1,518
Views
3
CrossRef citations to date
0
Altmetric
Research Article

Hopf bifurcation in delayed nutrient-microorganism model with network structure

, , &
Pages 1-13 | Received 26 Apr 2021, Accepted 23 Nov 2021, Published online: 10 Jan 2022

Abstract

In this paper, we introduce and deal with the delayed nutrient-microorganism model with a random network structure. By employing time delay τ as the main critical value of the Hopf bifurcation, we investigate the direction of the Hopf bifurcation of such a random network nutrient-microorganism model. Noticing that the results of the direction of the Hopf bifurcation in a random network model are rare, we thus try to use the method of multiple time scales (MTS) to derive amplitude equation and determine the direction of the Hopf bifurcation. It is showed that the delayed random network nutrient-microorganism model can exhibit a supercritical or subcritical Hopf bifurcation. Numerical experiments are performed to verify the validity of the theoretical analysis.

1. Introduction

The existence of the Hopf bifurcation in ecological systems is a hot investigation field, since its practical significance in biological control. Therefore, there are many results concerning the Hopf bifurcation in ecological systems that have been reported, see Refs. [Citation5–7,Citation12]. One of the most important results induced by the Hopf bifurcation is periodic solutions, including spatially homogeneous and spatially non-homogeneous periodic solutions. As a result, it is necessary to study or determine the stability of the periodic solution. As we know, there are two common techniques to study the direction of the Hopf bifurcation in a delayed differential equation, they are centre manifold reduction (CMR) and MTS, see Refs. [Citation2,Citation8,Citation11].

It is noticed that there are few results about the direction of the Hopf bifurcation in a delayed reaction-diffusion system with network structure have been reported [Citation9,Citation10], where the direction of the Hopf bifurcation is determined by employing the method of CMR. However, different from the technique is adopted in [Citation9,Citation10], we mainly attempt to use MTS to compute amplitude equation, and determine the direction of the Hopf bifurcation in a delayed nutrient-microorganism model with random network structure. More precisely, the system we consider takes the form (1) {duidt=d1j=1NΔijui+βui(uiviui+K1),dvidt=d2j=1NΔijvi+αvi(tτ)ui2viui+K,ui(0)>0,vi(0)>0.(1) This model is called nutrient-microorganism system in the sediment, and first proposed by Baurmann and Feudel in [Citation1](continuous form). All parameters d1,d2,α,β,K in (Equation1) are positive constants; τ0 is time delay, it implies that some time delay is required for microorganism to consume nutrient in the sediment.

For system (Equation1), we assume that it is defined on an undirected network with N nodes and there are no self-loops; ui and vi are the densities of the microorganism and the nutrient on node i, respectively; Δ is the N×N discrete Laplacian matrix of network with its elements Δij=kiδijLij, where L is the adjacency matrix encoding the network connection, this indicates it satisfies Lij=1 if there is a link connecting from patch i to patch j. If not, Lij=0 when there is no any link connecting from patch i to patch j, and δij is the Kronecker's delta [Citation4]. Moreover, the degree of the ith node is defined by ki=j=1NLij, and the connection probability between node i and node j for ij is p(0p1).

This paper is structured as follows. In Section 2, we establish the existence results of the Hopf bifurcation of the model (1). In Section 3, the amplitude equation of the Hopf bifurcation is deduced. As a result, the supercritical or the subcritical Hopf bifurcation can be yield by analysing the amplitude equation. We perform the numerical simulations to verify the theoretical analysis in Section 4, and some discussions are made in Section 5.

2. Existence of the Hopf bifurcation

In this section, we give some conditions to ensure the existence of the Hopf bifurcation of the networked model (1). To this end, we first consider the existence and stability of positive equilibria for model (1).

Lemma 2.1

[Citation3]

The possible positive equilibria of model (1) can be found as follows

(i)

if α<1+2K, model (1) has no positive equilibrium;

(ii)

if α>1+2K, model (1) has two positive equilibria E1=(u1,v1) and E2=(u2,v2). Here v1=αu1,v2=αu2 and u1=12(α1+(α1)24K),u2=12(α1(α1)24K),

(iii)

if α=1+2K, model (1) has one positive equilibrium E3=(u3,v3)=(K,1+K).

Lemma 2.2

[Citation3]

When diffusion and delay are absent, the following statements are valid.

(i)

If α=1+2K, the positive equilibrium E3 is unstable when β>2+1K;

(ii)

If α>1+2K, the positive equilibrium E2 is unstable and the positive equilibrium E1 is locally asymptotically stable when β<β0H and unstable when β>β0H, where β0H=αu1K.

Thereby, in view of Lemma 2.1 and Lemma 2.2, we only focus on the dynamical behaviours near E1, and we denote it by E1E=(u,v) for simplicity. Now, we shall perform the linear stability analysis of system (1) near the positive equilibrium E=(u,v). For this purpose, let Γui=uiu and Γvi=viv be the small perturbations, then the linear system of (1) evaluated at E=(u,v) can be written as follows (2) (Γu˙iΓv˙i)=J1(ΓuiΓvi)+J2(ΓuiτΓviτ)+D(j=1NΔijΓujj=1NΔijΓvj),(2) where ui=ui(t),uiτ=uiτ(tτ) and J1=(βKu+Kβuv1Ku+Kuv),J2=(0001),D=(d100d2).Let 0=Λ1>Λ2>>ΛN be the eigenvalues of the discrete Laplacian matrix Δ, and suppose that Lϕ={ϕi}i=1N is the subspace generated by the eigenfunctions associated to the topological eigenvalue Λi. Then, the general solution of system (2) can be rewritten as (ΓuiΓvi)=i=1N(ξ1ξ2)ϕieλit,with j=1NΔijϕj=Λiϕi. Inserting them into (2), one has (βKu+K+d1Λiβuvi1Ku+Kuveλiτ+d2Λi)(ξ1ξ2)=λi(ξ1ξ2).It then follows that (3) λi2+AΛiλi+(λi+BΛi)eλiτ+CΛi=0,(3) where AΛi=(d1+d2)Λi+uvβKu+K,BΛi=d1ΛiβKu+K,CΛi=d1d2Λi2(ud1vd2βKu+K)Λi+βuv.Let λ=iω(ω>0) be the solution of Equation (3), we have ω4+(AΛi22CΛi1)ω2+CΛi2BΛi2=0.Thereby, if one of the conditions in the following (H0)C2B2<0,(H1)C2B2=0,A22C1<0,(H2)C2B2>0,A22C1<0,(A22C1)24(C2B2)=0,is satisfied, one yields (4) ωΛi2=12(1+2CΛiAΛi2+(AΛi22CΛi1)24(CΛi2BΛi2)).(4) We thus obtain the critical value τ of the Hopf bifurcation is (5) τΛij={1ωΛiarccos{(BΛiAΛi)ωΛi2BΛiCΛiωΛi2+BΛi2}+2jπωΛi,sinωΛiτΛij>0,1ωΛiarccos{(BΛiAΛi)ωΛi2BΛiCΛiωΛi2+BΛi2}+2(j+1)πωΛi,sinωΛiτΛij<0,(5) where jN0={0,1,2,}. In addition, a straightforward calculation shows that Re{(dλdτ)1}τ=τΛij=Re{(AΛi+2λ)eλτ+1λ(λ+BΛi)τλ}τ=τΛij=Re{(AΛi+2λ)eλτ+1λ(λ+BΛi)}τ=τΛij=Re{(AΛi+2iωΛi)(cosωΛiτΛij+isinωΛiτΛij)+1iωΛi(iωΛi+BΛi)}=AΛi22CΛi1+2ωΛi2ωΛi2+BΛi2=(AΛi22CΛi1)24(CΛi2BΛi2)ωΛi2+BΛi2>0.Moreover, note that Sign{Re{(dλdτ)1}τ=τΛij}=Sign{Re{dλdτ}τ=τΛij},this implies that the transversality condition of the Hopf bifurcation is satisfied. Combine this with (Equation5) we claim that the delayed network model (1) undergoes the Hopf bifurcation at E when τ=τΛij for 0=Λ1>Λ2>>ΛN and jN0. Especially, the delayed network model (1) admits a Hopf bifurcation at E when τ=τ00, and in this case the periodic solution bifurcated from the Hopf bifurcation is spatially homogeneous.

3. Direction of the Hopf bifurcation

In this section, we shall employ MTS to derive amplitude equations and determine the direction of the Hopf bifurcation. We first define τ00τ,ω0ω and τt/τ, where ω0 and τ00 can be found in (Equation4) and (Equation5), respectively. Now, a rewritten form of the networked system (1) can be read as (6) U˙i=τLtUi+τLτUiτ+τF(Ui),(6) where we set Ui=(ui,vi)T,Uiτ=(uiτ(t1),viτ(t1))T, and Lt=(d1j=1NΔij+a11a12a21d2j=1NΔij+a22),Lτ=(0001),with a11=βKu+K,a12=βuv,a21=1Ku+K,a22=uv.In addition, the nonlinear term F(Ui)=(F(1)(Ui),F(2)(Ui))T takes the form F(1)(Ui)=fuiuiui2+fuiviuivi+fuiuiuiui3+fuiuiviui2vi+O(4),F(2)(Ui)=guiuiui2+guiviuivi+guiuiuiui3+guiuiviui2vi+O(4),with fuiui=βvK2(u+K)3,fuiuiui=βvK2(u+K)4,fuiuivi=βK2(u+K)3,fuivi=βu(u+2K)(u+K)2,guiuiui=vK2(u+K)4,guiui=vK2(u+K)3,guiuivi=K2(u+K)3,guivi=u(u+2K)(u+K)2.To employ the technique of MTS, let ε be a small perturbation parameter. Then introducing the time scales T0=t,T2=ε2t, this induces that (7) t=T0+ε2T2.(7) Solution Ui can be decribed as (8) Ui(T0,T2)=(ui(T0,T2)vi(T0,T2))=ε(ui(1)(T0,T2)vi(1)(T0,T2))+ε2(ui(2)(T0,T2)vi(2)(T0,T2))+ε3(ui(3)(T0,T2)vi(3)(T0,T2))+O(4).(8) We thus obtain a fact that (9) Uiτ(T01,T2)=(uiτ(T01,T2)viτ(T01,T2))=j=13εj(uiτ(j)(T01,T2)viτ(j)(T01,T2))ε3T2(uiτ(1)(T01,T2)viτ(1)(T01,T2))+.(9) By a similar manner, we can write the nonlinear term F as follows (10) F=ε2F(2)+ε3F(3)+O(ε4),(10) where F(2)=(fuiui(ui(1))2+fuiviui(1)vi(1)guiui(ui(1))2+guiviui(1)vi(1)),and F(3)=(2fuiuiui(1)ui(2)+fuivi(ui(1)vi(2)+ui(2)vi(1))+fuiuiui(ui(1))3+fuiuivi(ui(1))2vi(1)2guiuiui(1)ui(2)+guivi(ui(1)vi(2)+ui(2)vi(1))+guiuiui(ui(1))3+guiuivi(ui(1))2vi(1)),Next, we introduce the small perturbation of the Hopf bifurcation parameter τ=τ+ε2δ with δ>0. Keep this in mind, denote Lt|τ=τ=Lt,Lτ|τ=τ=Lτ and put (7)-(10) into (6), one has

O(ε): (11) T0(ui(1)vi(1))τLt(ui(1)vi(1))τLτ(uiτ(1)viτ(1))=0.(11) O(ε2): (12) T0(ui(2)vi(2))τLt(ui(2)vi(2))τLτ(uiτ(2)viτ(2))=τF(2).(12) O(ε3): (13) T0(ui(3)vi(3))τLt(ui(3)vi(3))τLτ(uiτ(3)viτ(3))=T2(ui(1)vi(1))τLτT2(uiτ(1)viτ(1))+δLt(ui(1)vi(1))+δLτ(uiτ(1)viτ(1))+τF(3).(13) Considering the solution of perturbation Equation (Equation11) near the Hopf mode, we have (14) (ui(1)vi(1))=H(T2)peiωτT0+H¯(T2)p¯eiωτT0,(14) where we assume that H(T2) is the complex amplitude and p=(1iωa11a12)(p11p12).We thus obtain (ui(1))2=H2(T2)e2iωτT0+2|H(T2)|2+H¯2(T2)e2iωτT0,ui(1)vi(1)=p12H2(T2)e2iωτT0+2Re{p12}|H(T2)|2+p¯12H¯2(T2)e2iωτT0.This means that the perturbation Equation (Equation12) has a particular solution with the form (15) (ui(2)vi(2))=H2(T2)(q11q12)e2iωτT0+(q21q22)|H(T2)|2+H¯2(T2)(q¯11q¯12)e2iωτT0,(15) where the vectors (q11,q12)T and (q21,q22)T are unknown. Thence, to determine them we should insert (Equation15) into (Equation12). Then equating like terms of e2iωτT0H2(T2) and |H(T2)|2, we have q11=a12(guiui+p12guivi)+(2iωa22+e2iωτ)(fuiui+p12fuivi)(2iωa11)(2iωa22+e2iωτ)a12a21,q12=(2iωa11)q11(fuiui+p12fuivi)a12,q21=2a12(guiui+Re{p12}guivi)2(a221)(fuiui+Re{p12}fuivi)a11(a221)a12a21,q22=a11q21+2(fuiui+Re{p12}fuivi)a12.As a result, it follows from (Equation14) and (Equation15) that ui(1)ui(2)=q11H3(T2)e3iωτT0+(q21+q11)H(T2)|H(T2)|2eiωτT0+(q¯21+q¯11)H¯(T2)|H(T2)|2eiωτT0+q¯11H¯3(T2)e3iωτT0,ui(1)vi(2)=q12H3(T2)e3iωτT0+(q22+q12)H(T2)|H(T2)|2eiωτT0+(q¯22+q¯12)H¯(T2)|H(T2)|2eiωτT0+q¯12H¯3(T2)e3iωτT0,ui(2)vi(1)=q11p12H3(T2)e3iωτT0+(q21p12+q11p¯12)H(T2)|H(T2)|2eiωτT0+q¯11p¯12H¯3(T2)e3iωτT0+(q¯21p¯12+q¯11p12)H¯(T2)|H(T2)|2eiωτT0,vi(1)vi(2)=q12p12H3(T2)e3iωτT0+(q22p12+q12p¯12)H(T2)|H(T2)|2eiωτT0+q¯12p¯12H¯3(T2)e3iωτT0+(q¯22p¯12+q¯12p12)H¯(T2)|H(T2)|2eiωτT0,(ui(1))3=H3(T2)e3iωτT0+3H(T2)|H(T2)|2eiωτT0+3H¯(T2)|H(T2)|2eiωτT0+H¯3(T2)e3iωτT0,(ui(1))2vi(1)=p12H3(T2)e3iωτT0+(p¯12+2p12)H(T2)|H(T2)|2eiωτT0+p¯12H¯3(T2)e3iωτT0+(2p¯12+p12)H¯(T2)|H(T2)|2eiωτT0.Now inserting all of them and (Equation14)–(Equation15) into the perturbation (Equation13), one obtains (16) T0(ui(3)vi(3))τLt(ui(3)vi(3))τLτ(uiτ(3)viτ(3))=weiωτT0+NST+c.c.,(16) where c.c. is the conjugation term, NST represents the non-secular term and w=(w(1),w(2))T with w(1)=H(T2)T2+δ(a11+a12p12)H(T2)+τψ1H(T2)|H(T2)|2,w(2)=(τeiωτ1)p12H(T2)T2+δ(a21+a22p12p12eiωτ)H(T2)+τψ2H(T2)|H(T2)|2,where ψ1=(q21+q11)fuiui+(q22+q12+q21p12+q11p¯12)fuivi+3fuiuiui+(p¯12+2p12)fuiuivi,ψ2=(q21+q11)guiui+(q22+q12+q21p12+q11p¯12)guivi+3guiuiui+(p¯12+2p12)guiuivi.Then perturbation Equation (Equation16) possesses its solution when the Fredholm alternative condition is satisfied. It implies that the following orthogonal condition should be satisfied (17) (p11p12),(w(1)w(2))=0,(17) where we take p11=eiωτiωa22eiωτ2iωa11a22,and p12=a12eiωτ2iωa11a22,the inner product satisfies <a,b>=a¯Tb. Then using (Equation17) and noticing p¯11+p12p¯12=1, we obtain (18) H(T2)T2=δη1H(T2)+η2H(T2)|H(T2)|2,(18) where we denote η1=(a11+a12p12)p¯11+(a21+a22p12p12eiωτ)p¯121p12p¯12τeiωτ,η2=τ(ψ1p¯11+ψ2p¯12)1p12p¯12τeiωτ.In what follows, for the purpose of investigating the Hopf bifurcation near τ=τ, we shall find the amplitude equation on the centre manifold, it is useful for investigating the Hopf bifurcation. To this end, let H(T2)=zeiωτT2 with z = x + iy, then it follows from (Equation18) that (19) {x˙=ωτy+δ(Re{η1}xIm{η1}y)+(Re{η2}xIm{η2}y)(x2+y2),y˙=ωτy+δ(Re{η1}yIm{η1}x)+(Re{η2}yIm{η2}x)(x2+y2),(19) where Re{} and Im{} represent the real part and imaginary part of •. Now, we let x=ρcosθ and y=ρsinθ, then (Equation19) becomes (20) {ρ˙=ρ(δK1+K2ρ2)+O(ρ4),θ˙=ωτ+O(|δ|,ρ2),(20) where K1=Re{η1} and K2=Re{η2}.

By virtue of (Equation20), we have the following results about the direction of the Hopf bifurcation.

Theorem 3.1

The direction of the Hopf bifurcation depends on (Equation20). More precisely

(i)

if K1K2>0, then there is no Hopf bifurcation in the networked model (1);

(ii)

if K1K2<0, then the Hopf bifurcation exists, and K1 determines the direction of the Hopf bifurcation. Namely,

  • (ii-a) it is supercritical when K1>0, and periodic solution bifurcated from the Hopf bifurcation is stable;

  • (ii-b) it is subcritical when K1<0, and periodic solution bifurcated from the Hopf bifurcation is unstable;

Proof.

It is easy to verify that (Equation20) has a unique positive equilibrium (corresponding to the Hopf bifurcation) ρ=δK1K2, and thus its existence condition is K1K2<0. It indicates (i) is true. Moreover, Equation (Equation20) has a unique eigenvalue, say λ, and λ=2δK1. By employing linear stability analysis theory, we know that (ii) is valid. The proof is completed.

4. Numerical simulations

In this section, we main verify the effectiveness of Theorem 3.1 by numerical simulations. We assume that the numbers of the node are 100, namely we take N = 100 in the random network model (1). Moreover, we fix the connecting probability p = 0.35 between different nodes ui and uj for ij. We now choose the parameters in system (1) are α=1.5,β=2.5,K=0.05,d1=2 and d2=0.5, then one obtains the positive equilibrium E=(0.3618,1.1585),ω=1.5364,τ=1.1585,η1=0.5465+1.2577i and η2=0.324016.3341i. This means that K1=Re{η1}=0.5465>0 and K1K2=0.1771<0 are satisfied in Theorem 3.1.

Figure  shows that the distribution of solutions ui(1i100) with the development of moments. When taking time delay 0.95=τ<τ, we find that the positive equilibrium E is locally asymptotically stable, see Figure . Figure  suggests that there is a supercritical Hopf bifurcation, and the periodic solution with the spatial homogeneity bifurcated from the Hopf bifurcation is stable in the delayed network model (1), where we choose 1.1587=τ>τ. As such, the results in Theorem 3.1 are valid.

Figure 1. The distribution of solution ui with the numbers i=1,2,100 and the connection probability p = 0.35.

Figure 1. The distribution of solution ui with the numbers i=1,2,…100 and the connection probability p = 0.35.

Figure 2. Positive equilibrium E is stable when taking 0.95=τ<τ and the probability p = 0.35.

Figure 2. Positive equilibrium E∗ is stable when taking 0.95=τ<τ∗ and the probability p = 0.35.

Figure 3. The delayed network model (1) admits a stable spatially homogeneous periodic solution when choosing 1.1587=τ>τ and the probability p = 0.35.

Figure 3. The delayed network model (1) admits a stable spatially homogeneous periodic solution when choosing 1.1587=τ>τ∗ and the probability p = 0.35.

5. Conclusions

We deal with a nutrient-microorganism model with time delay and random network structure in this paper. By employing time delay τ as the critical parameter of the Hopf bifurcation, we explore its occurrence conditions. For the direction of the Hopf bifurcation, we try to adopt MTS to derive amplitude equation near τ=τ, it is found that the sign of K1K2 determines the existence of the Hopf bifurcation. Namely the Hopf bifurcation exists when K1K2<0, and there is no Hopf bifurcation when K1K2>0. Moreover, the sign of K1 determines the direction of the Hopf bifurcation with hypothesis K1K2<0. More precisely, the Hopf bifurcation is supercritical (resp. subcritical) and the periodic solution is stable (resp. unstable) when K1>0 (resp. K1<0). Numerical simulations indicate that our theoretical analysis is valid, and compared with the works done in [Citation9,Citation10] we claim that MTS is a easier technique to determine the direction of the Hopf bifurcation in a delayed network model than CMR. For more interesting results about this networked model, for example, resonant/nonresonant Hopf bifurcation, will be further considered.

Disclosure statement

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

Additional information

Funding

This work was supported by the National Natural Science Foundation of China [grant number 11971032], and Young Talent Support Project of Henan [grant number 2020HYTP012].

References

  • M. Baurmann and U. Feudel, Turing patterns in a simple model of a nutrient-microorganism system in the sediment, Ecol. Complex. 1 (2004), pp. 77–94.
  • M. Bentounsi, I. Agmour, and N. Achtaich, The Hopf bifurcation and stability of delayed predator-prey system, Comput. Appl. Math. 37(5) (2018), pp. 5702–5714.
  • M.X. Chen, R.C. Wu, B. Liu, and L.P. Chen, Hopf-Hopf bifurcation in the delayed nutrient-microorganism model, Appl. Math. Model. 86 (2020), pp. 460–483.
  • Y. Ide, H. Izuhara, and T. Machidac, Turing instability in reaction diffusion models on complex networks, Physica A. 457 (2016), pp. 331–347.
  • A. Kumar and B. Dubey, Modeling the effect of fear in a prey-predator system with prey refuge and gestation delay, Int. J. Bifurcat. Chaos. 29(14) (2019), pp. 1950195.
  • Y.Q. Liu, D.F. Duan, and B. Niu, Spatiotemporal dynamics in a diffusive predator prey model with group defense and nonlocal competition, Appl. Math. Lett. 103 (2020), pp. 106175.
  • K. Manna and M. Banerjee, Stability of Hopf-bifurcating limit cycles in a diffusion-driven prey-predator system with Allee effect and time delay, Math. Biosci. Eng. 16(4) (2019), pp. 2411–2446.
  • B. Roy, S.K. Roy, and D.B. Gurung, Holling-Tanner model with Beddington-DeAngelis functional response and time delay introducing harvesting, Math. Comput. Simul. 142 (2017), pp. 1–14.
  • Y.L. Shi, Z.H. Liu, and C.R. Tian, Hopf bifurcation in an activator-inhibitor system with network, Appl. Math. Lett. 98 (2019), pp. 22–28.
  • C.R. Tian, Z. Ling, and L. Zhang, Delay-driven spatial patterns in a network-organized semiarid vegetation model, Appl. Math. Comput. 367 (2020), pp. 124778.
  • T.H. Zhang, Y.L. Song, and H. Zang, The stability and Hopf bifurcation analysis of a gene expression model, J. Math. Anal. Appl. 395 (2012), pp. 103–113.
  • H.Y. Zhao and D.Y. Wu, Point to point traveling wave and periodic traveling wave induced by Hopf bifurcation for a diffusive predator-prey system, Discrete Cont. Dyn. Syst. S. 13(11) (2020), pp. 3271–3284.