Processing math: 100%
Research article

Agriculture and tourism: economic evaluation of sustainable land management

  • Received: 04 September 2021 Revised: 24 December 2021 Accepted: 13 February 2022 Published: 28 February 2022
  • The ways in which the agricultural landscape has been used and managed by man has resulted in substantial changes over time in relation to the economic changes and social needs of local communities. In recent times, thanks to the multifunctional vision of agriculture, growing interest has focused on the recreational aspects of the landscape as a function of its usability. This interest derives both from its importance, highlighted by numerous studies on this aspect, and from its link with rural tourism. The latter phenomenon is growing rapidly and is capable of triggering important processes of development and local growth. In this context, the present study highlights some preliminary considerations on the relationships that, from the point of view of sustainable local development, exist between possible types of tourism and methods of landscape management. To this end, first explore some features of the agricultural landscape and their possible economic evaluations. The study shows that an agricultural landscape in which man is present with agricultural activity, and where the service sector offers adequate opportunities for receptivity, it is possible to create growth and development paths for the local economy. The empirical analysis carried out in the Madonie shows that the resilience of the agricultural landscape is strictly connected to the presence of man in the territory.

    Citation: Filippo Sgroi. Agriculture and tourism: economic evaluation of sustainable land management[J]. AIMS Environmental Science, 2022, 9(1): 83-94. doi: 10.3934/environsci.2022006

    Related Papers:

    [1] Xiaowan Liu, Qin Yue . Stability property of the boundary equilibria of a symbiotic model of commensalism and parasitism with harvesting in commensal populations. AIMS Mathematics, 2022, 7(10): 18793-18808. doi: 10.3934/math.20221034
    [2] A. E. Matouk . Chaos and hidden chaos in a 4D dynamical system using the fractal-fractional operators. AIMS Mathematics, 2025, 10(3): 6233-6257. doi: 10.3934/math.2025284
    [3] Binhao Hong, Chunrui Zhang . Bifurcations and chaotic behavior of a predator-prey model with discrete time. AIMS Mathematics, 2023, 8(6): 13390-13410. doi: 10.3934/math.2023678
    [4] Xin Xu, Yanhong Qiu, Xingzhi Chen, Hailan Zhang, Zhiyuan Liang, Baodan Tian . Bifurcation analysis of a food chain chemostat model with Michaelis-Menten functional response and double delays. AIMS Mathematics, 2022, 7(7): 12154-12176. doi: 10.3934/math.2022676
    [5] Rami Amira, Mohammed Salah Abdelouahab, Nouressadat Touafek, Mouataz Billah Mesmouli, Hasan Nihal Zaidi, Taher S. Hassan . Nonlinear dynamic in a remanufacturing duopoly game: spectral entropy analysis and chaos control. AIMS Mathematics, 2024, 9(3): 7711-7727. doi: 10.3934/math.2024374
    [6] Ibraheem M. Alsulami . On the stability, chaos and bifurcation analysis of a discrete-time chemostat model using the piecewise constant argument method. AIMS Mathematics, 2024, 9(12): 33861-33878. doi: 10.3934/math.20241615
    [7] Ning Cui, Junhong Li . A new 4D hyperchaotic system and its control. AIMS Mathematics, 2023, 8(1): 905-923. doi: 10.3934/math.2023044
    [8] Junhong Li, Ning Cui . A hyperchaos generated from Rabinovich system. AIMS Mathematics, 2023, 8(1): 1410-1426. doi: 10.3934/math.2023071
    [9] Yingyan Zhao, Changjin Xu, Yiya Xu, Jinting Lin, Yicheng Pang, Zixin Liu, Jianwei Shen . Mathematical exploration on control of bifurcation for a 3D predator-prey model with delay. AIMS Mathematics, 2024, 9(11): 29883-29915. doi: 10.3934/math.20241445
    [10] Wei Li, Qingkai Xu, Xingjian Wang, Chunrui Zhang . Dynamics analysis of spatiotemporal discrete predator-prey model based on coupled map lattices. AIMS Mathematics, 2025, 10(1): 1248-1299. doi: 10.3934/math.2025059
  • The ways in which the agricultural landscape has been used and managed by man has resulted in substantial changes over time in relation to the economic changes and social needs of local communities. In recent times, thanks to the multifunctional vision of agriculture, growing interest has focused on the recreational aspects of the landscape as a function of its usability. This interest derives both from its importance, highlighted by numerous studies on this aspect, and from its link with rural tourism. The latter phenomenon is growing rapidly and is capable of triggering important processes of development and local growth. In this context, the present study highlights some preliminary considerations on the relationships that, from the point of view of sustainable local development, exist between possible types of tourism and methods of landscape management. To this end, first explore some features of the agricultural landscape and their possible economic evaluations. The study shows that an agricultural landscape in which man is present with agricultural activity, and where the service sector offers adequate opportunities for receptivity, it is possible to create growth and development paths for the local economy. The empirical analysis carried out in the Madonie shows that the resilience of the agricultural landscape is strictly connected to the presence of man in the territory.



    The study of interactions within microbial ecological communities is one of the most important issues in microbial ecology [1, 2, 3, 4, 5]. An interaction has an impact (neutral, beneficial, or detrimental) on the partner microorganisms involved. Commensalism and syntrophy are among the interactions that benefit the species. A commensalistic association is a relationship in which one microbe derives benefits from the other, while the other microbe is not affected (neither detrimental nor beneficial) by the association. A food chain can be considered a commensal interaction, when the second organism in the food chain lives off the waste products of the first, which in turn is unaffected by the interaction [6, 7]. In a syntrophic association, both organisms benefit from the presence of the other. A food chain can be seen as a syntrophic interaction when the first organism in the food chain is inhibited by the accumulation of its own waste in the environment. This inhibition is diminished by the second organism, which uses the wastes of the first one as a food source [8, 9, 10, 11, 12, 13, 14].

    The aim of this work is to present a unified graphical approach of the models of two species having a commensalistic or a syntrophic relationship. Our objective is threefold. First, we show that the graphical concepts introduced by Tilman [15, 16] to study patterns of competition allow us to study these patterns of commensalism and syntrophy in a unified graphical way. Second, we determine the operating diagram of the model, thus obtaining an extension of all the results on the operating diagrams of the literature. Third, we consider the case including mortality, which has not been addressed in previous studies.

    Concerning the first objective, we show that the concepts and graphical methods of Tilman [15, 16], complemented by Ballyk and Wolkowicz [17], to determine the outcome of competition between two species for two resources are very useful to understand the condition of existence and stability of the equilibria of commensalistic and syntrophic models. Although these systems are not competitive systems, it appears that the concepts introduced in [15, 16, 17] permit a unifying graphical approach of their study.

    Concerning the second objective, we recall that the operating diagram has the operating parameters as its coordinates and the various regions defined within it correspond to qualitatively different asymptotic behaviors. Our model has three operating parameters that are the input concentrations of substrate and the dilution rate D. These parameters are control parameters since they are under the control of the experimenter. Apart from these three parameters, which can vary, all other parameters have biological meaning and are fitted using experimental data from ecological and/or biological observations of organisms and substrates. Therefore, the operating diagram is the bifurcation diagram that shows how the system behaves when we vary the control parameters. The importance of the operating diagram for bioreactors was emphasized in [18]. Although this diagram was not considered in the classic monograph on the chemostat [19], its importance was mentioned by Smith and Waltman, who stated that the operating diagram is probably the most useful tool for discussing the behavior of the model in relation to the parameters [19, p. 252]. The operating diagram is introduced in the recent book on the mathematical theory of the chemostat [20]. It is often constructed both in biological literature [14, 21, 22, 23] and in mathematical literature, in the study of anaerobic digestion [24, 25, 26, 27], commensalistic and syntrophic systems [12, 28, 29, 30, 31], microbial food-webs [23, 32], inhibition [33, 34], chemostats in series [35], and density-dependent models [36, 37, 38].

    Concerning the third objective, when there is no mortality, by using the theory of asymptotically autonomous systems, we can reduce the study to that of a model in dimension two so that the study of the stability is easy. In the general case, this reduction cannot be made and the Routh Hurwitz conditions must be used to determine whether the equilibrium is stable or not.

    This paper is organized as follows: In Section 2, we present the mathematical model and give the assumptions made on the growth functions. In Section 3, we describe Tilman's graphical method and we show how the existence and stability condition of the equilibria of the model can be easily read on the position of the projections of equilibria into the feasible set. We also construct in this section the operating diagram of the model. In Section 4, we show that the graphical method also applies in the case where the species can be inhibited by their limiting substrate. Some complements on the operating diagrams are given in Section 5. Finally, some conclusions are drawn in Section 6. Global asymptotic stability results are given in Appendix A, and bifurcation diagrams with respect to one of the operating parameters are shown in Appendix B. The technical proofs are reported in Appendix C. In Appendix D, we propose a review of the results of the literature on commensalism and syntrophy.

    We consider the two-species system modeled by

    ˙S1=D(Sin1S1)k1μ1(S1,S2)X1,˙X1=(μ1(S1,S2)D1)X1,˙S2=D(Sin2S2)+k3μ1(S1,S2)X1k2μ2(S1,S2)X2,˙X2=(μ2(S1,S2)D2)X2, (2.1)

    where Xi, i=1,2, represents the concentration of species i; Sj, j=1,2, is the concentration of chemical j and Sinj its inlet concentration; ki, i=1,2,3, are yield factors, referring to the amount of chemical that is produced or consumed by the growth of a unit amount of the biomass of microbial species; and Di, i=1,2, represents the disappearance rates of species i, and are modeled by

    Di=αiD+ai, (2.2)

    where D=1/HRT is the dilution rate, HRT being the hydraulic retention time; the non-negative death (or decay) rate parameters a1 and a2 are taken into consideration; and the coefficient αi[0,1], i=1,2, represents the biomass proportion that leaves the bioreactor. This coefficient models the decoupling between solids and hydraulic residence time [30, 31].

    Well-mixed continuous bioreactors (i.e., chemostats) for the culture of multiple species are modeled with the commonly used Monod-type kinetics. For example, in [2], the growth rates of the two species, μi, are modeled by multiplying substrate limitation terms described as Monod kinetics and inhibition terms as follows:

    μ1(S1,S2)=m1S1K1+S111+L1S2,μ2(S1,S2)=m2S2K2+S211+L2S1, (2.3)

    where mi is the maximum growth rate, Ki is the half-velocity constant, and Li quantifies the strength of inhibition of substrate Sj, ji, on species i. If Li=0, then there is no inhibition, and so μi depends only on Si.

    In Figure 1, we illustrate (2.1) by using the notations and representations of microbial communities proposed by Di and Yang [2]. Species X2 is produced by consuming chemical S2, which itself is produced by species X1 through consuming chemical S1, that acts as the substrate of the overall system. System C1, which corresponds to L1=L2=0 in (2.3), is an example of pure commensalism, where the second population (the commensal population) depends for its growth on the first population (the host), and thus benefits from its interaction, while the host population is not affected by the growth of the commensal population. System C2, which corresponds to L1=0 and L2>0 in (2.3), is also an example of commensalism since it has a cascade structure and the first population is not affected by the growth of the second population.

    Figure 1.  Commensalistic and syntrophic systems without self-inhibition.

    On the other hand, systems S1, which corresponds to L1>0 and L2=0 in (2.3), and S2, which corresponds to L1>0 and L2>0 in (2.3), are not commensal since the first population is affected by the growth of the second population. For instance, in S1, the first organism is inhibited by the substrate S2 produced by the degradation of S1 by X1. Hence, the extent to which the substrate S1 is degraded by the organism X1 to produce the substrate S2, which is necessary for the growth of X2, depends on the efficiency of the removal of the product S2 by the bacteria X2. Therefore, each population needs the other one for its growth, that is, there is a syntrophic relationship between them.

    Remark 1. Case C1 corresponds to the base system depicted on [2, Figure 1]. Cases C2, S1, and S2 correspond, respectively, to cases 1-4, 1-1, and 2-3 shown in [2, Figures 2 and 3]. Notice that [2] also considered inhibitions between species, see cases 1-2 and 1-3 in [2, Figure 2], and five other systems obtained by combining them, see [2, Figure 3]. Although very important for microbial ecology, these species inhibitions will not be considered in this paper. The models C1, C2, S1, and S2 have been extensively studied in the literature, see Tables A3 and A4 in Appendix D.

    There are many ways of representing inhibition other than (2.3). For example, instead of decreasing the maximum growth rate mi of the Monod function as in (2.3), we can increase its half-velocity constant (or substrate affinity) Ki:

    μ1(S1,S2)=m1S1K1+L1S2+S1,μ2(S1,S2)=m2S2K2+L2S1+S2, (2.4)

    where Li quantifies the strength of inhibition of Sj on species i. See also the Kreikenbohm and Bohl function KB1 in Table A7.

    In this work, instead of considering particular growth functions such as (2.3), (2.4), or KB1, we consider generalized growth functions characterized by their qualitative behaviors. Assume that μ1,μ2:R2+R+ are of class C1 such that μ1(0,S2)=0 for S20 and μ2(S1,0)=0 for S10. Hence, no growth takes place for species i=1,2 without substrate Si. This hypothesis is made throughout the paper and will not be repeated. The solutions of (2.1), with prescribed initial conditions, exist, are unique, are bounded, and the non-negative cone is positively invariant. The positive cone is also positively invariant. We make the following assumptions:

    H1. For S1>0 and S20, we have μ1S1(S1,S2)>0 and μ1S2(S1,S2)0.

    H2. For S10 and S2>0, we have μ2S2(S1,S2)>0 and μ2S1(S1,S2)0.

    Hypotheses H1 and H2 signify that the growth for species i=1,2 increases with Si, while it is inhibited by the other substrate Sj, ji : the first organism is inhibited by its product S2 and the second organism is inhibited by the food S1 of the first organism.

    Remark 2. Since the partial derivative μiSj(S1,S2), for ji, can be equal to zero, the inhibition of species i by substrate ji is not mandatory. Hence, the assumptions H1 and H2 cover the cases C1, C2, S1, and S2 of Figure 1. For example, C1 corresponds to μ1 depending only on S1 and increasing for all S1>0 and μ2 depending only on S2 and increasing for all S2>0.

    To reduce the number of parameters, model (2.1) is further converted to a simpler one where the yield factors ki are fixed to 1. We scale system (2.1) using the following change of variables and notations:

    s1=k3S1/k1,x1=k3X1,s2=S2,x2=k2X2,sin1=k3Sin1/k1,sin2=Sin2.

    We obtain the system of differential equations

    ˙s1=D(sin1s1)f1(s1,s2)x1,˙x1=(f1(s1,s2)D1)x1,˙s2=D(sin2s2)+f1(s1,s2)x1f2(s1,s2)x2,˙x2=(f2(s1,s2)D2)x2, (2.5)

    where the yield factors ki are fixed to 1 and the growth functions f1,f2:R2+R+ are given by

    f1(s1,s2)=μ1(k1s1/k3,s2)andf2(s1,s2)=μ2(k1s1/k3,s2). (2.6)

    Since μ1 and μ2 in (2.1) satisfy the conditions of H1 and H2, then the growth functions f1 and f2 in (2.5) have the same qualitative properties.

    In this section, we explain the graphical method of [15, 16, 17].

    The "feasible set" F is the set of points (s1,s2)R2+ where the (s1,s2)-coordinate of any equilibrium point must be located. We have the following result.

    Lemma 1. If E=(s1,x1,s2,x2) is an equilibrium point of (2.5), then 0<s1sin1, 0<s2sin1+sin2s1 and

    x1=DD1(sin1s1),x2=DD2(sin1+sin2s1s2). (3.1)

    Proof. The equilibria of the system are the solutions of the following set of equations obtained by setting the right-hand sides of the equations in (2.5) equal to zero

    0=D(sin1s1)f1(s1,s2)x1, (3.2)
    0=(f1(s1,s2)D1)x1, (3.3)
    0=D(sin2s2)+f1(s1,s2)x1f2(s1,s2)x2, (3.4)
    0=(f2(s1,s2)D2)x2. (3.5)

    At an equilibrium point we necessarily have s1>0 and s2>0. Using (3.2)+(3.3) and (3.4)+(3.5)-(3.3), we obtain the equations

    0=D(sin1s1)D1x1,0=D(sin2s2)+D1x1D2x2.

    By solving these equations, we obtain x1 and x2 as functions of s1 and s2, as given by (3.1). Consequently, the requirement that the components xi are non-negative imposes that the coordinates (s1,s2) of an equilibrium point must satisfy s1sin1 and s1+s2sin1+sin2.

    Therefore, the feasible set is given by

    F={(s1,s2)R2+:0<s1sin1,0<s2sin1+sin2s1}.

    The boundary of F consists of two portions of the coordinate axes, together with two curves, namely the feasible set boundary FSBi for population i, defined as follows:

    FSB1={(s1,s2)R2+:s1=sin1,0<s2sin2},FSB2={(s1,s2)R2+:0<s1sin1,s1+s2=sin1+sin2}. (3.6)

    These line segments are plotted in green and red, respectively, in Figures 2 and 3.

    Figure 2.  The feasible set F, the ZNGIs and the regions R1, R+1, R+2, and R2, for the commensalistic cases C1 and C2.
    Figure 3.  The feasible set F, the ZNGIs and the regions R1, R+1, R+2, and R2, for the syntrophic Cases S1 and S2.

    Remark 3. The formulas (3.1) show that the x1 and x2 components of an equilibrium point are uniquely determined by its s1 and s2 components. This is why we can, without risk of confusion, denote by the same notation an equilibrium point and its projection (s1,s2) on the feasible set, as we do in Figures 2 and 3. We will use this abuse of notation in all the figures.

    Then we construct the ZNGI for each population and plot them in (s1,s2)-plane using the same color used for population i as used for its FSBi. The ZNGI for population i, denoted ZNGIi, is the curve of substrate concentrations along which the decline in biomass density is balanced by its growth. Therefore,

    ZNGIi={(s1,s2)R2+:fi(s1,s2)=Di},i=1,2. (3.7)

    The following step is to identify equilibria: Each intersection of a green curve and a red curve in the feasible set corresponds to an equilibrium point, i.e., is the projection on the feasible set of an equilibrium point, as depicted in Figures 2 and 3. More precisely, we have the following result.

    Lemma 2.The intersection point (sin1,sin2)=FSB1FSB2 corresponds to the washout equilibrium E0=(sin1,0,sin2,0) where both species are extinct.

    Any point (ˉs1,ˉs2)ZNGI1FSB2 corresponds to a boundary equilibrium E1=(ˉs1,ˉx1,ˉs2,0), where species 2 is absent and species 1 is present.

    Any point (˜s1,˜s2)FSB1ZNGI2 corresponds to a boundary equilibrium E2=(˜s1,0,˜s2,˜x2), where species 1 is absent and species 2 is present.

    Any point (s1,s2)ZNGI1ZNGI2 lying in the interior Fo of the feasible set F corresponds to a coexistence equilibrium Ec=(s1,x1,s2,x2), where both species are present.

    Proof. Recall that an equilibrium point E=(s1,x1,s2,x2) is a solution of Eqs (3.2)–(3.5). Four cases must be distinguished:

    1) Assume that x1=x2=0, which corresponds to the washout equilibrium point E0. Then, from (3.2) and x1=0 it is deduced that s1=sin1 and from (3.4) and x2=0, it is deduced that s2=sin2. Hence (s1,s2)=(sin1,sin2)=FSB1FSB2.

    2) Assume that x1>0 and x2=0, which corresponds to a boundary equilibrium point E1. Then, from (3.3) and x1>0 it is deduced that f1(s1,s2)=D1. Therefore, (s1,s2)ZNGI1. Now using (3.2)+(3.4), we obtain

    0=D(sin1+sin2s1s2)f2(s1,s2)x2. (3.8)

    From x2=0 and (3.8), we deduce that s1+s2=sin1+sin2. Now using x1>0, from (3.1) we deduce that s1<sin1. Therefore, (s1,s2)FSB2. Hence, (s1,s2)ZNGI1FSB2.

    3) Assume that x1=0 and x2>0, which corresponds to a boundary equilibrium point E2. Then, from x1=0 and (3.2), we deduce that s1=sin1. Now using x2>0, from (3.1) we deduce that s2<sin2. Therefore, (s1,s2)FSB1. From x2>0 and (3.5), we have f2(s1,s2)=D2. Therefore, (s1,s2)ZNGI2. Hence, (s1,s2)FSB1ZNGI2.

    4) If x1>0 and x2>0, which corresponds to a coexistence equilibrium point Ec, then, from x1>0 and (3.3), we have f1(s1,s2)=D1 and from x2>0 and (3.5), we have f2(s1,s2)=D2. Therefore, (s1,s2)ZNGI1ZNGI2. The equilibrium point Ec exists if and only if x1 and x2 are positive, i.e., (s1,s2)Fo.

    The last step is to determine the conditions of existence and stability of equilibria, which result from their location with respect to the ZNGIs. A stable equilibrium is plotted with a solid circle and an unstable one with a circle, as depicted in Figures 2 and 3. This convention will be used in all figures.

    The washout equilibrium E0 is unique and always exists. The number and nature of the boundary and positive equilibria depend on the relative positions of the ZNGIs. The stability of E0, as well as the existence, uniqueness, and stability of the other equilibrium points, require the use of assumptions H1 and H2 on the growth functions μ1 and μ2. Since the growth functions f1 and f2 are given by (2.6), they satisfy the following conditions, where we use the notations fij=fi/sj, i,j=1,2, for the partial derivatives.

    For s1>0,s20,f11(s1,s2)>0 and f12(s1,s2)0. (3.9)
    For s10,s2>0,f21(s1,s2)0 and f22(s1,s2)>0. (3.10)

    In order to describe the ZNGIs, it is convenient to use the concept of break-even concentration.

    Definition 1. Let

    m1(s2)=f1(+,s2)=sups1>0f1(s1,s2).

    For s20 and D[0,m1(s2)), the break-even concentration is the unique solution s1=λ1(s2,D) of equation f1(s1,s2)=D.

    Let

    m2(s1)=f2(s1,+)=sups2>0f2(s1,s2).

    For s10 and D[0,m2(s1)), the break-even concentration is the unique solution s2=λ2(s1,D) of equation f2(s1,s2)=D.

    Notice that the existence and uniqueness of s1=λ1(s2,D) follows from (3.9), since f1/s1>0 for all s1>0. Using the implicit function theorem, one sees that λ1/s2=f12/f110. Therefore, λ1(s2,D) is increasing in s2. Similarly, s2=λ2(s1,D) is well defined and satisfies λ2/s1=f21/f220. Therefore, it is increasing in s1.

    Using the break-even concentrations, the ZNGIs (3.7) are given by

    ZNGI1={(s1,s2)R2+:s1=λ1(s2,D1)}, (3.11)
    ZNGI2={(s1,s2)R2+:s2=λ2(s1,D2)}. (3.12)

    Since ZNGI1 is the graph of the increasing function s2s1=λ1(s2,D1), it divides the positive cone R2+ into two connected regions

    R1={(s1,s2)R2+:f1(s1,s2)<D1},
    R+1={(s1,s2)R2+:f1(s1,s2)>D1}.

    The region R1 contains the origin, and the region R+1 is possibly empty. Similarly, since ZNGI2 is the graph of the increasing function s1s2=λ2(s1,D2), it divides the positive cone R2+ into two connected regions

    R2={(s1,s2)R2+:f2(s1,s2)<D2},
    R+2={(s1,s2)R2+:f2(s1,s2)>D2}.

    The region R2 contains the origin, and the region R+2 is possibly empty.

    We say that an equilibrium is stable if it is locally exponentially stable, i.e., the Jacobian matrix has eigenvalues with strictly negative real parts. We can now give the main result.

    Proposition 3. Assume that (3.9) and (3.10) are satisfied. The conditions of existence and stability of the equilibria of (2.5) are given in Table 1.

    Table 1.  Conditions of existence and stability of the equilibria of the system (2.5). Here, (ZNGI1,ZNGI2) is the signed angle between ZNGI1 and ZNGI2 at Ec.
    Existence condition Stability condition (local)
    E0=FSB1FSB2 Always exists E0R1R2
    E1=ZNGI1FSB2 E0R+1 E1R2
    E2=FSB1ZNGI2 E0R+2 E2R1
    EcZNGI1ZNGI2 EcFo (ZNGI1,ZNGI2)<0

     | Show Table
    DownLoad: CSV

    Proof. The proof is given in Appendix C.1.

    In Table 1, (ZNGI1,ZNGI2) is the signed angle between ZNGI1 and ZNGI2 at their intersection point Ec. As it is usual, we define the signed angle between two intersecting curves to be the signed angle between the tangents at the point of intersection.

    We have thus obtained a complete description of the existence and stability conditions of the equilibria of (2.5), simply by considering the location of their projections on the feasible set F. Existence of the equilibria is determined from the intersections of the ZNGIs and FSBs plotted with different colors. Stability of the equilibria is determined from the location of the projections of the equilibria E0, E1, E2, and Ec, in the regions R1, R+1, R+2 and R2 of the feasible set F. Let us illustrate these results in the cases shown in Figures 2 and 3.

    ● On these figures, E0 is unstable since E0R1R2 and E1 exists since E0R+1. On Figures 2 and 3 (Case S1), E1 is unstable since E1R2. On Figure 3 (Case S2), E1 is stable since E1R2.

    ● On Figures 2 and 3 (Case S1), E2 exists since E0R+2 and is unstable since E2R1. On Figure 3 (Case S2), E2 does not exist since E0R+2.

    ● On Figures 2 and 3 (Case S1), Ec is unique and stable since, at Ec, (ZNGI1,ZNGI2)<0.

    ● On Figure 3 (Case S2), E1c is stable since, at E1c, (ZNGI1,ZNGI2)<0, and E2c is unstable since, at E2c, (ZNGI1,ZNGI2)>0.

    In this section, we describe the operating diagram of (2.5). Since system (2.5) has three operating parameters, and it is not easy to visualize regions in the three-dimensional operating parameter space, we fix the dilution rate D and we show the operating diagram in the operating plane (sin1,sin2). The effects of D can be shown in a series of operating diagrams. In Section 5, we fix the operating parameter sin2 and we show the operating diagram in the operating plane (sin1,D). The effects of sin2 can be shown in a series of operating diagrams.

    We fix the growth functions fi and the parameters αi and ai, i=1,2. We begin with the more simple Cases C1, C2, and S1. We see in Figure 2 and Figure 3 (Case S1) that the ZNGIs have a unique intersection point (s1(D),s2(D)):=ZNGI1ZNGI2.

    We consider the curves

    Γi={(sin1,sin2)R2+:fi(sin1,sin2)=Di},i=1,2. (3.13)

    Even though the Γi are defined by the same equations as the ZNGIs, see (3.7), it should be noted that the Γi are curves in the plane of operating parameters (sin1,sin2), while the ZNGIi are curves in the phase plane (s1,s2). The curves Γ1 and Γ2, defined by (3.13), together with the curve Γ3 given by

    Γ3={(sin1,sin2)R2+:sin1+sin2=s1(D)+s2(D) and sin1>s1(D)}, (3.14)

    divide the set of operating parameters (sin1,sin2) into five regions denoted Ik, k=0,,4, see Figure 4. For the C1 case, we have:

    Γ1 is the vertical line with equation sin1=λ1(D1),

    Γ2 is the horizontal line with equation sin2=λ2(D2),

    Γ3 is the oblique line with equation sin1+sin2=λ1(D1)+λ2(D2).

    Figure 4.  The operating diagram in the operating plane (sin1,sin2) and 0<D<min(δ1,δ2), where the δi are defined by (3.16). The asymptotic behavior in the various regions of the operating diagram is shown in Table 2 for Cases C1 and C2, and in Table 3 for case S1. The case Dmin(δ1,δ2) is shown in Figure 6.

    The novelty when we consider the C2 case is that Γ2 becomes the curve with equation sin2=λ2(sin1,D2). We have the following result:

    Proposition 4. The condition of existence and stability of the equilibria of the C1 and C2 models in the regions Ik, k=0,,4, of Figure 4 (Case C1 or C2) are given in Table 2.

    Table 2.  Existence and stability of equilibria of (2.5) in the regions of the operating diagram depicted in Figure 4 (Cases C1 and C2). The letter U means that the equilibrium is unstable, and GAS means that it is globally asymptotically stable. No letter means that it does not exist.
    E0 E1 E2 Ec Color
    I0 GAS Red
    I1 U GAS Yellow
    I2 U GAS Blue
    I3 U U GAS Green
    I4 U U U GAS Green

     | Show Table
    DownLoad: CSV

    Proof. The proof consists simply of looking at the location of the equilibria in the feasible set and applying Proposition 3. We give the details for the Case C1. Assume that (sin1,sin2)I4. We see in Figure 5 that E0 is unstable since E0R1R2, and E1 exists since E0R+1 and is unstable since E1R2. Moreover, E2 exists since E0R+2 and is unstable since E2R1. On the other hand, Ec exists, is unique, and, stable since, at Ec, we have (ZNGI1,ZNGI2)<0. The proof for global asymptotic stability is given in Appendix A.1. This proves the results depicted in the last row of Table 2. The proofs for the other regions are illustrated in Figure 5. The details of the proof for the C2 case are given in Appendix C.2.

    Figure 5.  Proof of Proposition 4 in the Case C1: feasible sets and ZNGIs, showing equilibria and their stability, for the four regions of the operating diagram, shown in Figure 4 (Case C1).

    Although the regions I3 and I4 are different in terms of the existence of equilibria, they are colored green because they correspond to the stability of the coexistence equilibrium Ec.

    In addition to the curves Γ1 and Γ2, defined by (3.13), and the curve Γ3, defined by (3.14), consider the line segment defined by

    Γ4={(sin1,sin2)R2+:sin1=s1(D) and sin2>s2(D)}. (3.15)

    The novelty when we consider model S1 is that the curve Γ1 becomes the curve with equation sin1=λ1(sin2,D1), and is therefore distinct from Γ4, which is the vertical line with equation sin1=s1(D). Therefore, in the S1 case, the curves Γi, i=1,,4, defined by (3.13)–(3.15), divide the set of operating parameters (sin1,sin2) into six regions denoted Ik, k=0,,5, see Figure 4 (Case S1). We have the following result:

    Proposition 5. The conditions of existence and stability of the equilibria of the S1 model in the regions Ik, k=0,,5, of Figure 4 (Case S1) are given in Table 3.

    Table 3.  Existence and stability of equilibria of (2.5) in the regions of the operating diagram depicted in Figure 4 (Case S1). The letter S means that the equilibrium is locally exponentially stable. If D1=D2=D, the letter S should be replaced by GAS.
    E0 E1 E2 Ec Color
    I0 S Red
    I1 U S Yellow
    I2 U S Blue
    I3 U U S Green
    I4 U U U S Green
    I5 U U S Green

     | Show Table
    DownLoad: CSV

    Proof. The proof is given in Appendix C.3.

    Although the regions I3, I4, and I5 are different in terms of the existence of equilibria, they are colored green because all three correspond to the stability of the coexistence equilibrium Ec.

    The number of regions in the operating plane (sin1,sin2) depends on D. More precisely, let us define

    For C1:δ1=(m1a1)/α1,δ2=(m2a2)/α2.For C2:δ1=(m1a1)/α1,δ2=(m2(0)a2)/α2.For S1:δ1=(m1(0)a1)/α1,δ2=(m2a2)/α2. (3.16)

    If 0<D<min(δ1,δ2), then all regions Ik, k=0,,4, appear as shown in Figure 4. If δ1<δ2 and δ1D<δ2, only I0 and I2 appear. If δ2<δ1 and δ2D<δ1, only I0 and I1 appear. Finally, if Dmax(δ1,δ2), then the region I0 invades the whole plane, see Figure 6.

    Figure 6.  The operating diagram of model S1 in the operating plane (sin1,sin2) and Dmin(δ1,δ2), where the δi are defined by (3.16).

    To simplify the analysis, we assume that the ZNGIs intersect in two points, as in Figure 3 (Case S2). Cases with more than two intersections can be studied in the same way. More precisely, we make the following assumption.

    H3. There exists δ0>0 such that, for all D(0,δ0), the ZNGIs have two distinct intersection points (s11(D),s12(D)) and (s21(D),s22(D)); for D=δ0 the ZNGIs are tangent, and for D>δ0 there is no intersection.

    In this case, the line segments Γ3 and Γ4 defined in (3.14) and (3.15), respectively, must be defined for each intersection point of the ZNGIs. More precisely, we consider the curves

    Γj3={(s21(D),s22(D))R2+:sin1+sin2=sj1(D)+sj2(D) and sin1sj1(D)}Γj4={(s21(D),s22(D))R2+:sin1=sj1(D) and sin2sj2(D)},j=1,2. (3.17)

    The curves Γ1 and Γ2 defined by (3.13) and Γj3, Γj4, j=1,2, defined by (3.17) divide the set of operating parameters (s21(D),s22(D)) in nine regions denoted Ik, k=0,,8, see Figure 7. According to the value of D, some of the regions can be empty, as shown in the typical example studied in Figure 9. These regions are corresponding to different system behaviors, as stated in the following result.

    Figure 7.  The operating diagram of (2.5) in the (sin1,sin2) operating plane and D(0,δ0). The case Dδ0 is shown in Figure 9.
    Figure 8.  Proof of Proposition 6: feasible sets and ZNGIs, showing equilibria and their stability, for the nine regions of the operating diagram, shown in Figure 7.
    Figure 9.  The operating diagram of (2.5) for the parameter values given in Table 4. Here, δ01.602, δ13.111, and δ2=4.875. When D>δ2, region I0 invades the whole plane. The table showing the equilibria and their stability is the same as in Figure 7.

    Proposition 6. The conditions of existence and stability of the equilibria of (2.5) in the regions Ik, k=0,,8, shown in Figure 7 are given in the table shown in this figure.

    Proof. The proof consists simply in looking at the location of the equilibria in the feasible set and applying Proposition 3. Assume that (s21(D),s22(D))I8. We see in Figure 8 that E0 is unstable since E0R1R2, and E1 does not exist since E0R+1. Moreover, E2 exists since E0R+2 and is stable since E2R1. On the other hand, E1c is stable since, at E1c, we have (ZNGI1,ZNGI2)<0 and E2c is unstable since, at E2c, we have (ZNGI1,ZNGI2)>0. This proves the results depicted in the last row of the table in Figure 7. Similarly, if (s21(D),s22(D))I7, we see in Figure 8 that E0 is stable since E0R1R2, E1 does not exist since E0R+1, and E2 does not exist since E0R+2. On the other hand, E1c is stable since, at E1c, we have (ZNGI1,ZNGI2)<0, and E2c is unstable since, at E2c, we have (ZNGI1,ZNGI2)>0. This proves the results described in the penultimate row of the table in Figure 7. The proofs for the other regions are shown in Figure 8.

    Although the regions I3, I4, and I5 are different in terms of the existence of equilibria, they are colored green because all three correspond to the stability of the coexistence equilibrium E1c.

    We now illustrate how the operating diagram behaves when D is varied.

    To do this, we consider the growth functions defined by (2.3), and the values of the biological parameters used are listed in Table 4. The operating diagrams corresponding to various values of D are shown in Figure 9. This figure shows that, as D is increased, the green regions I3, I4, and I5, corresponding to the stability of E1c, shrink until they disappear completely at D=δ0. At this value of D, a saddle-node bifurcation occurs, in which E1c and E2c collide and annihilate each other. For D>δ0, the red region I0, corresponding to the stability of E0, increases in size until it invades the entire operating parameter plane.

    Table 4.  The parameter values for the growth functions (2.3) used in Figures 9, 17, and 18. These values have no particular biological significance. They have been chosen to illustrate possible behaviors.
    Parameters values
    m1 K1 L1 m2 K2 L2 α1 a1 α2 a2
    4 1 0.3 3 1 0.2 0.8 0.1 0.9 0.2

     | Show Table
    DownLoad: CSV

    In Figure 1, we considered only the case without self inhibition of a species i by its limiting substrate Si, as shown by the examples (2.3) and (2.4), or the general assumptions H1 and H2. In this section, our aim is to show that our graphical method applies without added difficulty to the case where self-inhibition is possible. For example, to represent self-inhibition, we can replace the Monod functions in (2.3) with Haldane functions and obtain

    μ1(S1,S2)=m1S1K1+S1+S21/KI111+L1S2,
    μ2(S1,S2)=m2S2K2+S2+S22/KI211+L2S1,

    or in (2.4) and obtain

    μ1(S1,S2)=m1S1K1+L1S2+S1+S21/KI1,
    μ2(S1,S2)=m2S2K2+L2S1+S2+S22/KI2.

    See also the Kreikenbohm and Bohl function KB2 in Table A7.

    In Figure 10, we illustrate the commensalistic systems with one self inhibition, the syntrophic systems with one self inhibition, and the systems with two self inhibitions. Some of the twelve systems shown in Figure 10 were considered in the literature, see Tables A5 and A6 in Appendix D.

    Figure 10.  Commensalistic and syntrophic systems with self inhibition.

    We will not try to propose a general definition of inhibition, as such a definition would probably not cover all the cases of interest for applications. We will not make a full description of every case in Figure 10 as we have done when there is no self-inhibition. Our aim is above all to show that our qualitative graphical method is applicable in each of these cases, and we will illustrate this in Cases C12 and S12 which have been particularly studied in the literature, see Tables A5 and A6.

    We consider Case C12 in Figure 10. The model takes the form

    ˙s1=D(sin1s1)f1(s1)x1,˙x1=(f1(s1)D1)x1,˙s2=D(sin2s2)+f1(s1)x1f2(s2)x2,˙x2=(f2(s2)D2)x2. (4.1)

    We assume that f1 satisfies the following condition:

    For all s1>0,f1(s1)>0. (4.2)

    We assume that f2 satisfies the following condition:

    There exists sm2>0 such that f2(s2)>0 for 0<s2<sm2 and f2(s2)<0 for s2>sm2. (4.3)

    We now define the break-even concentrations λ1(D), λ12(D), and λ22(D) of (4.1).

    Definition 2. Let

    m1=f1(+):=sups1>0f1(s1).

    For D[0,m1), the break-even concentration is the unique solution s1=λ1(D) of equation f1(s1)=D.

    Let

    m2=f2(sm2):=sups2>0f1(s2).

    For D[0,m2), the break-even concentration are the solutions λ12(D) and λ22(D) of equation f2(s2)=D, such that 0<λ12(D)<sm2<λ22(D)+.

    Note that for D close to m2, we necessarily have two solutions, i.e.,

    0<λ12(D)<sm2<λ22(D)<+.

    However, for D small enough, there may be only one solution, λ12(D), and the second solution λ22(D) does not exist. The ZNGIs of (4.1) are given by

    ZNGI1={(s1,s2)R2+:s1=λ1(D1)}

    and

    ZNGI2={(s1,s2)R2+:s2=λ12(D2)}{(s1,s2)R2+:s2=λ22(D2)}. (4.4)

    Note that, in this case, the R2 region has two connected components: R2=R21R22, where

    R21={(s1,s2)R2+:s2<λ12(D2)},
    R22={(s1,s2)R2+:s2>λ22(D2)}.

    Moreover, we have

    R+2={(s1,s2)R2+:λ12(D2)<s2<λ22(D2)}.

    What is new compared to the C1 case (see Figure 2, Case C1) is that, since ZNGI2 has two components, we have a multiple crossing of ZNGI2 with ZNGI1 and ZNGI2 with FSB1. The system can have up to six equilibria, see Figure 11. The existence and stability conditions of these equilibria are summarized in Table 5.

    Figure 11.  The feasible set and the ZNGIs for the commensalistic Case C12 and the syntrophic Case S12.
    Table 5.  Conditions of existence and stability of the equilibria of (4.1) or (4.5).
    Existence condition Stability condition (local)
    E0 Always exists E0R1R2
    E1 E0R+1 E1R2
    E12 E0R+2¯R22 E12R1
    E22 E0R22 Unstable if it exists
    E1c E1cFo Stable if it exists
    E2c E2cFo Unstable if it exists

     | Show Table
    DownLoad: CSV

    To construct the operating diagram, we consider

    ● the vertical line Γ1 of equation sin1=λ1(D1),

    ● the horizontal lines Γj2 of equations sin2=λj2(D2), j=1,2,

    ● the oblique lines Γj3 of equations sin1+sin2=λ1(D1)+λj2(D2), j=1,2.

    These lines divide the set of operating parameters (sin1,sin2) in nine regions denoted Ik, k=0,,8, as depicted in Figure 12 (Case C12).

    Figure 12.  The operating diagram of (4.1) or (4.5) in the (sin1,sin2) operating plane and D(0,δ0). The existence and stability of the equilibria in the regions Ik are given in Table 6 for the Case C12 and Table 7 for the Case S12.

    Proposition 7. The conditions of existence and stability of the equilibria of (4.1) in the regions Ik of Figure 12 (Case C12) are given in Table 6.

    Table 6.  Existence and stability of equilibria of (2.5) in the regions of the operating diagram depicted in Figure 12 (Case C12).
    E0 E12 E22 E1 E1c E2c Color
    I0 GAS Red
    I1 U GAS Blue
    I2 S S U Cyan
    I3 U GAS Yellow
    I4 U U GAS Green
    I5 U S S U Pink
    I6 U U U GAS Green
    I7 U U S S U Pink
    I8 U U U S S U Pink

     | Show Table
    DownLoad: CSV

    Proof. The proof is given in Figure 13. Assume that (sin1,sin2)I8. We see in Figure 13 that E0 is unstable since E0R1R2, E1 exists since E0R+1 and is stable since E1R2, E12 and E22 exist since E0R22, and E12 is unstable since E12R1. Moreover, E1c and E2c exist, E1c is stable, and E2c is unstable. The proofs for all other regions are similar. The proof for global asymptotic stability is given in Appendix A.1.

    Figure 13.  Proof of Proposition 7: feasible sets and ZNGIs, showing equilibria and their stability, for the nine regions of the operating diagram, shown in Figure 12 (Case C12).

    For a more detailed study of model C12 and information on how the operating diagram changes when the biological parameters are changed, as well as operating diagrams in the operating plane (sin1,D), where sin2 is kept constant, we refer the reader to [31, 39].

    We consider the Case S12 in Figure 10. The model takes the form

    ˙s1=D(sin1s1)f1(s1,s2)x1,˙x1=(f1(s1,s2)D1)x1,˙s2=D(sin2s2)+f1(s1,s2)x1f2(s2)x2,˙x2=(f2(s2)D2)x2. (4.5)

    We assume that f1(s1,s2) and f2(s2) satisfy (3.9) and (4.3), respectively.

    The break-even concentrations λ1(s2,D), λ12(D), and λ22(D) of (4.5) are defined by Definitions 1 and 2, respectively. The ZNGI1 of (4.1) is given by ZNGI1={(s1,s2)R2+:s1=λ1(s2,D1)}, while ZNGI2 is given by (4.4). As for the commensalistic case, the system can have up to six equilibria, see Figure 11, whose conditions of existence and stability are also given by Table 5.

    What is new compared to the C12 case is that the vertical line Γ1 of the commensalistic model becomes now the curve with equation sin1=λ1(sin2,D1). The horizontal lines Γj2 are not changed, and the oblique lines Γj3 now have the equations sin1+sin2=λ1(λj2(D2),D1)+λj2(D2), j=1,2. We must also consider the vertical lines Γj4 with equations s1=λ1(λj2(D2),D1), j=1,2. All these lines divide the set of operating parameters (sin1,sin2) into twelve regions denoted Ik, k=0,,11, depicted in Figure 12 (Case S12).

    Proposition 8. The existence and stability of the equilibria of (4.5) in the regions Ik, k=0,,11 of Figure 12 (Case S12) are given in Table 7.

    Table 7.  Conditions of existence and stability of equilibria of (2.5) in the regions of the operating diagram depicted in Figure 12 (Case S12).
    E0 E12 E22 E1 E1c E2c Color
    I0 S Red
    I1 U S Blue
    I2 S S U Cyan
    I3 U S Yellow
    I4 U U S Green
    I5 U S S U Pink
    I6 U U U S Green
    I7 U U S S U Pink
    I8 U U U S S U Pink
    I9 U U S Green
    I10 S U U S White
    I11 S U U S U White

     | Show Table
    DownLoad: CSV

    Proof. The proof for the regions Ik, k=0,,8 is the same as the proof of Proposition 7 and is given in Appendix C.1. The proof for the new regions Ik, k=9,10, and 11, which do not exist in the commensalistic case, are given in Figure 14. Assume that (sin1,sin2)I11. We see in Figure 14 that E0 is stable since E0R1R2, and E1 does exist since E0R+1. Moreover, E12 and E22 exist since E0R22, and E12 is unstable since E12R1. On the other hand, E1c and E2c exist, E1c is stable, and E2c is unstable. The proofs for the other regions is similar.

    Figure 14.  Proof of Proposition 8 in the regions Ik, k=9,10,11 of Figure 12 (Case S12).

    The most important novelty on the asymptotic behavior of S12 compared to C12 is the appearance of the white regions I10 and I11 of bistability of the washout equilibrium E0 and the coexistence equilibrium E1c. For a more detailed study of model S12 and information on how the operating diagram changes when the biological parameters are changed, as well as operating diagrams in the operating plane (sin1,D), where sin2 is kept constant, we refer the reader to [29].

    In the previous section, we determined the operating diagram when the operating parameter D is kept constant and the operating parameters sin1 and sin2 vary. For the results to be useful in practice, it is necessary to determine the operating diagram when D is one of the operating parameters that can vary, as this operating parameter is the most commonly used control parameter in laboratories. Since sin1 is the inlet concentration of the substrate of the overall system, a useful representation is to keep constant the inlet concentration sin2 of the substrate s2 produced by the first reaction and to describe the operating diagram in the operating plane (sin1,D). The effects of sin2 are shown in a series of operating diagrams.

    The construction of this diagram can be easily deduced from the construction of the operating diagram in the operating plane (sin1,sin2), and D is fixed. The study carried out in the plane (sin1,sin2) showed the existence of regions such that the system behaves in a certain way when the operating parameters are chosen in these regions. It is then sufficient to see how the boundaries of these regions, i.e., the Γi and Γji curves, are written in the operating plane (sin1,D) and which regions of this plane they delimit. In the following sections we illustrate this construction in models C1, C2, and S1, as well as in a typical example of model S2.

    Figure 15 shows the regions Ik, previously identified in Figure 4, in the operating plane (sin1,D), where sin2 is kept constant. The number of regions depends on sin2. Let us describe the operating diagram in the case δ1>δ2, where δ1 and δ2 are defined by (3.16). The case δ1δ2 is similar and is left to the reader.

    Figure 15.  The operating diagram of Cases C1, C2, and S1, in the operating plane (sin1,D), when δ1>δ2, where δi are defined by (3.16).

    Note that for C1, Γ1 is the curve with equation D=(f1(sin1)a1)/α1 and Γ2 is the horizontal line with equation D=(f2(sin2)a2)/α2, which is not empty if and only if sin2>λ2(a2). On the other hand, Γ3 is the curve with equation sin1+sin2=λ1(D1)+λ2(D2). The novelty when we consider model C2 is that Γ2 becomes the curve of equation f2(sin1,sin2)=D2. This curve is not empty if and only if f2(0,sin2)>a2, which is equivalent to sin2>λ2(0,a2). The novelty when we consider model S1 is that the curve Γ1 becomes the curve with equation sin1=λ1(sin2,D1), and is therefore distinct from Γ4, which is the curve with equation sin1=λ1(λ2(D2),D1). For the study of the intersection point of Γi curves of the model C2, we need to define the following function: D(0,δ2)ξ(D):=λ2(λ1(D1),D2). Using (3.9) and (3.10), we see that ξ(D)>0. Hence, ξ is an increasing function from ξ(0)=λ2(λ1(a1),a2) to ξ(δ2)=+. It then has an inverse function ξ1. The curves Γi intersect at P(sin2) defined by

    P(sin2)={(λ1(α1D0+a1),D0)where D0=f2(sin2)a2α2 for model C1,(λ1(α1D0+a1),D0)where D0=ξ1(sin2) for model C2,(λ1(sin2,α1D0+a1),D0)where D0=f2(sin2)a2α2 for model S1.

    Therefore, for C1 and sin2>λ2(a2), the five regions Ik, k=0,,4, appear as shown in Figure 15. Similarly for C2 and sin2>λ2(λ1(a1),a2), the five regions Ik appear, while for S1 and sin2>λ2(a2), the sixth region I5 appears, see Figure 15 (Case S1).

    If 0sin2λ2(a2), for models C1 and S1, only the I0, I1, and I3 regions appear, as depicted in Figure 15 (Cases C1 and S1). For C2, two cases must be distinguished: If 0sin2λ2(0,a2), only the I0, I1, and I3 regions appear, as shown in Figure 15 (Case C2). If λ2(0,a2)<sin2λ2(λ1(a1),a2), a small I2 region also appears close to the origin, as shown in Figure 16.

    Figure 16.  An enlargement near the origin of the operating diagram in Figure 15 (Case C2), showing the appearance of the I2 region, when sin2>λ2(0,a2).

    We consider the model S2 of Table 1. The curves Γ1, Γ2, Γi3, and Γi4, i=1,2, which are the boundaries of the various regions, have been derived analytically. Indeed, since sin2 is kept fixed, and using (2.2), we have the following equations for these curves:

    Γ1 is the curve with equation D=f1(sin1,sin2)a1α1.

    Γ2 is the curve with equation D=f2(sin1,sin2)a2α2.

    Γi3 and Γi4, i=1,2, as defined by (3.17), are also curves in the operating plane (sin1,D).

    In addition to these curves we must also consider the horizontal line defined by

    Γ0={(sin1,D)R2+:D=δ0}, (5.1)

    where δ0 is the value of D for which the ZNGIs are tangent, see Assumption H3 and Figure 9. These curves divide the operating plane (sin1,D) into nine regions corresponding to the nine regions depicted in Figures 7 and 9. Therefore, the operating diagrams can be drawn by plotting all these curves.

    It is not possible to give a general qualitative description of the operating diagram as we did in Section 3.2.1 for models C1, C2, and S1. In Figures 17 and 18, we present the specific example given in Table 4. Increasing sin2 from sin2=0 to sin2=7, we draw the curves Γj, j=0,1,2, Γi3, and Γi4, i=1,2, and color the Ik regions they delimit with the colors already used in Figure 7.

    Figure 17.  The operating diagram of (2.5), for the growth functions given in Table 4, in the (sin1,D) operating plane, with sin2 kept constant.
    Figure 18.  The operating diagram of (2.5), for the growth functions given in Table 4, in the (sin1,D) operating plane, with sin2 kept constant.

    We briefly explain how these figures are obtained. For sin2=0, the curves Γ2, Γ14, and Γ24 are empty and only regions I0, I1, I3, and I6 appear, as shown in the panel sin2=0 in Figure 17. When sin2=1, the curves Γ2 and Γ14 are not empty and the regions I2, I4, and I5 also appear. See the panel sin2=1 and its enlargement in Figure 17. If sin2 is increased to s2(δ0), corresponding to the tangency of the ZNGIs, then no new region appears. If sin2 is increased further, then the curve Γ24 appears, while the curve Γ13 disappears, defining two new regions I7 and I8. Therefore, the nine regions of the operating diagram in Figure 7 appear, as shown in the panel sin2=4, and its enlargement, in Figure 17 and the panel sin2=5, in Figure 18. If sin2 is increased to sin2=6.315, then the nine regions continue to appear. The value sin2=6.315 corresponds to the tangency of the curves Γ0 and Γ1, and is given by the equation f1(+,sin2)=α1δ0+a1. If sin2 is increased further, then the region I1 disappears, see the panel sin2=7 in Figure 18.

    In this work we have studied the general model (2.5) of commensalism or syntrophy. This model contains a large number of models in the existing literature. The case where f1 does not depend on s2, or f2 does not depend on s1, were studied in the literature when the removal rates Di are not equal to the dilution rate D, see [12, 28, 29, 39, 40]. However, the case where f1 depends also on s2, and f2 depends also on s1 were studied in the literature only when D1=D2=D so that the system can be reduced to a planar system, see [11]. Our mathematical analysis of the model has revealed several possible behaviors: Proposition 3 provides a complete theoretical description of asymptotic behavior of the system. In order that the results can be useful in practice, one must have a description of the system with respect to the operating parameters. The study of bifurcations according to the operating parameters D, sin1, and sin2 is the most meaningful one for the laboratory model since the experimenter can easily vary these parameters.

    Our results contain all the results from the literature as special cases, and more importantly present them in a unified way. The advantage of this unified presentation is that it also shows the similarities between the models and emphasizes the new behaviors that can emerge when new assumptions are introduced. For example:

    ● Extending the model from Case C1 (i.e., fi(si) depends only on si, i=1,2) to Case C2 (i.e., f2(s1,s2) is allowed to depend on s1 as well) introduces no new system behaviors: the same number of equilibria, the same asymptotic behaviors. The only modification is that the regions of the operating diagram are slightly modified, see Figure 4 (Cases C1 and C2) and Table 2.

    ● However, extending the model from Case C1 to Case S1 (i.e., f1(s1,s2) is allowed to depend on s2 as well) introduces new system behaviors: although the system retains the same number of equilibria, a new region appears in the operating diagram, see Figure 4 (Case S1) and Table 3.

    ● On the other hand, extending the model from the C2 case to the S2 case (i.e., f1(s1,s2) is allowed to depend on s2 as well and f2(s1,s2) is allowed to depend on s1 as well) introduces the most novelties in the system's behavior, the most important being the possibility of multiple coexistence equilibria, as well as a variety of asymptotic system behavior, including bistability, see Figure 7.

    ● When self-inhibition is allowed in the model, our graphical method makes it easy to compare the system's asymptotic behaviors and to highlight the contribution of syntrophy (Case S12) versus commensalism (Case C12). Figure 12 and Table 7 show how syntrophy brings out the possibility of bistability between the washout equilibrium E0 and the coexistence equilibrium E1c (see the white regions). This bistability does not occur in the commensal (Case C12) model, as shown in Figure 12 and Table 6.

    In the case without self-inhibition, the bistability phenomenon necessarily requires that the function f1 depends on s2 and the function f2 depends on s2. Indeed, the bistability of a positive equilibrium and a boundary equilibrium, where one of the species (or both) is extinct, is possible only for the S2 model and not for the S1, C1, or C2 models. If one of the species is unaffected by the food of its partner, the system is unable to undergo bistability.

    The growth functions KB1 and KB2 of Kreikenbohm and Bohl [9, 41] do not satisfy our assumptions since they are not C1, but only piecewise C1, see Table A7. It should be interesting to extend our graphical method to those cases. Moreover, the KB1 or KB2 functions are identically 0 until a threshold value and then become positive. Such functions often appear in biological literature since the growth of a species requires that the limiting substrate exceeds a certain threshold. The extension of our method in these cases, will be considered in a future work.

    As we pointed out in Remark 1, Di and Yang [2] also considered models of the form (2.1) with inhibition by the other species, in which, for example, μ1(S1,S2) is replaced by μ1(S1,x2) and μ2(S1,S2) is replaced by μ2(x1,S2), where μi is increasing in Si and decreasing in xj, for i,j=1,2 (ji). Such so-called density-dependent growth functions have been considered in the context of competition between species for a single limiting substrate, see [20, Chapter 4] and [36, 37, 38, 42]. When considering the combined effects of several interacting substrates and species density dependence, Tilman's graphical method developed in this article may prove useful in clarifying and unifying the many studies on the subject.

    The author declares he have not used Artificial Intelligence (AI) tools in the creation of this article.

    The author would like to thank Radhouane Fekih-Salem and Claude Lobry for their remarks which improved the presentation considerably. The author thanks the TREASURE Euro-Mediterranean research network for partial financial support. This work was presented as part of the CIMPA summer school Digital Green: mathematical biology and theoretical ecology. The author would like to thank the three anonymous reviewers for their careful reading and constructive comments.

    The author declares that he have no competing interest.

    Local stability analysis only implies that the solutions starting near an asymptotically stable equilibrium converge toward this equilibrium. Hence, one cannot make assertions about the eventual outcome that are global in the sense that they are independent of the initial conditions. However, in some cases, one can reduce the four-dimensional system (2.5) to a two-dimensional system, whose global study is in general more easy. Thanks to Thieme's theory [43, 44], we can deduce the global asymptotic stability of the initial four-dimensional system from the global asymptotic stability of the reduced two-dimensional one. For details and complements on how to use Thieme's theory, see [20, Appendix A] or [19, Appendix F]. This reduction is possible for the commensalistic models in the general case when the removal rates are not equal to the dilution rate and also in the syntrophic models when they are (i.e., D1=D2=D).

    We consider the commensalistic model C212 in Figure 10. System (2.5) becomes

    ˙s1=D(sin1s1)f1(s1)x1,˙x1=(f1(s1)D1)x1,˙s2=D(sin2s2)+f1(s1)x1f2(s1,s2)x2,˙x2=(f2(s1,s2)D2)x2, (A.1)

    where f1 is not assumed to be necessarily increasing in s1, and similarly f2 is not assumed to be necessarily increasing in s2. This system contains all commensalistic models C1, C2, C11, C12, C21, C22, and C112 as particular cases. For example, C12 is obtained when f1 satisfies (4.2) and f2 depends only on s2 and satisfies (4.3). The important thing is that system (A.1) has a cascade structure. Indeed, if (s1(t),x1(t),s2(t),x2(t)) is a solution of (A.1), then (s1(t),x1(t)) is a solution of the two-dimensional system

    ˙s1=D(sin1s1)f1(s1)x1,˙x1=(f1(s1)D1)x1, (A.2)

    and (s2(t),x2(t)) is a solution of the non autonomous two-dimensional system

    ˙s2=D(sin2s2)+f1(s1(t))x1(t)f2(s1(t),s2)x2,˙x2=(f2(s1(t),s2)D2)x2. (A.3)

    System (A.2) is a classical chemostat system. Every solution (except for a set of initial conditions of measure 0) converges toward an equilibrium (s1,x1), possibly the washout equilibrium (sin1,0). Therefore, system (A.3) is an asymptotically autonomous system whose limiting system is

    ˙s2=D(sin2s2)+f1(s1)x1f2(s1,s2)x2,˙x2=(f2(s1,s2)D2)x2. (A.4)

    System (A.4) is also a classical chemostat system and its solutions (except for a set of initial conditions of measure 0) converge toward an equilibrium (s2,x2). Using Thieme's theory we can conclude on the global asymptotic stability of the system (A.1). This proves the global asymptotic behavior shown in Tables 2 and 6. For details and complements on how these kinds of arguments are conducted, we refer the reader to [31, 39] for the C12 case in Figure 10.

    In this section, we consider the case where D1=D2=D. System (2.5) becomes

    ˙s1=D(sin1s1)f1(s1,s2)x1,˙x1=(f1(s1,s2)D)x1,˙x2=(f2(s1,s2)D)x2,˙s2=D(sin2s2)+f1(s1,s2)x1f2(s1,s2)x2. (A.5)

    We consider the change of variables z1=s1+x1 and z2=s2x1+x2. System (A.5) becomes

    ˙x1=(f1(z1x1,z2+x1x2)D)x1,˙x2=(f2(z1x1,z2+x1x2)D)x2,˙z1=D(sin1z1),˙z2=D(sin2z2). (A.6)

    Since z1(t) and z2(t) exponentially converge toward sin1 and sin2, respectively, (A.6) is an asymptotically autonomous system whose limiting system is

    ˙x1=(ϕ1(x1,x2)D)x1,˙x2=(ϕ2(x1,x2)D)x2, (A.7)

    where ϕ1 and ϕ2 are defined by

    ϕ1(x1,x2)=f1(sin1x1,sin2+x1x2),
    ϕ2(x1,x2)=f2(sin1x1,sin2+x1x2).

    Using Thieme's theory, the asymptotic behaviour of the solutions of the reduced model (A.7) is informative for the complete system (A.6). For details and complements on how these kinds of arguments are conducted, we refer the reader to [11] for the S2 model when D1=D2=D.

    Along Γ0, defined by (5.1), i.e., when D=δ0, a saddle-node bifurcation, in which E1c and E2c collide and annihilate each other, can occur, see Figures 9, 17, and 18. The transcritical bifurcations occurring along Γ1, Γ2, Γj3, and Γj4, j=1,2, are summarized in Table A1. The result is a straightforward consequence of Proposition 3.

    Table A1.  Codimension-one bifurcations of equilibria along the boundaries of the Ik regions. Only transcritical bifurcations (TB) and saddle-node bifurcations (SNB) occur.
    Boundary Bifurcation
    Γ0=(¯I1¯I6)(¯I2¯I8)(¯I0¯I7) E1c=E2c (SNB)
    Γ1=(¯I0¯I1)(¯I4¯I5)(¯I6¯I7) E0=E1 (TB)
    Γ2=(¯I0¯I2)(¯I3¯I4)(¯I7¯I8) E0=E2 (TB)
    Γ13=¯I1¯I3 E1=E1c (TB)
    Γ23=¯I3¯I6 E1=E2c (TB)
    Γ14=¯I2¯I5 E2=E1c (TB)
    Γ24=¯I5¯I8 E2=E2c (TB)

     | Show Table
    DownLoad: CSV

    We illustrate bifurcations by constructing in Figure A1 the one-parameter bifurcation diagram, showing the equilibria as a function of the parameter D when sin1 and sin2 are fixed. We use the growth functions given in Table 4 and take sin1=10, while the value for sin2 is given in the figure. The behavior of the system is easily deduced from the operating diagrams shown in Figures 17 and 18. Indeed, it corresponds to the behavior on the vertical line sin1=10 of these diagrams. For example, if we take sin2=4, we see in Figure 17 (panel sin2=4) that there exist four bifurcation values d0<d1<d2<d3 such that:

    Figure A1.  The bifurcation diagram of (2.5) showing the s2-component of the equilibria as a function of D for the growth functions given in Table 4 and sin1=10. An equilibrium is drawn as a bold line when it is stable and as a dotted line when it is unstable. Here, d2=δ01.602. For sin2=0, we have d11.335 and d34.420. For sin2=4, we have d0=2/3, d11.130 and d31.941. For sin2=7, we have d0=0.75, d11.002, and d31.341.

    ● If D>d3, then (sin1=10,D)I0. Hence, E0 is the only equilibrium and is stable.

    ● If d3>D>d2, then (sin1=10,D)I1. Hence, E0 and E1 are the only equilibria, E1 is stable, and E0 unstable.

    ● At D=d3, a transcritical bifurcation occurs in which E0 and E1 collide and exchange stability.

    ● If d2>D>d1, then (sin1=10,D)I6. Hence, E0, E1, E1c, and E2c are the equilibria, E1 and E1c are stable, and E0 and E2c are unstable.

    ● At D=d2, a saddle-node bifurcation occurs in which E1c and E2c collide and annihilate.

    ● If d1>D>d0, then (sin1=10,D)I3. Hence, E0, E1, and E1c are the only equilibria, E1c is stable, and E0 and E1 are unstable.

    ● At D=d1, a transcritical bifurcation occurs in which E1 and E2c collide and E1 becomes unstable.

    ● If d0>D>0, then (sin1=10,D)I4. Hence, E0, E1, E2, and E1c are the equilibria, E1c is stable, and E0, E1, and E2 are unstable.

    ● At D=d0, a transcritical bifurcation occurs in which E2 and E0 collide and E2 disappears.

    Recall that an equilibrium is said to be stable if it is locally exponentially stable, i.e., the Jacobian matrix has eigenvalues with strictly negative real parts. We begin by proving the following result.

    Proposition 9. System (2.5) can have four types of equilibria:

    1) The washout equilibrium E0=(sin1,0,sin2,0), which always exists. It is stable if and only if

    f1(sin1,sin2)<D1andf2(sin1,sin2)<D2. (C.1)

    2) A boundary equilibrium E1=(ˉs1,ˉx1,ˉs2,0), where ˉs1 is a solution of the equation

    f1(s1,sin1+sin2s1)=D1, (C.2)

    and

    ˉs2=sin1+sin2ˉs1,ˉx1=DD1(sin1ˉs1). (C.3)

    It is unique if it exists. It exists if and only if

    f1(sin1,sin2)>D1. (C.4)

    It is stable if and only if

    f2(ˉs1,ˉs2)<D2. (C.5)

    3) A boundary equilibrium E2=(˜s1,0,˜s2,˜x2), where

    ˜s1=sin1,˜s2=λ2(sin1,D2),˜x2=DD2(sin2˜s2). (C.6)

    It exists if and only if

    f2(sin1,sin2)>D2. (C.7)

    It is stable if and only if

    f1(˜s1,˜s2)<D1. (C.8)

    4) Coexistence equilibria Ec=(s1,x1,s2,x2), where (s1,s2) is a solution of the system of equations

    f1(s1,s2)=D1,f2(s1,s2)=D2, (C.9)

    and

    x1=DD1(sin1s1),x2=DD2(sin1+sin2s1s2). (C.10)

    It exists if and only if (s1,s2)Fo. It is stable if and only if

    (f11f22f12f21)(s1,s2)>0. (C.11)

    Proof. We begin by the conditions for the existence of equilibria. Using Lemma 2, for a boundary equilibrium E1=(ˉs1,ˉx1,ˉs2,0), we have (ˉs1,ˉs2)=ZNGI1FSB2. This condition is equivalent to

    f1(ˉs1,ˉs2)=D1 and ˉs1+ˉs2=sin1+sin2. (C.12)

    From the second formula in (C.12), we have ˉs2=sin1+sin2ˉs1, which is the first formula in (C.3). Replacing ˉs2 in the first formula of (C.12), we have

    f1(ˉs1,sin1+sin2ˉs1)=D1.

    Therefore, ˉs1 is a solution of Eq (C.2). The x1 component is then given by (3.1), which proves the second formula in (C.3). Eq (C.2) is equivalent to ψ1(s1)=D1, where

    ψ1(s1)=f1(s1,sin1+sin2s1).

    We have

    ψ1(s1)=(f11f12)(s1,sin1+sin2s1).

    Using (3.9), we have ψ1(s1)>0 for all s1>0. Therefore, Eq (C.2) has at most one solution. Hence, if it exists, E1 is unique. E1 exists if and only if equation ψ1(s1)=D1 has a solution in the interval (0,sin1). Since ψ1(0)=0 and ψ1(sin1)=f1(sin1,sin2), the solution exists if and only if f1(sin1,sin2)>D1, which proves (C.4).

    Using Lemma 2, for a boundary equilibrium E2=(˜s1,0,˜s2,˜x2) we have (˜s1,˜s2)=FSB1ZNGI2. This condition is equivalent to

    ˜s1=sin1 and f2(˜s1,˜s2)=D2. (C.13)

    The first formula in (C.13) is the first formula in (C.6). Replacing ˜s1 in the second formula of (C.13), we have f2(sin1,˜s2)=D2. Therefore, using Definition 1 we have ˜s2=λ2(sin1,D2), which proves the second formula in (C.6). The x2 component is then given by (3.1), which proves the third formula in (C.6). E2 exists if and only if the equation f2(sin1,s2)=D2 has a solution in the interval (0,sin2). Since f2(sin1,0)=0, the solution exists if and only if f2(sin1,sin2)>D2, which proves (C.7).

    Using Lemma 2, for Ec=(s1,x1,s2,x2) we have (s1,s2)ZNGI1ZNGI2Fo. This condition is equivalent to

    f1(s1,s2)=D1,f2(s1,s2)=D2, and (s1,s2)Fo.

    Therefore, (s1,s2) is a solution of (C.9), lying in the interior Fo of the feasible set. The x1 and x2 components are then given by (3.1), which proves (C.10).

    This ends the proof of the existence conditions in the proposition. The local stability of an equilibrium point of (2.5) depends on the sign of the real parts of the eigenvalues of the corresponding Jacobian matrix of system (2.5). The Jacobian matrix is the matrix of the partial derivatives of the right-hand side of system (2.5), with respect to the state variables, evaluated at the given equilibrium point (x1,x2,s1,s2):

    J=[f1D10f11x1f12x10f2D2f21x2f22x2f10Df11x1f12x1f1f2f11x1f21x2D+f12x1f22x2], (C.14)

    where fi and fij=fisj are evaluated at the components of the equilibrium point.

    At E0, x1=0 and x1=0. Hence, the Jacobian matrix (C.14) becomes

    J0=[f1(sin1,sin2)D10000f2(sin1,sin2)D200f1(sin1,sin2)0D0f1(sin1,sin2)f2(sin1,sin2)0D].

    Its eigenvalues are f1(sin1,sin2)D1, f2(sin1,sin2)D2 and D. Therefore, E0 is stable if and only if f1(sin1,sin2)<D1 and f2(sin1,sin2)<D2 which proves (C.1).

    At E1, x2=0 and x1>0 so that f1=D1. Evaluated at E1, the Jacobian matrix (C.14) becomes

    J1=[00f11(¯s1,¯s2)x1f12(¯s1,¯s2)x10f2(¯s1,¯s2)D200D10Df11(¯s1,¯s2)x1f12(¯s1,¯s2)x1D1f2(¯s1,¯s2)f11(¯s1,¯s2)x1D+f12(¯s1,¯s2)x1].

    Its characteristic polynomial is P1(λ)=(λ+D)(λf2(¯s1,¯s2)+D2)(λ2+c1λ+c2), where c1=D+(f11f12)(¯s1,¯s2)x1 and c2=D1(f11f12)(¯s1,¯s2)x1. The eigenvalues of J1 are D and f2(¯s1,¯s2)D2, together with the roots of the quadratic polynomial in P1(λ). Since f11>0 and f120, one has c1>0 and c2>0. Hence, the roots of the quadratic polynomial have negative real parts. Therefore, E1 is stable if and only if f2(¯s1,¯s2)<D2, which proves (C.5).

    At E2, x1=0 and x2>0 so that f2=D2. Evaluated at E2, the Jacobian matrix (C.14) becomes

    J2=[f1(˜s1,˜s2)D100000f21(˜s1,˜s2)x2f22(˜s1,˜s2)x2f1(˜s1,˜s2)0D0f1(˜s1,˜s2)D2f21(˜s1,˜s2)x2Df22(˜s1,˜s2)x2].

    Its characteristic polynomial is P2(λ)=(λ+D)(λf1(˜s1,˜s2)+D1)(λ2+c1λ+c2), where c1=D+f22(˜s1,˜s2)x2 and c2=D2f22(˜s1,˜s2)x2. The eigenvalues of J2 are D and f1(˜s1,˜s2)D1, together with the roots of the quadratic polynomial in P2(λ). Since f22>0, one has c1>0 and c2>0. Hence, the roots of the quadratic polynomial have negative real parts. Therefore, E2 is stable if and only if f1(˜s1,˜s2)<D1, which proves (C.8).

    At Ec, x1>0 and x2>0 so that f1=D1 and f2=D2. Evaluated at Ec, the Jacobian matrix (C.14) becomes

    Jc=[00f11x1f12x100f21x2f22x2D10Df11x1f12x1D1D2f11x1f21x2D+f12x1f22x2],

    where fi and fij are evaluated at (s1,s2). Its characteristic polynomial is Pc(λ)=λ4+c1λ3+c2λ2+c3λ+c4, where

    c1=2D+(f11f12)x1+f22x2,c2=D2+(D+D1)(f11f12)x1+(D+D2)f22x2+(f11f22f12f21)x1x2,c3=DD1(f11f12)x1+DD2f22x2+(D1+D2)(f11f22f12f21)x1x2,c4=D1D2(f11f22f12f21)x1x2.

    The eigenvalues of Jc have negative real parts if and only if the Routh-Hurwitz conditions

    c1>0,c3>0,c4>0 and r1=c1c2c3c21c4c23>0, (C.15)

    are satisfied. We use the following notations:

    A=f22,B=f11f22f12f21f22,C=f12(f21f22)f22.

    Using (3.9) and (3.10), we have A>0, C0, and B+C=f11f12>0. The coefficients ci can be written as follows:

    c1=2D+(B+C)x1+Ax2,c2=D2+(D+D1)(B+C)x1+(D+D2)Ax2+ABx1x2,c3=DD1(B+C)x1+DD2Ax2+(D1+D2)ABx1x2,c4=D1D2ABx1x2.

    Note that c1>0. The condition c4>0 is equivalent to B>0. If B>0, then we have c3>0. Straightforward computations show that r1 can be written r1=pq+r, where

    p=(D1Bx1D2Ax2)2,q=D2+D(Bx1+Ax2)+ABx1x2,

    and

    r=A2B2(B+C)(D1+D2)x31x22+A3B2(D1+D2)x21x32+AB(D(2D1+D2)(B+C)2+CD21(2B+C))x31x2+A2B(D(D1+D2)(5B+3C)+C(D21+D22))x21x22+A3BD(D1+2D2)x1x32+DD1(D(B+C)3+D1(C3+3BC2+3B2C))x31+AD((B+C)((7D1+4D2)B+C(2D1+D2))D+CD1((D1+2D2)C+2BD1))x21x2+A2D(D(B(4D1+7D2)+C(D1+2D2))+CD2(2D1+D2))x1x22+A3D2D2x32+D2D1(3D(B+C)2+D1C(2+C))x21+3D3A2D2x22+AD2(D(D1+D2)(5B+3C)+2CD1D2)x1x2+2D4D1(B+C)x1+2D4AD2x2.

    Hence, r1>0 if B>0. Therefore, the conditions (C.15) are satisfied if and only if B>0, which is equivalent to (f11f22f12f21)(s1,s2)>0. This proves (C.11).

    The conditions of existence and stability of equilibria are summarized in Table A2. Let us now prove Proposition 3.

    Table A2.  Conditions of existence and stability of the equilibria of (2.5).
    Existence condition Stability condition (local)
    E0 Always exists f1(sin1,sin2)<D1 and f2(sin1,sin2)<D2
    E1 f1(sin1,sin2)>D1 f2(ˉs1,ˉs2)<D2
    E2 f2(sin1,sin2)>D2 f1(˜s1,˜s2)<D1
    Ec (s1,s2)Fo (f11f22f12f21)(s1,s2)>0

     | Show Table
    DownLoad: CSV

    Proof of Proposition 3. The conditions f1(sin1,sin2)<D1 and f1(sin1,sin2)<D2 of stability of E0 are equivalent to E0R1R2. The condition f1(sin1,sin2)>D1 of existence of E1 is equivalent to E0R+1. Its condition f2(ˉs1,ˉs2)<D2 of stability is equivalent to E1R2. The condition f2(sin1,sin2)>D2 of existence of E2 is equivalent to E0R+2. Its condition f1(˜s1,˜s2)<D1 of stability is equivalent to E2R1. The condition of existence of Ec is (s1,s2)Fo. Using λ1/s2=f12/f11 and λ2/s1=f21/f22, the condition (f11f22f12f21)(s1,s2)>0 of stability of Ec is equivalent to

    λ1s2(s1,s2)λ2s1(s1,s2)<1.

    This condition means that the signed angle between between the tangent of ZNGI1 and the tangent of ZNGI2 at the point of intersection (s1,s2) is negative.

    The proof is given in Figure A2. Assume that (sin1,sin2)I4. We see in Figure A2 that E0 is unstable since E0R1R2, and E1 exists since E0R+1 and is unstable since E1R2. Moreover, E2 exists since E0R+2 and is unstable since E2R1. On the other hand, Ec exists, is unique, and is stable since, at Ec, (ZNGI1,ZNGI2)<0. The proof for global asymptotic stability is given in Section 6. This proves the results depicted in the last row of Table 2. The proofs for the other regions are illustrated in Figure A2.

    Figure A2.  Proof of Proposition 4 in Case C2: Feasible sets and ZNGIs, showing equilibria and their stability, for the five regions of the operating diagram, shown in Figure 4 (Case C2).

    The proof is given in Figure A3. Assume that (sin1,sin2)I5. We see in Figure A3 that E0 is unstable since E0R1R2, E1 does not exist since E0R+1, and E2 exists since E0R+2, and is unstable since E2R1. Moreover, Ec exists, is unique and is stable since, at Ec, (ZNGI1,ZNGI2)<0. This proves the results depicted in the last row of Table 3. The proofs for the other regions are illustrated in Figure A3. When D1=D2=D, the proof for global asymptotic stability is given in Appendix A.2.

    Figure A3.  Proof of Proposition 5: Feasible sets and ZNGIs, showing equilibria and their stability, for the six regions of the operating diagram, shown in Figure 4 (Case S1).

    The proof is given in Figure A4. Assume that (sin1,sin2)I8. We see in Figure A4 that E0 is unstable since E0R1R2, E1 exists since E0R+1 and is stable since E1R2, E12 and E22 exist since E0R22, E12 is unstable since E12R1, and E1c and E2c exist. The proofs for all other regions are similar.

    Figure A4.  Proof of Proposition 8: Feasible sets and ZNGIs, showing equilibria and their stability, for the nine regions of the operating diagram, shown in Figure 12 (Case S12).

    In this section, we review the main results of the existing literature concerning the systems described in Figures 1 and 10 and briefly summarize their contributions.

    Table A3 presents the main studies of the commensalistic models C1 and C2, while Table A4 presents the studies for the syntrophic models S1 and S2. The studies for the model including self-inhibition are presented in Tables A5 and A6. The growth functions used in these tables are defined in Table A7. For details and complements, the reader can consult the review papers [12, 29, 45].

    Table A3.  Models of commensalistic relationship. The growth functions μ1 and μ2 are defined in Table A7.
    μ1 μ2 Sin2 Di Year Ref. Case
    Monod Monod 0 D 1974 Reilly [6] C1
    M-function M-function 0 D 1981 Stephanopoulos [7] C1
    Monod Monod 0 D+ai 2003 Simeonov and Stoyanov [40] C1
    Monod Monod 0 D 2019 Di and Yang [2] C1
    Monod I-Monod 0 D 2019 Di and Yang [2] C2
    M-function I-M-function 0 D 2019 Ben Ali [57] C2

     | Show Table
    DownLoad: CSV
    Table A4.  Models of syntrophic relationship. The growth functions μ1 and μ2 are defined in Table A4.
    μ1 μ2 Sin2 Di Year Ref. Case
    Monod-I Monod 0 D 1974 Wilkinson et al. [13] S1
    KB1 Monod 0 D 1986 Kreikenbohm and Bohl [9] S1
    M-I-function M-function 0 D 1994 Burchard [8] S1
    Monod-I Monod 0 D+ai 2011 Xu et al. [14] S1
    M-I-function M-function 0 D 2010 El Hajji et al. [10] S1
    M-I-function M-function 0 D+ai 2016 Sari and Harmand [12] S1
    M-I-function M-function 0 D+ai 2018 Daoud et al. [28] S1
    Monod-I Monod 0 D 2019 Di and Yang [2] S1
    M-I-function M-function 0 D 2023 Albargi and El Hajji [58] S1
    M-I-function I-M-function 0 D 2012 Sari et al. [11] S2
    Monod-I I-Monod 0 D 2019 Di and Yang [2] S2

     | Show Table
    DownLoad: CSV
    Table A5.  Models of commensalistic relationship with self inhibitions. The growth functions μ1 and μ2 are defined in Table A4.
    μ1 μ2 Sin2 Di Year Ref. Case
    H-function M-function 0 D 1981 Stephanopoulos [7] C11
    M-function H-function 0 D 1981 Stephanopoulos [7] C12
    Monod Haldane 0 αD 2001 Bernard et al. [47] C12
    Monod Haldane 0 D 2010 Simeonov and Diop [48] C12
    M-function H-function 0 D 2010 Sbarciog et al [49] C12
    M-function H-function 0 αD 2012 Benyahia et al. [39] C12
    M-function H-function 0 D 2012 Weedermann [50] C12
    M-function H-function 0 αD 2018 Bayen and Gajardo [51] C12
    M-function H-function 0 αD 2020 Sari and Benyahia [31] C12
    M-function H-function 0 αiD+ai 2022 Sari [30] C12
    H-function H-function 0 D 1981 Stephanopoulos [7] C112

     | Show Table
    DownLoad: CSV
    Table A6.  Models of syntrophic relationship with self inhibitions. The growth functions μ1 and μ2 are defined in Table A4.
    μ1 μ2 Sin2 Di Year Ref. Case
    KB2 I-Monod 0 D 1988 Kreikenbohm and Bohl [41] S21
    HGG H-function 0 D 2014 Harvey et al. [54] S12
    M-I function H function 0 D+ai 2017 Fekih-Salem et al. [55] S12
    M-I function H function 0 αiD+ai 2020 Fekih-Salem et al. [29] S12

     | Show Table
    DownLoad: CSV
    Table A7.  Growth functions used in Tables A3-A6.
    Function Definition
    Monod μi(Si)=miSiKi+Si
    Monod-I μ1(S1,S2)=m1S1K1+S111+L2S2
    I-Monod μ2(S1,S2)=m2S2K2+S211+L1S1
    KB1 μ1(S1,S2)={m1(S1S2/K2)K1+S1+L1S2 if S1S2/K2>00 otherwise 
    Haldane μ2(S2)=m2S2K2+S2+S22/KI
    KB2 μ1(S1,S2)={m1(S1S2/K2)K1+S1+L1S2+S21/K1 if S1S2/K2>00 otherwise 
    M-function μ(0)=0 and for S>0, μ(S)>0
    H-function μ(0)=0 and { there is Sm(0,+] such that μ(S)>0 for S<Sm and μ(S)<0 for S>Sm. (Note that if Sm=+, we obtain an M-function). 
    M-I function μ1(0,S2)=0 and, for S1,S2>0, μ1S1(S1,S2)>0, μ1S2(S1,S2)0
    I-M function μ2(S1,0)=0 and, for S1,S2>0, μ2S2(S1,S2)>0, μ2S1(S1,S2)0
    HHG μ1(S1,S2)=f(S1)I(S2) with {f(0)=0,f(S1)>0 for S1>0 and I(S2)<0 for S2>0.

     | Show Table
    DownLoad: CSV

    To our knowledge, Reilly [6] was the first to propose a mathematical study of the pure commensalistic model C1, with Monod growth functions and without decay terms of the species. He also considered more complicated commensalistic systems when feedback inhibition and feedforward activation occur to explain oscillations observed in experimental data.

    The work of Stephanopoulos [7] is an important contribution to commensalism. He considered general growth functions and investigated the case of nonmonotone growth functions. In the case of equal removal rates, he reduced the system to a planar system and gave a complete qualitative description of the Cases C1, C11, C12, and C112, see Figures 3 and 4 in [7].

    The classical two-step (acidogenesis-methanisation) dynamical representation of anaerobic digestion processes, see [46, Eq (1.85)], has the structure of the pure commensalistic model C1 or its extension C12, obtained by adding an inhibition by the second substrate. It was used by Simeonov and Stoyanov [40] for the control of the process who considered C1, with Monod growth functions and including decay terms.

    An important contribution to the modeling of anaerobic digestion as the commensalistic system C12 is the model of Bernard et al. [47] with a Monod function for μ1 and a Haldane function for μ2. Their model, sometimes referred to as AMOCO or AM2, included the term α which represents the fraction of the biomass in the liquid phase. Simeonov and Diop [48] studied this model for α=1. Sbarciog et al. [49] and Weedermann [50] studied the model with general growth function and α=1, while the interesting case where 0<α1 was studied by Benyahia et al. [39]. Bayen and Gajardo [51] studied the steady state optimization of the biogas production of AM2, while the operating diagram of the model is described in [31]. For more details and complements on anaerobic digestion, we refer the reader to the recent review [52].

    System S1 with a syntrophic relationship between the species was considered by Wilkinson [13] with an inhibited Monod growth function for μ1 and a Monod growth function for μ2, while Kreikenbohm and Bohl [9] considered another form of feedback of the second substrate on the first species, see Table A7. Burchard [8] extended the results of [9, 13] to a large class of general growth functions. He highlighted conditions under which there is persistence or extinction. El Hajji et al. [10] also obtained results for general growth functions. The studies in [8, 9, 10, 13] have shown that, if it exists, the positive coexistence equilibrium is unique and stable, and there is no other stable equilibrium.

    Another model of syntrophic relationship in anaerobic digestion was considered by El Hajji [53].

    All these studies did not include the decay terms of the species. Xu et al. [14] considered the effects of the decay terms and studied the model S1 with an inhibited Monod growth function for μ1 and a Monod growth function for μ2, and Sin2=0. These authors leaved unanswered the question of the stability of the positive steady state as long as it exists. Sari and Harmand [12] considered S1 with general growth functions and proved that the positive steady state is stable whenever it exists. Daoud et al. [28] extended the results of [12] to the case Sin2>0, showing the appearance of a new boundary equilibrium where the species X1 is absent, while the species X2 is present.

    System S12, which includes an inhibition of the second species by the second substrate, was considered by Harvey et al. [54] in the case without decay terms, where μ1(S1,S2) is the product of an increasing function of S1 and a decreasing function of S2, which is a particular case of M-I-functions, while μ2(S2) is an H-function, see [54, Figure 2]. The more general case of M-I-function, also including decay terms of the species, was considered by Fekih-Salem et al. [29, 55]. In this case, the stability analysis is much more delicate since the system cannot be reduced to a planar system.

    System S2, with two inhibitions, was considered by Sari et al. [11], who showed that, in contrast with the case of S1 with only an inhibition of the second substrate on the first species, a multiplicity of positive equilibria can occur.

    An example of system S21 with three different inhibitions was considered by Kreikenbohm and Bohl [41]. The mathematical analysis of this model shows the occurrence of bistability as in the case of the system S2, without the additional inhibition of the first species by its limiting substrate.

    Other models for which μ1 and μ2 depend both on (S1,S2) and, in addition, X1 and X2 are in competition on substrate S1, exhibiting the multiplicity of positive equilibrium points, can be found in [56]. An important and interesting extension should be mentioned here: [26] proposed an 8-dimensional mathematical model, which includes syntrophy and inhibition, and both mechanisms considered by [47] and by [10]. The effects of decay terms are considered by [27].



    [1] Sauro I (2013) Turismo sostenibile: la Sicilia tra parchi e riserve naturali, Territorio paesaggio e comunità locali: sviluppo integrato e sostenibilità, tesi di dottorato, XXV ciclo. Università degli studi di Catania.
    [2] Mexa A, Coccossis H (2004) The Challenge of Tourism Carrying Capacity Assessment: Theory and Practice. Ashgate Publishing, Farnham.
    [3] Pan SY, Gao M, Kim H, et al. (2018) Advances and challenges in sustainable tourism toward a green economy. Sci Total Environ 6351: 452-469. https://doi.org/10.1016/j.scitotenv.2018.04.134 doi: 10.1016/j.scitotenv.2018.04.134
    [4] Godfrey K, Clarke J (2000) The tourism development handbook: a practical approach to planning and marketing. Thompson Learning, London, UK.
    [5] Paletto A (2002) Il valore economico totale come strumento di valutazione della multifunzionalità forestale. Un'analisi teorica e applicazione ai boschi del comune d'Oulx. Ph.D. thesis, University of Trento, Italy.
    [6] Distaso M (1998) L'economia del paesaggio rurale. Agribusiness Paesaggio e Ambiente 1: 22-39.
    [7] Pigliaru F (1996) Economia del turismo: note su crescita, qualità ambientale e sostenibilità. Contributi di Ricerca CRENoS 12.
    [8] Novelli S (2005) Aspetti economici e politici della conservazione del paesaggio rurale. Definizione dello strumento d'indagine per una valutazione economica nell'astigiano. Ph.D. thesis (XVI cycle), University of Turin, Italy
    [9] Tabet D (1963) La rendita fondiaria nell'agricoltura italiana, Editori Riuniti, Rome, Italy.
    [10] Vos W, Meekes M (1999) Trends in European cultural landscape development: perspectives for a sustainable future. Landscape Urban Plan 46: 3-14. doi: 10.1016/S0169- 2046(99)00043-2. https://doi.org/10.1016/S0169-2046(99)00043-2
    [11] Dillman BL, Bergstrom JC (1991) Measuring environmental amenity benefits of agricultural land. In: "Farming in the countryside: an economic analysis of costs and benefits" (Hanley N ed). CAB International, Wallingford, UK. r r
    [12] Price C (1978) Landscape economics. Macmillan, London, UK. https://doi.org/10.1007/978-1-349-03747-6
    [13] Van den Berg AE, Wlek CAJ (1998) The influence of planned-change context on the evolution of natural landscapes. Landscape Urban Plan 43: 1-10. doi: 10.1016/S0169- 2046(98)00102 9. https://doi.org/10.1016/S0169-2046(98)00102-9
    [14] Kaplan S (1979) Concerning the power of content identifying methodologies. In: "Assessment of amenity resources values" (Daniel TC, Zube EH eds). USDA Forest Service Rocky Mountains Station, USA.
    [15] Tempesta T, Thiene M (2006) Percezione e valore del paesaggio. Franco Angeli, Milan, Italy.
    [16] FAO (2021) The State of the World's Forests.
    [17] Sicilian Regional Law n°98/1981
    [18] Crescimanno ML, Licata F (1994) IL PARCO DELLE MADONIE - Natura e Paesi.
    [19] ISTAT (2020) Popolazione residente, Roma.
    [20] ISTAT (2011) Censimento Generale dell'Agricoltura Italiana, Roma
    [21] Sgroi F (2021) Food products, gastronomy and religious tourism: The resilience of food landscapes. Int J Gastronomy Food Sci 26: 100435. https://doi.org/10.1016/j.ijgfs.2021.100435 doi: 10.1016/j.ijgfs.2021.100435
    [22] Sgroi F (2021) Territorial development models: A new strategic vision to analyze the relationship between the environment, public goods and geographical indications. Sci Total Environ 787: 147585. https://doi.org/10.1016/j.scitotenv.2021.147585 doi: 10.1016/j.scitotenv.2021.147585
    [23] Sgroi F (2020) Forest resources and sustainable tourism, a combination for the resilience of the landscape and development of mountain areas. Sci Total Environ 736: 139539. https://doi.org/10.1016/j.scitotenv.2020.139539 doi: 10.1016/j.scitotenv.2020.139539
    [24] Della Corte V (2000) La gestione dei sistemi locali di offerta turistica. Cedam, Padua, Italy.
    [25] Rispoli M, Tamma M (1995) Risposte strategiche alla complessità: le forme di offerta dei prodotti alberghieri. Giappichelli, Turin, Italy.
    [26] Tamma M (2002) Destination management: gestire prodotti e sistemi locali di offerta turistica. In: "Destination management. Governare il turismo tra locale e globale" (Franch M ed). Giappichelli, Torino, Italy, 11-38.
    [27] Clauser O, Franch M, Gios G (2001) Compatibilità tra sviluppo della domanda di turismo invernale nelle Dolomiti e sostenibilità ambientale. In: "La Sardegna nel mondo mediterraneo: Quinto convegno internazionale di studi Turismo e Ambiente" (Scanu G ed). Sassari (Italy) 28-30 October 1998. Patron editore, Bologna, Italy.
    [28] Ejarque J (2003) La destinazione turistica di successo. Marketing e management. Hoepli, Milan, Italy.
  • This article has been cited by:

    1. Hanan H. Almuashi, Nada A. Almuallem, Miled El Hajji, The Effect of Leachate Recycling on the Dynamics of Two Competing Bacteria with an Obligate One-Way Beneficial Relationship in a Chemostat, 2024, 12, 2227-7390, 3819, 10.3390/math12233819
  • Reader Comments
  • © 2022 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(2091) PDF downloads(174) Cited by(3)

Figures and Tables

Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog