will create more complexity to investigate the global dynamics.

    Recently, Tang and Zhou [17] considered the situation where a new or exotic species invades such advective environment. They investigated the competitive consequence of two species model in nonhomogeneous environment, to be more specific, the following two-species Lotka-Volterra competition-diffusion-advection system:

    {ut=d1uxxα1ux+u[m1(x)uv],0<x<L,t>0,vt=d2vxxα2vx+u[m2(x)uv],0<x<L,t>0,d1ux(x,t)α1u(x,t)=d2ux(x,t)α2u(x,t)=0,x=0,L,t>0,u(x,0)=u0,0,v(x,0)=v0,0,0<x<L, (1.2)

    where u and v stand for the population densities of two competing aquatic species with their inter-specific and intra-specific competition intensities all equal to 1. d1, d2 and α1, α2 are the diffusion and advection rates respectively. The functions m1(x) and m2(x) represent the intrinsic growth rates at location x of these two competitors, which also reflect the distribution of resources. The parameter L>0 measures the length of the one-dimensional habitat. And at the boundary of the habitat, it is assumed that there is no net flux across for either of the two species which implies that the environment is closed. In their study, they imposed that m2=M where M is a constant, that is, the distribution of resources for the species u is spatially uneven while for the species v is homogeneous. By employing principal spectral theory, they investigated the global dynamics of system (1.2) on the condition that: (H)α1/d1=α2/d2=:k>0 and found that the competitive result was depended on d1,d2 and m1. However, for the general situation where m1m2 and m1 together with m2 is the function of spatial variation x, the global dynamics of system (1.2) is far from being completely understood. It is common to ask if the outcome of competition when m2 is a function with respect to x is more complex than when m2 is a constant. We will pursue further in this direction.

    For the spatially homogeneous case m1=m2m0 with m0 being a positive constant, a lot of researchers are interested in it and have investigated it qualitatively [18,19,20,21,22]. In the case that m1=m2=:m(x), non-constant, that is, spatially nonhomogeneous, system (1.2) is more difficult to handle and has been explored in many works. Lam et al. [23] seemed the first to try to discuss the case d1d2 and α1=α2, which directed at the existence and diversity of evolutionarily stable strategies applying some limiting arguments (in the sense of both diffusion and advection rates are sufficiently small and comparable). Zhao and Zhou [24], considering on the special case d1d2 and α1=0<α2, which meant that one species merely suffered random while the other one underwent both random and advective movements, attempted to uncover some various phenomena. For the general case d1d2 and α1α2, recently, Lou et al. [25], by developing new techniques to surmount the difficult caused by non-self-adjoint operators, obtained a profound understanding on the global dynamics. The more general case m1m2 now is far from being understood completely. Zhou and Xiao [26] established a classification of all possible long time behaviors for a more general competitive system in the condition of (H). Indeed, they discussed it in higher spatial dimensions.

    Motivated by Tang and Zhou [17], in this paper, we mainly investigate the dynamical behaviors of system (1.2). Firstly, we need to make the following basic assumptions:

    (H1)α1/d1=α2/d2=:k>0 (or equivalently, d2/d1=α2/α1=:k>0);

    (H2)miC1+γ([0,L]) with some γ(0,1) is nonconstant, mi(x)0 in [0,L], m1(x)m2(x), mi(x)c0ekx, where c0=MLk/(ekL1)(i=1,2) and

    1LL0m1(x)dx=1LL0m2(x)dx=M.

    Condition (H1), biologically, means that the movement strategies(random diffusion and advection rates) of two competitors are proportional, and it is a mathematically technical condition in the main body of this paper. In assumption (H2), it means the two competing populations have the same amount of total resources and the distributions of resources are spatially heterogeneous as well as nonidentical. Moreover, we exclude the ideal free distribution introduced in [27].

    According to the above hypotheses, system (1.2) changes to

    {ut=d1[uxxkux]+u[m1(x)uv],0<x<L,t>0,vt=d2[vxxkvx]+u[m2(x)uv],0<x<L,t>0,ux(x,t)ku(x,t)=vx(x,t)kv(x,t)=0,x=0,L,t>0,u(x,0)=u0,0,v(x,0)=v0,0,0<x<L, (1.3)

    where, the variables in the system (1.3) have the same meanings as in the system (1.2). Moreover, as we know, system (1.3) has one trivial steady state (0,0) (always linearly unstable) and two semi-trivial steady states respectively denoted by (θd1,k,m1,0) and (0,θd2,k,m2) in the sequel.

    Following the approach in [22], we first give some notations:

    Γ:=R+×R+andR+=(0,),

    and

    Σu : = {(d1,d2)Γ:(θd1,k,m1,0) is linearly stable},

    Σv : = {(d1,d2)Γ:(0,θd2,k,m2) is linearly stable},

    ˜Σu : = {(d1,d2)Γ:(θd1,k,m1,0) is neutrally stable},

    ˜Σv : = {(d1,d2)Γ:(0,θd2,k,m2) is neutrally stable},

    ˜Σvu : = ˜Σu˜Σv,

    Σo : = {(d1,d2)Γ: (θd1,k,m1,0) and (0,θd2,k,m2) are linearly unstable}.

    The definitions of the linear stability/instability and the neutrally stability of a steady state of system (1.3) will be given precisely in Section 2.

    Due to assumption (H1), the following complete classification on the global dynamics of system (1.3) can be obtained directly from (Theorem 1.2 [26]). It is a special case of (Theorem 1.2 [26]), where the advective direction P(x)=x, inter-specific competition ability b=c=1, and the habitat Ω=[0,L].

    Assume that (H1) and (H2) hold. Then for system (1.3), we have the following mutually disjoint decomposition of Γ:

    Γ=(Σu˜Σu˜Σvu)(Σv˜Σv˜Σvu)Σo˜Σvu. (1.4)

    Moreover,

    (ⅰ') for all (d1,d2)(Σu˜Σu˜Σvu), (θd1,k,m1,0) is g.a.s;

    (ⅱ') for all (d1,d2)(Σv˜Σv˜Σvu), (0,θd2,k,m2) is g.a.s;

    (ⅲ') for all (d1,d2)Σo, system (1.3) has a coexistence steady state that is g.a.s;

    (ⅳ') for all (d1,d2)˜Σvu, θd1,k,m1=θd2,k,m2 in (0,L) and system (1.3) has a compact global attractor consisting of a continuum of steady states

    {(ρθd1,k,m1,(1ρ)θd2,k,m2:ρ[0,L]}

    connecting the two semi-trivial steady states;

    where g.a.s means that the steady state is globally asymptotically stable among all non-negative and nontrivial initial conditions.

    A basic classification on all possible long time behaviors of system (1.3) is exhibited by the above statements. However, to obtain a transparent picture of the global dynamics of system (1.3), it remains to know explicitly when Σu,˜Σu,Σv,˜Σv and Σo will happen. Equivalently speaking, is it possible to provide a sharp division of these sets by using certain variable parameters? As mentioned in [26], each component of these sets could be empty, and more challengingly, it is hard to give a criteria guaranteeing the dynamics in these sets. Even for the non-advective case, it is not yet completely solved [4].

    In this paper, we study the global dynamics of system (1.3) which contains two competing species. We assume that the two species both are in heterogeneous environment and denote their intrinsic growth rates by the functions m1(x) and m2(x), respectively. As we known, the distribution of resources is uneven in the natural environment, hence, this case is of more realistic significance. In the condition that m1m2 and total resources of the two species are fixed at the same level(L0m1(x)dx=L0m2(x)dx), we find that the sign of L0(m1m2)ekxdx plays an extremely important role in determining the global dynamics. This result indicates that the values of m1 and m2 influence the global dynamics of the system. Moreover, we obtain the limiting behaviour of the coexistence steady state when the diffusion rates(d1 and d2) of the two species tend to zero.

    The rest of this paper is organized as follows. Section 2 contains some preliminaries which are useful in later analysis and our main results. In Section 3, we will give the proof of Theorem 1.1. We prove Theorem 1.2 in Section 4. Finally, we give a short discussion.

    This section is aim to display our main results and exhibit some fundamental results which will be utilized in later sections.

    We obtain the following result when one of d1 and d2 tends to zero.

    Theorem 1.1. Assume that (H1) and (H2) hold. The following statements are valid:

    (ⅰ) Fix d2>0.

    (i1) If L0(m1m2)ekxdx0, then for small d1>0, (θd1,k,m1,0) is linearly unstable;

    (i2) If L0(m1m2)ekxdx>0, then for small d1>0, (θd1,k,m1,0) is g.a.s provided d2>1/μ1(m2m1) and linear unstable provided d2<1/μ1(m2m1).

    (ⅱ) Fix d1>0.

    (ii1) If L0(m1m2)ekxdx0, then for small d2>0, (0,θd2,k,m2) is linearly unstable;

    (ii2) If L0(m1m2)ekxdx<0, then for small d2>0, (0,θd2,k,m2) is g.a.s provided d1>1/μ1(m1m2) and linear unstable provided d1<1/μ1(m1m2);

    where μ(h) is the unique nonzero principal eigenvalue of problem (2.4).

    Remark 1.1. In [17], the authors found that the outcome of competition in general heterogeneous distribution is very complicate: either u wins, or v wins, or uv coexists, depending on the size of diffusion rates d1, d2 and m1. Similar to [17], Theorem 1.1 indicates that the outcome of competition depends on the size of diffusion rates d1 and d2; however, in this paper, because the resource functions of the two competing species u and v are both spatially nonhomogeneous, then the outcome of competition depends not only on the distribution of m1 but also the distribution of m2. Moreover, the competition outcome in this situation will be more abundant.

    Remark 1.2. According to Theorem 1.1, we find that the sign of the quantity L0(m1m2)ekxdx plays a significant role in understanding the global dynamics of system (1.3) which is similarly to the description mentioned in [16]. Then we give some sufficient conditions about determining the sign of the quantity L0(m1m2)ekxdx on the condition that m1 and m2 are monotonic with respect to spatial variable x. In view of the monotonicity of m1 and m2, we can divide into six situations to talk about and the graphs of m1 and m2 are as follows Figure 1. From Figure 1, the following statements are true:

    Figure 1.  x0(0,L) is the unique point such that m1(x)=m2(x). (a) describes the situation where m1(x)>0, m2(x)>0 and m1(x)>m2(x) when x>x0; (b) shows that m1(x)>0, m2(x)>0 and m2(x)>m1(x) when x>x0; (c) displays that m1(x)>0 and m2(x)<0; (d) declares that m1(x)<0 and m2(x)>0; (e) indicates that m1(x)<0, m2(x)<0 and m1(x)<m2(x) when x>x0; (f) describes that m1(x)>m2(x) when x>x0 and m1(x)<0, m2(x)<0.

    (ⅰ) In cases of a, c and f, it holds that L0(m1m2)ekxdx<0;

    (ⅱ) In cases of b, d and e, we can obtain L0(m1m2)ekxdx>0.

    In fact, since m1 and m2 are monotonic, there is a unique x0(0,L) such that m1(x)=m2(x). Define

    ˜q(s):=L0(m2m1)es(xx0)dx.

    Then

    ˜q(s)=L0(m2m1)(xx0)es(xx0)dx
    {<0,m1(x)>0,m2(x)>0,m1(x)<m2(x)ifx<x0,>0,m1(x)>0,m2(x)>0,m1(x)<m2(x)ifx>x0,<0,m1(x)>0,m2(x)<0,>0,m1(x)<0,m2(x)>0,>0,m1(x)<0,m2(x)<0,m1(x)>m2(x)ifx<x0,<0,m1(x)<0,m2(x)<0,m1(x)<m2(x)ifx<x0,

    which, by q(0)=L0(m2m1)dx=0, implies

    ˜q(s){<0,m1(x)>0,m2(x)>0,m1(x)<m2(x)ifx<x0,>0,m1(x)>0,m2(x)>0,m1(x)<m2(x)ifx>x0,<0,m1(x)>0,m2(x)<0,>0,m1(x)<0,m2(x)>0,>0,m1(x)<0,m2(x)<0,m1(x)>m2(x)ifx<x0,<0,m1(x)<0,m2(x)<0,m1(x)<m2(x)ifx<x0.

    The desired result follows by letting s=k.

    When d1 and d2 tend to zero meanwhile, the following result is true.

    Theorem 1.2. Assume that (H1) and (H2) hold. Then we have:

    (ⅰ) As d1 and d2 tend to 0, system (1.3) has a coexistence steady state that is g.a.s;

    (ⅱ) Let us denote by (ud1,d2,vd1,d2) the unique coexistence stable state of system (1.3). Then we have

    limd1,d20(ud1,d2,vd1,d2)=(S(x),H(x)),

    uniformly on compact subsets of [0,L]{x[0,L]:m1(x)=m2(x)}, where

    S(x)={m1(x),ifm1(x)>m2(x),0,ifm1(x)<m2(x),

    and

    H(x)={0,ifm1(x)>m2(x),m2(x),ifm1(x)<m2(x).

    Remark 1.3. From a biological perspective, Remark 1.2, together with Theorems 1.1 and 1.2 implies several interesting explanations. Back to Theorem 1.1 (i1), if the resource m1 of species u is distributed decreasingly in spatial variable x and m2 of species v is distributed increasingly in spatial variable x or m1 and m2 increase in spatial variable meanwhile m1 changes faster than m2 or m1 and m2 decrease in spatial variable meanwhile m1 changes slower than m2, then d1k is not good for u no matter how small d1 is, however, in view of Theorem 1.1 (ii2), it is good for v if d2 is suitably small. Otherwise, from Theorem 1.1 (ii1), it is unfavorable for v no matter how small d2 is. In view of Theorem 1.1 (i2), it is favorable for u if d1 is suitably small. Therefore, we can infer Theorem 1.2 (i), which says that both species will coexist if they nearly do not move (d1,d20), regardless of the shape of m1 and m2.

    Let us look at the following single-species model corresponding to system (1.3)

    {θt=d[θxxkθx]+θ[m(x)θ],0<x<L,t>0,θx(x,t)kθ(k,t)=0,x=0,L,t>0,θ(x,0)=θ0,0,0<x<L, (2.1)

    where d,k>0, m is in C1+γ([0,L]) and m>0 in [0,L]. Under these conditions, it is easy to see that problem (2.1) admits a unique positive steady state (see, e.g., [3]), denoted by θd,k,m. By applying this fact to system (1.3), there are always two semi-trivial steady states (θd1,k,m1,0) and (0,θd2,k,m2) in the condition of (H2).

    About the positive steady solution θd,k,m of Eq (2.1), we have the following limiting profile. The limiting profile can be seen as an extension of the non-advective case (i.e., k=0); see [28,29].

    Lemma 2.1. The following statement on θd,k,m is true:

    limd0θd,k,mmL(0,L)=0.

    Proof. Denote θd,k,m by θd for simplicity. Let βd(x)=ekxθd. Then βd(x) satisfies

    {d[βdxx+kβdx]+βd(m(x)βdekx)=0,0<x<L,βdx(x)=0,x=0,L. (2.2)

    For any given ϵ>0, since m is in C1+γ([0,L]), there is some function ξ in C2([0,L)] such that

    ξx(0)=ξx(L)=0andξmekxL([0,L])<ϵ.

    Define β+:=ξ+2ϵ and β:=ξ2ϵ. Then for d sufficiently small, it is not difficult to prove that β+ and β are, respectively, super- and sub-solution of problem (2.2). That is to say, if d>0 is small, we have

    ξ2ϵβdξ+2ϵ,x[0,L],

    and therefore,

    mekx3ϵβdmekx+3ϵ,x[0,L],

    due to the arbitrariness of ϵ>0, the desired result is proved.

    Next we consider an auxiliary eigenvalue problem. It is helpful for studying the existence and stability of semi-trivial steady states:

    {d[ϕxxkϕx]+hϕ+λϕ=0,0<x<L,ϕxxkϕx=0,x=0,L, (2.3)

    where d,k>0 and h is in L([0,L]). By employing the Krein-Rutman Theorem [30], it is simple to prove that problem (2.3) admits a principal eigenvalue(simple and has the least real part among all eigenvalues) which is denoted by λ1(d,k,h) in the sequel, and its corresponding eigenfunction ϕ1(d,k,h) can be chosen strictly positive in [0,L]. Moreover, the linear stabilities of (θd1,k,m1,0) and (0,θd2,k,m2) are determined, respectively, by the sign of λ1(d2,k,m2θd1,k,m1) and λ1(d1,k,m1θd2,k,m2). Specifically, (θd1,k,m1,0) is linearly stable (resp. linearly unstable) if λ1(d2,k,m2θd1,k,m1)>0 (resp. λ1(d2,k,m2θd1,k,m1)<0); (0,θd2,k,m2) is linearly stable (resp. linearly unstable) if λ1(d1,k,m1θd2,k,m2)>0 (resp. λ1(d1,k,m1θd2,k,m2)<0). We say (θd1,k,m1,0)((0,θd2,k,m2)) is neutrally stable if λ1(d2,k,m2θd1,k,m1)=0(λ1(d1,k,m1θd2,k,m2) = 0).

    By the variational approach, λ1(d,k,h) can be characterized by

    λ1=inf0ψH1(0,L)L0dψ2xekxdxL0hψ2ekxdxL0ekxψ2dx (2.4)

    The following Lemma 2.2 collects some important properties of λ1(d,k,h). The proof can be found in [1,17,31,32].

    Lemma 2.2. The following statements on λ1(d,k,h) are true:

    (ⅰ) λ1(d,k,h) and ϕ1(d,k,h) depend continuously and differentially on parameters d and k;

    (ⅱ) λ1(d,k,h) is strictly increasing in the parameter d provided h is non-constant;

    (ⅲ) limd0λ1(d,k,h)=maxx[0,L]h(x) and limdλ1(d,k,h)=L0h(x)ekxdx/L0ekxdx.

    For later research, it is needed to introduce the following eigenvalue problem with indefinite weight

    {ϕxxkϕx+μhϕ=0,0<x<L,ϕxkϕ=0,x=0,L, (2.5)

    where h0, could change sign in (0,L). If there is a positive solution for problem (2.5) we say μ1=μ1(h) is a principal eigenvalue.(0 is always a principal eigenvalue.) The following result collects some important properties of μ1(h) in connection with λ1(d,k,h). The eigenvalue problem (2.5) is a special case of a general problem studied in [33,34]. Indeed, by the transformation ϱ=ekxϕ, one can change problem (2.5) to a Neumann type problem.

    Lemma 2.3. The problem (2.5) has a nonzero principal eigenvalue μ1=μ1(h) if and only if L0hekxdx0 and h changes sign in (0,L). More precisely, if h changes sign in (0,L), then

    (ⅰ) L0hekxdx=00 is the only principal eigenvalue;

    (ⅱ) L0hekxdx>0μ1(h)<0;

    (ⅲ) L0hekxdx<0μ1(h)>0; and in this case, μ1(h) can be characterized by

    μ1(h)=infψH1((0,L)),L0hψ2ekxdx>0L0ekxψ2xdxL0hψ2ekxdx; (2.6)

    (ⅳ) L0hekxdx0λ1(d,k,h)<0 for all d>0, and

    L0hekxdx<0{λ1(d,k,h)<0,ifd<1μ1(h),λ1(d,k,h)=0,ifd=1μ1(h),λ1(d,k,h)>0,ifd>1μ1(h). (2.7)

    For clarity, we proceed with the proof in the from of several claims.

    Claim 1: There are two functions ˆd2 = ˆd2(d1) and ˆd1 = ˆd1(d2) defined by

    ˆd2(d1)={0,ifmaxx[0,L](m2θd1,k,m1)0,,ifL0(θd1,k,m1m2)ekxdxL0ekxdx0,1μ1(m2θd1,k,m1)(0,),otherwise, (3.1)

    and

    ˆd1(d2)={0,ifmaxx[0,L](m1θd2,k,m2)0,,ifL0(θd2,k,m2m1)ekxdxL0ekxdx0,1μ1(m1θd2,k,m2)(0,),otherwise, (3.2)

    such that

    {Σu:={(d1,d2)Γ:d2>ˆd2(d1)}Σv:={(d1,d2)Γ:d1>ˆd1(d2)}Σo:={(d1,d2)Γ:d2<ˆd2(d1)andd1<ˆd1(d2)} (3.3)

    and the following statements are valid:

    (a) for all (d1,d2)(Σu˜Σu), (θd1,k,m1,0) is g.a.s;

    (b) for all (d1,d2)(Σv˜Σv), (0,θd2,k,m2) is g.a.s;

    (c) for all (d1,d2)Σo, system (1.3) has a coexistence steady state that is g.a.s.

    Where μh is the nonzero principal eigenvalue of Eq (2.5).

    We first assert that ˜Σvu=. On the contrary, suppose that ˜Σvu. In view of statement (iv), one finds that for some d01 and d02>0, θd01,k,m1θd02,k,m2 in [0,L].

    If (d02d01)=0, by the equations of θd01,k,m1 and θd02,k,m2, it holds that

    m1(x)θd01,k,m1d01m2(x)θd02,k,m2d02,x[0,L].

    From the above equation, one can derive that m1m2 in [0,L], which contradicts our assumption (H2).

    If (d02d01)0, by the equations of θd01,k,m1 and θd02,k,m2 again, one arrives at

    L0m1(x)θd01,k,m1d01dx=L0m2(x)θd02,k,m2d02dx,x[0,L].

    Because of the condition (H2), the above equality can be rewritten as

    (d02d01)L0m1(x)dx=(d02d01)L0θd01,k,m1dx,

    which implies that

    L0[m1(x)θd01,k,m1]dx=0.

    Dividing the first equation of Eq (2.1) by ekxθd,k,m and integrating over [0,L], one deduces that

    dL0ekx[ekxθd,k,m]2x[ekxθd,k,m]2dx+L0[m(x)θd,k,m]ekxdx=0,

    so the positive steady solution θd,k,m of problem (2.1) satisfies

    L0[m(x)θd,k,m]ekxdx0,

    and the equality is valid if and only if m(x)=cekx for some c>0.

    Due to the assumption (H2), there is a contradiction. This contradiction gives that ˜Σvu=. Then statements (a),(b) and (c) follow directly from Theorem 1.1.

    It is sufficient to prove Eqs (3.1) and (3.2) can be verified in the same manner so we omit here. Given any d1>0, as shown above, the sign of λ1(d2,k,m2θd1,k,m1) determines the linear stability of (θd1,k,m1,0), which, by Lemma 2.2 (ⅲ), we have

    limd20λ1(d2,k,m2θd1,k,m1)=maxx[0,L](m2θd1,k,m1),

    and

    limd2λ1(d2,k,m2θd1,k,m1)=L0(m2θd1,k,m1)ekxdxL0ekxdx.

    For the first case, if maxx[0,L](m2θd1,k,m1)0, by Lemma 2.2 (ⅱ) and the above first limit, λ1(d2,k,m2θd1,k,m1)>0 for every d2, so we can define ˆd2(d1)=0; similarly, for the second case, using Lemma 2.2 (ⅱ) again and the above second limit, one can define ˆd2(d1)= if L0(m2θd1,k,m1)ekxdx/L0ekxdx0; for the rest case, from Lemma 2.3 (ⅳ) it can be defined that ˆd2(d1)=1/μ1(m2θd1,k,m1)(0,).

    By the above statements, the formula in Eq (3.1) is established.

    In view of the definition of linear stability and unstability, it is easy to understand the sets described in Eq (3.3).

    Claim 2: The following limiting behaviors of ˆd2(d1) and ˆd1(d2) hold true:

    For ˆd2(d1), we have

    limd10ˆd2(d1)={,ifL0(m1m2)ekxdx0,1μ1(m2m1),otherwise; (3.4)

    For ˆd1(d2), we have

    limd20ˆd1(d2)={,ifL0(m1m2)ekxdx0,1μ1(m1m2),otherwise. (3.5)

    As mentioned above, the sign of λ1(d2,k,m2θd1,k,m1) determines the linear stability of (θd1,k,m1,0), which, from Lemma 2.1, satisfies

    limd10λ1(d2,k,m2θd1,k,m1)=λ1(d2,k,m2m1).

    Using Lemma 2.2 (ⅲ), one has

    limd20(limd10λ1(d2,k,m2θd1,k,m1))=maxx[0,L](m2m1)<0, (3.6)

    and

    limd2(limd10λ1(d2,k,m2θd1,k,m1))=L0(m1m2)ekxdxL0ekxdx. (3.7)

    Then we can infer the limit Eq (3.4) from Eqs (3.6) and (3.7) and Lemmas 2.2 (ⅱ) and 2.3 (ⅳ).

    The limit Eq (3.5) can be obtained by the above ideas as in the proof of Eq (3.4). This finishes claim 2.

    Now we are in a position to prove Theorem 1.1. We first prove statement (ⅰ) by fixing d2>0. If L0(m1m2)ekxdx0, by Eq (3.4), for small d1>0 and d2<ˆd2(d1), it can be obtained that (θd1,k,m1,0) is linearly unstable. If L0(m1m2)ekxdx>0, using the second limit in Eqs (3.4) and (3.2), for small d1>0, it is easy to conclude the second result of statement (ⅰ).

    Similar to the above argument, from Eqs (3.1) and (3.5), statement (ⅱ) could be obtained.

    In order to prove statement (ⅰ), the continuous dependence of the principal eigenvalue λ1(d,k,h) in the parameter d as well as the weight function h plays a vital role.

    Specifically, in view of Lemmas 2.1 and 2.3 (ⅳ), one obtains

    limd10,d20λ1(d2,k,m2θd1,k,m1)=maxx[0,L](m2m1)<0,

    and

    limd10,d20λ1(d1,k,m1θd2,k,m2)=maxx[0,L](m1m2)<0.

    From the first inequality, (θd1,k,m1,0) is linearly unstable and (0,θd2,k,m2) is linearly unstable by the second inequality. Due to the statement (c), it is easy to deduce that system 1.3 has a coexistence steady state that is g.a.s.

    Next we prove the limiting behavior of coexistence steady state of system (1.3). Following the ideas in [35], we obtain the proof of this statement.

    Lemma 4.1. (ⅰ) Consider the sequences {¯un}n=n=1 and {v_n}n=n=1 defined by

    ¯u1=θd1,k,m1,v_n=θd2,k,m2¯un,¯un+1=θd1,k,m1v_n,(n1).

    Then, for small d1,d2>0, denote the coexistence state of system (1.3) by (u,v), the following hold:

    u¯un+1¯un,vv_n+1v_n,(n1); (4.1)

    (ⅱ) Analogously for sequences {u_n}n=n=1 and {¯vn}n=n=1 defined by

    ¯v1=θd2,k,m2,u_n=θd1,k,m1¯vn,¯vn+1=θd2,k,m2u_n,(n1).

    We have

    uu_n+1u_n,v¯vn+1¯vn,(n1). (4.2)

    Proof. It is enough to prove Eq (4.1) as the proof of Eq (4.2) follows by the symmetry. The proof is obtained by induction argument.

    From the equation of u, we obtain

    d1[uxxkux]=u[m1uv]u[m1u],x(0,L),

    that is to say, u is a sub-solution of

    {d1[wxxkwx]+w[m1w]=0,x(0,L),wx(x)kw(x)=0,x=0,L.

    Therefore,

    u¯u1:=θd1,k,m1,x(0,L).

    Now, substitute this inequality into the equation of v

    d2[vxxkvx]=u[m2uv]v[m2¯u1v],x(0,L),

    that is, v is a super-solution of

    {d2[zxxkzx]+z[m2¯u1v]=0,x(0,L),zx(x)kz(x)=0,x=0,L,

    so

    vv_1:=θd2,k,m2¯u1,x(0,L).

    Back to the equations of u and v, one can obtain

    u¯u2:=θd1,k,m1v_1,x(0,L).

    and

    vv_2:=θd2,k,m2¯u2,x(0,L).

    For small d1 and d2>0, we have v_1>0 in (0,L), so

    ¯u2:=θd1,k,m1v_1¯u1:=θd1,k,m1,x(0,L),

    which further implies

    v_2:=θd2,k,m2¯u2v_1:=θd2,k,m2¯u1,x(0,L).

    This completes the proof of Eq (4.1) for n=1.

    Now suppose that Eq (4.1) is true for some N>1. Then it suffices to show that

    u¯uN+2¯uN+1,vv_N+2v_N+1,x(0,L).

    From vv_N+1 in (0,L),

    d1[uxxkux]=u[m1uv]u[m1uv_N+1],x(0,L),

    so

    u¯uN+2:=θd1,k,m1v_N+1,x(0,L).

    Similarly, one can derive

    vv_N+2:=θd2,k,m2¯uN+2,x(0,L).

    Thus it is sufficient to show that

    ¯uN+2¯uN+1andv_N+2v_N+1,x(0,L).

    Since v_N+1v_N in (0,L), we get

    ¯uN+2:=θd1,k,m1v_N+1¯uN+1:=θd1,k,m1v_N,x(0,L).

    Similarly,

    v_N+2:=θd2,k,m2¯uN+2v_N+1:=θd2,k,m2¯uN+1,x(0,L).

    This completes the proof.

    We now analyze the behavior of the scheme introduced by Lemma 4.1, Eqs (4.1) and (4.2) as d1 and d2 tend to 0.

    In consideration of Lemma 4.1, for each n1, one can define the following limits in the topology of C([0,L]):

    ¯Un=limd1,d20¯un,U_n=limd1,d20u_n,¯Vn=limd1,d20¯vn,V_n=limd1,d20v_n.

    As a matter of fact, these limits can be described detailedly.

    Lemma 4.2. The following identities hold:

    {¯U1=m1,¯Un+1=[m1(m2¯Un)+]+,x[0,L],¯V1=m2,¯Vn+1=[m2(m1¯Vn)+]+,x[0,L],U_1=(m1m2)+,U_n+1=[m1(m2U_n)+]+,x[0,L],V_1=(m2m1)+,V_n+1=[m2(m2V_n)+]+,x[0,L], (4.3)

    where n1.

    Proof. First, using the same arguments as in proof of Lemma 2.1, if L0m(x)dx>0 and the set {x:m(x)<0} has a positive measure, we can obtain θd,k,mm+L(0,L)0 as d0 with m+(x)=max{m(x),0}.

    Due to assumption (H2), Eq (4.3) is hold when n=1. Suppose that Eq (4.3) is valid for every 1nN. Then, by definition, ¯uN+2:=θd1,k,m1v_N+1. Using the argument above again, one obtains

    ¯UN+2=limd1,d20θd1,k,m1v_N+1=[m1(m2¯uN+1)+]+.

    We can demonstrate the rest ones in the similar manner. Hence, the proof is finished.

    By the above result, it is allowable to get explicit formulas for each of the sequences ¯Un, U_n, ¯Vn and V_n. The following lemma makes these precisely.

    Lemma 4.3. The following formulas for ¯Un,U_n,¯Vn and V_n are true:

    ¯Un+1={m1(x),ifm2(x)m1(x),(n+1)m1(x)nm2(x),ifm1(x)<m2(x)<n+1nm1(x),0,ifm2(x)n+1nm1, (4.4)
    U_n+1={m1(x),ifm1(x)n+1nm2(x),(n+1)[m1(x)m2(x)],ifm2(x)m1<n+1nm2(x),0,ifm1(x)<m2(x), (4.5)
    ¯Vn+1={m2(x),ifm1(x)m2(x),(n+1)m2(x)nm1(x),ifm2(x)<m1(x)<n+1nm2(x),0,ifm1(x)n+1nm2(x), (4.6)
    V_n+1={m2(x),ifm2(x)n+1nm1(x),(n+1)[m2(x)m1(x)],ifm1(x)m2<n+1nm1(x),0,ifm2(x)<m1(x), (4.7)

    where n1.

    Proof. We only prove Eq (4.4) by the induction argument and Eqs (4.5), (4.6) and (4.7) can be verified similarly.

    From Lemma 4.2, it can be inferred that

    ¯U2={m1(x),ifm1(x)m2(x),2m1(x)m2(x),ifm1(x)<m2(x)<2m1(x),0,ifm2(x)2m1(x).

    Hence for n=1, Eq (4.4) is hold.

    Now suppose that Eq (4.4) is true for every 1nN, then one has

    ¯UN+1={m1(x),ifm2(x)m1(x),(N+1)m1(x)Nm2(x),ifm1(x)<m2(x)<N+1Nm1(x),0,ifm2(x)N+1Nm1(x).

    By Lemma 4.2 again, it holds that

    ¯UN+2=[m1(m2¯UN+1)+]+,x[0,L].

    Due to the above two equalities, the following can be obtained that

    ¯UN+2={m1(x),ifm2(x)m1(x),(N+2)m1(x)(N+1)m2(x),ifm1(x)<m2(x)<N+2N+1m1(x),0,ifm2(x)N+2N+1m1(x).

    This completes the proof.

    Now it is in a position to prove

    limd1,d20(ud1d2,vd1,d2)=(S(x),H(x)).

    By Lemmas 4.1 and 4.2, for small d1,d2>0,

    u_nud1,d2¯unandv_nvd1,d2¯vn,for alln1.

    Setting d1,d20, we have

    U_nlim infd1,d20ud1,d2lim supd1,d20ud1,d2¯Un,for alln1,

    and

    V_nlim infd1,d20vd1,d2lim supd1,d20vd1,d2¯Vn,for alln1.

    By further letting n and by Lemma 4.3, one obtains

    limnU_n(x)=limn¯Un(x)={m1(x),ifm1(x)>m2(x),0,ifm1(x)<m2(x),

    and

    limnV_n(x)=limn¯Vn(x)={m2(x),ifm2(x)>m1(x),0,ifm2(x)<m1(x),

    which completes the proof.

    Remark 4.1. If mi(x) changes sign in [0,L](i=1,2), we have

    limd1,d20(ud1,d2,vd1,d2)=(S(x),H(x)),

    uniformly on compact subsets of [0,L]{x[0,L]:m1(x)=m2(x)}, where

    S(x)={m1(x),ifm2(x)<m1(x)and0<m1(x),0,if0<m1(x)<m2(x)orm1(x)0,

    and

    H(x)={0,ifm20or0<m2(x)<m1(x),m2(x),ifm1(x)<m2(x)and0<m2(x).

    In this paper, motivated by Tang and Zhou [17], we studied a classical two-species Lotka-Volterra competition-diffusion-advection system in which the diffusion rates, advection rates and intrinsic growth rates are allowed to take on different values in the space(heterogeneity).

    In the condition that total resources for two populations are fixed at the same level, we consider both species u and v are both in heterogeneous environment. We assume that the two competing species have different resource functions and the distributions of resources are uneven, which is different from the literature [17], where the authors supposed that one spatial distribution is even across space while the other one not. We find that the outcome of competition in this situation is very abundant: either one of the two competitors becomes the final single winner or both populations coexist eventually, which is dependent on the diffusion rates of both species and the specific shapes of m1 and m2; see Theorem 1.1. By limiting arguments, we investigate further the population dynamics when d1 and d2 tend to zero and give the asymptotic behaviour of coexistence steady state for small diffusion; see Theorem 1.2. These results partially generalizes Tang and Zhou [17]. In comparison with Tang and Zhou [17], we study a more general case. In their paper, they assume that the resource function of v expressed as m2 is a constant M. In our research, m2 is a function of the spatial variable x. This case is more realistic. Moreover, our results indicate that the values of m1 and m2 have a significant impact on the global dynamics of (1.3) and the limiting behaviour of the coexistence steady state. Therefore a change in the value of m2 will produce more complicated spatial population dynamics.

    From the above research, we know that both m1 and m2 have effects on global dynamics of system (1.3). It seems that the spatial population dynamics will appear to be more abundant. Moveover, assumption (H1) plays an important role in the proof process. When (H1) fails, it is far away from a complete understanding and extremely challenging to deal with. We will continue to explore these problems in the future.

    The authors express their gratitude to the anonymous reviewers and editors for their valuable comments and suggestions which led to the improvement of the original manuscript.

    This work was partially supported by NNSF of China(No. 11971185).

    All authors declare no conflicts of interest in this paper.



    [1] R. S. Cantrell, C. Cosner, Spatial Ecology via Reaction-Diffusion Equations, John Wiley and Sons, Chichester, UK, 2003.
    [2] R. S. Cantrell, C. Cosner, Y. Lou, Movement toward better environments and the evolution of rapid diffusion, Math. Biosci., 204 (2006), 199–214. doi: 10.1016/j.mbs.2006.09.003
    [3] W. M. Ni, The Mathematics of Diffusion, Society for Industrial and Applied Mathematics, Philedelphia, 2011.
    [4] X. He, W. M. Ni, Global dynamics of the Lotka-Volterra competition-diffusion system: diffusion and spatial heterogeneity Ⅰ, Commun. Pure Appl. Math., 69 (2016), 981–1014. doi: 10.1002/cpa.21596
    [5] X. He, W. M. Ni, Global dynamics of the Lotka-Volterra competition-diffusion system with equal amount of total resources, Ⅱ, Calc. Var. Partial Differ. Equations, 55 (2016), 25.
    [6] X. He, W. M. Ni, Global dynamics of the Lotka-Volterra competition-diffusion system with equal amount of total resources, Ⅲ, Calc. Var. Partial Differ. Equarions, 56 (2017), 132.
    [7] F. Belgacem, C. Cosner, The effects of dispersal along environmental gradients on the dynamics of population in heterogeneous environments, Canad. Appl. Math. Quart., 3 (1995), 379–397.
    [8] Y. Du, S. B. Hsu, Concentration phenomena in a nonlocal quasi-linear problem modeling phytoplankton: Ⅰ. Existence, SIAM J. Math. Anal., 40 (2008), 1419–1440. doi: 10.1137/07070663X
    [9] Y. Du, S. B. Hsu, Concentration phenomena in a nonlocal quasi-linear problem modeling phytoplankton: Ⅱ. Limiting profile, SIAM J. Math. Anal., 40 (2008), 1441–1470. doi: 10.1137/070706641
    [10] Y. Du, S. B. Hsu, On a nonlocal reaction-diffusion problem arising from the modeling phytoplankton growth, SIAM J. Math. Anal., 42 (2012), 1305–1333.
    [11] Y. Du, L. F. Mei, On a nonlocal reaction-diffusion-advection equation modelling phytoplankton dynamics, Nonlinearity, 24 (2011), 319–349. doi: 10.1088/0951-7715/24/1/016
    [12] F. Lutscher, M. A. Lewis, E. McCauley, Effects of heterogeneity on spread and persistence in rivers, Bull. Math. Biol., 68 (2006), 2129–2160. doi: 10.1007/s11538-006-9100-1
    [13] F. Lutscher, E. McCauley, M. A. Lewis, Spatial patterns and coexistence mechanisms in systems with unidirectional flow, Theor. Popul. Biol., 71 (2007), 267–277. doi: 10.1016/j.tpb.2006.11.006
    [14] F. Lutscher, E. Pachepsky, M. A. Lewis, The effect of dispersal patterns on stream populations, SIAM Rev., 47 (2005), 749–772. doi: 10.1137/050636152
    [15] H. Berestycki, O. Diekmann, C. J. Nagelkerke, P. A. Zegeling, Can a species keep pace with a shifting climate? Bull. Math. Biol., 71 (2009), 399–429.
    [16] D. C. Speirs, W. S. Gurney, Population persistence in rivers and estuaries, Ecology, 82 (2001), 1219–1237. doi: 10.1890/0012-9658(2001)082[1219:PPIRAE]2.0.CO;2
    [17] D. Tang, P. Zhou, On a Lotka-volterra competition-diffusion-advection system: Homogeneity vs heterogeneity, J. Differ. Equations, 268 (2020), 1570–1599. doi: 10.1016/j.jde.2019.09.003
    [18] J. Dockery, V. Hutson, K. Mischaikow, M. Pernarowski, The evolution of slow dispersal rates: a reaction-diffusion model, J. Math. Biol., 37 (1998), 61–83. doi: 10.1007/s002850050120
    [19] A. Hasting, Can spatial variation alone lead to selection for dispersal? Theor. Popul., 24 (1983), 244–251. doi: 10.1016/0040-5809(83)90027-8
    [20] Y. Lou, D. M. Xiao, P. Zhou, Qualitative analysis for a Lotka-Volterra competition system in advective homogeneous environment, Discrete Contin. Dyn. Syst. Ser. A, 36 (2016), 953–969.
    [21] Y. Lou, P. Zhou, Evolution of dispersal in advective homogeneous environment: the effect of boundary conditions, J. Differ. Equations, 259 (2015), 141–171. doi: 10.1016/j.jde.2015.02.004
    [22] P. Zhou, On a Lotka-Volterra competition system: diffusion vs advection, Calc. Var. Partial Differ. Equations, 55 (2016), 1570–1599.
    [23] K. Y. Lam, Y. Lou, F. Lutscher, Evolution of dispersal in closed advective environments, J. Biol. Dyn., 9 (2015), 188–212. doi: 10.1080/17513758.2014.969336
    [24] X.Q. Zhao, P. Zhou, On a Lotka-Volterra competition model: the effects of advection and spatial variation, Calc. Var. Partial Differ. Equations, 55 (2016), 73.
    [25] Y. Lou, X. Q. Zhao, P. Zhou, Global dynamics of a Lotka-Volterra competition-diffusion-advection system in heterogeneous environments, J. Math. Pures Appl., 121 (2019), 47–82. doi: 10.1016/j.matpur.2018.06.010
    [26] P. Zhou, D. M. Xiao, Global dynamics of a classical Lotka-Volterra competition-diffusionadvection system, J. Funct. Anal., 275 (2018), 356–380. doi: 10.1016/j.jfa.2018.03.006
    [27] S. D. Fretwell, H. L. Lucas, On territorial behavior and other factors influencing habitat selection in bird, Acta Biotheor., 19 (1969), 45–52. doi: 10.1007/BF01601955
    [28] V. Huston, Y. Lou, K. Mischaikow, Convergence in competition models with small diffusion coefficients, J. Differ. Equations, 211 (2005), 135–161. doi: 10.1016/j.jde.2004.06.003
    [29] K. Y. Lam, N. W. Ni, Uniqueness and complete dynamics in the heterogeneous competitiondiffusion system, SIAM J. Appl. Math., 72 (2012), 1695–1712. doi: 10.1137/120869481
    [30] M. G. Krein, M. A. Rutman, Linear operators leaving invariant a cone in a Banach space, Usp. Mat. Nauk, 3 (1948), 3–95.
    [31] X. F. Chen, Y. Lou, Principal eigenvalue and eigenfunctions of an elliptic operator with large advection and its application to a competition model, Indiana Univ. Math. J., 57 (2008), 627–658. doi: 10.1512/iumj.2008.57.3204
    [32] X. F. Chen, Y. Lou, Effects of diffusion and advection on the smallest eigenvalue of an elliptic operator and their applications, Indiana Univ. Math. J., 61 (2012), 45–80. doi: 10.1512/iumj.2012.61.4518
    [33] P. Hess, S. Senn, Another approach to elliptic eigenvalue problems with respect to indefinite weight functions, Nonlinear Analysis and Optimization, Springer, Berlin, Heidelberg, 1984,106–114.
    [34] S. Senn, P. Hess, On positive solutions of a linear elliptic eigenvalue problem with Neumann boundary conditions, Math. Ann., 258 (1982), 459–470. doi: 10.1007/BF01453979
    [35] V. Huston, J. López-Gómez, K. Mischaikow, G. Vickers, Limit behaviors for a competing species problem with diffusion, Dyn. Syst. Appl., 4 (1995), 343–358.
  • This article has been cited by:

    1. Jinyu Wei, Bin Liu, Coexistence in a competition-diffusion-advection system with equal amount of total resources, 2021, 18, 1551-0018, 3543, 10.3934/mbe.2021178
    2. Jinyu Wei, Bin Liu, Global dynamics of a competition-diffusion-advection system with general boundary conditions, 2024, 54, 0018-2079, 10.32917/h2022019
    3. Yong Yang, Zunxian Li, Chengyi Xia, Forced waves and their asymptotic behaviors in a Lotka-Volterra competition model with spatio-temporal nonlocal effect under climate change, 2023, 20, 1551-0018, 13638, 10.3934/mbe.2023608
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2956) PDF downloads(193) Cited by(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog