826
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Dynamics of autotroph–mixotroph interactions with the intraguild predation structure

, &
Pages 186-206 | Received 04 Nov 2021, Accepted 07 Apr 2022, Published online: 25 Apr 2022

Abstract

A mathematical model with the intraguild predation structure is proposed to describe the interactions of autotrophs and mixotrophs containing light and nutrients in a well-mixed aquatic ecosystem. The dissipation, existence and stability of equilibria of the model are proved, and the ecological reproductive indexes for the extinction, survival and coexistence of autotrophs and mixotrophs are established. We also consider the influence of Holling type functional responses and abiotic factors on the coexistence and biomass of autotrophs and mixotrophs. It is shown that the intraguild predation structure is beneficial to phytoplankton biodiversity and provides an explanation for the phytoplankton paradox.

1. Introduction

Algae, as typical phytoplankton, are the primary producer of organic compounds in the whole aquatic ecosystem. They form the basis of energy flow and material circulation in lakes or oceans. Algae mainly consist of two parts: autotrophs (autotrophic algae) and mixotrophs (mixotrophic algae). Autotrophs synthesize organic matters through photosynthesis to supply their own growth [Citation1, Citation11, Citation17, Citation18]. Mixotrophs as the combination of autotrophic and heterotrophic nutrition not only use photosynthesis to combine inorganic matters into organic matters, but also feed on microorganisms to meet their growth [Citation2, Citation3].

The interaction of autotrophs and mixotrophs is complicated containing competition and predation (see Figure ). The growth of autotrophs is affected by nutrition and light [Citation1, Citation10, Citation23, Citation29]. There is no obvious dichotomy for mixotrophs as producers or consumers. In general, autotrophic activities and heterotrophic activities of mixotrophs are carried out at the same time. In the autotrophic situation, the metabolism of mixotrophs is also limited by light and nutrients. This indicates that in this case, autotrophs and mixotrophs compete for light [Citation2, Citation14] or nutrients [Citation15, Citation16, Citation24, Citation25]. Mixotrophs in the heterotrophic nutrition can ingest micro-autotrophs [Citation14, Citation24, Citation25].

Figure 1. The interactions between autotrophs and mixotrophs.

Figure 1. The interactions between autotrophs and mixotrophs.

Intraguild predation is a ubiquitous ecological phenomenon in natural ecosystems [Citation5, Citation20]. It describes a relationship in which the predator not only consumes the prey, but also shares the same resource with the prey. From the relationship between autotrophs and mixotrophs, they form an intraguild predation structure. Previous studies have shown that autotrophs and mixotrophs compete only for nutrients [Citation15, Citation16, Citation24, Citation25] or only for light [Citation14]. Light comes from the water surface and gradually weakens with the depth of the water [Citation4, Citation19, Citation23]. Nutrients come from the bottom of the lake, and spread to the whole lake with the turbulence [Citation6, Citation27, Citation28]. Due to the importance of nutrients and light on the metabolism of autotrophs and mixotrophs [Citation1, Citation23, Citation29], it is essential to investigate the intraguild predation structure between autotrophs and mixotrophs with nutrients and light.

According to the above considerations and the existing research results, we propose and analyse a mathematical model of autotroph–mixotroph interactions with the intraguild predation structure. Our model contains the effect of light and nutrients. Through the analysis of the model, we investigate the dynamical behaviours for the survival, extinction and coexistence of autotrophs and mixotrophs, and consider Holling type functional responses, light and nutrients on the coexistence of autotrophs and mixotrophs. Here, we also try to study the ‘Paradox of the Plankton’ based on the model. This paradox, proposed by Hutchinson [Citation9], generally describes that the diversity of phytoplankton deviates from the limitation of the available abiotic resources in the same ecological niche.

Algal blooms are a natural ecological phenomenon of algae mass-breeding in the water source. It seriously affects aquatic ecological communities, water quality, and even threatens people's health. Algal blooms are caused by many reasons, including biotic and abiotic factors [Citation1, Citation27, Citation29]. The motivation of this paper is to explore the effect of biotic and abiotic factors on the blooms of autotrophic algae and mixotrophic algae. Based on our model, we will consider the biomass density of autotrophs and mixotrophs with the change of some important environmental parameters.

We construct this paper as follows. In Section 2, a dynamical model describing the interaction of autotrophs and mixotrophs is established in the intraguild predation structure. Then, we study dynamical features of this model containing dissipation of the solutions, the existence and stability of non-negative stable solutions in Section 3. In Section 4, by using some numerical simulations, we investigate the effect of abiotic and biotic factors on the coexistence and density of autotrophs and mixotrophs at a steady state with biologically reasonable parameters, and it explains their effects in phytoplankton biodiversity and blooms. Section 5 summarizes our results and proposes some research prospects to be studied in the future.

2. The model

We establish a dynamical model with the intraguild predation structure to describe the interactions of autotrophs and mixotrophs containing light and nutrients. Here, we consider a well-mixed aquatic environment such as a water column [Citation7, Citation8, Citation14] or epilimnion in lakes [Citation23, Citation29]. Let z represent the water depth coordinate. Then z = 0 denotes the water surface, and z = L denotes the bottom of the water column or epilimnion. Our model consists of three ordinary differential equations and one algebraic equation, characterizing the rate of change for autotrophs (A), mixotrophs (M) and dissolved nutrients (N) and light intensity (I(x,A,M)). Table  lists the information on variables and parameters of the model.

Table 1. The variables and parameters of model (Equation3).

Light from the water surface is absorbed by the water body and phytoplankton. The light intensity at depth z can be described by the Lambert–Beer law [Citation7] as I(z,A,M)=Iinexp(k0z(k1A+k2M)z),0<z<L.The growth of autotrophs or mixotrophs in the autotrophic situation are limited by the light intensity I(z,A,M) and dissolved nutrients N. It takes the form of multiplication of the two [Citation8, Citation23, Citation29] and is expressed as riHi(Iin,A,M)Fi(N)=ri1L0LH~i(I(z,A,M))dzFi(N),i=1,2.Here, we assume that H~i,Fi satisfy (1) H~i(0)=0,0H~i<1,H~iA<0andH~iM<0,i=1,2,Fi(0)=0,0Fi<1,Fi(N)>0andlimNFi(N)=1,i=1,2.(1) Typical expressions for H~i and Fi are Monod functions [Citation8, Citation23]: (2) H~i(I(z,A,M))=I(z,A,M)I(z,A,M)+hi,Fi(N)=NN+γi,i=1,2.(2) In the heterotrophic condition, mixotrophs supply their own growth by ingesting autotrophic organisms [Citation14, Citation24, Citation25]. The predation rate of mixotrophs to autotrophs is G(A), where G(A) is assumed to satisfy G(0)=0,G(A)>0.Based on the principle of ecological stoichiometry [Citation11], the conversion rate satisfies emin{θ1/θ2,1}.

The reduction of biomass of autotrophs and mixotrophs mainly contains lost biomass m1A,m2M due to death and respiration, sinking biomass v1A/L,v2M/L and water exchange biomass DA/L,DM/L [Citation23]. In addition, the predation of mixotrophs will also lead to the decline of autotrophic biomass.

The change of dissolved nutrients N consists of three parts. The first part depends on consumption by autotrophs and mixotrophs with a consumption rate θiriHi(Iin,A,M)Fi(N),i=1,2.The second part is nutrient recycling from the loss of autotrophic and mixotrophic with recycling proportion δ1,δ2 and the nutrients to carbon quota θ1,θ2. The last part is nutrient exchange (D/L)(NinN) at the bottom of the water column or epilimnion, where Nin is a fixed nutrient input concentration.

According to the above discussion, we have the following autotroph–mixotroph interaction model in a well-mixed aquatic environment: (3) dAdt=r1F1(N)H1(Iin,A,M)Agrowthofautotrophsm1Alossv1+DLAsinkingandexchangeaG(A)M,predationofmixotrophsdMdt=r2F2(N)H2(Iin,A,M)M+eaG(A)Mgrowthofmixotrophsm2Mlossv2+DLM,sinkingandexchangedNdt=DL(NinN)nutrientinputandexchange+δ1θ1m1A+δ2θ2m2Mnutrientrecyclingθ1r1F1(N)H1(Iin,A,M)Aconsumptionbyautotrophsθ2r2F2(N)H2(Iin,A,M)Mconsumptionbymixotrophs.(3) Considering the biological significance of (Equation3), we assume that all the parameters in Table  are positive and δ1,δ2[0,1],0<emin{θ1/θ2,1}. By using rigorous mathematical arguments, we can draw the conclusion that for any positive initial value, (Equation3) admits a unique positive solution defined for all t0.

3. Qualitative analysis

In this section, we explore the dynamics of (Equation3) containing dissipation of the solutions, the existence and stability of non-negative steady-state solutions. The following theorem concludes the positive invariance and dissipation of the solutions of model (Equation3).

Theorem 3.1

The system (Equation3) is dissipative.

Proof.

Let W=N+θ1A+θ2M, which is the total nutrients in system (Equation3). A direct calculation gives dWdt=DL(NinW)θ1v1A+θ2v2ML(1δ1)θ1m1A(1δ1)θ2m2Ma(θ1eθ2)G(A)MDL(NinW).This implies that limsuptW(t)Nin and the system (Equation3) is dissipative.

Let Δ:={(A,M,N)R+3|N+θ1A+θ2MNin}.It follows from Theorem 3.1 that Δ is a global attracting region of system (Equation3) and positively invariant.

3.1. Semi-trivial steady states

It follows from (Equation3) that the possible non-negative semi-trivial steady states are listed below:

  1. Nutrient-only semi-trivial steady state E0:(0,0,Nin).

  2. Nutrient-autotroph semi-trivial steady state Ea:(Aa,0,Na), where Aa,Na satisfy (4) r1F1(N)H1(Iin,A,0)m1v1+DL=0,DL(NinN)+δ1θ1m1Aθ1r1F1(N)H1(Iin,A,0)A=0.(4)

  3. Nutrient-mixotroph semi-trivial steady state Em:(0,Mm,Nm), where Mm,Nm satisfy r2F2(N)H2(Iin,0,M)m2v2+DL=0,DL(NinN)+δ2θ2m2Mθ2r2F2(N)H2(Iin,0,M)M=0.

We define the following ecological reproductive indexes for autotrophs and mixotrophs: R0A:=r1F1(Nin)H1(Iin,0,0)m1+(v1+D)/L,R0M:=r2F2(Nin)H2(Iin,0,0)m2+(v2+D)/Land R1A:=r1F1(Nm)H1(Iin,0,Mm)aG(0)Mmm1+(v1+D)/L,R1M:=r2F2(Na)H2(Iin,Aa,0)+eaG(Aa)m2+(v2+D)/L.As an indicator of viability for autotrophs or mixotrophs, RiA(RiM), i = 0, 1, describe the average number of new autotrophs (mixotrophs) produced by one cubic metre of autotrophs (mixotrophs) in a life cycle of autotrophs (mixotrophs). In the following discussion, it is shown that RiA and RiM, i = 0, 1 are four critical values for autotrophs and mixotrophs to invade aquatic ecosystems, respectively.

The following result shows that (Equation3) always has a unique nutrient-only semi-trivial steady state E0, which is also stable in terms of the ecological reproductive index max{R0A,R0M}<1.

Theorem 3.2

E0 always exists and it is globally asymptotically stable if max{R0A,R0M}<1, while E0 is unstable if max{R0A,R0M}>1.

Proof.

It is clear that E0 always exists. The Jacobian matrix at E0 is J(E0)=(a11000a220a31a32D/L),where a11=r1F1(Nin)H1(Iin,0,0)m1v1+DL,a31=δ1θ1m1θ1r1F1(Nin)H1(Iin,0,0),a22=r2F2(Nin)H2(Iin,0,0)m2v2+DL,a32=δ2θ2m2θ2r2F2(Nin)H2(Iin,0,0).If max{R0A,R0M}<1, then real parts of all the eigenvalues of the characteristic equation of J(E0) are negative. This implies that E0 is locally asymptotically stable. On the contrary, if max{R0A,R0M}>1, then E0 is unstable.

It follows from Theorem 3.1 that limsuptN(t)Nin. From (Equation3), we have dAdtr1F1(Nin)H1(Iin,0,0)Am1Av1+DLAfor sufficiently large t. This means that limsuptA(t)=0 if R0A<1. The second Equation of (Equation3) becomes dMdt=r2F2(N)H2(Iin,0,M)Mm2Mv2+DLMr2F2(Nin)H2(Iin,0,0)Mm2Mv2+DLM.By the comparison theorem, limsuptM(t)=0 if R0M<1. By using the theory of asymptotical autonomous systems [Citation13], model (Equation3) can reduce to a limiting system dNdt=DL(NinN).This implies that limtN(t)=Nin and E0 is globally attractive. Then E0 is globally asymptotically stable.

Remark 3.1

max{R0A,R0M} is a critical value for algae to survive in a well-mixed aquatic environment. It is affected by the nutrient input concentration Nin and the water surface light intensity Iin. It can be observed directly that max{R0A,R0M} is increasing with respect to Nin and Iin. This means that the high light intensity and nutrient input concentration are conducive to the invasion of algae.

Theorem 3.3

  1. Ea exists and is unique if and only if R0A>1 .

  2. If R1M<1, then Ea is locally asymptotically stable, while Ea is unstable if R1M>1.

Proof.

From (Equation4), we have Λ1(A):=Nin[L(1δ1)θ1m1+(v1+D)θ1D]A,Λ2(A):=F1(m1+(v1+D)/Lr1H1(Iin,A,0)).Note that Λ1(A) is strictly monotone decreasing with respect to A with limA0+Λ1(A)=Nin. It follows from (Equation1) that Λ2(A) is strictly monotone increasing with respect to A with limA0+Λ2(A)=F1(m1+(v1+D)/Lr1H1(Iin,0,0)).Then there exists a unique positive Aa such that Λ1(Aa)=Λ2(Aa) if and only if R0A>1. From the first equality of (Equation4), Na can be explicitly solved.

The Jacobian matrix at Ea is J(Ea)=(a11a12a130a220a31a32a33),where a11=r1F1(Na)H1AAa,a12=r1F1(Na)H1MAaaG(Aa),a13=r1F1(Na)H1Aa,a22=r2F2(Na)H2m2v2+DL+eaG(Aa),a31=δ1θ1m1θ1r1F1(Na)(H1AAa+H1),a33=DLθ1r1F1(Na)H1Aa,a32=δ2θ2m2θ1r1F1(Na)H1MAaθ2r2F2(Na)H2.From (Equation1), we have H1(0,0,0)=0,0H1(Iin,A,M)<1,H1(Iin,A,M)A<0,H1(Iin,A,M)M<0.A direct calculation gives a11+a33=r1F1(Na)H1AAaDLθ1r1F1(Na)H1Aa<0,a11a33a13a31=DLr1F1(Na)H1AAa+θ1r1((1δ1)m1+DL)F1(N)H1Aa>0.If R1M<1, then a22<0. To conclude, we have a11+a33<0,a11a33a13a31>0,a22<0.This shows that the real parts of all the eigenvalues of the characteristic equation of J(Ea) are negative. Hence we conclude that if R1M<1, then Ea is locally asymptotically stable; Conversely, if R1M>1, then Ea is unstable.

Remark 3.2

The conclusion in Theorem 3.3 reveals that R0A is a critical value for autotrophs from extinction to survival. If R1M<1, then autotrophs win the competition in the intraguild predation and dominate the whole aquatic ecosystem. This implies that R1M is a threshold value for mixotrophs to survive in the water ecosystem in the presence of autotrophs.

Theorem 3.4

  1. Em exists and is unique if and only if R0M>1 .

  2. If R1A<1, then Em is locally asymptotically stable, while Em is unstable if R1A>1.

Proof.

The proof of (i) has many similarities with Theorem 3.3. Then, we omit the proof process here. The Jacobian matrix at Em is J(Em)=(a1100a21a22a23a31a32a33),where a11=r1F1(Nm)H1m1v1+DLaG(0)Mm,a22=r2F2(Nm)H2MMm,a21=r2F2(Nm)H2AMm+eaG(0)Mm,a23=r2F2(Nm)H2Mm,a32=δ2θ2m2θ2r2F2(Nm)(H2MMm+H2),a33=DLθ2r2F2(Nm)H2Mm,a31=δ1θ1m1θ1r1F1(Nm)H1θ2r2F2(Nm)H2AMm.Similar to the proof of Theorem 3.3, we prove that the real parts of eigenvalues of J(Em) are all negative if R1A<1, which means that Em is locally asymptotically stable. On the contrary, Em is unstable if R1A>1.

Remark 3.3

Theorem 3.4 shows that Em exists if R0M>1 and R0M is the index of threshold determining the survival of mixotrophs. Here, we only obtain the local stability of Em. Actually, our numerical simulations indicate that Em should be globally asymptotically stable. This means that if R1A<1, then mixotrophs win the competition in the intraguild predation structure. Autotrophs invade successfully only when R1A>1.

3.2. Coexistence steady states

This subsection is to explore the existence of the coexistence solution Ec:(Ac,Mc,Nc) by using the method of repellers and persistence in Zhao [Citation30].

We first consider the following autotroph-nutrient model and mixotroph-nutrient model: (5) dAdt=r1F1(N)H1(Iin,A,0)Am1Av1+DLA,dNdt=DL(NinN)+δ1θ1m1Aθ1r1F1(N)H1(Iin,A,0)A(5) and (6) dMdt=r2F2(N)H2(Iin,0,M)Mm2Mv2+DLM,dNdt=DL(NinN)+δ2θ2m2Mθ2r2F2(N)H2(Iin,0,M)M.(6) From Theorem 3.1, we find that the sets Δ1={(A,N)R+2|N+θ1ANin},Δ2={(M,N)R+2|N+θ2MNin}are global attracting regions and positively invariant sets of (Equation5) and (Equation6), respectively. We next show the global asymptotic stability of the positive steady states of (Equation5) and (Equation6).

Lemma 3.5

  1. If R0A>1, then model (Equation5) holds a unique positive equilibrium E¯a(Aa,Na), which is globally asymptotically stable;

  2. If R0M>1, then model (Equation6) holds a unique positive equilibrium E¯m(Mm,Nm), which is globally asymptotically stable.

Proof.

The uniqueness, existence and local asymptotic stability of E¯a,E¯m can be obtained from Theorems 3.3 and 3.4. We show that there is no periodic orbit of (Equation5) in Δ1 by using Dulac's criterion. Let K1(A,N)=r1F1(N)H1(Iin,A,0)Am1Av1+DLA,K2(A,N)=DL(NinN)+δ1θ1m1Aθ1r1F1(N)H1(Iin,A,0)A.Choose D(A,N)=1/A as a Dulac function. Then (DK1(A,N))A+(DK2(A,N))N=r1F1(N)H1(Iin,A,0)ADALθ1r1F1(N)H1<0.This implies that (Equation5) has no positive periodic orbit in Δ1. When R0A>1, E¯ is globally asymptotically stable for (Equation5) by using the Poincare–Bendixson criterion. Similarly, E¯m is also globally asymptotically stable for (Equation6) when R0M>1.

In order to study the existence of Ec, we assume that Φ(t):ΔΔ is the semiflow generated by system (Equation3) and Φ(t)(A(0),M(0),N(0))=(A(t),M(t),N(t)) for all t0 with the initial value (A(0),M(0),N(0))R+3. Let Δ0:={(A,M,N)Δ:A>0andM>0},Δ0:=ΔΔ0={(A,M,N)Δ:A=0orM=0}and Λ:={(A0,M0,N0)Δ0:Φ(t)(A0,M0,N0)Δ0foralltR+}.From Theorem 3.1 and uniqueness of solutions, Δ and Δ0 are positively invariant for system (Equation3).

We are in a position to establish the existence of the coexistence steady-state solution Ec.

Theorem 3.6

Assume that R0A>1, R0M>1. If R1A>1 and R1M>1, then system (Equation3) is uniformly persistent with respect to (Δ0,Δ0) in the sense that there is a positive constant ϵ>0 such that lim inftA(t)>ϵandlim inftM(t)>ϵ.Moreover, system (Equation3) has at least one positive steady state solution Ec.

Proof.

We first prove that the omega limit set ω(q)={E0,Ea,Em} of the orbit with the initial value q(A0,M0,N0)Δ0. For any q(A0,M0,N0)Δ0, we have Φ(t)qΔ0 and consider the following three cases: (1) A0=0,M0=0; (2) A0=0,M00; (3) A00,M0=0.

Case (1): A0=0,M0=0. Then, we get A(t,q)=0 and M(t,q)=0 for any t0. From Theorem 3.2, (A(t,q),M(t,q),N(t,q))(0,0,Nb) as t;

Case (2): A0=0,M00. Then, we obtain A(t,q)=0 for any t0. This shows that (M(t,q),N(t,q)) satisfies (Equation6). By Lemma 3.5, (M(t,q),N(t,q))(Mm,Nm) as t;

Case (3): A00,M0=0. Then, we have M(t,q)=0 for any t0 and (A(t,q),N(t,q)) satisfies (Equation5). From Lemma 3.5, we have (A(t,q),N(t,q))(Aa,Na) as t.

In view of cases (1)–(3), we have ω(q)={E0,Ea,Em} for any qΔ0. From R0A>1 and R1M>1, we let δ1=12(r1F1(Nin)H1(Iin,0,0)m1v1+DL)>0,δ2=12(r2F2(Na)H2(Iin,Aa,0)+eaG(Aa)m2v2+DL)>0.We then show that E0,Ea,Em are uniform weak repellers for Δ0. Then it follows that there exists a ε>0 so that (7a) lim suptdist(Φ(t)q,E0)ε,(7a) (7b) lim suptdist(Φ(t)q,Ea)ε,(7b) (7c) lim suptdist(Φ(t)q,Em)ε(7c) for any q(A0,M0,N0)Δ0 and t0. If (Equation7a) does not hold, then there exists a q0Δ0 such that lim suptdist(Φ(t)q0,E0)<δ1.This means that there exists a T1>0 such that (8) |N(t,q0)Nin|<δ1,|A(t,q0)|<δ1foralltT1.(8) Then r1F1(N)H1(Iin,A,0)m1v1+DL>r1F1(Nin)H1(Iin,0,0)m1v1+DLδ1=δ1and dA(t,q0)dt>δ1A(t,q0)for any tT1. This shows that A(t,q0) as t, which is a contradiction to (Equation8).

Assume that (Equation7b) does not hold, then there exists a qaΔ0 and T2>0 such that lim suptdist(Φ(t)qa,Ea)<δ2and (9) |N(t,qa)Na|<δ2,|A(t,qa)Aa|<δ2,|M(t,qa)|<δ2foralltT2.(9) Note that r2F2(N)H2(Iin,A,M)+eaG(A)m2v2+DL>r2F2(Na)H2(Iin,Aa,0)+eaG(Aa)m2v2+DLδ2=δ2for all tT2. Hence dM(t,qa)/dt>δ2M(t,qa) and M(t,qa) as t. This is a contradiction with (Equation9). Similarly, (Equation7c) holds.

It follows from Theorem 3.1 that Φ(t) is point dissipative and compact, and then Φ(t) has a global attractor in Δ. From the above analysis, we obtain the following conclusions: (i) {E0,Ea,Em} is disjoint, compact, and isolated invariant sets in Δ0; (ii) qΔ0ω(q)={E0,Ea,Em}; (iii) E0, Ea and Em are isolated in Δ; (iv) no subset of E0, Ea and Em forms a cycle in Δ0; (v) Ws(Ei)Δ0=, i = 0, a, m, where Ws(Ei) is the stable set of Ei. By Theorem 1.3.1 in Zhao [Citation30], we have Φ(t):ΔΔ is uniformly persistent with respect to (Δ0,Δ0). According to Theorem 1.3.6 in Zhao [Citation30], there exists a global attractor for Φ(t) in Δ0 and Φ(t) has at least one fixed point (Ac,Mc,Nc)Δ0, which is a coexistence steady state solution of system (Equation3). This completes the proof.

Remark 3.4

In the intraguild predation structure, autotrophs and mixotrophs can coexist in an aquatic ecosystem if the conditions of Theorem 3.6 hold. This suggests that the intraguild predation is an important ecological mechanism in aquatic communities. It is beneficial to phytoplankton biodiversity, and provides an explanation for the phytoplankton paradox.

Remark 3.5

RiA,RiM>1, i = 0, 1 are only sufficient conditions for the coexistence of autotrophs and mixotrophs. If this condition does not hold, they can also coexist, such as region Δ4Δ5 in Figure . In the next section, our numerical simulations show the coexistence form of autotrophs and mixotrophs is either a positive steady-state solution or a positive periodic solution.

Figure 2. The parameter regions in the (m1,m2)-plane about the survival of autotrophs and mixotrophs. Here, other parameters are from Table .

Figure 2. The parameter regions in the (m1,m2)-plane about the survival of autotrophs and mixotrophs. Here, other parameters are from Table 1.

4. Autotroph and mixotroph coexistence and density

Algae are widely distributed in various aquatic ecosystems, ranging from lakes, rivers to oceans. The coexistence of autotrophic algae and mixotrophic algae in the intraguild predation structure plays an important role in maintaining algal biodiversity and explaining the phytoplankton paradox [Citation14, Citation24, Citation25]. The coexistence mechanism can be significantly influenced by some biotic and abiotic factors, such as the structure of functional response functions, light intensity and nutrients input. It is meaningful to investigate the influence of these factors on the coexistence of autotrophs and mixotrophs. The algal biomass density is an essential indicator for forewarning algal blooms and evaluating water quality. Another purpose of the section is to consider the effects of environmental factors on the biomass density of autotrophs and mixotrophs.

In the following discussion, some numerical simulations are done to explore the coexistence and biomass density of autotrophs and mixotrophs based on parameters in biologically reasonable ranges listed in Table . Here, there are several parameter values that we assume. Due to the simultaneous autotrophic and heterotrophic activities of mixotrophs, they have lower nutrient and light absorption rates than autotrophs. Hence it is biologically reasonable to assume that γ2γ1, k2k1. The value of β is chosen from the autotrophic biomass density in our numerical simulations. The research background of the current paper is a well-mixed aquatic environment such as shallow lakes or epilimnion. This means that L = 10 m is also consistent with the realistic aquatic environment.

4.1. Autotroph and mixotroph coexistence

We first consider the effect of different functional responses on the coexistence of autotrophs and mixotrophs. To facilitate the following analysis, in view of Theorems 3.2, 3.3, 3.4 and 3.6, we denote m1:=r1F1(Nin)H1(Iin,0,0)(v1+D)/L,m2:=r2F2(Nin)H2(Iin,0,0)(v2+D)/L,m1:=r1F1(Nm)H1(Iin,0,Mm)aG(0)Mm(v1+D)/L,m2:=r2F2(Na)H2(Iin,Aa,0)+eaG(Aa)(v2+D)/Land divide the parameter space of (m1,m2) as follows: Δ0:={(m1,m2):m1>m1,m2>m2},Δ1:={(m1,m2):0<m1<m1,m2>m2},Δ2:={(m1,m2):0<m2<m2,m1>m1},Δ3:={(m1,m2):0<m1<min{m1,m1},0<m2<min{m2,m2}},Δ4:={(m1,m2):0<m1<m1,m2<m2<m2},Δ5:={(m1,m2):0<m1<m1,m2<m2<m2}.Here, mi,mi,i=1,2 are the critical loss rates for autotrophs and mixotrophs from extinction to survival in an aquatic ecosystem.

The functional responses are the most important component between predator and prey. Choose (Equation2) and let (10) G(A)=Aαβα+Aα.(10) We consider two classical functional responses: α=1 (Holling type II functional response) and α=2 (Holling type III functional response). Figure  shows the parameter regions in the (m1,m2)-plane about survival of autotrophs and mixotrophs for different Holling type functional responses. The extinction of both autotrophs and mixotrophs with nutrients reaching the fixed external dissolved nutrients Nin (E0) can occur if autotrophic and mixotrophic loss rates are very large (see Δ0 and Theorem 3.2). The coexistence of autotrophs and dissolved nutrients with the extinction of mixotrophs (E1), and the coexistence of mixotrophs and dissolved nutrients with the extinction of autotrophs (E2) are both possible outcomes of model (Equation3) for different parameter ranges (see Δ1, Δ2 and Theorems 3.3, 3.4). Autotrophs, mixotrophs and nutrients appear together in aquatic ecosystems (see Δ3 and Theorem 3.6). The numerical simulations indicate that there are bistable structures in regions Δ1Δ2 and Δ5, and autotrophs and mixotrophs also coexist in Δ4.

Model (Equation3) has extremely complex dynamic behaviours. From Figure , one can see that there are two bistable cases for different values α. The first case is that the semi-trivial steady states Ea and Em are bistable in Δ1Δ2 with respect to different initial values and α=1 (see Figure (a)). This is due to the competition between autotrophs and mixotrophs for light and nutrients. Another case is that Ea and Ec or an oscillatory periodic solution are bistable in Δ5 for α=2 (see Figure (b,c)). This suggests that the initial values of the model can also determine the coexistence of autotrophs and mixotrophs. From the point of view of actual ecosystems, the initial biomass of autotrophic algae and mixotrophic algae directly affects the final results of competition and coexistence between them.

Figure 3. Bistability of model (Equation3). (a) Ea and Em are bistable in Δ1Δ2 for Iin=400, m1=0.02, m2=0.04; (b) Ea and Ec are bistable in Δ5 for Iin=400, m1=0.05, m2=0.05; (c) Ea and a positive periodic solution are bistable in Δ5 for Iin=150, m1=0.05, m2=0.05. Here, other parameters are from Table .

Figure 3. Bistability of model (Equation3(3) dAdt=r1F1(N)H1(Iin,A,M)A⏟growthofautotrophs−m1A⏟loss−v1+DLA⏟sinkingandexchange−aG(A)M,⏟predationofmixotrophsdMdt=r2F2(N)H2(Iin,A,M)M+eaG(A)M⏟growthofmixotrophs−m2M⏟loss−v2+DLM,⏟sinkingandexchangedNdt=DL(Nin−N)⏟nutrientinputandexchange+δ1θ1m1A+δ2θ2m2M⏟nutrientrecycling−θ1r1F1(N)H1(Iin,A,M)A⏟consumptionbyautotrophs−θ2r2F2(N)H2(Iin,A,M)M⏟consumptionbymixotrophs.(3) ). (a) Ea and Em are bistable in Δ1∩Δ2 for Iin=400, m1=0.02, m2=0.04; (b) Ea and Ec are bistable in Δ5 for Iin=400, m1=0.05, m2=0.05; (c) Ea and a positive periodic solution are bistable in Δ5 for Iin=150, m1=0.05, m2=0.05. Here, other parameters are from Table 1.

By comparing (a) and (b) in Figure , one can observe that there are similarities and many differences between Holling type II functional response (α=1) and Holling type III functional response (α=2). They all have four situations: algae extinction, only autotrophs, only mixotrophs and coexistence. But there are also some differences between them. One difference is that there are a bistable region Δ1Δ2 of Ea and Em for α=1, and a bistable region Δ5 of Ea and Ec or a positive periodic solution for α=2. Another difference is that autotrophs and mixotrophs have a larger coexistence area Δ3Δ4 for α=2. This means that Holling III type functional response is conducive to the coexistence of autotrophs and mixotrophs under the framework of intraguild predation.

The water surface light intensity Iin and the nutrient input concentration Nin are two important abiotic factors in aquatic ecosystems. They will change with the seasons and climate. Figure  shows the influence of Iin and Nin on the ecological reproductive indexes RiA,RiM, i = 0, 1 for different Holling type functional responses. From Figure (a,c), it is found that R0A,R0M,R1A will increase and exceed the threshold 1 if Iin and Nin raise. While R1M exhibits a rapid increase above 1, followed by a rapid decrease below 1. This makes the region with R:=min{R0A,R0M,R1A,R1M}>1 very small. On the contrary, it can be seen from Figure (b) that there is a unique threshold value about light Iin so that R>1 if Iin>Iin. Similarly there also is a unique threshold value about nutrient Nin so that R>1 if Nin>Nin (see Figure (d)). The above analysis shows that high light intensity Iin and nutrient concentration Nin are beneficial to the existence of autotrophs and mixotrophs for Holling III type functional response. This research also suggests that the intraguild predation structure can be used to explain the phytoplankton paradox, especially when (Equation10) is Holling III type functional response.

Figure 4. Influence of the water surface light intensity Iin and the nutrient input concentration Nin on the ecological reproductive indexes RiA,RiM, i = 0, 1. (a,b) Nin=80; (c,d) Iin=400. Here, m2=0.6 and other parameters are from Table .

Figure 4. Influence of the water surface light intensity Iin and the nutrient input concentration Nin on the ecological reproductive indexes RiA,RiM, i = 0, 1. (a,b) Nin=80; (c,d) Iin=400. Here, m2=0.6 and other parameters are from Table 1.

4.2. Biomass density

The biomass density of autotrophic algae and mixotrophic algae as an important indicator of algal blooms is closely related to some important environmental parameters. For example, L and D are parameters relevant to water depth and movement; Iin and k0 are parameters about light; Nin,θ1,θ2 are parameters related to nutrients; δ1,δ2 are parameters related to water temperature and organic carbon degradation; Parameter a describes the heterotrophic activity of mixotrophs. In the following study, we use the Holling type II functional response (α=2) and explore the effects of environmental factors in system (Equation3) on the density of autotrophs and mixotrophs.

From Figure (a), one can observe that if the surface light density Iin is very low, both autotrophic algae and mixotrophic algae will become extinct (E0). As light intensity increases gradually, autotrophs can invade the aquatic ecosystems and occupy a dominant position (Ea). When the light intensity is high, there is a regime transition from autotrophs to mixotrophs, and mixotrophs dominate the whole aquatic ecosystems. In this process, the numerical bifurcation diagram implies that there exist two stability switches for Iin. Figure (b) shows a similar process for the nutrient input concentration Nin from E0, Ea to Ec. The reason for the above phenomenons is that although high light intensity and nutrient input concentration are conducive to the growth of autotrophs and mixotrophs, the high biomass of mixotrophic algae will cause the rapid decrease of the biomass of autotrophic algae due to the predation between the two. Therefore, mixotrophic algae are more prone to blooms for high light intensity and nutrient concentration because of intraguild predation.

Figure 5. Bifurcation diagrams of autotrophs and mixotrophs for Iin(0,600) and Nin(0,200). Here, other parameters are from Table .

Figure 5. Bifurcation diagrams of autotrophs and mixotrophs for Iin∈(0,600) and Nin∈(0,200). Here, other parameters are from Table 1.

The light attenuation coefficient k0 is used to describe the turbidity of water quality. From Figure (a), we find that the increase of k0 causes a positive effect on autotrophs and a negative effect on mixotrophs. This means that the clear water is more favourable for mixotrophs in the framework of intraguild predation. The attack rate a is an important parameter to assess the heterotrophic activity of mixotrophs. Figure (b) shows three stages: autotroph-only Ea, the bistability of Ea and Ec and the coexistence Ec. When the value of a is low, which implies that mixotrophs mainly carry out autotrophic activities, the biomass of autotrophs is high, but that of mixotrophs is low. This suggests that autotrophs are a powerful competitor and limit the growth of mixotrophs in an aquatic ecosystem since mixotrophs have a lower growth rate. When the attack rate a varies in high level, we observed the opposite phenomenon because of the enhancement of predation. In the bistable stage, the final biomass density of autotrophs and mixotrophs is directly determined by their initial biomass. Therefore, high heterotrophic activity raises the probability of mixotrophic algal blooms, and conversely, the possibility of autotrophic algal blooms will increase.

Figure 6. Bifurcation diagrams of autotrophs and mixotrophs for k0(0.1,0.9) and a(0,0.1). Here, other parameters are from Table .

Figure 6. Bifurcation diagrams of autotrophs and mixotrophs for k0∈(0.1,0.9) and a∈(0,0.1). Here, other parameters are from Table 1.

The parameter δ1 and δ2 describe the nutrient recycling proportions, which are closely related and proportional to the degradation rate of bacterial organic carbon and the internal temperature of the water. It has been shown that the increase of the nutrient recycling proportion causes a sharp increase in the biomass of algae, and finally leads to the occurrence of algal blooms [Citation27, Citation28]. Figure  exhibits a different situation from the past. The biomass density of autotrophs and mixotrophs does not change, or only slightly increases. This is mainly attributed to the intraguild predation structure between autotrophs and mixotrophs. Mixotrophs control autotrophic algae by predation. On the contrary, autotrophs suppress mixotrophic algae through competition.

Figure 7. Bifurcation diagrams of autotrophs and mixotrophs for δ1(0,1) and δ2(0,1). Here, other parameters are from Table .

Figure 7. Bifurcation diagrams of autotrophs and mixotrophs for δ1∈(0,1) and δ2∈(0,1). Here, other parameters are from Table 1.

The quota of nutrients to carbon in algal cells describes the quality of algae, and it is constantly changing [Citation10, Citation11, Citation17]. From Figure (a), we find that when θ1 varies in low level, the biomass density of mixotrophs is higher than that of autotrophs when they coexist. Although the growth of autotrophs needs lower nutrients, which means that they should have higher biomass, the biomass of autotrophs did not increase significantly. This is mainly due to the predation of mixotrophs. When θ1 increases in high level, we get an opposite result. Mixotrophs are extinct, and autotrophs have become a dominator in aquatic ecosystems. In this case, autotrophs consume more nutrients, leading to the extinction of mixotrophs through competition. In comparison, θ2 only causes little effect on the biomass density of autotrophs and mixotrophs (see Figure (b)).

Figure 8. Bifurcation diagrams of autotrophs and mixotrophs for θ1(0,0.04) and θ2(0,0.02). Here, other parameters are from Table .

Figure 8. Bifurcation diagrams of autotrophs and mixotrophs for θ1∈(0,0.04) and θ2∈(0,0.02). Here, other parameters are from Table 1.

The water depth L has an important effect on the biomass density of autotrophs and mixotrophs. With the increase of the water depth L, model (Equation3) has four situations: autotroph-only Ea, the bistability of Ea and Ec, the coexistence Ec, and the extinction E0 (see Figure (a)). Their total biomass is gradually decreasing. It also confirms once again that the shallow aquatic systems are more prone to algal blooms. The exchange rate D characterizes the water movement, which is mainly caused by wind and the density difference of water body. A low value of D means less nutrient input. Mixotrophs are extinct since they are at a disadvantage in competing with autotrophs for nutrients. The high exchange rate (D) is conducive to the coexistence of autotrophs and mixotrophs, and mixotrophs have more biomass (see Figure (b)).

Figure 9. Bifurcation diagrams of autotrophs and mixotrophs for D(0,0.04) and L(0,30). Here, other parameters are from Table .

Figure 9. Bifurcation diagrams of autotrophs and mixotrophs for D∈(0,0.04) and L∈(0,30). Here, other parameters are from Table 1.

In summary, we let BDA represents the biomass density of autotrophic algae, and BDM represents the biomass density of mixotrophic algae. PAB characterizes the probability of autotrophic algal blooms, and PMB describes the probability of mixotrophic algal blooms. The influence of partial parameters on the growth of autotrophs and mixotrophs can be summarized in Table .

Table 2. Example of a table showing that its caption is as wide as the table itself and justified.

5. Conclusion and discussion

In this paper, we propose and investigate a mathematical model (Equation3) describing the interactions of autotrophs (autotrophic algae) and mixotrophs (mixotrophic algae) with the intragulid predation structure in a well-mixed aquatic environment. Mixotrophs compete with autotrophs for light and nutrients, and also consume micro-autotrophs. The ecological reproductive indexes for the extinction, survival and coexistence of autotrophs and mixotrophs are rigorously derived. All possible results of system (Equation3) are described in terms of parameters (m1,m2) and summarized in Figure . This research indicates that the intraguild predation structure is beneficial to phytoplankton biodiversity, and provides an explanation for the phytoplankton paradox.

The different Holling type functional responses and some important environmental parameters could influence the coexistence and the biomass density of autotrophs and mixotrophs. Our results show that Holling III type functional response is conducive to the coexistence of autotrophs and mixotrophs under the intraguild predation structure. There exists the critical light intensity Iin and the critical nutrient input concentration Nin that affect the coexistence of autotrophs and mixotrophs (see Figure ). The effect of environmental factors on the metabolism of autotrophs and mixotrophs is complicated (see Table ). These results are helpful for protecting water quality and evaluating algae blooms.

The coexistence mechanism of autotrophs and mixotrophs is very complex. It has stable positive steady-state solutions, periodic solutions generated by Hopf bifurcation and double Hopf bifurcations, and multiple stability switches. In the theoretical analysis, we only prove the existence of positive steady-state solutions. Rigorously proving the remaining coexistence mechanism is an interesting and important question. Ecological stoichiometry is a very important method and tool, including energy flow and element cycle [Citation10, Citation11, Citation18, Citation23]. Previous studies have considered stoichiometric autotroph–mixotroph interactions without light [Citation15, Citation16, Citation24, Citation25]. It is of interest to add stoichiometric mechanism into system (Equation3) to investigate some new findings.

The turbulence of the lake changes with the water depth [Citation26]. In the upper layer, algae mixes better since the turbulence is intense, while the turbulence is relatively small in the lower layer, and algae exhibit spatial heterogeneity. The interesting problem is to explore the interactions of autotrophs and mixotrophs with light and nutrients in a water ecosystem with spatial heterogeneity. In view of the present analysis and model (Equation3), there are more interesting ecological problems worthy of careful discussion. For example, adding dissolved inorganic carbon [Citation28], zooplankton and fishes [Citation11, Citation12] and the effect of toxic plankton species [Citation6].

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This research is partially supported by the National Natural Science Foundation of China [grant numbers 11801052 and 11971088], the Natural Science Foundation of Liaoning Province [grant number 2019-ZD-1056], the Natural Science Foundation of Heilongjiang Province [grant number LH2019A022] and the Fundamental Research Funds for the Central Universities [grant number 3132022202].

References

  • M. Chen, M. Fan, R. Liu, X.Y. Wang, X. Yuan, and H.P. Zhu, The dynamics of temperature and light on the growth of phytoplankton, J. Theor. Biol. 385 (2015), pp. 8–19.
  • K.W. Crane and J.P. Grover, Coexistence of mixotrophs, autotrophs, and heterotrophs in planktonic microbial communities, J. Theor. Biol. 262 (2010), pp. 517–527.
  • K.F. Edwards, Mixotrophy in nanoflagellates across environmental gradients in the ocean, Proc. Natl. Acad. Sci. 116 (2019), pp. 6211–6220.
  • C.M. Heggerud, H. Wang, and M.A. Lewis, Transient dynamics of a stoichiometric cyanobacteria model via multiple-scale analysis, SIAM J. Appl. Math. 80 (2020), pp. 1223–1246.
  • R.D. Holt and G.A. Polis, A theoretical framework for intraguild predation, Am. Nat. 149 (1997), pp. 745–764.
  • S.B. Hsu, F.B. Wang, and X.Q. Zhao, A reaction-diffusion model of harmful algae and zooplankton in an ecosystem, J. Math. Anal. Appl. 451 (2017), pp. 659–677.
  • J. Huisman and F.J. Weissing, Light-limited growth and competition for light in well-mixed aquatic environments: An elementary model, Ecology 75 (1994), pp. 507–520.
  • J. Huisman and F.J. Weissing, Competition for nutrients and light in a mixed water column: A theoretical analysis, Am. Nat. 146 (1995), pp. 536–564.
  • G.E. Hutchinson, The paradox of the plankton, Am. Nat. 95 (1961), pp. 137–145.
  • X. Li, H. Wang, and Y. Kuang, Global analysis of a stoichiometric producer-grazer model with Holling type functional responses, J. Math. Biol. 63 (2011), pp. 901–932.
  • I. Loladze, Y. Kuang, and J.J. Elser, Stoichiometry in producer-grazer systems: Linking energy flow with element cycling, Bull. Math. Biol. 62 (2000), pp. 1137–1162.
  • D.Y. Lv, M. Fan, Y. Kang, and K. Blanco, Modeling refuge effect of submerged macrophytes in lake system, Bull. Math. Biol. 78 (2016), pp. 662–694.
  • K. Mischaikow, H. Smith, and H.R. Thieme, Asymptotically autonomous semiflows: Chain recurrence and Lyapunov functions, Trans. Am. Math. Soc. 347 (1995), pp. 1669–1685.
  • H.V. Moeller, M.G. Neubert, and M.D. Johnson, Intraguild predation enables coexistence of competing phytoplankton in a well-mixed water column, Ecology 100 (2019), Article ID e02874.
  • H. Nie, S.B. Hsu, and F.B. Wang, Steady-state solutions of a reaction-diffusion system arising from intraguild predation and internal storage, J. Differ. Equ. 266 (2019), pp. 8459–8491.
  • H. Nie, S.B. Hsu, and F.B. Wang, Global dynamics of a reaction-diffusion system with intraguild predation and internal storage, Discrete Cont. Dyn. Syst. Ser. B 25 (2020), pp. 877–901.
  • A. Peace, Effects of light, nutrients, and food chain length on trophic efficiencies in simple stoichiometric aquatic food chain models, Ecol. Model. 312 (2015), pp. 125–135.
  • A. Peace and H. Wang, Compensatory foraging in stoichiometric producer-grazer models, Bull. Math. Biol. 81 (2019), pp. 4932–4950.
  • R. Peng and X.Q. Zhao, A nonlocal and periodic reaction–diffusion–advection model of a single phytoplankton species, J. Math. Biol. 72 (2016), pp. 755–791.
  • G.A. Polis and R.D. Holt, Intraguild predation: The dynamics of complex trophic interactions, Trends Ecol. Evol. 7 (1992), pp. 151–154.
  • A.B. Ryabov, L. Rudolf, and B. Blasius, Vertical distribution and composition of phytoplankton under the influence of an upper mixed layer, J. Theor. Biol. 263 (2010), pp. 120–133.
  • F.R. Vasconcelos, S. Diehl, P. Rodríguez, P. Hedström, J. Karlsson, and P. Byström, Asymmetrical competition between aquatic primary producers in a warmer and browner world, Ecology97 (2016), pp. 2580–2592.
  • H. Wang, H.L. Smith, Y. Kuang, and J.J. Elser, Dynamics of stoichiometric bacteria-algae interactions in the epilimnion, SIAM J. Appl. Math. 68 (2007), pp. 503–522.
  • S. Wilken, J.M.H. Verspagen, S. Naus-Wiezer, E. Van Donk, and J. Huisman, Biological control of toxic cyanobacteria by mixotrophic predators: An experimental test of intraguild predation theory, Ecol. Appl. 24 (2014), pp. 1235–1249.
  • S. Wilken, J.M.H. Verspagen, S. Naus-Wiezer, E. Van Donk, and J. Huisman, Comparison of predator-prey interactions with and without intraguild predation by manipulation of the nitrogen source, Oikos 123 (2014), pp. 423–432.
  • A. Wüest and A. Lorke, Small-scale hydrodynamics in lakes, Annu. Rev. Fluid Mech. 35 (2003), pp. 373–412.
  • J.M. Zhang, J.P. Shi, and X.Y. Chang, A mathematical model of algae growth in a pelagic-benthic coupled shallow aquatic ecosystem, J. Math. Biol. 76 (2018), pp. 1159–1193.
  • J.M. Zhang, J.P. Shi, and X.Y. Chang, A model of algal growth depending on nutrients and inorganic carbon in a poorly mixed water column, J. Math. Biol. 83 (2021), p. 15.
  • J.M. Zhang, J.D. Kong, J.P. Shi, and H. Wang, Phytoplankton competition for nutrients and light in a stratified lake: A mathematical model connecting epilimnion and hypolimnion, J. Nonlinear Sci. 31 (2021), p. 35.
  • X.Q. Zhao, Dynamical Systems in Population Biology, Springer, New York, 2003.