Research article Special Issues

Synchronization of the neurons coupled with sequential developing electrical and chemical synapses

  • Received: 30 October 2021 Accepted: 09 December 2021 Published: 20 December 2021
  • There is some evidence representing the sequential formation and elimination of electrical and chemical synapses in particular brain regions. Relying on this feature, this paper presents a purely mathematical modeling study on the synchronization among neurons connected by transient electrical synapses transformed to chemical synapses over time. This deletion and development of synapses are considered consecutive. The results represent that the transient synapses lead to burst synchronization of the neurons while the neurons are resting when both synapses exist constantly. The period of the transitions and also the time of presence of electrical synapses to chemical ones are effective on the synchronization. The larger synchronization error is obtained by increasing the transition period and the time of chemical synapses' existence.

    Citation: Zhen Wang, Ramesh Ramamoorthy, Xiaojian Xi, Hamidreza Namazi. Synchronization of the neurons coupled with sequential developing electrical and chemical synapses[J]. Mathematical Biosciences and Engineering, 2022, 19(2): 1877-1890. doi: 10.3934/mbe.2022088

    Related Papers:

    [1] Christian Cortés García . Bifurcations in a discontinuous Leslie-Gower model with harvesting and alternative food for predators and constant prey refuge at low density. Mathematical Biosciences and Engineering, 2022, 19(12): 14029-14055. doi: 10.3934/mbe.2022653
    [2] Jinxing Zhao, Yuanfu Shao . Bifurcations of a prey-predator system with fear, refuge and additional food. Mathematical Biosciences and Engineering, 2023, 20(2): 3700-3720. doi: 10.3934/mbe.2023173
    [3] Peter A. Braza . A dominant predator, a predator, and a prey. Mathematical Biosciences and Engineering, 2008, 5(1): 61-73. doi: 10.3934/mbe.2008.5.61
    [4] Yuanfu Shao . Bifurcations of a delayed predator-prey system with fear, refuge for prey and additional food for predator. Mathematical Biosciences and Engineering, 2023, 20(4): 7429-7452. doi: 10.3934/mbe.2023322
    [5] Shuangte Wang, Hengguo Yu . Stability and bifurcation analysis of the Bazykin's predator-prey ecosystem with Holling type Ⅱ functional response. Mathematical Biosciences and Engineering, 2021, 18(6): 7877-7918. doi: 10.3934/mbe.2021391
    [6] Hongqiuxue Wu, Zhong Li, Mengxin He . Dynamic analysis of a Leslie-Gower predator-prey model with the fear effect and nonlinear harvesting. Mathematical Biosciences and Engineering, 2023, 20(10): 18592-18629. doi: 10.3934/mbe.2023825
    [7] Rajalakshmi Manoharan, Reenu Rani, Ali Moussaoui . Predator-prey dynamics with refuge, alternate food, and harvesting strategies in a patchy habitat. Mathematical Biosciences and Engineering, 2025, 22(4): 810-845. doi: 10.3934/mbe.2025029
    [8] Weijie Lu, Yonghui Xia, Yuzhen Bai . Periodic solution of a stage-structured predator-prey model incorporating prey refuge. Mathematical Biosciences and Engineering, 2020, 17(4): 3160-3174. doi: 10.3934/mbe.2020179
    [9] Tingting Ma, Xinzhu Meng . Global analysis and Hopf-bifurcation in a cross-diffusion prey-predator system with fear effect and predator cannibalism. Mathematical Biosciences and Engineering, 2022, 19(6): 6040-6071. doi: 10.3934/mbe.2022282
    [10] Rongjie Yu, Hengguo Yu, Chuanjun Dai, Zengling Ma, Qi Wang, Min Zhao . Bifurcation analysis of Leslie-Gower predator-prey system with harvesting and fear effect. Mathematical Biosciences and Engineering, 2023, 20(10): 18267-18300. doi: 10.3934/mbe.2023812
  • There is some evidence representing the sequential formation and elimination of electrical and chemical synapses in particular brain regions. Relying on this feature, this paper presents a purely mathematical modeling study on the synchronization among neurons connected by transient electrical synapses transformed to chemical synapses over time. This deletion and development of synapses are considered consecutive. The results represent that the transient synapses lead to burst synchronization of the neurons while the neurons are resting when both synapses exist constantly. The period of the transitions and also the time of presence of electrical synapses to chemical ones are effective on the synchronization. The larger synchronization error is obtained by increasing the transition period and the time of chemical synapses' existence.



    The predator-prey model is a model in biomathematics used, and modified, by many researchers to describe the dynamics between a prey and predator species in order to determine conditions in their parameters that allow for either the conservation or extinction of one or both species, where such models are generally posed by systems of ordinary [1,2,3,4,5] or partial differential equations [6,7,8,9,10].

    In particular, the Leslie-Gower models assumes that the intrinsic growth of both species is described by logistic functions, whose carrying capacity between prey and predators is subject to the capacity of the environment or directly proportional to the population size of the prey, respectively [11]. Similarly, this model assumes a functional response of the predator that is generally described by Holling-type functions [12,13,14,15,16].

    For example, assuming that the predator response functional is given by a Holling Ⅱ function f(x)=axx+d [17,18,19], where a>0 is the maximum per capita consumption rate of predators, and 0<d<K is the semi-saturating rate of capture, where predators can be harvested [20,21], the dynamics of population sizes of prey x(t)0 and predators y(t)0 is given by

    {˙x=rx(1xK)axyx+d˙y=sy[1ynx+c]qEy, (1.1)

    where K>0 is the carrying capacity of prey in the environment, which normally determines the resources available for prey survival, n>0 is the quantity of prey consumed by the predator and converted into new births, q>0 is the catchability coefficient, E>0 is the harvesting efforty, and r,s>0 are the intrinsic birth rates of prey and predators, respectively. In particular, if the diet of predators is based exclusively on prey consumption, then c=0 [22,23,24]; otherwise, c>0 describes the alternative food for predators, so their diet is not based exclusively on prey consumption [25,26].

    However, given the existence of prey that take refuge from the predator to subsist in the environment [1,29,30,31,32], and whose intrinsic growth is subject to an auxiliary parameter b>0 (half-saturation) that affects the hyperbolic function of the per capita birth rate curve of the prey [27,28], in addition to the maximum birth rate r>0, then model (1.1) is modified by

    {˙x=rxx+b(1xK)a(xP)y(xP)+d˙y=sy[1yn(xP)+c]qEy, (1.2)

    where P<K is the population size of refugia in prey. In particular, the intrinsic growth of prey approximates logistic growth for a large x; however, if x is small, the parameter b>0 influences its dynamics, of which it may represent examples such as inbreeding, ease of reproduction, difficulty in finding mates, or mating resistance at low temperatures.

    Observe that the model (1.2) is used if the initial prey population size is above the critical refuge population size, that is, if P<x(0). For the case where x(0)<P, the entire prey population size is refugeed from the predator whenever x<P, so there is no interaction between the two species and, hence, the functional response of the predator is deactivated when x<P. In this case, the dynamics of the prey is described by a logistic differential equation ˙x=rx(1xK) for the case in which there are either no external factors inhibiting its growth, or by an autonomous differential equation ˙x=rxx+b(1xK), until they exceed their critical refuge population size and the quantity xP becomes susceptible to predator consumption.

    On the other hand, if the predator's diet is based only on the consumption of prey, they are forced to change their diet due to the deficit of prey in order to survive in the environment [33,34,35], so the model (1.2) should be modified by

    Z0(x,y):{˙x=rxx+b(1xK)aϵ(xP)y(xP)+d˙y=sy[1ynϵ(xP)+(1ϵ)c]qEy, (1.3)

    with

    ϵ={1,ifx>P0,ifx<P. (1.4)

    However, if the predator's diet is not based exclusively on prey consumption, but has equal probability of consumption with the alternative food provided in the environment, the predator's diet is not affected and depends only on the alternative food provided by the environment in the absence of prey. In this case, the model (1.2) is modified by

    Zc(x,y):{˙x=rxx+b(1xK)aϵ(xP)y(xP)+d˙y=sy[1ynϵ(xP)+c]qEy, (1.5)

    with ϵ as described in (1.4).

    Consequently, and given the possible scenarios for predator growth and survival at low and high prey densities subject to their refuge, the objective of this paper is to perform a qualitative and bifurcation analyses to models (1.3) and (1.5), as shown in Sections 2 and 3, respectively, to determine similarities or difference in both models.

    Let x(t)0 and y(t)0 be considered as the population sizes of prey and predators, respectively, whose dynamics are given by the model (1.3), defined in the biological sense region

    Ω={(x,y)R2:0xK,0yM} (2.1)

    with M=max{n(KP),c}. In particular, the vector fields of the model (1.3) are given by

    Z0(x,y)={X(x,y)=(rxx+b(1xK)a(xP)y(xP)+dsy[1yn(xP)]qEy),x>PY(x,y)=(rxx+b(1xK)sy(1yc)qEy),x<P (2.2)

    where X and Y are defined in Σ+={(x,y)Ω:P<xK} and Σ={(x,y)Ω:0<x<P}, respectively.

    Since the model (2.2) is equivalent to a Filippov system [36,37,38,39], if f(x,y)=xP, Xf(p)=X(p),gradf(p) and Yf(p)=Y(p),gradf(p), then Xf(p)=Yf(p)=rPP+b(1PK)>0 for all pΣ={(x,y)Ω:x=P}. Therefore, since the vector field Y is composed of a system of autonomous and independent differential equations, the trajectories φZ0 with initial condition (x(0),y(0))Σ must cross Σ and remain in Σ+. Additionally, φZ0 with the initial condition p=(P,y(0))Σ is defined as φZ0(p,t)=φX(p,t) for all tI{t>0} and φZ0(p,t)=φY(p,t) for tI{t0}.

    Lemma 1. For an arbitrary initial condition (x(0),y(0))Ω, the model (2.2) has a unique φZ0 trajectory and remains in Ω.

    Proof. Since the vector fields X and Y are continuously differential and the trajectories φZ0 cross Σ or remain Σ+ for all initial conditions (x(0),y(0))Σ or (x(0),y(0))Σ+, by the existence and uniqueness theorem [40] for every vector field the uniqueness of the trajectories are guaranteed. Additionally, for all 0x<P, if x=0 then ˙x=0 for all y0. Similarly, if y=0, then ˙y=0, and if y=M, we have that ˙y0. Thus, the trajectories φZ0 do not cross Σ. Moreover, for all P<xK, if x=K, then ˙x=0; if y=0, then ˙y=0; and if y=n(KL)+c, then ˙y0, so φZ0 does not cross Σ+. The dynamics of the model at points (P,0) and (P,M) do not cross Ω and remain to be analyzed. In this case, since the trajectories φZ0 with the initial condition (x(0),y(0))ΣΣ+ cross Σ, we have that the trajectories φZ0 do not escape from the points (P,0) and (P,M).

    The equilibria in the model (2.2) over the coordinate axes are given by P0=(0,0)Σ, P1=(K,0)Σ+, and P2=(0,c(sqE)s)Σ if sqE>0. The local stability of the equilibria is summarized in the following result.

    Lemma 2. If sqE<0, then P0 and P1 are saddle points. If sqE>0, then P0 is a locally unstable node, P1 a locally stable node and P2 is a saddle point.

    Proof. The local stability of P1Σ+ and P0,P2Σ are guaranteed from the computation of the eigenvalues of the Jacobian matrix DX of the vector field X computed in P1Σ+ and of the Jacobian matrix DY of the vector field Y computed in P0,P2Σ.

    On the other hand, since the vector field Y has no interior equilibria, the intersections in Σ+ of nullclines

    y=f1(x):=nx(sqE)snP(sqE)s,y=f2(x):=rx(Pdx)(xK)aK(xP)(x+b), (2.3)

    are defined in the vector field X, and determine the number of interior equilibria in the model (2.2). In particular, note that the model (2.2) has no interior equilibria if sqE<0 since the nullclines (2.3) do not intersect in Σ+.

    For sqE>0, the possible positive roots of the polynomial

    Q(x)=Ax3+Bx2+Cx+D, (2.4)

    with

    A=rs+aKn(sqE)>0,B=rs(dKP)+aKn(b2P)(sqE),C=K[anP(2bP)(sqE)+rs(dP)],D=abnKP2(sqE)>0, (2.5)

    determine the number of interior equilibria in the model (2.2).

    The following result shows the existence of a unique interior equilibrium P3=(x,y) in the model (2.2), with x as the positive root of the polynomial (2.5) and y=f1(x), as observed in Figure 1.

    Figure 1.  Nullclines (2.3) with sqE>0. The green and margenta curves represent the nullclines y=f1(x) and y=f2(x), respectively. The brown and black dotted lines are x=Pd and x=P, respectively.

    Lemma 3. If sqE>0, the model (2.2) has a unique interior equilibrium P3.

    Proof. By Descartes' sign criterion, we have that the polynomial Q(x) (2.4) must have two positive roots: x and x+. However, since y=f1(x) is a straight line with a positive slope and intercept at (P,0), f2(x) when x± and f2(x)± for xP±, and by constraint, we have that x+<P and P<x. Therefore, the model (2.2) has a unique interior equilibrium.

    The local dynamics of the interior equilibrium P3 remain to be analyzed. In this case, since Ω is invariant and the equilibria P0, P1, and P2 are unstable, by the Poincare - Bendixson Theorem [41,42] we have that P3 cannot be a local saddle point, that is, det DX(P3)>0, with DX the Jacobian matrix of the vector field X computed in P3.

    On the other hand, the sign of the trace in the Jacobian matrix DX of the vector field X computed in P3, with

    tr DX(P3)=br(b+K)K(x+b)2+ad2n(sqE)s(x+dP)2adn(sqE)s(x+dP)(sqE)rK,

    which determines the local stability of the equilibrium P3, and is summarized in the following result. In particular, if P3 is locally unstable, by the Poincare -Bendixson Theorem it guarantees the existence of at least one stable ΓX limit cycle in Σ+, which does not collide with Σ.

    Lemma 4. For sqE>0 and :=[trDX(P3)]24detDX(P3)<0, then P3 is either a locally stable or unstable focus if trDX(P3)<0 or 0<trDX(P3), respectively. For sqE>0 and :=[trDX(P3)]24detDX(P3)>0, then P3 is either a locally stable or unstable node if trDX(P3)<0 or 0<trDX(P3), respectively.

    In the first instance, the following result shows conditions on the model parameters (2.2) for which guarantees the global stability of P1.

    Theorem 1. If sqE<0, then P1 is globally asymptotically stable in Ω.

    Proof. Since P0 is a saddle point, P2,P3Ω and ˙y=(sqE)ysy2n(xP)<0 for all (x,y)Σ+, then the model (2.2) has no limit cycles in Σ+. Furthermore, since the model (2.2) has no interior equilibria and no limit cycles in Σ, and the differential equations describing the vector field Y are autonomous and independent, all trajectories φZ0 with initial condition (x(0),y(0))Σ cross Σ and remain in Σ+, so the trajectories φZ0 converge to P1.

    On the other hand, by using the Bendixon-Dilac [40] criterion on the model (2.2), the following result shows a first sufficient condition on the model parameters (2.2) for which guarantees the global stability of P3.

    Theorem 2. If sqE>0 and b(b+P)<(b+K)(dP), then P3 is globally asymptotically stable in Ω.

    Proof. If

    g1(x,y)=rxx+b(1xK)a(xP)y(xP)+d,g2(x,y)=sy[1yn(xP)]qEy, (2.6)

    are functions describing the vector field X in Σ+, and

    h(x,y)=(xP)+dxy

    is a Dirac function, then

    hg1x(x,y)+hg2y(x,y)={αryK(x+b)2+aPx2+rKy+s[(xP)+d]nx(xP)}<0

    with α=b(dKP)+K(dP)b2>0 for all (x,y)Σ+, so the model (2.2) has no limit cycles in Σ+. Therefore, and analogous to the proof of Theorem 1, we have that P3 is globally asymptotically stable in Ω.

    From the Lyapunov criterion [40], the following result shows a second sufficient condition on the parameters of the model (2.2) for which the global stability of P3 is guaranteed in Ω.

    Theorem 3. If sqE>0 and d+2P>K, the equilibrium P3 is globally asymptotically stable in Ω.

    Proof. Similarly to what we proved in Theorem 1, it is sufficient to prove that P3 is globally asymptotically stable in Σ+. Indeed, if

    V(x,y)=xxux(uP)ρ(u)du+1sqEyyvyvdv

    for all (x,y)Σ+ and ρ(x)=a(xP)(xP)+d is a Lyapunov function, g(x)=rxx+b(1xK)ρ(x) and (2.3), we have that

    ˙V=xx(xP)ρ(x)˙x+yyy(sqE)˙y=xxxP[rxx+b(1xK)ρ(x)y]+yysqE[s(1yn(xP))qE]=xxxP[rxx+b(1xK)ρ(x)y](xx)(yy)xP+yysqE[syn(xP)syn(xP)]=xxxP[g(x)g(x)](xx)(yy)xP+s(yy)sqE[(yy)P+xyxyn(xP)(xP)]=xxxP[g(x)g(x)](xx)(yy)xP+yyy[(yy)P+xyxyxP]=xxxP[g(x)g(x)](xx)(yy)xP+(yy)2y[y(xx)(yy)(xP)xP]=xxxP[g(x)g(x)](xP)(yy)2y(xP)=1xP{(xx)[g(x)g(x)]s(yy)2n(sqE)}.

    Since P<x, a sufficient condition for ˙V<0 is to prove that g(x) is non-increasing. Indeed, by substituting ˉx=xP, then

    g(ˉx)=r(ˉx+P)(ˉx+d)(KˉxP)aKˉx(ˉx+P+b)

    and so

    g(ˉx)=r{ˉx+2ˉx(b+P)+ˉx[βb+d(KP)+P]+2dPˉx(KP)+d(b+P)(KP)}aˉx2K(ˉx+b+P)2<0,

    with β=dK+2P>0.

    In the first instance, the model (2.2) has a Hopf bifurcation if tr DX(P3)=0, as observed in the following result. In particular, and as observed in Lemma 4, if tr DX(P3)<0, then P3 is unstable with at least one stable limit cycle ΓX around P3, where ΓX collides with P3 if tr DX(P3)=0. However, if tr DX(P3)>0, then P3 is stable.

    Theorem 4. The model (2.2) has a Hopf bifurcation around P3 if trDX(P3)=0.

    Proof. Since

    tr DX(P3)a=dn(sqE)(xP)s(x+dP)2<0,

    the transversality condition for existence of the Hopf bifurcation is guaranteed [43].

    Moreover, the model (2.2) has a transcritical bifurcation when sqE=0, as summarized by the following result. In this case, if sqE>0, then P3Σ+ and P1 is a saddle point. For sqE=0, P3 collides with P1, so P1 is a stable saddle-node equilibrium since det DX(P1)=0 and tr DX(P1)=r<0. However, if sqE<0, then P3Σ+, and P1 is locally stable.

    Theorem 5. The model (2.2) has a transcritical bifurcation around P1 if sqE=0.

    Proof. Let Xs(x,y,s)=(0,y(1yn(xP)))T be the derivative of the vector field X(x,y,s) with respect to s, U=(a(KP)(b+K)r(d+KP),1)T is the eigenvector associated with the null eigenvalue in the Jacobian matrix DX calculated in P1, and W=(0,1)T is the eigenvector associated with the null eigenvalue in the Jacobian matrix DXT calculated in P1.

    If s=s0:=qE, by the transcritical bifurcation theorem [44], three conditions must be guaranteed: WTXs(P1,s0)=0, WT[DXs(P1,s0)U]0, and WT[D2X(P1,s0)(U,U)]0. Indeed,

    ● Since Xs(P1,s0)=(0,0)T, then WTXs(P1,s0)=0.

    ● Since DXs(P1,s0)U=(0,1)T, then WT[DXs(P1,s0)U]=1>0.

    ● If X=(g1(x,y,s),g2(x,y,s))T, with g1 and g2 as described in (2.6), and U=(u1,u2), then

    D2X(P1,s0)(U,U)=(2g1x2u1u1+22g1xyu1u2+2g1y2u2u22g2x2u1u1+22g2xyu1u2+2g2y2u2u2)=(2a2(KP)[bdPb(KP)2+dK2]rK(d+KP)32s0n(KP)),

    and so WT[D2f(P1,s0)(U,U)]=2s0n(KP)>0.

    To conclude, and as shown in Lemmas 2 and 4, Figure 2 shows the bifurcation diagram of the model (2.2) in the plane (P,E), realized by the numerical bifurcation package SlideCont [45], whose selected fixed parameters were chosen in order to show all possible dynamics of the model (2.2) in the bifurcation plane (P,E).

    Figure 2.  Bifurcation diagram of the model (2.2) in the plane (P,E) with fixed parameters: a=0.3, b=3, c=r=0.5, d=0.001, K=2, n=2.8, q=1 and s=0.1. Dotted line: =0. The Normal and dash dotted lines represent the Hodf and Transcritical bifurcation, respectively. Region Ⅰ: P2,P3Σ+ and P1 globally asymptotically stable. Region Ⅱa: P2,P3Σ+ with P3 globally asymptotically stable node. Region Ⅱb: P2,P3Σ+ with P3 globally asymptotically stable focus. Region Ⅲ: P2,P3Σ+ with P3 unstable focus and ΓX a globally asymptotically stable limit cycle.

    In particular, since the trajectories φX and φY in the model (2.2) do not collide in Σ, that is, the trajectories φZ0 with initial condition (x(0),y(0))Σ must cross Σ and remain in Σ+, the phase portraits characterizing each bifurcation region, shown in Figures 3, are drawn by the two vector fields X and Y, separated by the straight line x=P, and simulated in Matlab [46], where Y can be defined for all xP because the trajectory φZ0 with initial condition p=(P,y(0))Σ is defined as φZ0(p,t)=φX(p,t) for all tI{t>0} and φZ0(p,t)=φY(p,t) for tI{t0}. The bifurcation regions are described below.

    Figure 3.  Dynamics of the model (2.2) characterizing each bifurcation region shown in Figure 3. The vector fields X and Y are represented by blue and orange colors, respectively, and Σ is the black dashed line. Margenta curve: stable limit cycle ΓX around P3. Red point: P0. Pink point: P1. Blue point: P2. Black point: P3.

    ● If sqE<0, represented in region Ⅰ, the model (2.2) has no interior equilibria, so the φZ0 trajectories converge to equilibrium P1, as observed in Figure 3(b). However, if sqE>0, then P3Ω, as observed in regions Ⅱa, IIb and Ⅲ. Note that the transcritical bifurcation separates region Ⅰ from regions Ⅱa, IIb, and Ⅲ.

    ● In regions Ⅱa and IIb we have that P3 is a stable node or focus, as observed in Figures 3(c, d), respectively. In both regions, the trajectories φZ0 converge to the equilibrium P3. Moreover, the curve =0 separates regions Ⅱa and IIb.

    ● The model (2.2) has a stable limit cycle ΓX in Σ+, characterized in region Ⅲ, and which will not collide with Σ. Moreover, P3 is unstable, so the φZ0 trajectories converge to ΓX, as observed in Figure 3(e). In particular, the Hopf bifurcation separates region Ⅲ from regions Ⅱa and IIb.

    Let x(t)0 and y(t)0 be the population size of prey and predators, respectively, whose dynamics are given by the model (1.5) in the biological sense region:

    ˜Ω={(x,y)R2:0xK,0yn(KP)+c}. (3.1)

    In this case, the vector field of the model (1.5) is given by

    Zc(x,y)={˜X(x,y)=(rxx+b(1xK)a(xP)y(xP)+bsy[1yn(xP)+c]qEy),x>PY(x,y)=(rxx+b(1xK)sy(1yc)qEy),x<P (3.2)

    where the vector fields ˜X and Y are defined in ˜Σ+={(x,y)˜Ω:P<xK} and ˜Σ={(x,y)˜Ω:0x<P}, respectively.

    If ˜Σ={(x,y)˜Ω:x=P} and similarly to the dynamics of the model (2.2) in Σ, the trajectories φZc with initial condition p=(P,y)˜Σ is defined as φZc(p,t)=φ˜X(p,t) for tI{t>0} and φZc(p,t)=φY(p,t) for tI{t0} because ˜Xf(p)=Yf(p)=rP(1PK)>0.

    Analogous to the proof shown in Lemma 1, some results are guaranteed in the model (3.2).

    Lemma 5. ˜Ω is an invariant set and the trajectory (x(0),y(0))˜Ω exists and is unique.

    The equilibria of the model (3.2) on the coordinate axes (x,y) is given by P0,P2˜Σ, and ˜P1=(K,0)˜Σ+, where P0 and P2 are equivalent to the equilibria of the model (2.2), so that their local stability is described in Lemma 2. In addition, and equivalently to what is shown in Lemma 2, ˜P1 is either a locally stable node or saddle point if sqE<0 or sqE>0, respectively.

    On the other hand, the intersections in the nullclines of the vector field ˜X in ˜Σ+

    y=f2(x),y=f3(x):=nx(sqE)s(sqE)(nPc)s, (3.3)

    with y=f2(x) as represented in (2.3), and y=f3(x) a line with slope nx(sqE)s and intercept at (0,(sqE)(nPc)s) and (nPcn,0), determine the number of interior equilibria of the model (3.2), equivalent to determining the positive roots of the polynomial

    ˜Q(x)=˜Ax3+˜Bx2+˜Cx+˜D, (3.4)

    with

    ˜A=rs+aKn(sqE)>0,˜B=rs(dKP)+aKn(bn2nP+c)(sqE),˜C=K[ab(2nPP)+aP(cnP)](sqE)rsK(dP),˜D=abP(cnP)(sqE). (3.5)

    In this case, it is evident that the nullclines (3.3) do not intersect at ˜Σ+ if sqE<0. Similarly, if sqE>0 and nPc>0, it follows that there is a unique point of the intersection at the nullclines (3.3) in ˜Σ+, which corresponds to a unique interior equilibrium ˜P=(x1,y1)˜Σ+ in the model (3.2), with y1=f3(x1), as observed in Figure 4, and whose proof is analogous to what is shown in Lemma 3.

    Figure 4.  Nullclines (3.3) with sqE>0. Margenta and turquoise curves are the nullclines y=f2(x) and y=f3(x), respectively. Brown and dark cyan dotted lines are x=Pd and x=nPcn, respectivaly. Dotted black line is x=P.

    Lemma 6. If sqE<0, the model (3.2) has no interior equilibria. If sqE>0 and nPc>0, the model (3.2) has only one equilibrium.

    On the other hand, if sqE>0 and nPc<0, then y1=f3(x1)>0 for all x1>0, so the positive roots of the polynomial (3.4) are interior equilibria of the model (3.2). In this case, assuming that ˜P=(x1,y1) is an interior equilibrium that always exists in ˜Ω with sqE>0 and nPc<0, a division between the polynomial ˜Q(x) (3.4) and x1 leads to the polynomial

    ˉQ(x)=˜Ax2+(˜B+˜Ax1)x+˜C+x(˜B+˜Ax1) (3.6)

    as a factor of ˜Q(x) and the rest of the division is ˜Q(x1)=0.

    By clearing ˜B+˜Ax1 from (3.4) and substituting it in (3.6), we have

    ˉQ(x)=˜Ax2+(˜B+˜Ax1)x˜Dx1. (3.7)

    If sqE>0 and nPc<0, then the following result shows conditions on the coefficients of the polynomial ˉQ(x) that determine the number of interior equilibria in the model (3.2).

    Lemma 7. Let sqE>0, nPc<0 and ˜=(˜B+˜Ax1)2+4˜A˜Dx1.

    1) If ˜<0, the model (3.2) has only one interior equilibrium.

    2) If ˜=0, the model (3.2) has two interior equilibria.

    3) For ˜>0,

    (a) If ˜B+˜Ax1<0, the model (3.2) has three interior equilibria.

    (b) If ˜B+˜Ax10, the model (3.2) has only one interior equilibrium.

    In particular, if the model (3.2) has a unique interior equilibrium point ˜P, and since ˜Ω is an invariant with equilibria on the unstable coordinate axes, by constraint, we have that ˜P must be locally either a node or a focus, stable or unstable, so its local stability result is equivalent to what is shown in Lemma 4, where the trace in the Jacobian matrix D˜X computed on ˜P is given by:

    tr D˜X(˜P)=br(b+K)K(x1+b)2ad(cnd)(sqE)s(x1+dP)2adn(sqE)s(x1+dP)(sqE)rK. (3.8)

    On the other hand, if the model (3.2) has three different interior equilibria ˜P3=(x1,g3(x1)), ˜P4=(x,g3(x)) and ˜P5=(x+,g3(x+)), with x1<x<x+, the eigenvalues of the Jacobian matrix D˜X computed at each interior equilibrium ˜Pi,i=3,4,5, shows that ˜P4 is a saddle point; therefore, ˜P3 and ˜P5 must be locally stable or unstable nodes and/or focus, as summarized by the following result.

    Lemma 8. If the model (3.2) has three different interior equilibriums ˜P3=(x1,f3(x1)), ˜P4=(x,f3(x)), and ˜P5=(x+,f3(x+)), with x1<x<x+, then ˜P4 is a saddle point.

    For the case of local stability of the equilibria ˜P3 and ˜P5, the following section generalizes the cases of local stability for both equilibria.

    Let us consider the case where the model (3.2) has three interior equilibria ˜P3, ˜P4, and ˜P5, with ˜P3 stable and Ws+(˜P4), Wu+(˜P4) the upper stable and unstable varieties to the right in ˜Σ+ of the saddle point ˜P4, respectively, and whose dynamics is observed in Figure 5.

    Figure 5.  Relative positions of the varieties W+u(˜P4) and W+s(˜P4), with red and blue curves are W+u(˜P4) and W+s(˜P4), respectively. Margenta and turquoise curves are the nullclines (3.3) as shown in Figure 4.

    In this case, since ω-limit of Wu+(˜P4) could be the equilibrium ˜P5, when locally stable, or a stable limit cycle surrounding ˜P5 when locally unstable, or the equilibrium ˜P3 as observed in Figure 5, if (˜x0,˜ya)Wu+(˜P4) and (˜x0,˜yb)Ws+(˜P4), with P<˜x0<K, ˜ya=˜f1(a,b,c,d,E,K,P,q,r,s) and ˜yb=˜f2(a,b,c,d,E,K,P,q,r,s) continuous functions in X, then there exist (˜x0,˜ya),(˜xo,˜yb)˜Σ+ such that ˜ya=˜yb; therefore, the varieties Wu+(˜P4) and Ws+(˜P4) intersect and form a homoclinic curve ˜γ created by ˜P4 and surrounding ˜P5 as shown in the following result.

    Theorem 6. If ˜P3 is stable and Ws+(˜P4)Wu+(˜P4), then the model (3.2) has a homoclinic curve ˜γ surrounding ˜P5.

    In particular, since the trace of DX is calculated in the equilibria ˜P3, ˜P5 is equivalent to that shown in (3.8), replacing x1 by x or x+, respectively, the following result shows a generalization about the local stability of the equilibria ˜P3 and ˜P5.

    Theorem 7. If ˜P3 is unstable, that is, trD˜X(˜P3)>0, then

    1) ˜P5 is stable if trD˜X(˜P5)<0.

    2) ˜P5 is unstable if trD˜X(˜P5)>0. In this case, at least one stable cycle is formed surrounding ˜P3, ˜P4, and ˜P5.

    For ˜P3 stable, that is, trD˜X(˜P3)<0,

    1) If ˜yb<˜ya,

    (a) ˜P5 is unstable if trD˜X(˜P5)>0

    b) ˜P5 is stable if trD˜X(˜P5)<0. In this case, at least one unstable cycle is formed, surrounding ˜P5.

    2) If ˜yb>˜ya, ˜P5 is a stable node or focus.

    In addition, if ˉ(˜Pi):=[trD˜X(˜Pi)]24detD˜X(˜Pi), i=3,5, then ˜Pi is locally a node or a focus if ˉ(˜Pi)>0 or ˉ(˜Pi)<0, respectively.

    Proof. The Theorem is immediate in view of the results shown in Lemma 4 and the Poincare - Bendixson Theorem [40]. Moreover, since there is a stable limit cycle surrounding the equilibrium ˜P3, this cycle increases until it coincides with the homoclinic curve ˜γ that joins the equilibrium ˜P2. Upon breaking ˜γ, that is, if ˜yb<˜ya, the local stability of ˜P3 is not altered.

    In the first instance, if ˜P is a locally non saddle point equilibrium, then the model (3.2) has a Hopf bifurcation if tr D˜X(˜P)<0, the proof of which is analogous to what is shown in Theorem 4. Similarly, the model (3.2) has a transcritical bifurcation if sqE=0, that is, when the equilibrium ˜P collides with ˜P1, and whose proof is similar to that shown in Theorem 4, so it is omitted for brevity. On the other hand, since the model (3.2) has a homoclinic curve ˜γ if ˜P3 is unstable and the varieties Ws+(˜P4) and Wu+(˜P4) intersect, then such a model has a homoclinic bifurcation as shown in the following result, the proof of which is deduced from Theorems 6 and 7.

    Theorem 8. If the model (3.2) has three interior equilibria ˜P3, ˜P4 and ˜P5, with ˜P3 stable and ˜ya˜yb=0, the model (3.2) has a homoclinic bifurcation.

    Finally, if ˜P3,˜P4,˜P5˜Ω, the model (3.2) has a saddle-node bifurcation if the saddle point ˜P4 collides with ˜P3 or ˜P5, forming the saddle-node equilibria ˜P+=(˜x+,f2(˜x+)) with ˜x+=˜B+˜Ax12˜A or ˜P=(˜x,f3(˜x)) with ˜x=x, as observed in the following result. In particular, given that the Jacobian matrix D˜X in the vector field ˜X of the model (3.2) calculated in ˜P+ is

    D˜X(˜P+)=[an(sqE)(˜x+P)s[(˜x+P)+d]a(˜x+P)(˜x+P)+dn(sqE)2s(sqE)], (3.9)

    with det D˜X(˜P+)=0 and

    tr D˜X(˜P+)=an(sqE)(˜x+P)s[(˜x+P)+d](sqE),

    then ˜P+ is a stable or unstable saddle-node equilibrium if tr DX(˜P+)<0 or tr DX(˜P+)>0, respectively. Analogously for ˜P if ˜x+ is replaced by ˜x.

    Theorem 9. If ˜P3 and ˜P4, or ˜P5 and ˜P4, collide, the model (3.2) has a saddle-node bifurcation around ˜P+ o ˜P, respectively.

    Proof. The case in which ˜P4 collides with ˜P5 will be proved, similarly for the case in which ˜P3 and ˜P4 collide. Let ˜Xc(x,y,c) be the derivative of the vector field ˜X(x,y,c) with respect to c, that is, ˜Xc(x,y,c)=(0,sy2[n(xP)+c]2)T, and U=(sn(sqE),1)T, W=((sqE)[(˜x+P)+d]a(˜x+P),1) the eigenvectors associated with the null eigenvalue in D˜X and D˜XT calculated in ˜P+, respectively.

    If c0 satisface ˜(c0)=0, by the saddle-node bifurcation Theorem [44], two conditions must be guaranteed: WT˜Xc(˜P+,c0)0 and WT[D2˜X(˜P+,c0)(U,U)]0. Indeed,

    1) Since ˜Xc(˜P+,c0)=(0,(sqE)2s)T then WT˜Xc(˜P+,c0)(sqE)2s>0.

    2) If ˜X=(˜g1(x,y,c),˜g2(x,y,c))T, with ˜g1=rxx+b(1xK)a(xP)y(xP)+d and ˜g2=sy[1yn(xP)]qEy, and U=(u1,u2), then

    D2˜X(˜P+,c0)(U,U)=(2˜g1x2u1u1+22˜g1xyu1u2+2˜g1y2u2u22˜g2x2u1u1+22˜g2xyu1u2+2˜g2y2u2u2)=(2s(3˜Ax++˜B)n2K[(˜x+P)+d](sqE)20),

    and WT[D2˜X(˜P+,c0)(U,U)]=2s(3˜A˜x++˜B)n2K[(x+P)+b](sqE)20.

    Figure 6 shows the bifurcation curves in the (P,E) plane, and realized by the numerical bifurcation package SlideCont [45], which the transcritical bifurcation separates regions Ⅰ and Ⅱ, the homoclinic bifurcation separates regions Ⅵ and Ⅶ and the Hopf bifurcation divides regions Ⅱ and Ⅲ, Ⅴ and Ⅵ, and Ⅶ and Ⅷ. In this case, the stability of ˜P, ˜P3 and ˜P5 are altered in the curves separating regions Ⅱ-Ⅲ, Ⅴ-Ⅵ, and Ⅶ-Ⅷ, respectively.

    Figure 6.  Bifurcation diagram of the model (3.2) in the plane (P,E) with fixed parameters: a=0.3, b=3, c=5, d=0.001, K=2, n=2.8, q=1, r=0.5 and s=0.1. Dash dotted line: Transcritical bifurcation. Normal line: Hopf Bifurcation. Dash dot dotted line: Saddle-node bifurcation. Dashed line: Homoclinic bifurcation. Region Ⅰ: P2, ˜P, ˜Pi˜Ω, i=3,4,5, and ˜P1 globally asymptotically stable. Region Ⅱ: P2˜Ω and ˜P˜Ω globally asymptotically stable. Region Ⅲ: P2˜Ω, ˜P˜Ω unstable and existence of a globally asymptotically stable limit cycle Γ˜X around ˜P. Region Ⅳ: P2˜Ω, ˜Pi˜Ω unstables, i=3,4,5, and existence of a globally asymptotically stable limit cycle. Γ˜X around ˜Pi. Region Ⅴ: P2˜Ω, ˜Pi˜Ω, i=3,4,5, with ˜P3 unstable and ˜P5 globally asymptotically stable. Region Ⅵ: P2, ˜Pi˜Ω, i=3,4,5, with ˜P3 and ˜P5 locally asymptotically stable. Region Ⅶ: P2, ˜Pi˜Ω, i=3,4,5, with ˜P3, ˜P5 locally asymptotically stable and an unstable limit cycle Γ˜X around ˜P5. Region Ⅷ: P2, ˜Pi˜Ω, i=3,4,5, with ˜P3 globally asymptotically stable and ˜P5 unstable.

    Furthermore, in regions Ⅴ and Ⅷ, we observe the existence of three interior equilibria in the model (3.2), which the chair-node bifurcation formed between regions Ⅱ and Ⅴ shows the collision of P3 and ˜P4. However, the collision between ˜P4 and ˜P5 is shown by the curve between regions Ⅶ and Ⅷ. Similarly, the curve separating regions Ⅴ and Ⅲ represents a saddle-node bifurcation.

    On the other hand, since the selected fixed parameters were chosen in order to show all possible dynamics of the model (3.2) in the bifurcation plane (P,E) in Figure 6, Figure 7 shows the phase portraits of each bifurcation region, whose dynamics are described as:

    Figure 7.  Dynamics of the model (3.2) describing bifurcation regions Ⅰ to Ⅵ shown in 6. The vector fields ˜X and Y are represented by green and orange colors, respectively, and ˜Σ is the black dashed line. Blue line: φZc. Blue curve: unstable limit cycle Γ˜X. Margenta curve: stable limit cycle Γ˜X. Red point: P0. Blue point: P2. Pink point: ˜P1. Green point: ˜P3. Orange point: ˜P4. Gray point: ˜P5. Black point: ˜P.

    ● Regions Ⅰ-Ⅲ are equivalent to that shown in Figure 2, so there is a single equilibrium ˜P and its local stability is observed in Figures 7(a, b, c).

    ● For the case where ˜P3 and ˜P5 are unstable equilibria, we observe the formation of a limit cycle Γ˜X surrounding the three interior equilibria in region Ⅳ as observed in Figure 7(d), so that the trajectories φZc converge to Γ˜X.

    ● If ˜P3 is unstable and ˜P5 is stable, in region Ⅴ the trajectories φZc converge to ˜P3 as observed in Figure 7(e).

    ● For the case where ˜P3 and ˜P5 are stable, we have either the formation of a locally unstable limit cycle surrounding ˜P5, or a homoclinic curve joining ˜P4 and surrounding ˜P5, or the breaking of such a curve without altering the stability of ˜P5. In this case, in region Ⅶ we have a locally unstable limit cycle as observed in Figure 7(g), which vanishes as shown in region Ⅵ of Figure 7(f).

    ● In region Ⅷ we have that φZc converges to ˜P3 as observed in Figure 7(h).

    Since the predator-prey Leslie-Gower models proposed by several researchers which consider constant prey refuge are valid only when the initial population size of the prey is above their refuge [1,29], that is, when P<x(0), in this paper, we make a general modification to these models in order to show the dynamics of both species without restrictions on the initial condition of the proposed Leslie-Gower models.

    On the other hand, given the hypotheses on the dynamics of predators in the presence or absence of prey for their food, the proposed model (1.3) considers that the predator's diet depends only on prey; in the absence of prey, they are forced to obtain other types of food for their subsistence. However, the model (1.5) considers that the predator's diet is given by prey and alternative food provided by the environment, and in the absence of prey, predators are not forced to change their diet.

    In both proposed models (1.3) and (1.5), all trajectories with initial condition in the vector field Y cross the line x=P, and remain in the vector field X or ˜X, respectively, because the system of differential equations describing the vector field Y are independent and the only stable interior equilibrium does not belong to the biological sense region, though the trajectories try to make a convergence to this point. Similarly, the stable or unstable varieties of each interior equilibrium in both models do not cross the straight line x=P, unlike the unstable variety of P0, which connects to P1 or ˜P1, and P2. Biologically, the growth of the prey described by both models (1.3) and (1.5) cannot be controlled below P>0, so prey growth over time is higher than P.

    The model (1.3) has a unique interior equilibrium point if the intrinsic birth rates of the predators are greater than the product of the catchability coefficient and hasvesting effort of the predators, that is sqE>0; otherwise, the prey converge to their carrying capacity K>0 and the predators become extinct. Moreover, the model (1.3) has a stable limit cycle as long as the interior equilibrium is unstable, so both species stabilize over time if the interior equilibrium is stable. In this model we have two types of bifurcations: Transcritical and Hopf. These are what separates the extinction or not of predators and the long-term stabilization of the population size of both species, respectively.

    The bifurcation cases of model (1.3) can be transferred to model (1.5). However, model (1.5) has two additional bifurcations: saddle-node and homoclinic. In this case, the model can have between one or three interior equilibria as long as sqE>0; otherwise, analogous to model (1.3), predators die out and prey converge to the carrying capacity.

    For the case where model (1.5) has three interior equilibria, there exists a stable limit cycle surrounding all its interior equilibria if ˜P3 and ˜P5 are unstable. Similarly, the model (1.5) could have a locally unstable limit cycle if ˜P3 and ˜P5 are stable, and the homoclinic bifurcation is given by the formation of a homoclinic curve ˜γ given by ˜P4.

    In addition to the homoclinic bifurcation, the model (1.5) could have a heteroclinic bifurcation, whose bifurcation curve coincides with the homoclinic, from which a variety connection transition is made between ˜P1 and ˜P3 or ˜P1 and ˜P5, where the locally unstable limit cycle is given by the formation of the heteroclinic curve connecting ˜P1 and ˜P5.

    On the other hand, if the model (1.5) has a unique interior equilibrium, equivalently to the model (1.3), we have that the growth of both species stabilizes with time, or on the contrary has oscillatory solutions, when the equilibrium is stable or unstable, respectively. However, if the model (1.5) has three interior equilibria, the growth of the prey converges to its minimum quantity, regardless of the initial condition, if ˜P3 is stable and ˜P5 is unstable. However, if ˜P3 and ˜P5 are stable, and there is no unstable cycle surrounding ˜P5, then prey converges to its minimum quantity if the initial condition of the species is above the stable variety of ˜P4, otherwise prey converges to its maximum quantity. Similarly, if the model (1.5) has three equilibria where ˜P3 and ˜P5 are stable, with an unstable cycle surrounding ˜P5, prey converges to its minimum quantity if the initial condition is above the limit cycle.

    Similarly, if the model (1.5) has three interior equilibria, where ˜P3 is unstable and ˜P5 is stable, the prey converge to their maximum quantity regardless of their initial condition. However, if ˜P3 and ˜P5 are unstable, both species have oscillatory solutions. In particular, alternative food for predators could lead to a possible maximum or minimum choice in the amount of prey over time as long as the model (1.5) has three interior equilibria, with appropriate stability for each of these, and an appropriate selection in the initial condition of prey and predators.

    To conclude, if in both proposed models (1.3) and (1.5) no external factors inhibiting the intrinsic growth of prey are considered, that is, if ˙x=rx(1xK) in the absence of predators, the bifurcation cases for the two models (1.3) and (1.5) hold, and the interior equilibria still need to be computed computationally, so the alteration of ˙x=rx(1xK) or ˙x=rxx+b(1xK) in the absence of predators does not affect the overall mathematical results in both proposed models.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This research was supported by MCIN/AEI/10.13039/501100011033 through grant BADS, no. PID2019-109320GB-100, and the associated FPI contract PRE2019-088899 to Christian Cortés García. The Spanish MICINN has also funded the "Severo Ochoa" Centers of Excellence to CNB, SEV 2017-0712.

    The author declare there is no conflict of interest.



    [1] S. Boccaletti, J. Kurths, G. Osipov, D. Valladares, C. Zhou, The synchronization of chaotic systems, Phys. Rep., 366 (2002), 1–101, doi: 10.1016/S0370-1573(02)00137-0. doi: 10.1016/S0370-1573(02)00137-0
    [2] A. L. Barabási, Network science, Philos. Trans. R. Soc. London Ser. A, 371 (2013), 20120375, doi: 10.1098/rsta.2012.0375. doi: 10.1098/rsta.2012.0375
    [3] S. Boccaletti, V. Latora, Y. Moreno, M. Chavez, D.-U. Hwang, Complex networks: Structure and dynamics, Phys. Rep., 424 (2006), 175–308, doi: 10.1016/j.physrep.2005.10.009. doi: 10.1016/j.physrep.2005.10.009
    [4] L. M. Pecora, T. L. Carroll, Synchronization in chaotic systems, Phys. Rev. Lett., 64 (1990), 821, doi: 10.1103/PhysRevLett.64.821. doi: 10.1103/PhysRevLett.64.821
    [5] A. Akgül, K. Rajagopal, A. Durdu, M. A. Pala, Ö. F. Boyraz, M. Z. Yildiz, A simple fractional-order chaotic system based on memristor and memcapacitor and its synchronization application, Chaos Solitons Fractals, 152 (2021), 111306, doi: 10.1016/j.chaos.2021.111306. doi: 10.1016/j.chaos.2021.111306
    [6] F. Drauschke, J. Sawicki, R. Berner, I. Omelchenko, E. Schöll, Effect of topology upon relay synchronization in triplex neuronal networks, Chaos, 30 (2020), 051104, doi: 10.1063/5.0008341. doi: 10.1063/5.0008341
    [7] L. M. Pecora, T. L. Carroll, Master stability functions for synchronized coupled systems, Phys. Rev. Lett., 80 (1998), 2109, doi: 10.1103/PhysRevLett.80.2109. doi: 10.1103/PhysRevLett.80.2109
    [8] F. Parastesh, C.-Y. Chen, H. Azarnoush, S. Jafari, B. Hatef, Synchronization patterns in a blinking multilayer neuronal network, Eur. Phys. J. Spec. Top., 228 (2019), 2465–2474, doi: 10.1140/epjst/e2019-800203-3. doi: 10.1140/epjst/e2019-800203-3
    [9] S. Rakshit, S. Majhi, B. K. Bera, S. Sinha, D. Ghosh, Time-varying multiplex network: Intralayer and interlayer synchronization, Phys. Rev. E, 96 (2017), 062308, doi: 10.1103/PhysRevE.96.062308. doi: 10.1103/PhysRevE.96.062308
    [10] I. V. Belykh, V. N. Belykh, M. Hasler, Blinking model and synchronization in small-world networks with a time-varying coupling, Phys. D, 195 (2004), 188–206, doi: 10.1016/j.physd.2004.03.013. doi: 10.1016/j.physd.2004.03.013
    [11] A. Buscarino, M. Frasca, M. Branciforte, L. Fortuna, J. C. Sprott, Synchronization of two Rössler systems with switching coupling, Nonlinear Dyn., 88 (2017), 673–683, doi: 10.1007/s11071-016-3269-0. doi: 10.1007/s11071-016-3269-0
    [12] J. J. Torres, I. Elices, J. Marro, Efficient transmission of subthreshold signals in complex networks of spiking neurons, PLoS One, 10 (2015), e0121156, doi: 10.1371/journal.pone.0121156. doi: 10.1371/journal.pone.0121156
    [13] A. Calim, T. Palabas, M. Uzuntarla, Stochastic and vibrational resonance in complex networks of neurons, Philos. Trans. R. Soc. London Ser. A, 379 (2021), 20200236, doi: 10.1098/rsta.2020.0236. doi: 10.1098/rsta.2020.0236
    [14] J. Ma, J. Tang, A review for dynamics of collective behaviors of network of neurons, Sci. China Technol. Sci., 58 (2015), 2038–2045, doi: 10.1007/s11431-015-5961-6. doi: 10.1007/s11431-015-5961-6
    [15] L. M. Ward, Synchronous neural oscillations and cognitive processes, Trends Cognit. Sci., 7 (2003), 553–559, doi: 10.1016/j.tics.2003.10.012. doi: 10.1016/j.tics.2003.10.012
    [16] K. Rajagopal, S. Jafari, A. Karthikeyan, A. Srinivasan, Effect of magnetic induction on the synchronizability of coupled neuron network, Chaos, 31 (2021), 083115, doi: 10.1063/5.0061406. doi: 10.1063/5.0061406
    [17] M. Shafiei, S. Jafari, F. Parastesh, M. Ozer, T. Kapitaniak, M. Perc, Time delayed chemical synapses and synchronization in multilayer neuronal networks with ephaptic inter-layer coupling, Commun. Nonlinear Sci. Numer. Simul., 84 (2020), 105175, doi: 10.1016/j.cnsns.2020.105175. doi: 10.1016/j.cnsns.2020.105175
    [18] P. Zhou, Z. Yao, J. Ma, Z. Zhu, A piezoelectric sensing neuron and resonance synchronization between auditory neurons under stimulus, Chaos Solitons Fractals, 145 (2021), 110751, doi: 10.1016/j.chaos.2021.110751. doi: 10.1016/j.chaos.2021.110751
    [19] I. Belykh, E. de Lange, M. Hasler, Synchronization of bursting neurons: What matters in the network topology, Phys. Rev. Lett., 94 (2005), 188101, doi: 10.1103/PhysRevLett.94.188101. doi: 10.1103/PhysRevLett.94.188101
    [20] S. Ostojic, N. Brunel, V. Hakim, Synchronization properties of networks of electrically coupled neurons in the presence of noise and heterogeneities, J. Comput. Neurosci., 26 (2009), 369–392, doi: 10.1007/s10827-008-0117-3. doi: 10.1007/s10827-008-0117-3
    [21] X. Sun, M. Perc, J. Kurths, Effects of partial time delays on phase synchronization in Watts-Strogatz small-world neuronal networks, Chaos, 27 (2017), 053113, doi: 10.1063/1.4983838. doi: 10.1063/1.4983838
    [22] C. Batista, S. Lopes, R. L. Viana, A. M. Batista, Delayed feedback control of bursting synchronization in a scale-free neuronal network, Neural Networks, 23 (2010), 114–124, doi: 10.1016/j.neunet.2009.08.005. doi: 10.1016/j.neunet.2009.08.005
    [23] L. Zhang, Y. Zhu, W. X. Zheng, Synchronization and state estimation of a class of hierarchical hybrid neural networks with time-varying delays, IEEE Trans. Neural Networks Learn. Syst., 27 (2015), 459–470, doi: 10.1109/TNNLS.2015.2412676. doi: 10.1109/TNNLS.2015.2412676
    [24] W. Qing-Yun, L. Qi-Shao, Time delay-enhanced synchronization and regularization in two coupled chaotic neurons, Chin. Phys. Lett., 22 (2005), 543, doi: 10.1088/0256-307X/22/3/007. doi: 10.1088/0256-307X/22/3/007
    [25] V. P. Zhigulin, M. I. Rabinovich, R. Huerta, H. D. Abarbanel, Robustness and enhancement of neural synchronization by activity-dependent coupling, Phys. Rev. E, 67 (2003), 021901, doi: 10.1103/PhysRevE.67.021901. doi: 10.1103/PhysRevE.67.021901
    [26] E. V. Pankratova, A. I. Kalyakulina, S. V. Stasenko, S. Y. Gordleeva, I. A. Lazarevich, V. B. Kazantsev, Neuronal synchronization enhanced by neuron–astrocyte interaction, Nonlinear Dyn., 97 (2019), 647–662, doi: 10.1007/s11071-019-05004-7. doi: 10.1007/s11071-019-05004-7
    [27] H. Fan, Y. Wang, H. Wang, Y.-C. Lai, X. Wang, Autapses promote synchronization in neuronal networks, Sci. Rep., 8 (2018), 1–13, doi: 10.1038/s41598-017-19028-9. doi: 10.1038/s41598-017-19028-9
    [28] A. E. Pereda, Electrical synapses and their functional interactions with chemical synapses, Nat. Rev. Neurosci., 15 (2014), 250–263, doi: 10.1038/nrn3708. doi: 10.1038/nrn3708
    [29] M. Golan, J. Boulanger-Weill, A. Pinot, P. Fontanaud, A. Faucherre, D. Gajbhiye, et al., Synaptic communication mediates the assembly of a self-organizing circuit that controls reproduction, Sci. Adv., 7 (2021), eabc8475, doi: 10.1126/sciadv.abc8475. doi: 10.1126/sciadv.abc8475
    [30] M. V. Bennett, Gap junctions as electrical synapses, J. Neurocytol., 26 (1997), 349–366, doi: 10.1023/A:1018560803261. doi: 10.1023/A:1018560803261
    [31] N. T. Carnevale, D. Johnston, Electrophysiological characterization of remote chemical synapses, J. Neurophysiol., 47 (1982), 606–621, doi: 10.1152/jn.1982.47.4.606. doi: 10.1152/jn.1982.47.4.606
    [32] A. O. Martin, G. Alonso, N. C. Guérineau, Agrin mediates a rapid switch from electrical coupling to chemical neurotransmission during synaptogenesis, J. Cell Biol., 169 (2005), 503–514, doi: 10.1083/jcb.200411054. doi: 10.1083/jcb.200411054
    [33] T. A. Zolnik, B. W. Connors, Electrical synapses and the development of inhibitory circuits in the thalamus, J. Physiol., 594 (2016), 2579–2592, doi: 10.1113/JP271880. doi: 10.1113/JP271880
    [34] S. Jabeen, V. Thirumalai, The interplay between electrical and chemical synaptogenesis, J. Neurophysiol., 120 (2018), 1914–1922, doi: 10.1152/jn.00398.2018. doi: 10.1152/jn.00398.2018
    [35] K. Kandler, E. Thiels, Flipping the switch from electrical to chemical communication, Nat. Neurosci., 8 (2005), 1633–1634, doi: 10.1038/nn1205-1633. doi: 10.1038/nn1205-1633
    [36] A. Marin-Burgin, F. J. Eisenhart, S. M. Baca, W. B. Kristan, K. A. French, Sequential development of electrical and chemical synaptic connections generates a specific behavioral circuit in the leech, J. Neurosci., 25 (2005), 2478–2489, doi: 10.1523/JNEUROSCI.4787-04.2005. doi: 10.1523/JNEUROSCI.4787-04.2005
    [37] T. M. Szabo, D. S. Faber, M. J. Zoran, Transient electrical coupling delays the onset of chemical neurotransmission at developing synapses, J. Neurosci., 24 (2004), 112–120, doi: 10.1523/JNEUROSCI.4336-03.2004. doi: 10.1523/JNEUROSCI.4336-03.2004
    [38] K. L. Todd, W. B. Kristan, K. A. French, Gap junction expression is required for normal chemical synapse formation, J. Neurosci., 30 (2010), 15277–15285, doi: 10.1523/JNEUROSCI.2331-10.2010. doi: 10.1523/JNEUROSCI.2331-10.2010
    [39] Y. Wang, J. E. Rubin, Complex bursting dynamics in an embryonic respiratory neuron model, Chaos, 30 (2020), 043127, doi: 10.1063/1.5138993. doi: 10.1063/1.5138993
  • This article has been cited by:

    1. Christian Cortés García, Jasmidt Vera Cuenca, Additive Allee effect on prey in the dynamics of a Gause predator–prey model with constant or proportional refuge on prey at low or high densities, 2023, 126, 10075704, 107427, 10.1016/j.cnsns.2023.107427
    2. Christian Cortés-García, Population dynamics in a Leslie-Gower predator-prey model with proportional prey refuge at low densities, 2025, 543, 0022247X, 128993, 10.1016/j.jmaa.2024.128993
    3. Christian Cortés García, Population dynamics in a Leslie–Gower predator–prey model with predator harvesting at high densities, 2024, 0170-4214, 10.1002/mma.10359
    4. Christian Cortés-García, Dynamics in a Filippov–Gause Predator–Prey Model with Hunting Cooperation or Competition Among Predators for High or Low Prey Densities, 2025, 35, 0218-1274, 10.1142/S0218127425500142
  • 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(3501) PDF downloads(197) Cited by(27)

Figures and Tables

Figures(8)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog