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

Global stability of a tridiagonal competition model with seasonal succession

  • Received: 03 December 2022 Revised: 05 January 2023 Accepted: 16 January 2023 Published: 18 January 2023
  • In this paper, we investigate a tridiagonal three-species competition model with seasonal succession. The Floquet multipliers of all nonnegative periodic solutions of such a time-periodic system are estimated via the stability analysis of equilibria. Together with the Brouwer degree theory, sufficient conditions for existence and uniqueness of the positive periodic solution are given. We further obtain the global dynamics of coexistence and extinction for three competing species in this periodically forced environment. Finally, some numerical examples are presented to illustrate the effectiveness of our theoretical results.

    Citation: Xizhuang Xie, Meixiang Chen. Global stability of a tridiagonal competition model with seasonal succession[J]. Mathematical Biosciences and Engineering, 2023, 20(4): 6062-6083. doi: 10.3934/mbe.2023262

    Related Papers:

    [1] Yilei Tang, Dongmei Xiao, Weinian Zhang, Di Zhu . Dynamics of epidemic models with asymptomatic infection and seasonal succession. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1407-1424. doi: 10.3934/mbe.2017073
    [2] Jie Hu, Juan Liu, Peter Yuen, Fuzhong Li, Linqiang Deng . Modelling of a seasonally perturbed competitive three species impulsive system. Mathematical Biosciences and Engineering, 2022, 19(3): 3223-3241. doi: 10.3934/mbe.2022149
    [3] Mahmudul Bari Hridoy . An exploration of modeling approaches for capturing seasonal transmission in stochastic epidemic models. Mathematical Biosciences and Engineering, 2025, 22(2): 324-354. doi: 10.3934/mbe.2025013
    [4] Lale Asik, Jackson Kulik, Kevin Long, Angela Peace . Dynamics of a stoichiometric producer-grazer system with seasonal effects on light level. Mathematical Biosciences and Engineering, 2019, 16(1): 501-515. doi: 10.3934/mbe.2019023
    [5] Yufei Wang, Huidong Cheng, Qingjian Li . Dynamic analysis of wild and sterile mosquito release model with Poincaré map. Mathematical Biosciences and Engineering, 2019, 16(6): 7688-7706. doi: 10.3934/mbe.2019385
    [6] Zhenyao Sun, Da Song, Meng Fan . Dynamics of a stoichiometric phytoplankton-zooplankton model with season-driven light intensity. Mathematical Biosciences and Engineering, 2024, 21(8): 6870-6897. doi: 10.3934/mbe.2024301
    [7] Chentong Li, Jinyan Wang, Jinhu Xu, Yao Rong . The Global dynamics of a SIR model considering competitions among multiple strains in patchy environments. Mathematical Biosciences and Engineering, 2022, 19(5): 4690-4702. doi: 10.3934/mbe.2022218
    [8] Salamida Daudi, Livingstone Luboobi, Moatlhodi Kgosimore, Dmitry Kuznetsov . Dynamics for a non-autonomous fall armyworm-maize interaction model with a saturation functional response. Mathematical Biosciences and Engineering, 2022, 19(1): 146-168. doi: 10.3934/mbe.2022008
    [9] Robert Stephen Cantrell, Brian Coomes, Yifan Sha . A tridiagonal patch model of bacteria inhabiting a Nanofabricated landscape. Mathematical Biosciences and Engineering, 2017, 14(4): 953-973. doi: 10.3934/mbe.2017050
    [10] Harry J. Dudley, Zhiyong Jason Ren, David M. Bortz . Competitive exclusion in a DAE model for microbial electrolysis cells. Mathematical Biosciences and Engineering, 2020, 17(5): 6217-6239. doi: 10.3934/mbe.2020329
  • In this paper, we investigate a tridiagonal three-species competition model with seasonal succession. The Floquet multipliers of all nonnegative periodic solutions of such a time-periodic system are estimated via the stability analysis of equilibria. Together with the Brouwer degree theory, sufficient conditions for existence and uniqueness of the positive periodic solution are given. We further obtain the global dynamics of coexistence and extinction for three competing species in this periodically forced environment. Finally, some numerical examples are presented to illustrate the effectiveness of our theoretical results.



    The far-reaching consequences of ecological interactions in the dynamics of biological communities remain an intriguing subject. The interaction between two species can basically be of three different kinds: competition, mutualism and predator-prey. Competition and predator-prey are, perhaps, the most well-known types of interactions. They embody the natural dispute for resources in ecology. Competition is characterized by a decrease of the growth rate as the density of the other species increases, which has been extensively studied (see [1,2,3,4]). One special competition example occurs in the water column of an ocean or on a steep mountain side or on island groups, where each species dominates a species zone (depth, altitude or different island) but is obliged to interact with other species in the narrow overlap of their zones of dominance. Taking three species as an example, one can think of a hierarchy of species x1, x2 and x3, where xi is the density or biomass of the i-th species. In this hierarchy, x1 only competes with x2, x3 only with x2, and x2 competes with x1 and x3. The concrete form is as follows,

    {dx1dt=x1(r1a11x1a12x2),dx2dt=x2(r2a21x1a22x2a23x3),dx3dt=x3(r3a32x2a33x3),(x1(0),x2(0),x3(0))=x0R3+, (1.1)

    where ri and aij are all positive real numbers. Such system is called tridiagonal competitive system and there have also been a number of theoretical works (see [5,6,7,8]).

    As we know, interactive species often live in a fluctuating environment. Due to seasonal or daily variations, species experience a periodic dynamical environment such as temperature, rainfall, humidity, wind, and other resources. A typical example here is the phytoplankton and zooplankton of the temperate lakes, where the species grow during the warmer months and die off or form resting stages in winter. The phenomenon is described as seasonal succession. It has been a fascinating subject for ecologists and mathematicians to explore the dynamics of periodic models by means of seasonal succession numerically and analytically. Recently, Klausmeier [9] proposed a novel approach, called successional state dynamics (SSD), to modeling seasonally forced food webs. The SSD approach treats succession as a series of state transitions driven by both the internal dynamics of species interactions and external forcing, and can uncover unexpected phenomena such as multiple stable annual trajectories and year-to-year irregularity in successional trajectories (chaos). Steiner et al. [10] later validated the utility of the SSD approach as a framework for predicting the effects of altered seasonality on the structure and dynamics of multitrophic communities by using controlled laboratory experiments.

    However, there are few analytic results on these ecological models with seasonal succession. One of the major reasons is that the vector fields of these models are discontinuous and periodic in time t. Hsu and Zhao [11] first studied the global dynamics of a Lotka-Volterra two-species competition model with seasonal succession and obtained a complete classification for the global dynamics and the effects of season succession on the competition outcomes via the theory of monotone dynamical systems. Recently, Niu et al. [12] were concerned with an n-dimensional Lotka-Volterra competition model with seasonal succession and obtained the existence of a carrying simplex. Based on this, they reconsidered the two-dimensional model proposed by Hsu and Zhao [11] and simplified their approach to acquire the complete classification of global dynamics. In [13], Xie and Niu analyzed a three-dimensional Lotka-Volterra cooperation model with seasonal succession and derived a completed dynamics of global coexistence and extinction, which extends previous results with respect to three-dimensional cooperation models. There are other works on seasonal succession, such as [14,15,16,17,18] and references therein.

    Yet, to our knowledge, there are few works on the global stability for high dimensional competitive systems with seasonal succession due to such obstacles as the estimates for the Floquet multipliers, the existence and uniqueness of the positive periodic solutions, etc. Therefore, it is interesting for us to introduce the seasonal succession into the three-speices tridiagonal competition model and study the global stability for such a periodically forced system. Motivated by the modelling methods in Klausmeier [9], we propose a tridiagonal three-species competition model with seasonal succession as follows:

    {dxidt=λixi, mωtmω+(1φ)ω,i=1,2,3,dx1dt=x1(r1a11x1a12x2), mω+(1φ)ωt(m+1)ω,dx2dt=x2(r2a21x1a22x2a23x3), mω+(1φ)ωt(m+1)ω,dx3dt=x3(r3a32x2a33x3),  mω+(1φ)ωt(m+1)ω,(x1(0),x2(0),x3(0))=x0R3+,  m=0,1,2, (1.2)

    where mZ+, φ[0,1] and ω, λi, ri, aij (i,j=1,2,3) are all positive constants.

    In particular, if φ=0, then the system (1.2) become the linear system

    {dxidt=λixi,i=1,2,3,(x1(0),x2(0),x3(0))=x0R3+. (1.3)

    While, if φ=1, system (1.2) turns out to be the tridiagonal competitive system (1.1).

    Obviously, system (1.2) is a time periodic system in a season alternate environment. Overall period is ω, and φ stands for the switching proportion of a period between two subsystems (1.1) and (1.3). Biologically, φ is used to describe the proportion of the period in the good season in which the species follow system (1.1), while (1φ) represents the proportion of the period in the bad season in which the species die exponentially according to system (1.3).

    In addition, when we write

    ri(t)={λi,[mω,mω+(1φ)ω),bi,[mω+(1φ)ω,(m+1)ω],aij(t)={0,[mω,mω+(1φ)ω),aij,[mω+(1φ)ω,(m+1)ω],

    where i,j=1,2,3, system (1.2) can be expressed as a three-dimensional time ω-periodic tridiagonal competitive system with discontinuous ω-periodic coefficients,

    {dx1(t)dt=x1(t)(r1(t)a11(t)x1(t)a12(t)x2(t)),dx2(t)dt=x2(t)(r2(t)a21(t)x1(t)a22(t)x2(t)a23(t)x3(t)),dx3(t)dt=x3(t)(r3(t)a32(t)x2(t)a33(t)x3(t)),(x1(0),x2(0),x3(0))=x0R3+. (1.4)

    From system (1.2), it can easily be seen that system (1.2) admits a unique nonnegative global solution x(t,x0) on [0,+) for any x0R3+. Since the system is ω-periodic, it suffices to consider the Poincarˊe map S on R3+, that is,

    S(x0)=x(ω,x0),  x0R3+.

    Let us first define a linear map L by

    L(x1,x2,x3)=(eλ1(1φ)ωx1,eλ2(1φ)ωx2,eλ3(1φ)ωx3),(x1,x2,x3)R3+. (1.5)

    We also let {Qt}t0 represent the solution flow associated with the tridiagonal competitive system (1.1). Then, Qt(x0) is the unique global solution of the system (1.1) on [0,+). Thus, we have

    S(x0)=Qφω(Lx0),x0R3+,i.e.,S=QφωL.

    We only need to focus on the dynamics of the discrete-time system {Sn}n0.

    The purpose of this paper is to investigate the global stability for system (1.2). Firstly, the Floquet multipliers of all non-negative periodic solutions, including trivial periodic solution, semi-trivial periodic solutions and the positive periodic solutions, are estimated via the stability analysis of equilibria (see Lemmas 3.1–3.8). To obtain the existence and uniqueness for the positive periodic solution, we provide the index theory of the fixed points for the Poincaré map S (see Lemma 4.1). Compared to the proof of the existence and uniqueness for the positive periodic solution in Xie and Niu [13] and [Lemmas 4.1 and 4.2], our approach is much more general which avoids Brouwer fixed point theorem and the connecting orbits theorem and can be applied to more mappings, such as the the Poincaré maps associated with competition and cooperation models with seasonal succession. Together with the existence of carrying simplex, we obtain that system (1.2) has a unique positive periodic solution under appropriate conditions, and moreover the positive periodic solution is globally asymptotically stable in IntR3+ (see Theorem 4.4). In addition, sufficient conditions for two-species coexistence, two-species extinction and global extinction are given (see Theorems 4.5–4.9). Some numerical examples are provided to illustrate our theoretical results (see Figures 15).

    Figure 1.  Three species will be coexistence.
    Figure 2.  Species 1 and Species 3 will be coexistence.
    Figure 3.  Species 1 and Species 2 will be coexistence.
    Figure 4.  Species 2 will win the competition.
    Figure 5.  Three species will be extinction.

    The paper is organized as follows. In Section 2, we introduce some notations, relevant definitions and preliminaries. Section 3 is devoted to analyzing the local dynamics of all nonnegative periodic solutions of system (1.2). In Section 4, the global dynamics of coexistence and extinction for three competing species in terms of system (1.2) are obtained. In Section 5, we present some numerical simulations to illustrate our analytic results. The paper ends with a discussion in Section 6.

    In this section, we first introduce some definitions and describe some results which are essential tools for the later sections.

    Define Z+ to be the set of nonnegative integers. Let R3+={xR3:xi0, iΛ} and IntR3+={xR3:xi>0, iΛ}, where Λ:={1,2,3}. Let LΛ and ˉL=ΛL be its complementary set in Λ. We define the sets HL={xR3:xj=0  for  jˉL}, H+L=R3+HL, and IntH+L={xH+L:xi>0  for all  iL}. For two vectors x,yR3, we write xy if xiyi for all iΛ, and xy if xi<yi for all iΛ. If xy but xy, we write x<y. If x,yR3 and xy, let [x,y]={zR3:xzy} be a closed order interval, and if xy, let [[x,y]]={zR3:xzy} be a open order interval. For an n×n matrix A, we write A0 iff A is a nonnegative matrix (i.e., all the entries are nonnegative) and A0 iff A is a positive matrix (i.e., all the entries are positive).

    Let XR3+ and S:XX be the Poincarˊe (period) map. The orbit of a state x for S is γ(x)={Sn(x), nZ+}. A fixed point x of S is a point xX such that S(x)=x. A point yX is called a k-periodic point of S if there exists some positive integer k>1, such that Sk(y)=y and Sm(y)y for every positive integer m<k. The k-periodic orbit of the k-periodic point y, γ(y)={y, S(y), S2(y), ..., Sk1(y)}, is often called a periodic orbit for short. The ω-limit set of x is defined by ω(x)={yR3+:Snkxy (k) for some sequence nk+ in Z+}. A set VX is positively invariant under S if SVV, and invariant if SV=V. Note that if the orbit γ(x) of x has compact closure, ω(x) is nonempty, compact and invariant. If S is a differentiable map, we write DS(x) as the Jacobian matrix of S at the point x, and r(DS(x)) stands for the spectral radius of DS(x).

    A continuous map S:R3+R3+ is said to be monotone if S(x)S(y) whenever xy with x,yR3+; S is called strictly monotone if x<y, then S(x)<S(y); strongly monotone if x<y, then S(x)S(y).

    A carrying simplex for the periodic map S is a subset ΣRn+{0} with the following properties [26]:

    (P1) Σ is compact, invariant and unordered;

    (P2) Σ is homeomorphic via radial projection to the (n1)-dim standard probability simplex Δn1:={xRn+| Σixi=1};

    (P3)  xRn+{0}, there exists some yΣ such that limn|SnxSny|=0.

    Lemma 2.1. ([11, Lemma 2.1]) Let x(t,x0) be the unique solution of the following system

    {dxdt=λx,mωtmω+(1φ)ω,m=0,1,2,dxdt=x(rax),mω+(1φ)ωt(m+1)ω,x(0)=x0R+, (2.1)

    where λ, r, a are all positive constants. Then the following two statements are valid:

    (ⅰ) If rφλ(1φ)0, then limtx(t,x0)=0 for all x0R+;

    (ⅱ) If rφλ(1φ)>0, then system (2.1) admits a unique positive ω-periodic solution x(t), and limt(x(t,x0)x(t))=0 for all x0R+{0}.

    Lemma 2.2. (Boundness) The Poincarˊe map S associated with system (1.2) is bounded in R3+. Moreover, every forward orbit of S is precompact in R3+.

    Proof. Firstly, we rewrite system ( 1.1 ) as

    {˙xi(t)=Fi(x),i=1,2,3,(x1(0),x2(0),x3(0))=x0R3+,

    where

    {F1(x)=x1(r1a11x1a12x2),F2(x)=x2(r2a21x1a22x2a23x3),F3(x)=x3(r3a32x2a33x3).

    Let b=max{r1a11+1,r2a22+1,r3a33+1}, we have B:=(b,b,b)0. For any positive number l1, whenever x:=(x1,x2,x3)[0,lB] satisfies x1=lb, then

    ˙x1(t)=F1(x)=lb(r1a11lba12x2)lb(r1a11lb)<0.

    Similarly, for any x[0,lB] satisfies x2=lb, then

    ˙x2(t)=F2(x)=lb(r2a22lba21x1a23x3)lb(r2a22lb)<0.

    while, for any x[0,lB] satisfies x3=lb, we also have

    ˙x3(t)=F3(x)lb(r3a33lba32x2)lb(r3a33lb)<0.

    Thus, [0,lB] is positive invariant for system ( 1.1 ).

    For any x0R3+, there exists some positive number l01 such that 0x0l0B. The positive invariance of [0,l0B] implies that 0x(t,x0)l0B for all t0 where x(t,x0)=(x1(t),x2(t),x3(t)) stands for the solution of system (1.1) with initial value x0 in R3+. Note that {Qt}t0 is the solution flow of system (1.1), it follows that 0Qt(x0)l0B for any t0. By the expression () of L(x1,x2,x3), we have Lx0x0, and then, 0Qt(Lx0)l0B for any t0. This implies that 0Qφω(Lx0)=S(x0)l0B, that is, S is bounded. Based on this, we easily see that every forward orbit of S is precompact. The proof is completed.

    Lemma 2.3. (Monotonicity) If x,yR3+ and S(x)<S(y), then x<y. In particular, if x,yIntR3+ and S(x)<S(y), then xy.

    Proof. For x,yR3+, if S(x)<S(y), then Qφω(Lx)<Qφω(Ly). Define u(t):=Qφωt(Lx) and v(t):=Qφωt(Ly). It is clear that u(t) and v(t) are two solutions of a three dimensional cooperative system for t[0,φω]. Since u(0)=Qφω(Lx)<Qφω(Ly)=v(0), it follows from the comparison theorem of cooperative systems that Lx=u(φω)<v(φω)=Ly. By the expression of the linear map L, we have x<y. For the strong monotonicity of S, see Smith [24, Chapter 3].

    Remark 2.1. By the expression of system (1.4), HL,H+L and IntH+L are positively invariant under the map S. So we can rewrite the Poincarˊe map S as:

    S(x1,x2,x3)=(x1G1(x),x2G2(x),x3G3(x)),   xR3+

    where

    Gi(x):={Si(x)xiif xi0,Sixi(x)if xi=0.

    Moreover, Gi(x) are continuous functions with Gi(x)0 for xR3+,i=1,2,3.

    In this section, we will analyze the local stabilities of all nonnegative fixed points of S in R3+. Clearly, O:=(0,0,0) is a trivial fixed point of S. Let hi:=(riφλi(1φ))ω,i=1,2,3. By Lemma 2.1, if hi>0 (i=1,2,3), then system (1.2) has three axial fixed points, say R1:=(x1,0,0), R2:=(0,x2,0) and R3:=(0,0,x3). By the second and fourth equations of system (1.2), it is not difficult to see that there admits a unique interior fixed point in coordinate planar {x2=0}, say E2=(ˆx1,0,ˆx3). Noticing that Hsu and Zhao [11, Theorems 2.3 and 2.4], system (1.2) has interior fixed points in coordinate planar {x1=0} and {x3=0} under certain conditions. If exist, we write them as E3=(ˉx1,ˉx2,0) and E1=(0,ˇx2,ˇx3), respectively.

    Lemma 3.1. (Stability of the fixed point O)

    (ⅰ) If hi<0 (i=1,2,3), then O is an asymptotically stable fixed point of S.

    (ⅱ) If one of h1,h2 and h3 is greater than 0, then O is an unstable fixed point of S. In particular, if hi>0 (i=1,2,3), then O is a hyperbolic repeller.

    Proof. Let F(x)=(F1,F2,F3)T, where

    {F1=r1x1a11x21a12x1x2,F2=r2x2a21x1x2a22x22a23x2x3,F3=r3x3a32x2x3a33x23.

    Then, DF(x)=

    (r12a11x1a12x2a12x10a21x2r2a21x12a22x2a23x3a23x20a32x3r3a32x22a33x3).

    For simplicity, we denote u(t,x):=Qt(x) and V(t,x):=Dxu(t,x)=DxQt(x). Then S(x)=Qφω(Lx)=u(φω,Lx). Thus,

    DS(x)=D(Qφω(Lx))D(Lx)=V(φω,Lx)D(Lx) (3.1)
    =V(φω,Lx)diag{eλ1(1φ)ω,eλ2(1φ)ω,eλ3(1φ)ω}. (3.2)

    Note that V(t,x) satisfies

    dV(t)dt=DF(u(t,x))V(t),V(0)=I. (3.3)

    Taking x=O, we have u(t,LO)=(0,0,0), which implies that DF(O)=diag{r1,r2,r3}. Then,

    V(φω,LO)=diag{eφω0r1dt,eφω0r2dt,eφω0r3dt},

    and hence, one has

    DS(O)=diag{e(r1φλ1(1φ))ω,e(r2φλ2(1φ))ω,e(r3φλ3(1φ))ω}=diag{eh1,eh2,eh3}.

    Consequently, the matrix DS(O) has three positive eigenvalues μ1,μ2 and μ3 given by

    μ1=eh1,μ2=eh2andμ3=eh3.

    Then, the conclusion is immediate.

    By Lemma 3.1, we assume that hi>0 (i=1,2,3) in following Lemmas 3.2–3.8.

    Lemma 3.2. (Stability of the axial fixed point R1)

    (ⅰ) If h2a21>h1a11, then R1 is a saddle point with one-dimensional stable manifold.

    (ⅱ) If h2a21<h1a11, then R1 is a saddle point with two-dimensional stable manifold.

    Proof. Let u(t,LR1):=(u1(t),0,0). By the proof of Lemma 3.1, it is not difficult to see that

    DF(u(t,LR1))=(r12a11u1(t)a12u1(t)00r2a21u1(t)000r3).

    Then,

    V(φω,LR1)=(eφω0r12a11u1(t)dt0eφω0r2a21u1(t)dt00eφω0r3dt).

    Note that u1(t) satisfies the equation du1(t)dt=u1(t)(r1a11u1(t)), it follows that

    φω0u1(t)dt=h1a11,

    and then,

    DS(R1)=(eh10e(h2a21a11h1)00eh3),

    where stands for unknown algebraic expression. Therefore, the matrix DS(R1) has three positive eigenvalues μ1,μ2 and μ3 given by

    μ1=eh1<1,μ2=e(h2a21a11h1)andμ3=eh3>1.

    Therefore, the conclusion is immediate.

    Similarly, we have the Lemmas 3.3 and 3.4.

    Lemma 3.3. (Stability of the axial fixed point R2)

    (ⅰ) If h2a22>max{h1a12,h3a32}, then R2 is an asymptotically stable fixed point of S.

    (ⅱ) If h2a22<min{h1a12,h3a32}, then R2 is a saddle point with one-dimensional stable manifold.

    (ⅲ) If h1a12<h2a22<h3a32 or h1a12>h2a22>h3a32, then R2 is a saddle point with two-dimensional stable manifold.

    Lemma 3.4. (Stability of the axial fixed point R3)

    (ⅰ) If h2a23>h3a33, then R3 is a saddle point with one-dimensional stable manifold.

    (ⅱ) If h2a23<h3a33, then R3 is a saddle point with two-dimensional stable manifold.

    Lemma 3.5. (Stability of the planar fixed point E3) If h2a21>(<)h1a11,h1a12>(<)h2a22, then there admits a unique interior fixed point, say E3:=(ˉx1,ˉx2,0), for S in the coordinate plane {x3=0}. Moreover,

    (i) If h2a21>h1a11,h1a12>h2a22 and h3a32<h1a21h2a11a12a21a11a22, then E3 is an asymptotically stable fixed point of S.

    (ii) If h2a21<h1a11,h1a12<h2a22 and h3a32<h1a21h2a11a12a21a11a22, then E3 is a saddle point with two-dimensional stable manifold.

    (iii) If h2a21<h1a11,h1a12<h2a22 and h3a32>h1a21h2a11a12a21a11a22, then E3 is a saddle point with one-dimensional stable manifold.

    Proof. By Remark 2.1, we can see that all coordinate axes and planes are positively invariant under the map S. Theorem 3.6 in [12] implies that if h2a21>(<)h1a11,h1a12>(<)h2a22, then system (1.2) admits a unique interior fixed point in the coordinate planar {x3=0}, say E3:=(ˉx1,ˉx2,0). Let F(x), u(t,x) and V(t,x) be denoted as in the proof of Lemma 3.1. We also define u(t,LE3):=(ˉu1(t),ˉu2(t),0). Then,

    DF(u(t,LE3))=(r12a11ˉu1a12ˉu2a12ˉu10a21r22a22ˉu2a21ˉu1a23ˉu200r3a32ˉu2).

    For convenience, we write DF(u(t,LE3)):=(A1B10r3a32ˉu2(t)), where

    B1=(0a23ˉu2(t))  andA1=(r12a11ˉu1(t)a12ˉu2(t)a12ˉu1(t)a21ˉu2(t)r22a22ˉu2(t)a21ˉu1(t)).

    Since DS(x)=V(φω,Lx)D(Lx), it follows that DS(E3) has a positive eigenvalue μ3 given by

    μ3=exp(φω0r3a32ˉu2(t)dtλ3(1φ)ω).

    Note that ˉu1(t) and ˉu2(t) satisfy the following equations

    {dˉu1(t)dt=ˉu1(t)(b1a11ˉu1(t)a12ˉu2(t)),dˉu2(t)dt=ˉu2(t)(b2a22ˉu2(t)a21ˉu1(t)). (3.4)

    Integrating the equation (3.4) for t from 0 to φω, one has

    φω0ˉu1(t)dt=h2a12h1a22a12a21a11a22andφω0ˉu2(t)dt=h1a21h2a11a12a21a11a22.

    Then,

    μ3=exp(h3(h1a21h2a11)a32a12a21a11a22).

    On the other hand, the other two eigenvalues μ1 and μ2 are determined by A1. Together with Lemma 2.4 in [11], the conclusion is immediate.

    Similarly, we also have the following Lemmas 3.6 and 3.7.

    Lemma 3.6. (Stability of the planar fixed point E2) There admits a unique interior fixed point, say E2:=(ˆx1,0,ˆx3), for S in the coordinate plane {x2=0}. Moreover,

    (i) If h2<h1a21a11+h3a23a33, then E2 is an asymptotically stable fixed point of S.

    (ii) If h2>h1a21a11+h3a23a33, then E2 is a saddle point with two-dimensional stable manifold.

    Lemma 3.7. (Stability of the planar fixed point E1) If h2a23>(<)h3a33,h3a32>(<)h2a22, then there admits a unique interior fixed point, say E1:=(0,ˇx2,ˇx3), for S in the coordinate plane {x1=0}. Moreover,

    (i) If h2a23>h3a33,h3a32>h2a22 and h1a12<a23h3a33h2a23a32a22a33, then E1 is an asymptotically stable fixed point of S.

    (ii) If h2a23<h3a33,h3a32<h2a22 and h1a12<a23h3a33h2a23a32a22a33, then E1 is a saddle point with two-dimensional stable manifold.

    (iii) If h2a23<h3a33,h3a32<h2a22 and h1a12>a23h3a33h2a23a32a22a33, then E1 is a saddle point with one-dimensional stable manifold.

    Assume that S has a positive fixed point, say P:=(˜x1,˜x2,˜x3), we will discuss the dynamics of P in the following. For simplicity, we write

    ˉA:=(a11a120a21a22a230a32a33).

    Lemma 3.8. (Stability of the positive fixed point P)

    (i) If detˉA<0, then P is an asymptotically stable fixed point of S.

    (ii) If detˉA>0, then P is an unstable fixed point of S.

    Proof. By the expression of L, it follows that

    LP=(eλ1(1φ)ω˜x1,eλ2(1φ)ω˜x2,eλ3(1φ)ω˜x3).

    Define Qt(LP):=(˜u1(t),˜u2(t),˜u3(t))=˜u(t), and V(t,x):=Dx˜u(t). Note that

    DF(u(t,LP))=(r12a11˜u1a12˜u2a12˜u10a21˜u2r22a22˜u2a21˜u1a23˜u3a23˜u20a32˜u3r32a33˜u3a32˜u2),

    the matrix function V(t)=V(t,LP) satisfies

    dV(t)dt=DF(˜u(t))V(t), V(0)=I.

    Let w(t)=P(t)V(t), where P(t)=diag{1˜u1(t),1˜u2(t),1˜u3(t)}. Then, one has

    dw(t)dt=A(t)w(t) (3.5)

    where

    A(t)=(a11˜u1(t)a12˜u2(t)0a21˜u1(t)a22˜u2(t)a23˜u3(t)0a32˜u2(t)a33˜u3(t)).

    Let W(t) be the monodromy matrix of the above Eq (3.5), then W(t) satisfies

    dW(t)dt=A(t)W(t),W(0)=I,

    and hence W(t)=P(t)V(t)P1(0). Thus V(t)=P1(t)W(t)P(0). In view of

    P(0)=diag{eλ1(1φ)ω˜x1,eλ2(1φ)ω˜x2,eλ3(1φ)ω˜x3}

    and

    P(φω)=diag{1˜u1(φω),1˜u2(φω),1˜u3(φω)}=diag{1˜x1,1˜x2,1˜x3},

    we have

    V(φω)=P1(φω)W(φω)P(0).

    Then,

    DS(P)=V(φω,LP)D(LP)=P1(φω)W(φω)P(0)D(LP)=diag{˜x1,˜x2,˜x3}W(φω)diag{1˜x1,1˜x2,1˜x3},

    and hence, DS(P)W(φω), that is, r(DS(P))=r(W(φω)).

    Let

    Z(t)=diag{1,1,1}W(t)diag{1,1,1}1,

    then

    dZ(t)dt=˜A(t)Z(t),Z(0)=I, (3.6)

    where

    ˜A(t)=(a11˜u1(t)a12˜u2(t)0a21˜u1(t)a22˜u2(t)a23˜u3(t)0a32˜u2(t)a33˜u3(t)).

    Note that Z(φω)W(φω) implies that DS(P)Z(φω), then,

    r(DS(P))=r(Z(φω)).

    Since the matrix ˜A(t) is cooperative and irreducible, it follows that Z(t)>Z(0) for each t>0, and then Z(t) is a positive matrix (see [25, Theorem B.3]). By Perron-Frobenius theorem, ρ3:=r(Z(φω)) is a simple eigenvalue of Z(φω) with a positive eigenvector e=(e1,e2,e3)T. If ρ1 and ρ2 represent the other two eigenvalues of Z(φω), then |ρi|<ρ3, i=1,2. By Liouville's formula, we also obtain that 0<ρ1ρ2ρ3=detZ(φω)=eφω0trace(˜A(s))ds<1.

    Let z(t):=Z(t)e=(z1(t),z2(t),z3(t)), then z(φω)=Z(φω)e=ρ3(e1,e2,e3) and z(0)=Z(0)e=(e1,e2,e3)>0. Noticing that z(t) satisfies the equation

    dz(t)dt=˜A(t)z(t),

    it follows that

    {˙z1(t)=a11˜u1(t)z1(t)+a12˜u2(t)z2(t),˙z2(t)=a21˜u1(t)z1(t)a22˜u2(t)z2(t)+a23˜u3(t)z3(t),˙z3(t)=a32˜u2(t)z2(t)a33˜u3(t)z3(t).

    Using the method of elimination, we have

    a21a33˙z1(t)+a11a33˙z2(t)+a23a11˙z3(t)=(a33(a12a21a11a22)+a11a23a32)˜u2(t)z2(t).

    Integrating the above equation for t from 0 to φω, we then obtain

    a21a33φω0dz1(t)+a11a33φω0dz2(t)+a23a11φω0dz3(t)
    =(a33(a12a21a11a22)+a11a23a32)φω0˜u2(t)z2(t)dt.

    Note that

    a33(a12a21a11a22)+a11a23a32=detˉA,

    it follows that

    (a21a33e1+a11a33e2+a11a23e3)×(ρ31)=detˉAφω0˜u2(t)z2(t)dt.

    Based on the fact that φω0˜u2(t)z2(t)dt>0, we have

    (i) If detˉA>0, then ρ3>1.

    (ii) If detˉA<0, then ρ3<1.

    This implies that the proof is completed.

    To obtain the existence and uniqueness of the positive fixed point for the Poincaré map S, we first give a lemma with respect to the index of fixed points. For the reader's convenience, we recall some known results on the fixed point index of a continuous map (see Amann [19] for a more detailed discussion).

    Let URn+ be open and S:URn+ be a continuous map such that Fix(S,U) is compact, where Fix(S,U) is defined by the set of all fixed points of S in the subset U. The fixed point index of S is denoted by

    i(S,U,Rn+):=deg(idS,U,0),

    where id is the identity map and deg(idS,U,0) is the Brouwer degree for idS. The fixed point index of S at an isolated fixed point θU is defined by

    i(S,θ):=i(S,Bδ(θ),Rn+)=deg(idS,Bδ(θ),0),

    where Bδ(θ):={xRn:xθ<δ} is an open ball in U such that Fix(S,Bδ(θ))={θ}. In particular, if S is differentiable at θ Fix(S,U) and 1 is not an eigenvalue of DS(θ), then

    i(S,θ)=(1)β,

    where β is the sum of the multiplicities of all the eigenvalues of DS(θ) which are greater than one. When S has only finitely many fixed points in U, one has

    i(S,U,Rn+)=θFix(S,U)i(S,θ).

    Lemma 4.1. Assume that each fixed point θ of the Poincaré map S deduced by system (1.2) is hyperbolic, then

    (i) i(S,R3+,R3+)=1;

    (ii) If DS(θ) has no eigenvalue whose modulus is larger than 1, then i(S,θ)=1;

    (iii) If θIntH+L, LΛ={1,2,3}, then DS(θ) has at least one eigenvalue with modulus larger than 1 i(S,θ)=0.

    Proof. (ⅰ) Since R3+ is nonempty closed convex set, it follows from Dugundjis theorem that i(S,R3+,R3+) is well-defined. Let x0R3+ be arbitrary and define a compact map

    g:[0,1]×R3+R3+  by  g(λ,x):=(1λ)x0+λS(x).

    Then g maps the product space into R3+ and has no fixed points on the boundary of R3+ (relative to R3+) because this boundary is empty. By the homotopy invariance and the normalization property, we have

    i(S,R3+,R3+)=i(x0,R3+,R3+)=1.

    (ⅱ), (ⅲ) See the proof of Lemma 4.2 for n=3 and U=R3+ in Liang and Jiang [20].

    Arguing as the proof of Theorem 2.3 in Niu et al. [12], it can easily be proved that the Poincaré map S also has a carrying simplex.

    Lemma 4.2. (The existence of the carrying simplex) Assume that hi>0,i=1,2,3, then the Poincaré map S admits a carrying simplex Σ which attracts every nontrivial orbit in R3+.

    Proof. See the proof of Theorem 2.3 for n=3 in Niu et al. [12].

    For each coordinate plane, we write Πi={xR3+:xi=0},i=1,2,3. We denote by S|Πi the restriction of S to Πi. To obtain the global stability of the positive fixed point, we also need the following lemma which is a special version of Theorem 2.4 in Balreira et al. [21] for maps defined on R3+.

    Lemma 4.3. ([21, Theorem 2.4]) Consider the map S=(x1G1(x),x2G2(x),x3G3(x)) defined on R3+ with Gi(x)0, i=1,2,3, which has a carrying simplex. Assume that

    (a) detDS(x)>0 for all xR3+;

    (b) DS(x)1>0 for all xIntR3+;

    (c) for each i=1,2,3, S|Πi has a unique interior fixed point E{i} that is globally asymptotically stable in IntΠi, but a saddle for S;

    (d) S has a unique positive fixed point pIntR3+.

    Then P is globally asymptotically stable in IntR3+ for S.

    Next, we will analyze the global dynamics of system (1.2). Applying Lemma 4.1 to Lemma 3.1-3.7, it is not difficult to calculate the index of each of axial and planar fixed points. If S has a positive fixed point, it follows from Lemma 3.8 that i(S,P)=1 under the condition detˉA<0. By using the index of each boundary fixed point and i(S,R3+,R3+)=1, we can choose appropriate parameter values to ascertain whether the positive fixed point exists. We can further obtain the global dynamics for system (1.2).

    Theorem 4.4. (Three species Coexistence) Suppose that system (1.2) satisfies

    (i) hi>0 (i=1,2,3), detˉA<0;

    (ii) h1a12>h2a22; h3a32>h2a22; h2>a21a11h1+a23a33h3;

    (iii) h1(a22a33a23a32)>a12(a33h2a23h3);

    (iv) h3(a11a22a12a21)>a32(a11h2a21h1);

    then there admits a unique positive fixed point, say P=(˜x1,˜x2,˜x3), for the Poincaré map S. Moreover, P is globally asymptotically stable in IntR3+ .

    Proof. By Lemma 2.1, S|H+i has a unique interior fixed point Ri that is globally asymptotically stable in IntH+i for each i=1,2,3 because hi>0. And yet, Ri is a saddle point for S due to the assumption (ⅱ) and Lemmas 3.2–3.4. Under the assumptions (ⅱ)–(ⅳ), by Lemmas 3.5–3.7 and Theorem 3.6 in [12], S|Πi has a unique interior fixed point E{i} for i=1,2,3 that is globally asymptotically stable in IntΠi, but a saddle point for S. Using Lemma 4.1(ⅲ), one has

    i(S,O)=i(S,Ri)=i(S,Ei)=0,  i=1,2,3.

    Since i(S,R3+,R3+)=1 (Lemma 4.1(ⅰ)), it is easy to see that there exists at least a positive fixed point for S, say P. Note that detˉA<0 and Lemma 3.8(ⅰ), so P is an asymptotically stable fixed point for S. Lemma 4.1(ⅱ) implies that i(S,P)=1. If there also exists another positive fixed point, say P, that is, PP, then P is also asymptotically stable, and hence i(S,P)=1, which contradicts the fact that i(S,R3+,R3+)=1. Consequently, the positive fixed point P is unique. By the expression (3.1) of DS(x) and the Eq (3.3), we can obtain that for any xR3+,

    detDS(x)=detV(φω,Lx)exp((λ1+λ2+λ3)(1φ)ω)=exp{φω0trace{DF(u(t,Lx))}dt}exp((λ1+λ2+λ3)(1φ)ω)>0.

    Besides, it follows from the proof of Theorem 2.3 in [12] that DS(x)1>0,   xIntR3+. So far, the conditions (a), (b), (c) and (d) of Lemma 4.3 are satisfied for S, which implies that P is globally asymptotically stable in IntR3+. We have completed the proof.

    Theorem 4.5. (Two species Coexistence) Suppose that system (1.2) satisfies

    (i) hi>0 (i=1,2,3), detˉA<0;

    (ii) h2a21<h1a11; h2a23<h3a33; h1a12>h2a22; h3a32>h2a22;h2<a21a11h1+a23a33h3;

    then there is no positive fixed point for the Poincaré map S. Moreover, the planar fixed point E2 is globally asymptotically stable in IntR3+.

    Proof. By Lemma 2.1 and Theorem 3.6 in [12], there exist three axial fixed points Ri (i=1,2,3) and one planar fixed point E2 under the assumptions (ⅰ) and (ⅱ). Moreover, Ri is a saddle point for S and E2 is an asymptotically stable fixed point of S. In view of Lemma 4.1(ⅱ) and (ⅲ), one has

    i(S,O)=i(S,Ri)=0, i=1,2,3  and  i(S,E2)=1.

    If there exists a positive fixed point for S, say P, it follows from detˉA<0 and Lemma 3.8(i) that P is an asymptotically stable fixed point of S. Lemma 4.1(ⅱ) implies that i(S,P)=1. This contradicts the fact that i(S,R3+,R3+)=1. So there is no positive fixed point for S. Lemma 4.2 states that the Poincaré map S has a carrying simplex Σ by which is homeomorphic to the probability simplex Δ2. We further get that Σ is a topological disk and S|Σ is an orientation preserving homeomorphism from Σ onto Σ (see [22]). By Corollary 2.1 in [23], every trajectory on Σ converges to some fixed point. In particular, R1 and R3 are saddle points, R2 is an unstable node point, while E2 is an asymptotically stable node point on Σ. Since S|Σ is an orientation preserving homeomorphism, every trajectory on Σ converges to E2. By using the property (P3) of the carrying simplex, one has

    limn+Sn(x)=E2, xIntR3+.

    Thus, E2 is globally asymptotically stable in IntR3+. The proof is completed.

    By suitable modification to the proof of Theorem 4.5, we can also obtain the following Theorems 4.6 and 4.7.

    Theorem 4.6. (Two species Coexistence) Suppose that system (1.2) satisfies

    (i) hi>0 (i=1,2,3), detˉA<0;

    (ii) h3a32<h2a22; h1a12>h2a22; h2>a21a11h1+a23a33h3;

    (iii) h3(a11a22a12a21)<a32(a11h2a21h1);

    then there is no positive fixed point for the Poincaré map S. Moreover, the planar fixed point E3 is globally asymptotically stable in IntR3+.

    Theorem 4.7. (Two species Extinction) Suppose that system (1.2) satisfies

    (i) hi>0 (i=1,2,3), detˉA<0;

    (ii) h1a12<h2a22; h3a32<h2a22; h2>a21a11h1+a23a33h3;

    then there is no positive fixed point for the Poincaré map S. Moreover, the axial fixed point R2 is globally asymptotically stable in IntR3+.

    Remark 4.1. By using our analytical approach in Theorems 4.5–4.7, we also take different sufficient conditions to obtain the global dynamics of system (1.2).

    Next, we will investigate the global extinction for system (1.2). For simplicity, we introduce the following notations

    ˜A=(a11a120a21a22a230a32a33),  ˜B=(h1h2h2)  and  y=(y1y2y2).

    Lemma 4.8. If S has a positive fixed point, say P, then y:=φω0Qt(LP)dt is a positive solution of the linear algebraic system ˜Ay=˜B. In other words, if the linear system ˜Ay=˜B has no positive solution, then S has no positive fixed point in IntR3+.

    Proof. We write the positive fixed point P:=(˜x1,˜x2,˜x3) as Lemma 3.8. Let Qt(LP)=u(t,LP):=(˜u1(t),˜u2(t),˜u3(t)), then (˜u1(t),˜u2(t),˜u3(t)) satisfies the following equations

    {˜u1(t)˜u1(t)=r1a11˜u1(t)a12˜u2(t),˜u2(t)˜u2(t)=r2a22˜u2(t)a21˜u1(t)a23˜u3(t),˜u3(t)˜u3(t)=r3a33˜u3(t)a32˜u2(t). (4.1)

    Integrating the above equations of (4.1) for t from 0 to φω, it follows that

    {a11φω0˜u1(t)dt+a12φω0˜u2(t)dt=(r1φλ1(1φ))ω,a21φω0˜u1(t)dt+a22φω0˜u2(t)dt+a23φω0˜u3(t)dt=(r2φλ2(1φ))ω,a32φω0˜u2(t)dt+a33φω0˜u3(t)dt=(r3φλ3(1φ))ω.

    Note that yi=φω0˜ui(t)dt and hi=(riφλi(1φ))ω,i=1,2,3, we have

    {a11y1+a12y2=h1,a21y1+a22y2+a23y3=h2,a32y2+a33y3=h3. (4.2)

    From the above equations, it is clear that ˜Ay=˜B with yi>0 (i=1,2,3). Therefore, the conclusion is immediate. The proof is completed.

    Theorem 4.9. (Global Extinction) Suppose that hi<0 (i=1,2,3). Then the trivial fixed point O is globally asymptotically stable in R3+.

    Proof. Since hi<0 (i=1,2,3), it follows that S has no interior fixed points on the coordinate axes of R3+ due to Lemma 2.1(i). By appealing to Lemma 2.4 in [11], there are also no interior fixed points on the coordinate plane of R3+. Lemma 4.8 implies that there is no positive fixed point in IntR3+ for S. Therefore, O is a unique fixed point for S in R3+. By the equations of system (1.4), one has

    dxi(t)dtxi(t)(ri(t)aii(t)xi(t)),  i=1,2,3.

    Since hi<0, by differential inequality theorem and Lemma 2.1(ⅰ), it follows that

    limt+xi(t,x0))=0, x0R3+,  i=1,2,3.

    Note that Lemma 3.1(ⅰ), O is an asymptotically stable fixed point of S, and hence, O is globally asymptotically stable in R3+. We have completed the proof.

    In this section, we provide some numerical examples to illustrate our analytic results.

    Example 1. (Three species Coexistence) Taking parameter values ω=10, ϕ=0.5, λ1=0.1, λ2=0.1, λ3=0.1, r1=0.3, r2=0.3, r3=0.3, a11=0.4, a12=0.2, a21=0.1, a22=0.5, a23=0.1, a32=0.2, a33=0.5, and initial values x0=(0.8,0.9,0.8), system (1.2) satisfies the conditions of Theorem 4.4. The numerical simulations for the solution of system (1.2) and the orbit of the Poincaré map S are shown in Figure 1, which imply that three species will be coexistence.

    Example 2. (Two species coexistence) Taking parameter values ω=10, ϕ=0.5, λ1=0.1, λ2=0.1, λ3=0.1, r1=0.3, r2=0.3, r3=0.3, a11=0.2, a12=0.2, a21=0.3, a22=1, a23=0.22, a32=0.1, a33=0.2, and initial values x0=(0.8,0.9,0.8), system (1.2) satisfies the conditions of Theorem 4.5. The numerical simulations for the solution of system (1.2) and the orbit of the Poincaré map S are shown in Figure 2, which imply that Species 1 and Species 3 will be coexistence, and Species 2 will go to extinction.

    Example 3. (Two species coexistence) Taking parameter values ω=10, ϕ=0.5, λ1=0.1, λ2=0.1, λ3=0.1, r1=0.3, r2=0.3, r3=0.3, a11=0.6, a12=0.1, a21=0.1, a22=0.2, a23=0.2, a32=0.3, a33=0.5, and initial values x0=(0.8,0.9,0.8), system (1.2) satisfies the conditions of Theorem 4.6. The numerical simulations for the solution of system (1.2) and the orbit of the Poincaré map S are shown in Figure 3, which imply that Species 1 and 2 will be coexistence, and Species 3 will go to extinction.

    Example 4. (Two species extinction) Taking parameter values ω=10, ϕ=0.5, λ1=0.1, λ2=0.1, λ3=0.1, r1=0.3, r2=0.3, r3=0.3, a11=0.6, a12=0.3, a21=0.1, a22=0.23, a23=0.2, a32=0.25, a33=0.5, and initial values x0=(0.8,0.9,0.8), system (1.2) satisfies the conditions of Theorem 4.7. The numerical simulations for the solution of system (1.2) and the orbit of the Poincaré map S are shown in Figure 4, which imply that Species 2 will win the competition and Species 1 and Species 3 will go to extinction.

    Example 5. (Global extinction) Taking parameter values ω=10, ϕ=0.5, λ1=0.3, λ2=0.3, λ3=0.3, r1=0.1, r2=0.1, r3=0.1, a11=0.6, a12=0.1, a21=0.1, a22=0.2, a23=0.2, a32=0.3, a33=0.5, and initial values x0=(0.8,0.9,0.8), system (1.2) satisfies the conditions of Theorem 4.9. The numerical simulations for the solution of system (1.2) and the orbit of the Poincaré map S are shown in Figure 5, which imply that three species will go to extinction.

    In this paper, we focus on a tridiagonal three-species competition model with time ω-periodic coefficients (called Seasonal Succession). By the stability analysis of equilibria, we estimate the Floquet multipliers of all nonnegative periodic solutions of system (1.2), and get the local dynamics of these periodic solutions. By using the Brouwer degree theory, we present an index result of the fixed points for the Poincaré map S. Based on this, we can verify the existence and uniqueness of the positive fixed point under appropriate conditions. Sufficient conditions of the global stability for coexistence and extinction of system (1.2) are provided via the local dynamics of all nonnegative fixed points. More precisely, three species will be existence under the assumptions of Theorem 4.4; Species 1 and 3 will be coexistence and Species 2 will go to extinction in the competition under the assumptions of Theorem 4.5; Species 1 and 2 will be coexistence and Species 3 will go to extinction in the competition under the assumptions of Theorem 4.6; Species 2 will win the competition and Species 1 and 3 will go to extinction under the assumptions of Theorem 4.7; three species will go to extinction under the assumptions of Theorem 4.9. From above analytic results, it is not difficult to see that the introduction of seasonal succession may lead to species' extinction.

    On the other hand, there is no explicit expression of the Poincaré map S for the time-periodic differential equations, even for the simplest form as system (1.2). This makes the researches on the dynamics the Poincaré map of the time-periodic Kolmogorov competitive systems become much more difficult and complicated. In future work, we will try to give a complete classification for the dynamics of system (1.2) and explore the influence of parameter values φ and λi,i=1,2,3 related to the seasonal succession on the dynamics of the system (1.2).

    We are very grateful to the anonymous referees for their valuable comments and suggestions, which led to an improvement of our original manuscript. This work was supported by the National Natural Science Foundation of China (No.11871231) and Natural Science Foundation Fund of Fujian Province (No.2022J01305).

    The authors declare there is no conflict of interest.



    [1] H. L. Smith, Monotone Dynamical Systems. An Introduction to the Theory of Competitive and Cooperative Systems, American Mathematical Society, 1995. https://doi.org/10.1090/surv/041
    [2] J. D. Murray, Mathematical Biology I: An Introduction. 3rd edition, Springer-Verlag, Berlin, 2002.
    [3] M. W. Hirsch, H. L. Smith, Monotone dynamical systems, in Handbook of Differential Equations: Ordinary Differential Equations, Elsevier, Amsterdam, 2005. https://doi.org/10.1016/S1874-5725(05)80006-9
    [4] X. Q. Zhao, Dynamical systems in population biology. 2nd edition, Springer, New York, 2017. https://doi.org/10.1007/978-3-319-56433-3
    [5] J. Smillie, Competitive and cooperative tridiagonal systems of differential equations, SIAM J. Math. Anal., 15 (1984), 530–534. https://doi.org/10.1137/0515040 doi: 10.1137/0515040
    [6] H. L. Smith, Periodic tridiagonal competitive and cooperative systems of differential equations, SIAM J. Math. Anal., 22(1991), 1102–1109. https://doi.org/10.1137/0522071 doi: 10.1137/0522071
    [7] M. Gyllenberg, Y. Wang, Periodic tridiagonal systems modeling competitive-cooperative ecological interactions, Discrete Contin. Dyn. Syst. B, 5 (2005), 289–298. https://doi.org/10.3934/dcdsb.2005.5.289 doi: 10.3934/dcdsb.2005.5.289
    [8] C. Fang, M. Gyllenberg, Y. Wang, Floquet bundles for tridiagonal competitive-cooperative systems and the dynamics of time-recurrent systems, SIAM J. Math. Anal., 45 (2013), 2477–2498. https://doi.org/10.1137/120878021 doi: 10.1137/120878021
    [9] C. A. Klausmeier, Successional state dynamics: A novel approach to modeling nonequilibrium foodweb dynamics, J. Theor. Biol., 262 (2010), 584–595. https://doi.org/10.1016/j.jtbi.2009.10.018 doi: 10.1016/j.jtbi.2009.10.018
    [10] C. F. Steiner, A. S. Schwaderer, V. Huber, C. A. Klausmeier, E. Litchman, Periodically forced food chain dynamics: model predictions and experimental validation, Ecology, 90 (2009), 3099–3107. https://doi.org/10.1890/08-2377.1 doi: 10.1890/08-2377.1
    [11] S. B. Hsu, X. Q. Zhao, A Lotka-Volterra competition model with seasonal succession, J. Math. Biol., 64 (2012), 109–130. https://doi.org/10.1007/s00285-011-0408-6 doi: 10.1007/s00285-011-0408-6
    [12] L. Niu, Y. Wang, X. Xie, Carrying simplex in the Lotka-Volterra competition model with seasonal succession with applications, Discrete Contin. Dyn. Syst. B, 26 (2021), 2161–2172. https://doi.org/10.3934/dcdsb.2021014 doi: 10.3934/dcdsb.2021014
    [13] X. Xie, L. Niu, Global stability in a three-species Lotka-Volterra cooperation model with seasonal succession, Math. Meth. Appl. Sci., 44 (2021), 14807–14822. https://doi.org/10.1002/mma.7744 doi: 10.1002/mma.7744
    [14] Y. Zhang, X. Q. Zhao, Bistable travelling waves for reaction and diffusion model with seasonal succession, Nonlinearity, 26 (2013), 691–709. https://doi.org/10.1088/0951-7715/26/3/691 doi: 10.1088/0951-7715/26/3/691
    [15] D. Xiao, Dynamics and bifurcations on a class of population model with seasonal constant-yield harvesting, Discrete Contin. Dyn. Syst. B, 21 (2016), 699–719. https://doi.org/10.3934/dcdsb.2016.21.699 doi: 10.3934/dcdsb.2016.21.699
    [16] P. G. Barrientos, J. A. Rodriguez, A. Ruiz-Herrera, Chaotic dynamics in the seasonally forced SIR epidemic model, J. Math. Biol., 75 (2017), 1655–1668. https://doi.org/10.1007/s00285-017-1130-9 doi: 10.1007/s00285-017-1130-9
    [17] M. Han, X. Hou, L. Sheng, C. Wang, Theory of rotated equations and applications to a population model, Discrete Contin. Dyn. Syst., 38 (2018), 2171–2185. https://doi.org/10.3934/dcds.2018089 doi: 10.3934/dcds.2018089
    [18] Y. Tang, D. Xiao, W. Zhang, D. Zhu, Dynamics of epidemic models with asymptomatic infection and seasonal succession, Math. Biosci. Eng., 14 (2017), 1407–1424. https://doi.org/10.3934/mbe.2017073 doi: 10.3934/mbe.2017073
    [19] H. Amann, Fixed point equations and nonlinear eigenvalue problems in Ordered Banach Spaces, SIAM Rev., 18 (1976), 620–709. https://doi.org/10.1137/1018114 doi: 10.1137/1018114
    [20] X. Liang, J. Jiang, On the finite-dimensional dynamical systems with limited competition, Trans. Am. Math. Soc., 354 (2002), 3535–3554. https://doi.org/10.1090/S0002-9947-02-03032-5 doi: 10.1090/S0002-9947-02-03032-5
    [21] E. C. Balreira, S. Elaydi, R. Luís, Global stability of higher dimensional monotone maps, J. Differ. Equation Appl., 23 (2017), 2037–2071. https://doi.org/10.1080/10236198.2017.1388375 doi: 10.1080/10236198.2017.1388375
    [22] A. Ruiz-Herrera, Exclusion and dominance in discrete population models via the carrying simplex, J. Differ. Equation Appl., 19 (2013), 96–113. https://doi.org/10.1080/10236198.2011.628663 doi: 10.1080/10236198.2011.628663
    [23] A. Ruiz-Herrera, Topological criteria of global attraction with applications in population dynamics, Nonlinearity, 25 (2012), 2823–2841. https://doi.org/10.1088/0951-7715/25/10/2823 doi: 10.1088/0951-7715/25/10/2823
    [24] H. L. Smith, Competing subcommunities of mutualists and a generalized Kamke theorem, SIAM J. Math. Anal., 46 (1986), 856–874. https://doi.org/10.1137/0146052 doi: 10.1137/0146052
    [25] H. L. Smith, P. Waltman, The Theory of the Chemostat, Cambridge University Press, 1995. https://doi.org/10.1017/CBO9780511530043
    [26] Y. Wang, J. Jiang, Uniqueness and attractivity of the carrying simplex for discrete-time competitive dynamical systems, J. Differ. Equations, 186 (2002), 611–632. https://doi.org/10.1016/S0022-0396(02)00025-6 doi: 10.1016/S0022-0396(02)00025-6
  • This article has been cited by:

    1. Meixiang Chen, Xizhuang Xie, Global dynamics of a three-species mutualist and competitor model with seasonal succession, 2025, 1023-6198, 1, 10.1080/10236198.2025.2466476
  • Reader Comments
  • © 2023 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(1515) PDF downloads(102) Cited by(1)

Figures and Tables

Figures(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog