Research article Special Issues

Mathematical modeling and analysis of harmful algal blooms in flowing habitats

  • In this paper, we survey recent developments of mathematical modeling and analysis of the dynamics of harmful algae in riverine reservoirs. To make the models more realistic, a hydraulic storage zone is incorporated into a flow reactor model and new mathematical challenges arise from the loss of compactness of the solution maps. The key point in the study of the evolution dynamics is to prove the existence of global attractors for the model systems and the principal eigenvalues for the associated linearized systems without compactness.

    Citation: Sze-Bi Hsu, Feng-Bin Wang, Xiao-Qiang Zhao. Mathematical modeling and analysis of harmful algal blooms in flowing habitats[J]. Mathematical Biosciences and Engineering, 2019, 16(6): 6728-6752. doi: 10.3934/mbe.2019336

    Related Papers:

    [1] Wenjie Yang, Qianqian Zheng, Jianwei Shen, Linan Guan . Bifurcation and pattern dynamics in the nutrient-plankton network. Mathematical Biosciences and Engineering, 2023, 20(12): 21337-21358. doi: 10.3934/mbe.2023944
    [2] Xuehui Ji, Sanling Yuan, Tonghua Zhang, Huaiping Zhu . Stochastic modeling of algal bloom dynamics with delayed nutrient recycling. Mathematical Biosciences and Engineering, 2019, 16(1): 1-24. doi: 10.3934/mbe.2019001
    [3] Abdelheq Mezouaghi, Salih Djillali, Anwar Zeb, Kottakkaran Sooppy Nisar . Global proprieties of a delayed epidemic model with partial susceptible protection. Mathematical Biosciences and Engineering, 2022, 19(1): 209-224. doi: 10.3934/mbe.2022011
    [4] Lizhong Qiang, Ren-Hu Wang, Ruofan An, Zhi-Cheng Wang . A stage-structured SEIR model with time-dependent delays in an almost periodic environment. Mathematical Biosciences and Engineering, 2020, 17(6): 7732-7750. doi: 10.3934/mbe.2020393
    [5] Jean-Jacques Kengwoung-Keumo . Competition between a nonallelopathic phytoplankton and an allelopathic phytoplankton species under predation. Mathematical Biosciences and Engineering, 2016, 13(4): 787-812. doi: 10.3934/mbe.2016018
    [6] Da Song, Meng Fan, Ming Chen, Hao Wang . Dynamics of a periodic stoichiometric model with application in predicting and controlling algal bloom in Bohai Sea off China. Mathematical Biosciences and Engineering, 2019, 16(1): 119-138. doi: 10.3934/mbe.2019006
    [7] Kazuo Yamazaki, Xueying Wang . Global stability and uniform persistence of the reaction-convection-diffusion cholera epidemic model. Mathematical Biosciences and Engineering, 2017, 14(2): 559-579. doi: 10.3934/mbe.2017033
    [8] Lin Zhao, Zhi-Cheng Wang, Liang Zhang . Threshold dynamics of a time periodic and two–group epidemic model with distributed delay. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1535-1563. doi: 10.3934/mbe.2017080
    [9] Yun Kang, Sourav Kumar Sasmal, Amiya Ranjan Bhowmick, Joydev Chattopadhyay . Dynamics of a predator-prey system with prey subject to Allee effects and disease. Mathematical Biosciences and Engineering, 2014, 11(4): 877-918. doi: 10.3934/mbe.2014.11.877
    [10] Shengqiang Liu, Lin Wang . Global stability of an HIV-1 model with distributed intracellular delays and a combination therapy. Mathematical Biosciences and Engineering, 2010, 7(3): 675-685. doi: 10.3934/mbe.2010.7.675
  • In this paper, we survey recent developments of mathematical modeling and analysis of the dynamics of harmful algae in riverine reservoirs. To make the models more realistic, a hydraulic storage zone is incorporated into a flow reactor model and new mathematical challenges arise from the loss of compactness of the solution maps. The key point in the study of the evolution dynamics is to prove the existence of global attractors for the model systems and the principal eigenvalues for the associated linearized systems without compactness.


    In the past decades, harmful algal blooms (HAB) have become important water quality issues in both coastal and inland waters, and the frequency and intensity of HAB are apparently increasing worldwide. Blooms of the haptophyte algae Prymnesium parvum have become more and more common in the world, and it is referred to as golden algae [1,2], which were documented to cause large fish kills [3,4]. Recent studies suggest possible potential techniques for managing and mitigating harmful algal blooms through flow manipulations in some river systems [5,6,7,8]. This motivates the theoretical exploration of harmful algal dynamics in riverine reservoirs. To understand longitudinal patterns arising along the axis of flow, advection-dispersion-reaction systems were employed to study the effects of spatial variations of harmful algae and its toxin production and decay in riverine reservoirs [9,10,11,12,13]. The models are one-dimensional systems with simple habitat geometry and transport processes. The flow reactor model with transport of nutrient and organisms by both advection and diffusion was first proposed in [14]. Recently, the flow reactor model in [14] has been further incorporated with a hydraulic storage zone for persistence and coexistence of competing populations [9]. Such systems with/without a hydraulic storage zone become more and more attractive since they can be regarded as a surrogate model for riverine reservoirs with strong advective flows [10].

    It should be pointed out that the flow reactor system presented in [14] was used to model the influences of bacterial motility, fluid advection and other spatial variations on the competition between different strains of bacteria for the limiting nutrient in the large intestine of animals. Differently, our main purpose here is to use the flow reactor system to describe the dynamics/interactions of harmful algae and nutrient(s) in the river/stream. In [15], the author extended the model in [14] by considering two species competition for two essential/complementary nutrients, such as nitrogen and phosphorus. The complementary nutrient model is highly relevant since the limiting nutrient(s) in many ecosystems should be multiple, and hence, the single-nutrient model can be seen as a special case. Since the environment of the plug-flow reactor is the intestine or a river, it is much more realistic to assume that the input nutrient concentration is time-dependent. Thus, the periodically varying input nutrient concentration is incorporated into the model of [14] and the model of [15] in [16] and [17], respectively.

    There is a persistence paradox in the river ecology, namely, rapid advective flow in such habitats can prevent persistence of one species for realistic parameters. This motivates us to incorporate the factor of hydraulic storage zones in flowing water habitats [9,11] since it might resolve this persistence paradox [18]. Introducing a hydraulic storage zone into the flow reactor model not only makes sense biologically but also causes mathematical challenges. Some equations in the flow reactor with a hydraulic storage zone have no diffusion terms, and hence, the associated solution maps are not compact. In order to obtain the existence of global attractor, we show that solution maps are asymptotically smooth by using the Kuratowski measure of noncompactness. Note that the existence of global attractor is assumed in the theory of uniform persistence and coexistence states (see, e.g., [19]).

    Another problem is about the local stability of the trivial and semi-trivial solutions of the model, which are usually determined by the sign of the principal eigenvalue(s) of the associated linearized system at these states. Although the associated linearized system is cooperative, the "compactness" is required when one uses the classical Krein–Rutman theory to obtain the existence of the principal eigenvalue. Very recently, Wang and Zhao developed some sharp abstract results (see [20,Theorem 2.3] and [20,Remark 2.2]) on the existence of principal eigenvalues for an elliptic eigenvalue problem with some zero diffusion coefficients. Two closely related applications can be found in [12,Lemma 3.3] and [21,Theorem 2.1]. The authors in [22] further studied the existence of the principal eigenvalue for degenerate periodic reaction-diffusion systems (see [22,Theorem 2.16 and Remark 2.21]). An alternative approach is to directly utilize the generalized Krein–Rutman Theorem developed by Nussbaum [23], see, e.g., [12,Lemma 4.4].

    The rest of the paper is organized as follows. In section 2, a flow reactor model proposed by Kung and Baltzis [14] and its extensions are briefly reviewed. Section 3 is devoted to the survey of a model of a flowing water habitat with a hydraulic storage zone in which no diffusive and advective flow occurs. The input nutrient concentration can be a constant [9] or time-dependent [11]. In section 4, we review a model of interactions of a single limiting nutrient, harmful algae, toxins, and zooplankton [10,12]. Coexistence of harmful algae and zooplankton was also investigated in [12]. We further discuss a model of algal toxins and nutrient recycling (see [10] and [13]), which is based on the fact that many cyanotoxins produced by cyanobacteria can get recycled back into the system as a nutrient resource after they decompose. Finally, a brief discussion section completes the paper.

    We first review a model of microbial competition for a single limited nutrient in a riverine reservoir occupying a simple channel of longitudinally invariant cross-section, which was formulated by Kung and Baltzis in [14] and analyzed by the authors in [24,25]. The channel is assumed to have constant cross-sectional area A and length L, yielding volume V. A flow of water enters at the upstream end (x=0), with discharge F (dimensions length3 / time). An equal flow exits at the downstream end (x=L), which is assumed to be a dam. Based on this flow, a dilution rate D (dimensions time1) is defined as F/V. The advective flow within the channel is set to maintain water balance, by transporting water with a net velocity ν=DL.

    The reactor occupies the portion of the channel from x=0 to x=L in which the microbial populations Ni, i=1,2, compete for nutrient R. A flow of medium in the channel with velocity ν in the direction of increasing x brings fresh nutrient at a constant concentration R(0) into the reactor at x=0 and carries medium, unutilized nutrient and organisms out of the reactor at x=L. Nutrient and organisms are assumed to diffuse throughout the vessel with the same diffusivity δ.

    Both advective and diffusive transport occur at the upstream boundary (x=0). The inflow contains dissolved nutrient R(x,t) at a concentration R(0). The downstream boundary is assumed to be a dam, over which there is advective flow but through which no diffusion can take place. These assumptions lead to boundary conditions for nutrient:

    νR(0,t)δRx(0,t)=νR(0), Rx(L,t)=0. (2.1)

    Boundary conditions for the nutrient are given by equations (2.1), while those for population densities are

    νNi(0,t)δNix(0,t)=Nix(L,t)=0,i=1,2.

    These boundary conditions mean that no inflow of the populations occurs, and there is no dispersive transport over the dam at the downstream end. Under these assumptions, we have the following reaction-diffusion system describing the densities R(x,t), N1(x,t) and N2(x,t):

    {Rt=δ2Rx2νRxq1f1(R)N1q2f2(R)N2,x(0,L), t>0,N1t=δ2N1x2νN1x+f1(R)N1,x(0,L), t>0,N2t=δ2N2x2νN2x+f2(R)N2,x(0,L), t>0, (2.2)

    with boundary conditions

    {νR(0,t)δRx(0,t)=νR(0), Rx(L,t)=0,νNi(0,t)δNix(0,t)=Nix(L,t)=0, i=1,2, (2.3)

    and initial conditions

    R(x,0)=R0(x)0, Ni(x,0)=N0i(x)0,0<x<L,i=1,2, (2.4)

    where qi is the constant nutrient quota for species i. The nonlinear functions fi(R) describes the nutrient uptake rate and the growth rate of the organisms Ni at nutrient concentration R. We assume that these functions satisfy

    fi(0)=0, fi(R)>0 R>0, fiC2,i=1,2.

    The usual example is the Monod function

    fi(R)=μmax, iRKi+R,

    where μmax, i (resp. Ki) represents the maximal growth rate (resp. the half saturation constant) of species i. In [25,Chapter 8], the author showed that both species in (2.2)-(2.4) can coexist under suitable conditions. The condition (2.3) is called as the Danckwerts' boundary condition by Aris [26]. For a detailed derivation of it, we refer to a review paper [27].

    The authors in [24] extended (2.2)-(2.4) to the following system

    {Rt=δ02Rx2νRxq1f1(R)N1q2f2(R)N2,x(0,L), t>0,N1t=δ12N1x2νN1x+[f1(R)m1]N1,x(0,L), t>0,N2t=δ22N2x2νN2x+[f2(R)m2]N2,x(0,L), t>0, (2.5)

    with boundary conditions (2.3), and initial conditions (2.4). Here δ0 and δi stand for the random motility coefficients of nutrient and species i, respectively; mi is the death rate of species i. The effects of random motility on the extinction/persistence of a single population model, and the influences of random motility on the competition outcomes between two species were investigated in [24]. Note that if we assume δ0=δ1=δ2=δ and m1=m2=0 in (2.5), then it becomes (2.2). The authors in [16] further incorporated a periodically varying input nutrient concentration into system (2.5) with boundary conditions (2.3) and initial conditions (2.4), where the input concentration R(0) is replaced by a τ-periodic function R(0)(t). Then they used the theories of monotone dynamical systems and uniform persistence to obtain some analytic results about the extinction/persistence of a single population model and coexistence of two species system in terms of the principal eigenvalue(s) of the associated periodic-parabolic eigenvalue problem(s).

    To address the multiple nutrients in ecosystems, the author in [15] also generalized model (2.2)-(2.4) to an evolution system of two species competition for two essential nutrients with constant input concentrations. Later on, a model of two species competition for two essential nutrients with periodically varying input concentrations was studied in [17].

    This section is devoted to the survey of models with hydraulic storage zones, which partition the cross-section of the channel into a flowing zone of area A, and a static zone of area AS. Exchange of nutrient and populations between the flowing and storage zones occurs by Fickian diffusion with rate α. Nutrient concentration and population densities vary with location in both the flowing channel and the storage zone, however, we assume that advective and diffusive transport occur only in the flowing zone, not the storage zone. Nutrient concentration and population densities in the flowing channel (resp. the storage zone) are denoted by R(x,t) and Ni(x,t) (resp. RS(x,t), NS,i(x,t)). Then we generalize system (2.2)-(2.4), by adding a hydraulic storage zone and including the seasonality of R(0), to the following form [9,11]:

    {Rt=δ2Rx2νRxq1f1(R)N1q2f2(R)N2+α(RSR),N1t=δ2N1x2νN1x+α(NS,1N1)+f1(R)N1,N2t=δ2N2x2νN2x+α(NS,2N2)+f2(R)N2,RSt=αAAS(RSR)q1f1(RS)NS,1q2f2(RS)NS,2,NS,1t=αAAS(NS,1N1)+f1(RS)NS,1,NS,2t=αAAS(NS,2N2)+f2(RS)NS,2, 0<x<L, t>0 (3.1)

    with boundary conditions

    {νR(0,t)δRx(0,t)=νR(0)(t),νNi(0,t)δNix(0,t)=0,Rx(L,t)=Nix(L,t)=0, t>0, i=1,2, (3.2)

    and initial conditions

    {R(x,0)=R0(x)0, Ni(x,0)=N0i(x)0, 0<x<L,RS(x,0)=R0S(x)0, NS,i(x,0)=N0S,i(x)0, i=1,2. (3.3)

    Here R(0)(t) satisfies

    {R(0)()C2(R+,R),R(0)(t)0 but R(0)()0 on R+:=[0,),R(0)(t+τ)=R(0)(t), for some real number τ>0. (3.4)

    We should point out that the authors in [9] considered system (3.1)-(3.3) under the assumption that R(0)(t)R(0) is a positive constant.

    By [33,Theorem 1 and Remark 1.1], we have the following results.

    Lemma 3.1. ([11,LEMMA 2.1]) System (3.1)-(3.3) has a unique noncontinuable solution and solutions of (3.1)-(3.3) remain non-negative on their interval of existence if they are non-negative initially.

    In the following, we demonstrate that system (3.1)-(3.3) have a mass conservation in the flow and storage zones. Let

    W(x,t)=R(x,t)+q1N1(x,t)+q2N2(x,t) and WS(x,t)=RS(x,t)+q1NS,1(x,t)+q2NS,2(x,t). (3.5)

    Then W(x,t) and WS(x,t) satisfy the following system

    {Wt=δ2Wx2νWx+αWSαW, 0<x<L, t>0,WSt=αAASWS+αAASW, 0<x<L, t>0,νW(0,t)δWx(0,t)=νR(0)(t), Wx(L,t)=0, t>0,W(x,0)=W0(x), WS(x,0)=WS0(x), 0<x<L. (3.6)

    The following result is concerned with the global dynamics of system (3.6).

    Lemma 3.2. ([11,LEMMA 2.3]) System (3.6) admits a unique positive τ-periodic solution (W(x,t),WS(x,t)) and for any (W0,WS0)C([0,L],R2), the unique mild solution (W(x,t),WS(x,t)) of (3.6) with (W(x,0),WS(x,0))=(W0(x),WS0(x)) satisfies limt((W(x,t),WS(x,t))(W(x,t),WS(x,t)))=(0,0) uniformly for x[0,L].

    We should point out that [11,Lemma 2.2] is based on the assumption that the associated eigenvalue problem admits a principal eigenvalue, and there is a gap in the arguments for the existence of the principal eigenvalue in the paragraph above [11,Lemma 2.2]. However, this gap can be easily filled by using [21,Theorem 2.1] or the arguments similar to those in [12,Lemma 3.3] combined with [20,Theorem 2.3 and Remark 2.2].

    By Lemma 3.1, the relation (3.5) and Lemma 3.2, we have the following result.

    Lemma 3.3. ([11,LEMMA 2.4]) Any solution of the system (3.1)-(3.3) exists globally on [0,). Moreover, solutions are ultimately bounded and uniformly bounded.

    This subsection is devoted to the investigation of the single population model. Mathematically, it means that we set (N1,NS,1)=(0,0) or (N2,NS,2)=(0,0) in the model system (3.1)-(3.3). In order to simplify notation, we drop all subscripts in the remaining equations and then consider

    {Rt=δ2Rx2νRxqf(R)N+α(RSR), 0<x<L, t>0,Nt=δ2Nx2νNx+α(NSN)+f(R)N, 0<x<L, t>0,RSt=αAAS(RSR)qf(RS)NS, 0<x<L, t>0,NSt=αAAS(NSN)+f(RS)NS, 0<x<L, t>0, (3.7)

    with boundary conditions

    {νR(0,t)δRx(0,t)=νR(0)(t), Rx(L,t)=0, t>0,νN(0,t)δNx(0,t)=Nx(L,t)=0, t>0, (3.8)

    and initial conditions

    {R(x,0)=R0(x)0, N(x,0)=N0(x)0, 0<x<L,RS(x,0)=R0S(x)0, NS(x,0)=N0S(x)0, 0<x<L, (3.9)

    where R(0)(t) satisfies (3.4).

    Let

    W(x,t)=R(x,t)+qN(x,t) and WS(x,t)=RS(x,t)+qNS(x,t).

    Then W(x,t) and WS(x,t) satisfy (3.6). By Lemma 3.2, we see that the limiting system of (3.7)-(3.9) takes the following form:

    {Nt=δ2Nx2νNx+α(NSN)+f(W(x,t)qN)N, 0<x<L, t>0,NSt=αAAS(NSN)+f(WS(x,t)qNS)NS, 0<x<L, t>0, (3.10)

    with boundary conditions

    νN(0,t)δNx(0,t)=0, Nx(L,t)=0, t>0, (3.11)

    and initial conditions

    N(x,0)=N0(x)0, NS(x,0)=N0S(x)0,0<x<L. (3.12)

    From the biological view of point, the feasible domain Λ(t) for (3.10)-(3.12) should be

    Λ(t)={(N,NS)C([0,L],R2+):qN()W(,t), qNS()WS(,t)}.

    We further have the following basic properties of the set Λ(t).

    Lemma 3.4. ([11,LEMMA 3.1]) For any ϕ:=(ϕ1,ϕ2)Λ(0), system (3.10)-(3.12) has a unique mild solution (N(,t),NS(,t)) with (N(,0),NS(,0))=ϕ and (N(,t),NS(,t))Λ(t), for all t0.

    By Lemma 3.4, we can define solution maps Ψt:Λ(0)Λ(t) associated with (3.10)-(3.12) by

    Ψt(P)=(N(,t,P),NS(,t,P)), P:=(N0(), N0S())Λ(0), t0.

    Note that Ψτ:Λ(0)Λ(τ)=Λ(0) is the Poincaré map associated with (3.10)-(3.12).

    For convenience, we let

    Y+=Λ(0), Y0=Y+{(0,0)}, Y0:=Y+Y0={(0,0)}.

    Since one equation in (3.10)-(3.12) has no diffusion term, its solution map Ψt is not compact. Due to the lack of compactness, we need to impose the following condition:

    αAAS>f(WS(x,t)),  x[0,L], t0. (3.13)

    Recall that the Kuratowski measure of noncompactness (see, e.g., [28]), κ, is defined by

    κ(B):=inf{r:B has a finite cover of diameter<r}, (3.14)

    for any bounded set B. We set κ(B)= whenever B is unbounded. Note that B is precompact(i.e., ˉB is compact) if and only if κ(B)=0.

    Lemma 3.5. ([11,LEMMA 3.2]) Let (3.13) hold. Then Ψτ is κ-contracting in the sense that limnκ(ΨnτB)=0 for any bounded set BY+.

    By Lemma 3.3, Lemma 3.5 and [29,Theorem 2.6], we have the following result.

    Theorem 3.1. ([11,THEOREM 3.1]) Ψτ admits a global attractor on Y+ that attracts each bounded set in Y+ provided that (3.13) holds.

    Note that (0,0) is the trivial solution of (3.10)-(3.12). Linearizing system (3.10)-(3.12) at (0,0), we have

    {Nt=δ2Nx2νNx+α(NSN)+f(W(x,t))N,NSt=αAAS(NSN)+f(WS(x,t))NS, 0<x<L, t>0,νN(0,t)δNx(0,t)=0, Nx(L,t)=0, t>0. (3.15)

    Substituting N(x,t)=eμtϕ1(x,t) and NS(x,t)=eμtϕ2(x,t), we obtain the associated eigenvalue problem

    {ϕ1t=δ2ϕ1x2νϕ1x+α(ϕ2ϕ1)+f(W(x,t))ϕ1+μϕ1, t>0, x(0,L),ϕ2t=αAAs(ϕ2ϕ1)+f(WS(x,t))ϕ2+μϕ2, t>0, x(0,L),νϕ1(0,t)δϕ1x(0,t)=ϕ1x(L,t)=0, t>0,ϕ1,ϕ2 are τ-periodic in t. (3.16)

    As in the proof of [11,Lemma 3.3], we let Πt:C([0,L],R2)C([0,L],R2) be the solution maps associated with (3.15). Then P:=Πτ is the Poincaré map associated with system (3.15). Let r(P) be the spectral radius of P. By the proof of [11,Lemma 3.3], we further see that Πτ is an κ-contraction on C([0,L],R2) in the sense that

    κ(ΠτB)er0τκ(B) (3.17)

    for any bounded set B in C([0,L],R2), where r0 is a positive number such that

    αAASf(WS(x,t))r0,  x[0,L], t0.

    By (3.17) and the arguments similar to those in [12,Lemma 4.4]) or [30,Lemma 3.1], we can use the generalized Krein-Rutman Theorem [23] to obtain the following result, which is a corrected version of [11,Lemma 3.3].

    Lemma 3.6. Define μ:=1τlnr(P) and let (3.13) hold. If r(P)1, then μ is the principal eigenvalue of the eigenvalue problem (3.16) with a strongly positive eigenfunction ϕ=(ϕ1,ϕ2)0.

    The following result is concerned with the global dynamics of system (3.10)-(3.12).

    Theorem 3.2. ([11,THEOREM 3.2]) Assume that (3.13) holds. Let (N(x,t),NS(x,t)) be the solution of (3.10)-(3.12) with initial data (N0(), N0S())Y+. Then the following statements are valid:

    (1)If μ>0, then limt|(N(x,t),NS(x,t))|=0 uniformly for x[0,L];

    (2)If μ<0, then (3.10)-(3.12) admit a unique positive τ-periodic solution

    (N(x,t),NS(x,t)) and for any (N0(),N0S())Y0, we have

    limt|(N(x,t),NS(x,t))(N(x,t),NS(x,t))|=0uniformly for x[0,L].

    By Theorem 3.2 and the theories of chain transitive sets (see, e.g., [19,31]), one can obtain a threshold type result on the global dynamics of the single population model (3.7)-(3.9).

    This subsection focuses on the investigation of the possibility of coexistence for system (3.1)-(3.3). In view of the relation (3.5) and Lemma 3.2, we see that the limiting systems of (3.1)-(3.3) take the forms:

    {N1t=δ2N1x2νN1x+α(NS,1N1)+f1(W(x,t)q1N1q2N2)N1,NS,1t=αAAS(NS,1N1)+f1(WS(x,t)q1NS,1q2NS,2)NS,1,N2t=δ2N2x2νN2x+α(NS,2N2)+f2(W(x,t)q1N1q2N2)N2,NS,2t=αAAS(NS,2N2)+f2(WS(x,t)q1NS,1q2NS,2)NS,2, (3.18)

    in (0,L)×(0,), with boundary conditions

    νNi(0,t)δNix(0,t)=0, Nix(L,t)=0, t>0, i=1,2, (3.19)

    and initial conditions

    Ni(x,0)=N0i(x)0, NS,i(x,0)=N0S,i(x)0,0<x<L, i=1,2. (3.20)

    From the biological view of point, the feasible domain D(t) for (3.18)-(3.20) should be

    D(t)={(N1,NS,1,N2,NS,2)C([0,L],R4+):q1N1()+q2N2()W(,t), q1NS,1()+q2NS,2()WS(,t)}.

    The following result indicates that D(t) is positively invariant for the solution maps associated with (3.18)-(3.20).

    Lemma 3.7. ([11,LEMMA 2.5]) For any ϕ:=(ϕ1,ϕ2,ϕ3,ϕ4)D(0), system (3.18)-(3.20) has a unique mild solution (N1(,t),NS,1(,t),N2(,t),NS,2(,t))D(t), for all t0, whenever (N1(,0),NS,1(,0),N2(,0),NS,2(,0))=ϕ.

    Since two equations in (3.18)-(3.20) have no diffusion terms, its solution maps are not compact. So we require the following conditions in this subsection:

    αAAS>fi(WS(x,t)), x[0,L], t0, i=1,2. (3.21)

    Fix i{1,2}, we consider the following linear system

    {Nt=δ2Nx2νNx+α(NSN)+fi(W(x,t))N,NSt=αAAS(NSN)+fi(WS(x,t))NS, 0<x<L, t>0,νN(0,t)δNx(0,t)=0, Nx(L,t)=0, t>0. (3.22)

    Then the associated eigenvalue problem takes the form

    {φt=δ2φx2νφx+α(ψφ)+fi(W(x,t))φ+μφ, t>0, x(0,L),ψt=αAAs(ψφ)+fi(WS(x,t))ψ+μψ, t>0, x(0,L),νφ(0,t)δφx(0,t)=φx(L,t)=0, t>0,φ,ψ are τ-periodic in t. (3.23)

    Let Pi, i=1,2, be the Poincaré map associated with system (3.22), and r(Pi) be the spectral radius of Pi. By the same arguments as in Lemma 3.6, we have the following result.

    Lemma 3.8. Define μi:=1τlnr(Pi) and let (3.21) hold. If r(Pi)1, then μi is the principal eigenvalue of the eigenvalue problem (3.23) with a strongly positive eigenfunction.

    Note that Then system (3.18)-(3.20) admits the following possible trivial/semi-trivial solutions:

    (i) Trivial solution ˆ0:=(0,0,0,0) always exists;

    (ii) Semi-trivial solution (N1(x,t),NS,1(x,t),0,0) exists provided that μ1<0;

    (iii) Semi-trivial solution (0,0,N2(x,t),NS,2(x,t)) exists provided that μ2<0;

    (iv) There may be additional τ-periodic solutions as well and these must be positive.

    Here (Ni(x,t),NS,i(x,t)) denotes the unique positive τ-periodic solution of (3.10)-(3.12) resulting from putting f=fi and q=qi. The two organisms can coexist if a positive τ-periodic solution exists.

    In view of Lemma 3.7, we let Φt:D(0)D(t) be the solution map of system (3.18)-(3.20). Let K=C([0,L],R2+)×(C([0,L],R2+)) and denote its induced order by K. Thus, the solution map Φt is monotone [25] with respect to the partial order K. Note that Φτ:D(0)D(τ)=D(0) and for the Poincaré map S:=Φτ, we have Sn(P)=Φnτ(P), for all nZ. Set Y+=D(0),

    Y0:={(N1,NS,1,N2,NS,2)Y+:(N1,NS,1)(0,0) and (N2,NS,2)(0,0)}

    and Y0:=Y+Y0. For convenience, we further set

    gi(t,x,u1,u2,v1,v2)=αAAS(viui)+fi(WS(x,t)q1v1q2v2)vi,i=1,2,

    and

    D={(t,x,u,v)R6+:x[0,L],q1u1+q2u2W(x,t),q1v1+q2v2WS(x,t)},

    where u:=(u1,u2)R2+ and v:=(v1,v2)R2+. With the assumption (3.21), it follows whenever αAAS is sufficiently large, there exists a constant r>0 such that

    zT[g(t,x,u,v)v]zrzTz,zR2,(t,x,u,v)D, (3.24)

    where g(t,x,u,v):=(g1(t,x,u1,u2,v1,v2),g2(t,x,u1,u2,v1,v2)).

    Lemma 3.9. ([11,LEMMA 4.1]) Let (3.21) and (3.24) hold. Then the map Φτ is κ-contracting in the sense that limnκ(Φnτ(B))=0 for any bounded set BY+, where κ is the Kuratowski measure of noncompactness as defined in (3.14).

    By Lemma 3.3, Lemma 3.9, and [29,Theorem 2.6], we have the following result.

    Theorem 3.3. ([11,THEOREM 4.1]) Let (3.21) and (3.24) hold. Then Φτ admits a global attractor on Y+ that attracts each bounded set in Y+.

    Fix i{1,2}, let ˆPi be the Poincaré map associated with system (3.22) when (fi(W(x,t)),fi(WS(x,t))) in (3.22) is replaced by

    (f3i(W(x,t)qiNi(x,t)),f3i(WS(x,t)qiNS,i(x,t))). (3.25)

    Let r(ˆPi) be the spectral radius of ˆPi. By the same arguments as in Lemma 3.6, we have the following observation.

    Lemma 3.10. Define ηi:=1τlnr(ˆPi) and let (3.21) hold. If r(ˆPi)1, then ηi is the principal eigenvalue of (3.23) with (fi(W(x,t)),fi(WS(x,t))) replaced by the one in (3.25).

    The following result is concerned with the coexistence of system (3.18)-(3.20),

    Theorem 3.4. Let (3.21) and (3.24) hold, and assume that μi<0 and ηi<0, i=1,2. Then system (3.18)-(3.20) admits at least one (componentwise) positive τ-periodic solution and there exists a positive constant ζ>0 such that for any solution (N1(x,t),NS,1(x,t),N2(x,t),NS,2(x,t)) of system (3.18)-(3.20) with the initial data in Y0 satisfies lim inftminx[0,L]Ni(x,t)ζ and lim inftminx[0,L]NS,i(x,t)ζ, for all i=1,2.

    We remark that Theorem 3.4 follows from [11,Theorem 4.2], where the theory of monotone dynamical systems have been used. Instead, we can also obtain Theorem 3.4 by using the theory of uniform persistence. In [11,Section 5], the authors further lifted the dynamics of the limiting system (3.18)-(3.20) to the full system (3.1)-(3.3) by the theory of chain transitive sets (see, e.g., [19,31]).

    In this section, we survey systems modeling the interactions of nutrient, harmful algae, toxins, and zooplankton, in which the input concentration R(0) is always a constant.

    This subsection is devoted to the study of the influences of spatial variations on the growth of harmful algae and the production/decay of their toxins in riverine reservoirs. Suppose R(x,t), N(x,t) and C(x,t) (resp. RS(x,t), NS(x,t) and CS(x,t)) denote dissolved nutrient concentration, algal abundance and dissolved toxin concentration at location x and time t in the flowing channel, respectively (resp. in the storage zone). The authors in [10] propose the following advection-dispersion-reaction system:

    {Rt=δ2Rx2νRxqN[f(R)m]N+α(RSR),Nt=δ2Nx2νNx+α(NSN)+[f(R)m]N,Ct=δ2Cx2νCx+α(CSC)+ϵp(R,N)kC,RSt=αAAS(RSR)qN[f(RS)m]NS,NSt=αAAS(NSN)+[f(RS)m]NS,CSt=αAAS(CSC)+ϵp(RS,NS)kCS, (4.1)

    in (x,t)(0,L)×(0,) with boundary conditions

    {νR(0,t)δRx(0,t)=νR(0),νN(0,t)δNx(0,t)=νC(0,t)δCx(0,t)=0,Rx(L,t)=Nx(L,t)=Cx(L,t)=0, (4.2)

    and initial conditions

    {R(x,0)=R0(x)0, N(x,0)=N0(x)0, C(x,0)=C0(x)0,RS(x,0)=R0S(x)0, NS(x,0)=N0S(x)0, CS(x,0)=C0S(x)0, (4.3)

    in x(0,L). Here the mortality of algae is assumed to be a constant rate m; qN represents the constant quota of algae. For simplicity, we have assumed that toxin degradation follows first order kinetics with a decay coefficient k. We point out that system (4.1)-(4.3) applies to many flagellate toxins [32].

    There are two types of productions for dissolved toxins [10]. The first assumes that the algae produce toxin more rapidly when there is little nutrient in the system,

    ϵp(R,N)=ϵ[μmaxf(R)]N=ϵμmaxKK+RN,

    where ϵ is a constant coefficient and μmax represents the maximal growth rate. It has been observed that toxins produced by Prymnesium parvum (toxic flagellates) are proportional to the degree of algal nutrient limitation. The second type of toxin production assumes that the toxin is produced proportional to the algal productivity,

    ϵp(R,N)=ϵf(R)N=ϵμmaxRK+RN.

    This case assumes that toxin is produced in proportion to other cellular products and released into the water at a constant rate. We refer to this as the case of cylindrospermopsin, which is a cyanotoxin produced by a variety of freshwater cyanobacteria.

    By [33,Theorem 1 and Remark 1.1], we have the following result.

    Lemma 4.1. ([12,LEMMA 3.1]) System (4.1)-(4.3) has a unique noncontinuable solution and solutions of (4.1)-(4.3) remain non-negative on their interval of existence if they are non-negative initially.

    In the following, we will demonstrate that mass conservation is satisfied in the flow and storage zones for the equations given by (4.1)-(4.3). Let

    W(x,t)=R(x,t)+qNN(x,t) and WS(x,t)=RS(x,t)+qNNS(x,t).

    Then W(x,t) and WS(x,t) satisfy the following system

    {Wt=δ2Wx2νWx+αWSαW, 0<x<L, t>0,WSt=αAASWS+αAASW, 0<x<L, t>0,νW(0,t)δWx(0,t)=νR(0), Wx(L,t)=0, t>0,W(x,0)=W0(x)0, WS(x,0)=W0S(x)0. (4.4)

    Then one can show that (e.g., [9] and [11,Lemma 2.3]) system (4.4) admits a unique positive steady-state solution (R(0),R(0)) and

    limt(W(x,t),WS(x,t))=(R(0),R(0)) uniformly for x[0,L].

    It is not hard to see that (R(0),0,0,R(0),0,0) is the trivial steady-state solution of (4.1)-(4.3). Linearizing system (4.1)-(4.3) around (R(0),0,0,R(0),0,0), we get the following cooperative system for the algae population:

    {Nt=δ2Nx2νNx+α(NSN)+[f(R(0))m]N, 0<x<L, t>0,NSt=αAAS(NSN)+[f(R(0))m]NS, 0<x<L, t>0,νN(0,t)δNx(0,t)=Nx(L,t)=0, t>0,N(x,0)=N0(x)0, NS(x,0)=N0S(x)0, 0<x<L. (4.5)

    Substituting N(x,t)=eλtϕ(x) and NS(x,t)=eλtϕS(x) into (4.5), we obtain the associated eigenvalue problem

    {λϕ(x)=δϕ(x)νϕ(x)+α(ϕS(x)ϕ(x))+[f(R(0))m]ϕ(x), 0<x<L,λϕS(x)=αAAS(ϕS(x)ϕ(x))+[f(R(0))m]ϕS(x), 0<x<L,νϕ(0)δϕ(0)=ϕ(L)=0. (4.6)

    Due to the noncompactness of the model system, we impose the following condition

    αAAS+m>f(R(0)). (4.7)

    By [20,Theorem 2.3] or [21,Theorem 2.1] (see also the arguments in [12,Lemma 3.3], it follows that the eigenvalue problem (4.6) has a principal eigenvalue, denoted by λ0.

    We are in a position to adopt the results developed in [20] to define the basic reproduction ratio for algae. Let S(t):C([0,L],R2)C([0,L],R2) be the C0-semigroup generated by the following system

    {Nt=δ2Nx2νNx+α(NSN)mN, 0<x<L, t>0,NSt=αAAS(NSN)mNS, 0<x<L, t>0,νN(0,t)δNx(0,t)=Nx(L,t)=0, t>0.

    Note that S(t) is a positive C0-semigroup on C([0,L],R2). We further assume that both algae individuals in the flow and storage zones are near the trivial steady-state solution of (4.5), and introduce fertile individuals at time t=0, where the distribution of initial algae individuals in the flow and storage zones is described by φ:=(φ2,φ5)C(ˉΩ,R2). Thus, S(t)φ represents the distribution of fertile algae individuals at time t0.

    Let L:C([0,L],R2)C([0,L],R2) be defined by

    L(φ)()=0(f(R(0))00f(R(0)))(S(t)φ)()dt.

    It then follows that L(φ)() represents the distribution of the total new population generated by initial fertile algae individuals φ:=(φ2,φ5), and hence, L is the next generation operator. We define the spectral radius of L as the basic reproduction ratio for algae, that is,

    R0:=r(L).

    By [34] or [20,Theorem 3.1 (ⅰ) and Remark 3.1], we have the following observation.

    Lemma 4.2. R01 and λ0 have the same sign.

    We first consider the following auxiliary system:

    {Nt=δ2Nx2νNx+α(NSN)+[f(R(0)qNN)m]N,NSt=αAAS(NSN)+[f(R(0)qNNS)m]NS, (4.8)

    in (x,t)(0,L)×(0,) with boundary conditions

    νN(0,t)δNx(0,t)=Nx(L,t)=0, t>0, (4.9)

    and initial conditions

    N(x,0)=N0(x)0, NS(x,0)=N0S(x)0, 0<x<L. (4.10)

    The biologically relevant domain for the system (4.8)-(4.10) is given by

    Y+={(N0,N0S)C([0,L],R2+):0N0()R(0)qN, 0N0S()R(0)qN}.

    For convenience, we let Y0=Y+{(0,0)}, Y0:=Y+Y0={(0,0)}. By Lemma 4.2 and the arguments similar to those in [11,Lemma 3.2,Theorems 3.1 and 3.2], we have the following result.

    Lemma 4.3. ([12,LEMMA 3.6])Assume that (4.7) holds. For any (N0(),N0S())Y+, let (N(,t),NS(,t)) be the solution of (4.8)-(4.10). Then the following statements are valid:

    (i) If R01, then limt(N(x,t),NS(x,t))=(0,0) uniformly for x[0,L];

    (ii) If R0>1, then (4.8)-(4.10) admit a unique positive steady-state solution (N(x),NS(x)) and for any (N0(), N0S())Y0, we have

    limt(N(x,t),NS(x,t))=(N(x),NS(x)), uniformly for x[0,L].

    Recall that X+=C([0,L],R6+) is the biologically relevant domain for the system (4.1)-(4.3). For convenience, we set X0:=X+{(R(0),0,0,R(0),0,0)}, X0:=X+X0={(R(0),0,0,R(0),0,0)}. By Lemma 4.3 and the theory of chain transitive sets (see, e.g., [19,31]), one can lift the threshold type result of (4.8)-(4.10) to the full system (4.1)-(4.3).

    Theorem 4.1. ([12,THEOREM 3.2]) Assume that (4.7) holds. Let

    (R(x,t),N(x,t),C(x,t),RS(x,t),NS(x,t),CS(x,t))

    be the solution of (4.1)-(4.3) with initial data in X+. Then the following statements are valid:

    (i) If R01, then

    limt(R(x,t),N(x,t),C(x,t),RS(x,t),NS(x,t),CS(x,t))=(R(0),0,0,R(0),0,0),

    uniformly for x[0,L].

    (ii) If R0>1, then (4.1)-(4.3) admit a unique positive steady-state solution (R(x),N(x),C(x),RS(x),NS(x),CS(x)), and for any

    (R0(),N0(),C0(),R0S(),N0S(),C0S())X0,

    we have

    limt(R(x,t),N(x,t),C(x,t),RS(x,t),NS(x,t),CS(x,t))       =(R(x),N(x),C(x),RS(x),NS(x),CS(x)), uniformly for x[0,L].

    Next, we consider a model incorporating nutrient recycling. Cyanobacteria excrete some toxins that contain nitrogen, a potential limiting nutrient for algae. Hence, chemical decomposition of the toxin results in nutrient recycling [10]. We assume that ϵ represents a dimensionless coefficient that specifies the allocation to toxin production [10]. Accordingly, the authors in [10] proposed another reaction-diffusion-advection system:

    {Rt=δ2Rx2νRxqN[f(R)m]N+α(RSR)+kqCC,Nt=δ2Nx2νNx+α(NSN)+[(1ϵ)f(R)m]N,Ct=δ2Cx2νCx+α(CSC)+ϵf(R)qNqCNkC,RSt=αAAS(RSR)qN[f(RS)m]NS+kqCCS,NSt=αAAS(NSN)+[(1ϵ)f(RS)m]NS,CSt=αAAS(CSC)+ϵf(RS)qNqCNSkCS, (4.11)

    for (x,t)(0,L)×(0,) with boundary conditions

    {νR(0,t)δRx(0,t)=νR(0),νN(0,t)δNx(0,t)=νC(0,t)δCx(0,t)=0,Rx(L,t)=Nx(L,t)=Cx(L,t)=0, (4.12)

    and initial conditions

    {R(x,0)=R0(x)0, N(x,0)=N0(x)0, C(x,0)=C0(x)0,RS(x,0)=R0S(x)0, NS(x,0)=N0S(x)0, CS(x,0)=C0S(x)0, (4.13)

    for x(0,L), where qN (qC) represents the nutrient quota of algae (toxin). The terms kqCC and kqCCS in (4.11) reflect that the toxin can get recycled back into the system as available nutrient. From the second and fifth equations of (4.11), we realize that only a part, (1ϵ), of the nutrient consumed is used for algal growth, which is discounted by the cost of toxin production.

    In [13], the extinction/persistence of system (4.11)-(4.13) is investigated in terms of a reproduction number by the comparison arguments and the theory of uniform persistence. Due to the introduction of nutrient recycling, the mathematics becomes more challenging. For example, the uniqueness and global attractivity of the positive steady state of system (4.11)-(4.13) are unclear in general. With an additional assumption, we can establish the uniqueness and global attractivity of the positive steady state, see [13,Section 4].

    In [12,Section 4], the zooplankton is further incorporated into system (4.1)-(4.3). Suppose Z and ZS represent the densities of zooplankton in the flow and storage zones, respectively; qZ is the constant nutrient quota for zooplankton; mZ is the mortality of zooplankton. Then the governing equations take the following form:

    {Rt=δ2Rx2νRxqN[f(R)m]N+α(RSR),Nt=δ2Nx2νNx+α(NSN)+[f(R)m]NqZg(N)eηCZ,Ct=δ2Cx2νCx+α(CSC)+ϵp(R,N)kC,Zt=δ2Zx2νZx+α(ZSZ)+[g(N)eηCmZ]Z,RSt=αAAS(RSR)qN[f(RS)m]NS,NSt=αAAS(NSN)+[f(RS)m]NSqZg(NS)eηCSZS,CSt=αAAS(CSC)+ϵp(RS,NS)kCS,ZSt=αAAS(ZSZ)+[g(NS)eηCSmZ]ZS, (4.14)

    in (x,t)(0,L)×(0,) with boundary conditions

    {νR(0,t)δRx(0,t)=νR(0),νN(0,t)δNx(0,t)=νC(0,t)δCx(0,t)=νZ(0,t)δZx(0,t)=0,Rx(L,t)=Nx(L,t)=Cx(L,t)=Zx(L,t)=0, (4.15)

    and initial conditions

    {R(x,0)=R0(x)0, N(x,0)=N0(x)0,C(x,0)=C0(x)0, Z(x,0)=Z0(x)0,RS(x,0)=R0S(x)0, NS(x,0)=N0S(x)0,CS(x,0)=C0S(x)0, ZS(x,0)=Z0S(x)0, (4.16)

    in x(0,L). Here η>0 is a constant and represents the effect of the inhibitor on zooplankton, the term eηC represents the degree of inhibition of C on the growth rate of zooplankton, and the function g(N) has the following form:

    g(N)=ˆμmaxNˆK+N.

    Let X+:=C([0,L],R8+). By comparison arguments, one can show that solutions of system (4.14)-(4.16) exist globally on [0,), and ultimately bounded and uniformly bounded in X+ (see Lemma 4.1 and Lemma 4.2 in [12]). Then we define the solution semiflow Θ(t):X+X+ of (4.14)-(4.16) by

    Θ(t)(ϕ)=u(,t,ϕ),  t0,ϕX+,

    where u(x,t,ϕ) is the solution of (4.14)-(4.16) with u(,0,ϕ)=ϕX+. We can further find a bounded set D in X+ and a t0>0 such that

    Θ(t)(ϕ)D,  tt0, ϕX+,

    and D is positively invariant for Θ(t) in the sense that

    Θ(t)(ϕ)D,  t0, ϕD.

    In view of the assumption (4.7), it follows whenever αAAS is sufficiently large, there exists a constant r>0 such that

    vTM(ϕ(x))vrvTv,ϕD,x[0,L],vR4, (4.17)

    where M(R,N,C,Z,RS,NS,CS,ZS)=

    (m11m1200f(RS)NSm22m23m24ϵp(RS,NS)RSϵp(RS,NS)NSαAASk00m42m43m44),

    and

    m11=αAASqNf(RS)NS, m12=qN[f(RS)m],m22=αAAS+[f(RS)m]qZg(NS)eηCSZS,m23=ηqZg(NS)eηCSZS, m42=g(NS)eηCSZS,m43=ηg(NS)eηCSZS, m24=qZg(NS)eηCS,m44=αAAS+g(NS)eηCSmZ.

    We note that the last four equations in system (4.14)-(4.16) have no diffusion terms, and hence, its solution map Θ(t) is not compact. By arguments similar to those in [11,Lemma 4.1], we have the following result.

    Lemma 4.4. ([12,LEMMA 4.3]) Let (4.7) and (4.17) hold. Then the solution semiflow Θ(t) is κ-contracting in the sense that limtκ(Θ(t)(B))=0 for any bounded set BX+, where κ is the Kuratowski measure of noncompactness.

    By [29,Theorem 2.6], we have the following result.

    Theorem 4.2. ([12,THEOREM 4.1]) Let (4.7) and (4.17) hold. Then Θ(t) admits a global attractor on X+ that attracts each bounded set in X+.

    We note that the system (4.14)-(4.16) admits the following trivial/semitrivial steady states: E0:=(R(0),0,0,0,R(0),0,0,0) and

    E1:=(R(x),N(x),C(x),0,RS(x),NS(x),CS(x),0) provided that R0>1,

    where R0 is the algal reproduction ratio for system (4.1)-(4.3), and

    (R(x),N(x),C(x),RS(x),NS(x),CS(x))

    is the unique positive steady-state solution of (4.1)-(4.3). Linearizing system (4.14)-(4.16) around the state E1, we get the following system for the zooplankton compartments (Z,ZS):

    {Zt=δ2Zx2νZx+α(ZSZ)                               +[g(N)eηCmZ]Z, 0<x<L, t>0,ZSt=αAAS(ZSZ)+[g(NS)eηCSmZ]ZS, 0<x<L, t>0,νZ(0,t)δZx(0,t)=0, Zx(L,t)=0, t>0,Z(x,0)=Z0(x)0, ZS(x,0)=Z0S(x)0, 0<x<L. (4.18)

    The eigenvalue problem associated with (4.18) takes the form:

    {Λψ(x)=δψνψ+α(ψSψ)                               +[g(N)eηCmZ]ψ(x), 0<x<L,ΛψS(x)=αAAS(ψSψ)+[g(NS)eηCSmZ]ψS, 0<x<L,νψ(0)δψ(0)=0, ψ(L)=0. (4.19)

    Due to the loss of compactness, we need to impose the following condition:

    αAAS+mZ>g(NS(x))eηCS(x),  x[0,L]. (4.20)

    The following result is a straightforward consequence of [21,Theorem 2.1].

    Lemma 4.5. Assume that condition (4.20) holds. Then the eigenvalue problem (4.19) admits the principal eigenvalue, denoted by Λ.

    We remark that in [12,Lemma 4.4], the authors used a generalized Krein-Rutman Theorem (see, e.g., [23]) to show that (4.19) admits the principal eigenvalue if (4.20) holds and one additional condition is satisfied. Combining [21,Lemmas 2.1-2.3] with [20,Theorem 2.3], one can obtain [21,Theorem 2.1] and hence Lemma 4.5. Thus, Lemma 4.5 is an improved version of [12,Lemma 4.4] since that additional condition is removed. Here we emphasize that Lemmas 2.1-2.3 in [21] hold true only for the autonomous system. So the arguments in [12,Lemma 4.4] are still useful for us to establish the existence of the principal eigenvalue for degenerate periodic reaction-diffusion systems.

    In the following, we shall adopt the theory developed in [20] to define the basic reproduction ratio for zooplankton. Let S(t):C([0,L],R2)C([0,L],R2) be the C0-semigroup generated by the following system

    {Zt=δ2Zx2νZx+α(ZSZ)mZZ, 0<x<L, t>0,ZSt=αAAS(ZSZ)mZZS, 0<x<L, t>0,νZ(0,t)δZx(0,t)=0, Zx(L,t)=0, t>0.

    Note that S(t) is a positive C0-semigroup on C([0,L],R2). Assume that both zooplankton individuals in the flow and storage zones are near the trivial steady-state solution (0,0) for (4.18), and introduce fertile individuals at time t=0, where the distribution of initial zooplankton individuals in the flow and storage zones is described by φ:=(φ4,φ8)C([0,L],R2). Thus, S(t)φ represents the distribution of fertile zooplankton individuals at time t0.

    Let L:C([0,L],R2)C([0,L],R2) be defined by

    L(φ)()=0(g(NS())eηCS()00g(NS())eηCS())(S(t)φ)()dt.

    It then follows that L(φ)() represents the distribution of the total new population generated by initial fertile zooplankton individuals φ:=(φ4,φ8), and hence, L is the next generation operator. We define the spectral radius of L the basic reproduction ratio of zooplankton compartments for system (4.14)-(4.16), that is,

    Rz0:=r(L).

    By [34] or [20,Theorem 3.1 (ⅰ) and Remark 3.1], we have the following observation.

    Lemma 4.6. Rz01 and Λ have the same sign.

    Recall that X+:=C([0,L],R8+). Let

    X0={(R,N,C,Z,RS,NS,CS,ZS)X+:Z()0 and ZS()0,  and (R,N,C,RS,NS,CS)(R(0),0,0,R(0),0,0)},

    and

    X0:=X+X0.

    Now we are in a position to state the main result of this subsection.

    Theorem 4.3. ([12,THEOREM 4.2]) Assume (4.7), (4.17) and (4.20) hold. Then the following statements are valid:

    (i) If R0<1, then the trivial solution E0 is globally attractive in X+ for (4.14)-(4.16).

    (ii) If R0>1 and Rz0>1, then system (4.14)-(4.16) admits at least one (componentwise) positive equilibrium

    (ˆR(),ˆN(),ˆC(),ˆZ(),ˆRS(),ˆNS(),ˆCS(),ˆZS()),

    and there is a positive constant ζ>0 such that every solution

    (R(,t),N(,t),C(,t),Z(,t),RS(,t),NS(,t),CS(,t),ZS(,t))

    of (4.14)-(4.16) with

    (R(,0),N(,0),C(,0),Z(,0),RS(,0),NS(,0),CS(,0),ZS(,0))X0

    satisfies lim inftminx[0,L]Z(x,t)ζ and lim inftminx[0,L]ZS(x,t)ζ.

    It is still an open problem whether E1 is globally attractive in X+ for system (4.14)-(4.16) in the case where R0>1 and Rz0<1.

    This paper surveys mathematical models describing the spatial variation of population dynamics of harmful algae and toxin production and decay in flowing-water habitats [9,10,14]. Previous mathematical models have been somewhat simplified, and raise many paradoxes [14,18]. One of the paradoxes is the persistence of harmful algae in the river/stream. Intuitively, phytoplankton populations in riverine reservoirs should be washed out by the strong flow, however, we did observe the occurrences of harmful algal blooms. This persistence paradox may be resolved by the complexity of the channel. In fact, the shoreline features and the bed of the channel can retard flow, producing slow-flowing regions. These slow-flowing regions constitute a hydraulic storage zone that may promote algal persistence [9,10]. The authors in [9] proposed and analyzed system (3.1)-(3.3) under the case where R(0)(t)R(0) is a positive constant. The analytical and numerical results in [9] confirm that the system with a storage zone can enhance the persistence of phytoplankton populations. More precisely, numerical work in [9] shows that persistence is possible at higher advective flows for biologically reasonable parameters in the system with a storage zone. The authors in [10] also proposed two-vessel gradostat models of algal dynamics, in which one compartment is a small cove connected to a larger lake. Incorporating seasonal temperature variations into two-vessel gradostat models, rigorous analysis of the time-periodic two-vessel gradostat models are given in [35], and their numerical simulations on the basic reproduction number also indicate that seasonality can play a central role in the extinction/persistence of harmful algae.

    Some previous mathematical models closely related to this survey paper, using ordinary or partial differential equations and integro-differential or integro-difference equations, can be found in [36,37,38,39,40,41,42]. Those works focus on the investigations of spatial spread and persistence of populations in the river/stream. Recently, the authors in [40,43] also studied reaction-diffusion-advection systems describing the growth of a single species where the species lives in both flowing water and river benthos, respectively. The next generation operator mapping the population from one generation to its next generation offsprings was also used to define three different measures that can determine the extinction/persistence of population in a river. The global dynamics and spreading properties were also investigated in [22,30] for time-periodic benthic-drift population models. In [44], the authors further studied a reaction-diffusion-advection system of two species competing in a river environment where the populations grow and compete in the benthic zone and disperse in the drifting water zone. The influences of advection rates, diffusion rates, river length, competition rates, transfer rates, and spatial heterogeneity on the persistence/coexistence of species were also numerically investigated. Comparing with the models reviewed in this paper, those in [40,41,43,44,45] neglect the classes of nutrient(s) and toxin(s). Mathematically, these models are similar to our limiting systems (3.10) and (3.18). Recently, the authors in [46,47,48,49,50,51,52] (and the related references therein) also considered two-species competition models in a one-dimensional advective environment, where the governing equations are restricted to Lotka-Volterra type reaction-diffusion-advection systems. Assuming that the two species share the same resources, these authors focused on the study of different evolution strategies reflected by their different random dispersal rates and/or advection rates.

    In a real ecosystem, the interactions of nutrients, harmful algae, toxins and zooplankton can be very complex. For example, in a real reservoir, P. parvum competes for nitrogen and phosphorus with cyanobacteria, which also excrete allelopathic cyanotoxins that inhibit the growth of P. parvum. A small-bodied zooplankton population consume both types of algae for growth, but the dissolved toxins produced by P. parvum also inhibits zooplankton ingestion, growth and reproduction. In order to understand such complex interactions and reactions in an ecosystem, the authors in [53] proposed a well-mixed chemostat system to explore the dynamics of nutrients, P. parvum, toxin(s) produced by P. parvum, cyanobacteria, cyanotoxin(s) produced by cyanobacteria, and zooplankton.

    In [54], the authors further modify the model in [53] to an unstirred chemostat model of the dynamics of P. parvum, cyanobacteria, and a zooplankton population, in which spatial variations are included, but the compartments of algal toxins produced by P. parvum and cyanobacteria are neglected. The strength of inhibition/allelopathy is directly determined by the densities of P. parvum and cyanobacteria, respectively, which reduces the numbers of the modeling equations. It turns out that this model system admits a coexistence steady state and is uniformly persistent provided that the trivial steady state, two semi-trivial steady states and a global attractor on the boundary are all weak repellers.

    The factors of seasonal temperature, salinity and vertical variations (due to light limitation in deeper riverine systems) also have been known to have crucial influences on the evolution dynamics of harmful algae. It will be challenging and interesting projects if the aforementioned mechanisms are added into the models reviewed in this paper.

    We are grateful to Professor James P. Grover for the continued discussions on the mathematical models of harmful algal blooms. Research of S.-B. Hsu is supported in part by Ministry of Science and Technology, Taiwan. Research of F.-B. Wang is supported in part by Ministry of Science and Technology, Taiwan, and National Center for Theoretical Sciences (NCTS), National Taiwan University and Chang Gung Memorial Hospital (CRRPD3H0011, BMRPD18, NMRPD5F0543 and CLRPG2H0041). Research of X.-Q. Zhao is supported in part by the NSERC of Canada. We would like to express our thanks to the anonymous reviewers for their careful reading and helpful suggestions which led to improvements of our original manuscript.

    The authors declared that they have no conflicts of interest to this work.



    [1] D. L. Roelke, J. P. Grover, B. W. Brooks, et al., A decade of fishkilling Prymnesium parvum blooms in Texas: roles of inflow and salinity, J. Plankton Res., 33 (2011), 243–253.
    [2] G. M. Southard, L. T. Fries and A. Barkoh, Prymnesium parvum: the Texas experience, J. Am. Water Resources Assoc., 46 (2010), 14–23.
    [3] T. L. James and A. De La Cruz, Prymnesium parvum Carter (Chrysophyceae) as a suspect of mass mortalities of fish and shellfish communities in western Texas, Texas J. Sci., 41 (1989), 429–430.
    [4] D. L. Roelke, A. Barkoh, B. W. Brooks, et al., A chronicle of a killer alga in the west: ecology, assessment, and management of Prymnesium parvum blooms, Hydrobiologia, 764 (2016), 29–50.
    [5] V. M. Lundgrena, D. L. Roelke, J. P. Grover, et al., Interplay between ambient surface water mixing and manipulatedhydraulic flushing: Implications for harmful algal bloom mitigation, Ecol. Eng., 60 (2013), 289–298.
    [6] C. G. R. Maier, M. D. Burch and M. Bormans, Flow management strategies to control blooms of the cyanobacterium, Anabaena circinalis, in the river Murray at Morgan, South Australia, Regul. Rivers Res. Mgmt., 17 (2001), 637–650.
    [7] S. M. Mitrovic, L. Hardwick, R. Oliver, et al., Use of flow management to control saxitoxin pro-ducing cyanobacterial blooms in the Lower Darling River, Australia, J. Plankton Res., 33 (2011), 229–241.
    [8] D. L. Roelke, G. M. Gable and T. W. Valenti, Hydraulic flushing as a Prymnesium parvum bloom terminating mechanism in a subtropical lake, Harmful Algae, 9 (2010), 323–332.
    [9] J. P. Grover, S. B. Hsu and F. B. Wang, Competition and coexistence in flowing habitats with a hydraulic storage zone, Math. Biosci., 222 (2009), 42–52.
    [10] J. P. Grover, K. W. Crane, J. W. Baker, et al., Spatial variation of harmful algae and their toxins in flowing-water habitats: a theoretical exploration, J. Plankton Res., 33 (2011), 211–227.
    [11] S. B. Hsu, F. B. Wang and X. Q. Zhao, Dynamics of a periodically pulsed bio-reactor model with a hydraulic storage zone, J. Dynam. Differ. Eq., 23 (2011), 817–842.
    [12] S. B. Hsu, F. B. Wang and X. Q. Zhao, Global dynamics of zooplankton and harmful algae in flowing habitats, J. Diff. Eqns., 255 (2013), 265–297.
    [13] F. B. Wang, S. B. Hsu and X. Q. Zhao, A reaction-diffusion-advection model of harmful algae growth with toxin degradation, J. Diff. Eqns., 259 (2015), 3178–3201.
    [14] C. M. Kung and B. Baltzis, The growth of pure and simple microbial competitors in a moving distributed medium, Math. Biosci., 111 (1992), 295–313.
    [15] F. B. Wang, A system of partial differential equations modeling the competition for two comple-mentary resources in flowing habitats, J. Diff. Eqns., 249 (2010), 2866–2888.
    [16] H. L. Smith and X. Q. Zhao, Dynamics of a periodically pulsed bio-reactor model, J. Diff. Eqns.,155 (1999), 368–404.
    [17] F. B. Wang and C. C. Huang, A reaction-advection-diffusion system modeling the competition for two complementary resources with seasonality in a flowing habitat, J. Math. Anal. Appl., 428 (2015), 145–164.
    [18] C. S. Reynolds, Potamoplankton: paradigms, paradoxes and prognoses, in Algae and the Aquatic Environment, F. E. Round, ed., Biopress, Bristol, UK, 1990.
    [19] X. Q. Zhao, Dynamical systems in population biology, second edition, Springer, New York, 2017.
    [20] W. Wang and X. Q. Zhao, Basic reproduction numbers for reaction-diffusion epidemic models, SIAM J. Appl. Dyn. Syst., 11 (2012), 1652–1673.
    [21] S. B. Hsu, J. López-Gómez, L. Mei, et al., A pivotal eigenvalue problem in river ecology, J. Diff. Eqns., 259 (2015), 2280–2316.
    [22] X. Liang, L. Zhang and X. Q. Zhao, The principal eigenvalue for degenerate periodic reaction-diffusion systems, SIAM J. Math. Anal., 49 (2017), 3603–3636.
    [23] R. D. Nussbaum, Eigenvectors of nonlinear positive operator and the linear Krein-Rutman theo-rem, in Fixed Point Theory (E. Fadell and G. Fournier, eds.), 309–331, Lecture Notes in Mathe-matics 886, Springer, New York/Berlin, 1981.
    [24] M. Ballyk, D. Le, D. A. Jones, et al., Effects of random motility on microbial growth and compe-tition in a flow reactor, SIAM J. Appl. Math., 59 (1998), 573–596.
    [25] H. L. Smith, Monotone Dynamical Systems:An Introduction to the Theory of Competitive and Cooperative Systems, Math. Surveys Monogr 41, American Mathematical Society Providence, RI, 1995.
    [26] R. Aris, Mathematical Modeling, a Chemical Engineer's Perspective, Academic Press, New York, 1999.
    [27] M. Ballyk, D. A. Jones and H. L. Smith, The Freter Model of Biofilm Formation: a review, a book chapter in "Structured Population Models in Biology and Epidemiology", eds P.Magal, S. Ruan, Lecture Notes in Mathematics, Mathematical Biosciences subseries, Springer, 2008.
    [28] K. Deimling, Nonlinear Functional Analysis, Springer-Verlag, Berlin, 1988.
    [29] P. Magal and X. Q. Zhao, Global attractors and steady states for uniformly persistent dynamical systems, SIAM J. Math. Anal., 37 (2005), 251–275.
    [30] X. Yu and X. Q. Zhao, A periodic reaction-advection-diffusion model for a stream population, J. Diff. Eqns., 258 (2015), 3037–3062.
    [31] W. M. Hirsch, H. L. Smith and X. Q. Zhao, Chain transitivity, attractivity, and strong repellers for semidynamical systems, J. Dynam. Differ. Eq., 13 (2001), 107–131.
    [32] M. Murata and T. Yasumoto, The structure elucidation and biological activities of high molec-ular weight algal toxins: maitotoxins, prymnesins and zooxanthellatoxins, Nat Prod. Rep., 17 (2000),293–314.
    [33] R. Martin and H. L. Smith, Abstract functional differential equations and reaction-diffusion sys-tems, Trans. Amer. Math. Soc., 321 (1990), 1–44.
    [34] H. R. Thieme, Spectral bound and reproduction number for infinite-dimensional population struc-ture and time heterogeneity, SIAM J. Appl. Math., 70 (2009), 188–211.
    [35] F. B. Wang, S. B. Hsu and W. Wang, Dynamics of harmful algae with seasonal temperature varia-tions in the cove-main lake, Discrete Cont. Dyn. S., 21 (2016), 313–335.
    [36] K. E. Bencala and R. A. Walters, Simulation of solute transport in a mountain pool-and-riffle stream: a transient storage model, Water Resour. Res., 19 (1983), 718–724.
    [37] Y. Jin, F. M. Hilker, P. M. Steffler, et al., Seasonal invasion dynamics in a spatially heterogeneous river with fluctuating flows, B. Math. Biol., 76 (2014), 1522–1565.
    [38] Y. Jin and M. A. Lewis, Seasonal influences on population spread and persistence in streams: critical domain size, SIAM J. Appl. Math., 71 (2011), 1241–1262.
    [39] F. Lutscher, E. Pachepsky and M. A. Lewis, The effect of dispersal patterns on stream populations, SIAM Rev., 47 (2005), 749–772.
    [40] H. W. McKenzie, Y. Jin, J. Jacobsen, et al., R0 Analysis of a spationtemporal model for a stream population, SIAM J. Appl. Dyn. Syst., 11 (2012), 567–596.
    [41] E. Pachepsky, F. Lutscher, R. M. Nisbet, et al., Persistence, spread and the drift paradox, Theor. Popul. Biol., 67 (2005), 61–73.
    [42] D. C. Speirs and W. S. C. Gurney, Population persistence in rivers and estuaries, Ecology, 82 (2001), 1219–1237.
    [43] Q. Huang, Y. Jin and M. A. Lewis, R0 analysis of a benthic-drift model for a stream population, SIAM J. Appl. Dyn. Syst., 15 (2016), 287–321. Erratum: 16(1), (2017), 770.
    [44] Y. Jin and F. B. Wang, Dynamics of a benthic-drift model for two competitive species, J. Math. Anal. Appl., 462 (2018), 840–860.
    [45] F. Lutscher, M. A. Lewis and E. McCauley, Effects of Heterogeneity on Spread and Persistence in Rivers, B. Math. Biol., 68 (2006), 2129–2160.
    [46] K. Y. Lam, Y. Lou and F. Lutscher, Evolution of dispersal in closed advective environments, J. Biol. Dynam., 9 (2015), 188–212.
    [47] K. Y. Lam, Y. Lou and F. Lutscher, The emergence of range limits in advective environments, SIAM J. Appl. Math., 76 (2016), 641–662.
    [48] Y. Lou and F. Lutscher, Evolution of dispersal in open advective environments, J. Math. Biol., 69 (2014), 1319–1342.
    [49] Y. Lou, X. Q. Zhao and P. Zhou, Global dynamics of a Lotka-Volterra competition-diffusion-advection system in heterogeneous environments, J. de Mathématiques Pures et Appliquées, 121 (2019), 47–82.
    [50] F. Lutscher, E. McCauley and M. A. Lewis, Spatial patterns and coexistence mechanisms in sys-tems with unidirectional flow, Theor. Popul. Biol., 71 (2007), 267–277.
    [51] X. Q. Zhao and P. Zhou, On a Lotka-Volterra competition model: the effects of advection and spatial variation, Calc. Var. Partial Dif., 55 (2016), 55–73.
    [52] P. Zhou and D. Xiao, Global dynamics of a classical Lotka-Volterra competition-diffusion-advection system, J. Funct. Anal., 275 (2018), 356–380.
    [53] J. P. Grover, D. L. Roelke and B. W. Brooks, Modeling of plankton community dynamics charac-terized by algal toxicity and allelopathy: A focus on historical Prymnesium parvum blooms in a Texas reservoir, Ecol. Model., 227 (2012), 147–161.
    [54] S. B. Hsu, F. B. Wang and X. Q. Zhao, A reaction-diffusion model of harmful algae and zooplank-ton in an ecosystem, J. Math. Anal. Appl., 451 (2017), 659–677.
  • This article has been cited by:

    1. Yu Zhao, Chunjin Wei, Dynamics of a Toxin Producing Plankton-Fish Model with Three-Dimensional Patch and Time Delay, 2022, 32, 0218-1274, 10.1142/S0218127422501838
    2. Cheng-Cheng Zhu, Jiang Zhu, Spread trend of COVID-19 epidemic outbreak in China: using exponential attractor method in a spatial heterogeneous SEIQR model, 2020, 17, 1551-0018, 3062, 10.3934/mbe.2020174
    3. Wang Zhang, Hua Nie, Jianhua Wu, A Reaction–Diffusion–Advection Chemostat Model in a Flowing Habitat: Mathematical Analysis and Numerical Simulations, 2023, 33, 0218-1274, 10.1142/S0218127423500736
    4. Wang Zhang, Xiao Yan, Yimamu Maimaiti, Dynamical behavior of a reaction-diffusion-advection chemostat model with Holling III function, 2024, 0924-090X, 10.1007/s11071-024-10366-8
  • Reader Comments
  • © 2019 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(4628) PDF downloads(737) Cited by(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog