Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Analysis for the hierarchical architecture of the heterogeneous FitzHugh-Nagumo network inducing synchronization

  • Synchronization is a key topic of research in neuroscience, medicine, and artificial neural networks; however, understanding its principle is difficult, both scientifically and mathematically. Specifically, the synchronization of the FitzHugh-Nagumo network with a hierarchical architecture has previously been studied; however, a mathematical analysis has not been conducted, owing to the network complexity. Therefore, in this paper, we saught to understand synchronization through mathematical analyses. In particular, we consider the most common types of hierarchical architecture and present a condition of the hierarchical architecture to induce synchronization. First, we provide mathematical analyses of a Lyapunov function for each layer, from which we obtain sufficient conditions guaranteeing synchronization and show that the Lyapunov function decreases exponentially. Moreover, we show that the internal connectivity critically affects synchronization in the first layer; however, in the second and subsequent layers, the internal connectivity is not important for synchronization, and the connectivity up to the first layer critically affects synchronization. We expect that the results and mathematical methodology can be applied to study other similar neural models with hierarchical architectures.

    Citation: Soo-Oh Yang, Jea-Hyun Park. Analysis for the hierarchical architecture of the heterogeneous FitzHugh-Nagumo network inducing synchronization[J]. AIMS Mathematics, 2023, 8(9): 22385-22410. doi: 10.3934/math.20231142

    Related Papers:

    [1] Canhong Long, Zuozhi Liu, Can Ma . Synchronization dynamics in fractional-order FitzHugh–Nagumo neural networks with time-delayed coupling. AIMS Mathematics, 2025, 10(4): 8673-8687. doi: 10.3934/math.2025397
    [2] Xintao Li, Lianbing She, Rongrui Lin . Invariant measures for stochastic FitzHugh-Nagumo delay lattice systems with long-range interactions in weighted space. AIMS Mathematics, 2024, 9(7): 18860-18896. doi: 10.3934/math.2024918
    [3] Xintao Li, Yunlong Gao . Random uniform attractors for fractional stochastic FitzHugh-Nagumo lattice systems. AIMS Mathematics, 2024, 9(8): 22251-22270. doi: 10.3934/math.20241083
    [4] Feifei Cheng, Ji Li, Qing Yu . The existence of solitary wave solutions for the neuron model with conductance-resistance symmetry. AIMS Mathematics, 2023, 8(2): 3322-3337. doi: 10.3934/math.2023171
    [5] Waleed Mohamed Abd-Elhameed, Omar Mazen Alqubori, Ahmed Gamal Atta . A collocation procedure for the numerical treatment of FitzHugh–Nagumo equation using a kind of Chebyshev polynomials. AIMS Mathematics, 2025, 10(1): 1201-1223. doi: 10.3934/math.2025057
    [6] Shuang Li, Xiao-mei Wang, Hong-ying Qin, Shou-ming Zhong . Synchronization criteria for neutral-type quaternion-valued neural networks with mixed delays. AIMS Mathematics, 2021, 6(8): 8044-8063. doi: 10.3934/math.2021467
    [7] Jingyang Ran, Tiecheng Zhang . Fixed-time synchronization control of fuzzy inertial neural networks with mismatched parameters and structures. AIMS Mathematics, 2024, 9(11): 31721-31739. doi: 10.3934/math.20241525
    [8] Pratap Anbalagan, Evren Hincal, Raja Ramachandran, Dumitru Baleanu, Jinde Cao, Chuangxia Huang, Michal Niezabitowski . Delay-coupled fractional order complex Cohen-Grossberg neural networks under parameter uncertainty: Synchronization stability criteria. AIMS Mathematics, 2021, 6(3): 2844-2873. doi: 10.3934/math.2021172
    [9] Xinxin Zhang, Yunpeng Ma, Shan Gao, Jiancai Song, Lei Chen . Robust synchronization analysis of delayed fractional order neural networks with uncertain parameters. AIMS Mathematics, 2022, 7(10): 18883-18896. doi: 10.3934/math.20221040
    [10] N. Jayanthi, R. Santhakumari, Grienggrai Rajchakit, Nattakan Boonsatit, Anuwat Jirawattanapanit . Cluster synchronization of coupled complex-valued neural networks with leakage and time-varying delays in finite-time. AIMS Mathematics, 2023, 8(1): 2018-2043. doi: 10.3934/math.2023104
  • Synchronization is a key topic of research in neuroscience, medicine, and artificial neural networks; however, understanding its principle is difficult, both scientifically and mathematically. Specifically, the synchronization of the FitzHugh-Nagumo network with a hierarchical architecture has previously been studied; however, a mathematical analysis has not been conducted, owing to the network complexity. Therefore, in this paper, we saught to understand synchronization through mathematical analyses. In particular, we consider the most common types of hierarchical architecture and present a condition of the hierarchical architecture to induce synchronization. First, we provide mathematical analyses of a Lyapunov function for each layer, from which we obtain sufficient conditions guaranteeing synchronization and show that the Lyapunov function decreases exponentially. Moreover, we show that the internal connectivity critically affects synchronization in the first layer; however, in the second and subsequent layers, the internal connectivity is not important for synchronization, and the connectivity up to the first layer critically affects synchronization. We expect that the results and mathematical methodology can be applied to study other similar neural models with hierarchical architectures.



    The phenomenon in which different oscillators (objects that show repetitive behavior over time) have rules in repetitive behavior is called synchronization. Similar to groups of fireflies that blink simultaneously or a number of pendulum clocks that sway in the same direction at the same time, the synchronization of various oscillators has been found in nature, in our daily lives [6], and in various fields such as neuroscience [16], biology [13,20], sociology [10], physics [1,33], chemistry [21,24], and computer science [36,42].

    Because biological neural network synchronization plays an important role in the brain and central nervous system activity, there is extensive and ongoing research among a diverse group of fields, especially in neuroscience and medicine [2,7,8,9,11,17,19,22,26,35,41,42,43].

    Studies have shown that rapid synchronization can cause functional disturbances in the nervous system and cause pathological patterns such as epilepsy and Parkinson's disease; however, it can also improve the function and performance of the central nervous system. Therefore, research on the mechanisms of synchronization/asynchronization is a key topic in the field of neuroscience. This topic is a core topic not only in the fields of neuroscience and medicine, but also in the field of artificial neural networks. Research on this subject is very interesting, but it is difficult to understand the principle scientifically and mathematically. Therefore, it is important to study synchronization and asynchronous mechanisms in controllable situations through mathematical modeling and analysis.

    As previously explained, synchronization is not a phenomenon caused by a single object, but rather a phenomenon caused by interactions between multiple objects. Therefore, the study of the synchronization phenomenon should also consider the interactions between objects. To understand and illustrate the relationship between objects, we formed a network topology using a graph that expresses objects as nodes and the interactions between objects as either links or arrows. The main difficulty in studying network systems is that the more objects and the more complex the relationships, the more complex the network. In particular, it is important to understand the synchronization phenomenon in a network structure because synchronization can occur differently depending on the network structure (i.e, connection relationship).

    The brain and central nervous system are among the most complex networks. It is well-known that neurons are embedded in assemblies and networks that influence each other through excitatory and inhibitory synaptic connections. Therefore, they are activated and inhibited rhythmically. We note that this rhythmicity is reflected in oscillations of the extracellular field potential that can be measured through recordings of local field potentials and through electroencephalography (EEG) (for more details, see [14] and references therein). There are approximately 86 billion neurons in the human nervous system, which are connected by approximately 1014 synapses. To study the complex interactions of these large-scale neural networks, the neural network is divided into several neural groups, each of which can be represented as a form of the interactions between the neurons within it.

    One of the most simplified models to represent the features of interacting neurons is the FitzHugh–Nagumo (FHN) network [15,27], which is described as the following system:

    ϵ˙u=uu33v, (1.1)
    ˙v=ubv+a, (1.2)

    where u and v represent the state variables of a neuron, that is, the membrane potential (activator) and the recovery variable (inhibitor), respectively, and ϵ, a, and b are positive constants. The FHN network is a reduced form of the Hodgkin-Huxley equation. Both equations are models that suggest how neural stimuli are transmitted. Each nerve cell stays still and is excited/ignited by external stimuli; thereafter, it quickly returns to its original state, which has some recovery periods and cannot be ignited for some time. The FHN network models only the important characteristics of the neuron properties. After introducing the FHN network, this work was extended in various directions, such as FHN networks coupled with either gap-junctions or spaced clamped [19,38,40], mean field models for the FHN network [4,7,32], an FHN type coupled reaction-diffusion network on a continuous domain [3], a photosensitive memristive FHN network [28], and concentration phenomena in FHN network [5] (for more topics, see [18,29,33] and therein). There is also a lot of research on synchronizing other types of models for heterogeneous neural networks such as the synchronization under Dirichlet boundary conditions [23], and the event-triggered synchronization [25,44]. In addition, we can find many results for delay-type models such as adaptive control [39], pinning control [34], and intermittent control [37].

    Among these studies, those of interest to us are [30] and [31]. In [31], Plotnikov et al. considered heterogeneous FHN networks with diffusive coupling and a greater number of neurons (nodes) as follows:

    ϵ˙ui=uiu3i3vi+nj=1cij(ujui) (1.3)
    ˙vi=uibvi+ai, (1.4)

    where cij is the coupling strength between the ith and jth neuron for i,j=1,,n. They presented analytic conditions of synchronization with precision levels (quasi-synchronization) for (1.3)-(1.4). They showed the importance of the coupling strengths for synchronization and presented two precision levels for ui and vi, which depend on the parameter ai, which is referred to as the natural frequency in this paper.

    In [30], Plotnikov et al. studied synchronization in heterogeneous FHN networks with hierarchical architecture for the case where b=0. According to [30], diffusion tensor magnetic resonance imaging (DT-MRI) studies have revealed the complexity of neuronal interconnections in human and mammalian brains. In particular, from analyzing DT-MRI, it has been shown that the connectivity of the neuron axon network is represented by a hierarchical architecture with fractal dimensions varying between 2.3 and 2.8, which is essential for the fast and optimal handling of information in the brain (for more details, see [30] and references therein). To study the synchronization of an FHN network with hierarchical architecture, they considered a leader system, in which all but one node are in complete synchrony (i.e. uiuj), and derived a sufficient condition for synchronization.

    The purpose of this study is to prove the synchronization phenomenon of a heterogeneous FHN network with a hierarchical architecture through a mathematical analysis of Lyapunov functions and to prove that the Lyapunov function decreases exponentially until synchronization. We also propose conditions for a hierarchical architecture to induce synchronization. In particular, it is noteworthy that the neurons of the first layer (i.e., leaders) are (i.e., not in a synchronized state). Therefore, the results of this study provide a more accurate analysis of how synchronization between neurons spreads. Moreover, we show that the structural connectivity of neurons in the first layer is very important for synchronization, and that in the other layers is not significant.

    The rest of this paper is organized as follows. In Section 2, we recall some definitions and properties of graph theory and present our main system, which is a heterogeneous FHN network with a hierarchical architecture that considers multiple leaders. In Section 3, we discuss the synchronization for the first layer. Section 4 presents results concerning the synchronization for the second layer, and Section 5 proves the main result of this paper. In Section 6, we provide a numerical experiment for the main results of this study. Finally, we discuss the conclusions in Section 7.

    We recall some definitions and properties of graph theory used throughout this paper.

    An (undirected) graph G=G(V,E) consists of a set V={pi}Ni=1 of vertices and a collection E of unordered pairs {pi,pj} of vertices, called an edge. For notational convenience, we denote piV by iV (or iG) and {pi,pj}E by {i,j}E or ij. A simple graph is a graph without multiple edges and loops. A complete graph is a simple graph G=G(V,E) satisfying pipj for all i,j=1,,N, (ij). A graph H=H(V,E) is said to be a subgraph of G=G(V,E) if V and E are subsets of V and E, respectively. A (vertex) induced subgraph consists of some of the vertices of the original graph and all the edges that connect them in the original graph.

    A digraph (or directed graph) is a graph in which the edges have orientations. In this case, we call the oriented edges a directed edge (or arc). For an induced subgraph H=H(V,E) of a graph G=G(V,E), the boundary H of H is the set of all vertices satisfying qVV, and there exists pV such that pq; that is,

    H:={qVV|qp for some pV}.

    Hence an induced subgraph H of a graph G with a boundary can be expressed as a directed graph whose vertex set is HH, and the edge set E satisfies {p,q}E for all pH; qH. ¯H, denotes an induced subgraph of G, whose vertices and edges are in H and vertices in H.

    A path P={p0,p1,,pn} between two vertices, p0 and pn, is a sequence of distinct vertices satisfying p0p1p2pn. A cycle is a path between a vertex and itself. A path between two vertices may or may not exist, according to the given graph. Hence, if there exists a path for all pairs of vertices, then we consider the graph G=G(V,E) to be connected. In particular, a component of an undirected graph is an induced subgraph in which any two nodes are connected to each other by paths, and it is connected to no additional vertices in the rest of the graph.

    A weighted graph G=G(V,E,ω) is a graph associated with a (coupling) weight function ω:V×V[0,) satisfying

    (i) ωij:=ω(pi,pj)>0 if ij, and ij,(ii) ωij=0 if not.

    In particular, if a weight function ω satisfies ωij=1 for ij=1,,N, then we call it the standard weight. We note that if ωij=ωji for all i, j, then the weighted graph is an undirected graph with weight; otherwise, the weighted graph is a directed graph with weight.

    Throughout this paper, all the subgraphs are assumed to be vertex-induced and connected subgraphs of a weighted graph.

    For a weighted subgraph H of a weighted graph G, the graph Laplacian L of functions ui:RR, i¯H is defined by

    Lui:=j¯Hωij(ujui),iH.

    It is well-known that there exists a real value λ and a non-zero vector ϕ:=(ϕ1,,ϕ|¯H|) satisfying the eigenvalue problem for H, that is,

    Lϕi=λϕi,iH,ϕi=0,iH,and iHϕ2i=1,

    where |A| is the cardinal number of a set A. In the case where H=, we call λ and ϕ an eigenvalue and eigenvector corresponding to λ, respectively. If H, we call these the Dirichlet eigenvalues and Dirichlet eigenvector. Because ωij is symmetric, if |H|=N, then there are N eigenvalues λiR, i=1,,N such that 0λ1λ2λN. Moreover, for a vertex-induced subgraph H, if H=, then 0=λ1<λ2λN, and if H, then 0<λ1λ2λN. In particular, it is well known that the first (Dirichlet) eigenvalue can be represented as the following Rayleigh quotient forms:

    λ2=infϕ0iHϕi=0i,jHωij(ϕjϕi)2iHϕ2i, if H=, (2.1)

    and

    λ1=infϕ0ϕ|H0i,j¯Hωij(ϕjϕi)2iHϕ2i, if H. (2.2)

    In general, λ2 is called the first eigenvalue, and λ1 is called the first Dirichlet eigenvalue. In this paper, for the sake of consistency, we denote the first (Dirichlet) eigenvalue for a subgraph H as λHω,1 without boundary (with boundary). For more details, see [12].

    In this subsection, we introduce the concept of an FHN network with a hierarchical architecture.

    A hierarchical architecture comprises multiple layers of components that reflect the relationships of neurons within each layer and the relationships between the layers. In this paper, we assume the following:

    (H1) two neurons inside each layer either interact with the same intensity (coupling strength) or do not interact at all, and

    (H2) the relationship between the layers is unilaterally influenced by the upper layer to the lower layer.

    This assumption reflects the hierarchical property of neurons. This hierarchical architecture can be expressed by an undirected graph. First, neurons and their interactions in the k-th layer are represented by nodes and edges. In particular, according to hypothesis (H1), each layer is represented by an undirected graph that may or may not be complete. It follows from hypothesis (H2) that the upper layer has no incoming directed edge from the lower layer, whereas the lower layer must have an incoming directed edge from the upper layer. Two examples of this structure are given in Figure 1. Figure 1 (a) is a simple hierarchical architecture in which we assume that the undirected graph for each layer is connected, and the k-th layer only affects the (k+1)-th layer. Figure 1 (b) shows a typical hierarchical architecture. An important difference between these two examples is the presence or absence of connectivity within each layer, except for the first layer, and the influence from the upper layer to the lower layer is not sequential.

    Figure 1.  Examples of heterogeneous FHN networks with hierarchical architecture. Green circles, red nodes, lines, and arrows indicate connected components in each layer, neurons, undirected edges, and directed edges, respectively. (a) heterogeneous FHN networks with a simple hierarchical architecture and (b) heterogeneous FHN networks with a general hierarchical architecture.

    In this paper, the hierarchical architecture that we deal with is a general hierarchical architecture as shown in Figure 1 (b). First, we study a simple type, as shown in Figure 1 (a). Based on the results, we discuss the general type in Figure 1 (b).

    For the simple type, we consider a total of m layers for neurons. In this case, because the first layer is not affected by any other layer, the FHN network for the first layer is described as follows:

    ϵ˙ui=uif(ui)vi+cjH1ωij(ujui) (2.3)
    ˙vi=uibvi+ai, (2.4)

    for iH1, where H1 is an undirected graph for the first layer, and f is a real-valued function on R satisfying fα for some α0. We note that H1 is connected, and thus the first eigenvalue λH1ω,1>0. In fact, this model is a simple generalization of the term u3i/3 in a typical FHN network (1.3)-(1.4) to the nonlinear term f(ui), and we will discuss the synchronization of this model in the next section.

    In the case of the second layer— because the second layer is affected by the first layer, while the layers after the 3rd layer do not affect the second layer— the model (2.3)-(2.4) for the first layer becomes the Robin-type boundary condition for the second layer. Hence, the FHN network for the second layer is modeled by the following:

    ϵ˙ui=uif(ui)vi+cjH1H2ωij(ujui) (2.5)
    ˙vi=uibvi+ai, (2.6)

    for iH2 subject to (2.3)-(2.4) as the Robin-type boundary condition. Using a similar argument, we model the dynamics of the third layer as follows:

    ϵ˙ui=uif(ui)vi+cjH2H3ωij(ujui)˙vi=uibvi+ai,

    for iH3 subject to (2.5)-(2.6) as the Robin-type boundary condition. In the first equation, the range of the summation is H2H3 (i.e., H1 is not included). Moreover, although the third layer appears to be independent of the first layer, it can be seen that the first layer indirectly affects the third layer, as the second layer is actually affected by the first layer.

    After the third layer, the dynamics of each layer are expressed by the same type of system as the third layer, as follows:

    ϵ˙ui=uif(ui)vi+cjHk1Hkωij(ujui)˙vi=uibvi+ai,

    for iHk, subject to the Robin-type boundary condition:

    ϵ˙ui=uif(ui)vi+cjHk2Hk1ωij(ujui)˙vi=uibvi+ai,

    for iHk1, k=4,5,,m.

    We now discuss the model for the general type of hierarchical architecture. Let Hk,l be the l-th component in the k-th layer and Hk,l be the boundary of Hk,l. Since the 1st layer is connected and has no boundary, the number of components in the 1st layer is one (simply denoted by H1,1=H1), and H1=. For k2, the boundary Hk,l can consist of several components (Hk,l)m, which are not empty. Since a component (Hk,l)m is the largest induced subgraph in the i-th layer for some i<k, if the component (Hk,l)m satisfies ((Hk,l)m)=, then (Hk,l)m=H1; otherwise, k3, and the component (Hk,l)m is a component in the i-th layer, 1<i<k. Therefore, the heterogeneous FHN network with a general hierarchical architecture is described as follows:

    ϵ˙ui=uif(ui)vi+cjHk,lHk,lωij(ujui)˙vi=uibvi+ai,

    for iHk,l, subject to the Robin-type boundary condition:

    ϵ˙ui=uif(ui)vi+cjHk,l(Hk,l)ωij(ujui)˙vi=uibvi+ai,

    for iHk,l, where (Hk,l) is the boundary of Hk,l. In particular, for the above system, we additionally assume that H1=, (H1)=, H2,l=H1, and (H2,l)= for l1, and Hk,l for k3, and l1. Then, the above system expresses the dynamics of each component Hk,l for all k1 and l1. We note that in the case where k3, it is possible that the boundary Hk,l of Hk,l can consist of several components. Thus, Hk,l is generally not connected. Moreover, the boundary between the two different boundaries of Hk,l may include the same component. In particular, because the first layer H1 is connected, the first eigenvalue λH1ω,1 is strictly positive, and because Hk,l is connected, and Hk,l for k2, l1, the first Dirichlet eigenvalue λHk,lω,1 is also strictly positive.

    Finally, because of the natural frequency ai (i.e., heterogeneities), the complete synchronization for the system is not possible except when ai0. Therefore, in this study, we will discuss the synchronization with precision levels (quasi-synchronization), which is defined as follows.

    Definition 2.1 ([31, Definition 2]). Let us consider the following n coupled nonlinear systems :

    ˙xi=Fi(x1,x2,,xn,t)

    where xi=(xi1,,xid)Rd and Fi are vector-valued functions, i=1,2,,n. Then, the n coupled nonlinear systems are synchronized with the precision levels Δ1,Δ2,,Δn if there exists t>0 such that |xij(t)ˉxj(t)|Δj, i=1,2,,n, j=1,2,,d, tt, where ˉxj:=1/nni=1xij(t) for j=1,2,,d.

    We begin this section with the heterogeneous FHN network without a boundary (i.e., the model for the first layer H1):

    ϵ˙ui=uif(ui)vi+cjH1ωij(ujui) (3.1)
    ˙vi=uibvi+ai (3.2)

    where c>0, and iH1. For the system (3.1)-(3.2), we show the synchronization with precision levels and present the precision levels.

    Theorem 3.1. Suppose that there exist σ>0 and α0 such that |aiaj|σ for i,j=1,2,,n and f(s)α for all sR. If c is sufficiently large to cλH1ω,1+α1>0, then the FHN network (3.1)-(3.2) synchronizes with precision levels. Moreover, the Lyapunov function E(t):=12iH1(ϵˆu2i(t)+ˆv2i(t)), t0, satisfies

    E(t)[E(0)Aη2bη]exp(2bηt)+Aη2bη

    for t0, where η is a constant in (0,1) satisfying η1bϵ(cλω,1+α1), and Aη:=|H1|σ2/(4b(1η)).

    Proof. First, we consider the fluctuations ˆui, ˆvi, and ˆai defined by ˆui=uiˉu, ˆvi=viˉv, and ˆai=aiˉa, where ˉu, ˉv, and ˉa are the averages of ui, vi, and ai, respectively. (i.e., ˉu:=1nni=1ui, ˉv:=1nni=1vi, and ˉa:=1nni=1ai). Subsequently, upon differentiating ˆui and ˆvi with respect to t, we find that

    ϵ˙ˆui=ˆui(f(ui)1njH1f(uj))ˆvi+cjH1ωij(ˆujˆui)˙ˆvi=ˆuibˆvi+ˆai,

    for i=1,,n. The Lyapunov function for the above systems is given by

    E(t)=12iH1(ϵˆu2i(t)+ˆv2i(t)),t0. (3.3)

    Differentiating with respect to t, we deduce

    ˙E=iH1ˆu2i+ci,jH1ωij(ˆujˆui)ˆui1ni,jH1(f(ui)f(uj))ˆuiiH1bˆv2i+iH1ˆaiˆvi (3.4)

    Since iˆui=0, by ωij=ωji and (2.1), we have

    ci,jH1ωij(ˆujˆui)=c2i,jH1ωij(ˆujˆui)2cλH1ω,1iH1ˆu2i

    Note that since we assume that H1 is a connected and undirected graph without a boundary, the first eigenvalue λH1ω,1 is strictly positive. Now, we discuss an estimate of the third term in (3.4). Applying the mean-value theorem, for i,j=1,,n, there exists ξij such that

    1ni,jH1(f(ui)f(uj))ˆui=12ni,jH1(f(ui)f(uj))(ˆuiˆuj)=12ni,jH1f(ξij)(ˆuiˆuj)2.

    Since f(s)α and iˆui=0, we obtain

    12ni,jH1f(ξij)(ˆuiˆuj)212ni,jH1α(ˆuiˆuj)2=αiH1ˆu2i.

    Hence, the Lyapunov function E satisfies

    ˙E(cλH1ω,1+α1)iH1ˆu2ibiH1ˆv2i+iH1ˆviˆai(cλH1ω,1+α1)iH1ˆu2ibiH1ˆv2i+b(1η)iH1ˆv2i+14b(1η)iH1ˆa2imin{1ϵ(cλH1ω,1+α1),bη}iH1(ϵˆu2i+ˆv2i)+14b(1η)iH1ˆa2imin{1ϵ(cλH1ω,1+α1),bη}iH1(ϵˆu2i+ˆv2i)+|H1|σ24b(1η)=:ληE+Aη (3.5)

    where λη=2min{1ϵ(cλH1ω,1+α1),bη}. Aη=|H1|σ2/(4b(1η)), and we employed Cauchy's inequality with η(0,1). Since η is arbitrarily chosen in (0,1), and cλω,1+α1>0, taking η satisfying η1bϵ(cλω,1+α1), we have λη=bη. Thus, by the Grönwall inequality, we obtain

    E(t)[E(0)Aη2bη]exp(2bηt)+Aη2bη (3.6)

    for t0. Therefore, there exist t>0 and Δ>0 such that E(t)Δ for t>t. This implies that |uiˉu|Δ and |viˉv|Δ for i=1,2,,n. Hence, we have the desired result.

    Remark 3.2. (ⅰ) From (3.6), we can see that the synchronization proceeds exponentially. In particular, in the case where ai0, complete synchronization occurs exponentially, (i.e., uiuj0 and vivj0 as t for all i and j).

    (ⅱ) Since exp(2bηt)0 as t, and |aiaj|σ, the precision level Δ is asymptotically equal to Aη2bη and satisfies

    Δ(18b2η(1η)iH1ˆa2i)12(nσ28b2η(1η))12=O(σ).

    (ⅲ) Since η(0,1) is an arbitrary real value satisfying η1bϵ(cλH1ω,1+α1), we obtain the minimum of Aη2bη as setting η=min{1bϵ(cλH1ω,1+α1),12}>0. Hence, it is clear that the connectivity of the first layer is the crucial condition that induces synchronization. If the first layer is not connected, then it is trivial that no synchronization is induced. Hence, the connectivity is an equivalent condition for synchronization.

    (ⅳ) In [31], the authors discussed the case where f(s)=s3/3 and presented two precision levels

    |uiˉu|Δ1=(3nσ2b)14=O(σ1/2),and|viˉv|Δ2=σ2(nb+1b)=O(σ),

    as t. Therefore, from the precision level in (ⅱ), the decay rate of Δ is better than that of Δ1.

    In this section, for the simple type of hierarchical architecture introduced in Section 2.2, we consider two kinds of boundary value problems. The first is a case where synchronization has already occurred at the boundary, and the second is a case where synchronization has not occurred. The first case can be represented by a heterogeneous FHN network with Robin-type boundary conditions:

    ϵ˙ui=uif(ui)vi+cjH1H2ωij(ujui), (4.1)
    ˙vi=uibvi+ai, (4.2)

    for iH2, which is an index of interior vertices and

    ϵ˙ui=uif(ui)vi, (4.3)
    ˙vi=uibvi+ai, (4.4)

    for iH1, which is an index of boundary vertices.

    We note that this system represents a situation where uj and vj, jH1, are neurons in the first layer, and ui and vi, iH2, are neurons in the second layer and are affected by the neurons in the first layer. In particular, we do not consider the Laplacian graph on the boundary because the neurons in the first layer are already synchronized with either a zero precision level or the number of neurons is 1, (i.e., |H1|=1). Therefore, we assume that the fluctuations of uj and vj on the boundary are bounded, that is, there exists σB>0 such that |ujˉuB|<σB and |vjˉvB|<σB, jH1, where ˉuB:=1|H1|iH1ui and ˉvB:=1|H1|iH1vi. Henceforth, we consider m:=|H1| and n:=|H2|.

    In fact, the case where b=0 and all but one node are in synchrony has already been discussed in [30], and Plotnikov et al. proved that the system synchronizes with a precision level Δ and showed that Δ=O(σ) when c>1/dmin, where dmin is the minimal degree of the network.

    As the main takeaway from this section, we introduce the Lyapunov function, which is slightly different from [30], and present the precision level through mathematical analysis. In particular, we prove the following lemma, which is very useful for dealing with the Lyapunov function:

    Lemma 4.1. For a real-valued function ui:[0,)R, iH1H2, we have

    iH2jH1H2ωij(ujui)uiλBω,1iH2u2i+iH2jH1ωiju2j+14i,jH1ωij(ujui)2.

    where λBω,1 is the first Dirichlet eigenvalue.

    Proof. First, we have

    iH2jH1H2ωij(ujui)ui=i,jH1H2ωij(ujui)uiiH1jH1H2ωij(ujui)ui. (4.5)

    The first term on the right-hand side of (4.5) satisfies

    ci,jH1H2ωij(ujui)ui=c2i,jH1H2ωij(ujui)2c2i,jH2ωij(ujui)2=ci,jH2ωij(ujui)ui. (4.6)

    Hence, (4.5) and (4.6) imply

    iH2jH1ωij(ujui)uiiH1jH1H2ωij(ujui)ui. (4.7)

    Since

    iH2jH1H2ωij(ujui)ui=i,jH2ωij(ujui)ui+iH2jH1ωij(ujui)ui, (4.8)

    applying (4.7) to (4.8), we find

    iH2jH1H2ωij(ujui)ui12i,jH2ωij(ujui)2+12iH2jH1ωij(ujui)ui12iH1jH2ωij(ujui)ui12i,jH1ωij(ujui)ui.

    Let us swap the indices i and j in the third term on the right-hand side of the inequality above and apply the same method used in deriving the first equality in (4.6). Then,

    12iH1jH2ωij(ujui)ui12i,jH1ωij(ujui)ui=12iH2jH1ωij(ujui)uj+14i,jH1ωij(ujui)2.

    After some elementary calculations, we obtain

    iH2jH1H2ωij(ujui)ui12i,jH2ωij(ujui)212iH2jH1ωij(0ui)212iH2jH1ωij(uj0)2+iH2jH1ωiju2j+14i,jH1ωij(ujui)2.

    Therefore, if we define a function u0i as

    u0i(t):={ui,for iH2,0,for iH1,

    then

    iH2jH1H2ωij(ujui)ui12i,jH1H2ωij(u0ju0i)2+iH2jH1ωiju2j+14i,jH1ωij(ujui)2.

    Finally, using the Rayleigh quotient (2.2) of the first Dirichlet eigenvalue λBω,1, we have the desired result that

    iH2jH1H2ωij(ujui)uiλBω,1iH2u2i+iH2jH1ωiju2j+14i,jH1ωij(ujui)2.

    We now show that the system (4.1)–(4.4) synchronizes with precision levels.

    Theorem 4.2. Suppose that there exists σ>0 and α0 such that |aiaj|σ for i,jH1H2 and f(s)α for all sR. If c>0 and δ>0 satisfy cλH2ω,1+α1δ>0, where λH2ω,1 is the first Dirichlet eigenvalue, then the FHN network with Robin-type boundary synchronizes with precision levels. Moreover, there exist KB>0 and σB>0 such that the functional E(t):=12iH2(ϵˆu2i(t)+ˆv2i(t)), t0, satisfies

    E(t)[E(0)ABη2bη]exp(2bηt)+ABη2bη

    for t0, where η is a constant in (0,1) satisfying η1bϵ(cλH2ω,1+α1δ), and ABη:=|H2|σ24b(1η)+nmKB4δσ2B.

    Proof. Let us define fluctuations ˆui=uiˉuB, ˆvi=viˉvB, and ˆai=aiˉaB for iH1H2, where ˉuB, ˉvB, and ˉaB are the averages of ui, vi, and ai on the boundary, respectively; that is,

    ˉuB:=1miH1ui,ˉvB:=1miH1vi,ˉaB:=1miH1ai.

    Then, by calculations similar to those used to establish (4.1)-(4.2), we establish that

    ϵ˙ˆui=ˆui(f(ui)1njH1f(uj))ˆvi+cjH1H2ωij(ˆujˆui), (4.9)

    and

    ˙ˆvi=ˆuibˆvi+ˆai,

    for iH2. The FHN network with a Robin-type boundary admits the following Lyapunov functional E, which can be viewed as energy functionals:

    E(t)=12iH2(ϵˆu2i(t)+ˆv2i(t)),t0.

    Differentiating it with respect to t, we deduce that

    ˙E=iH2ˆu2i+ciH2jH1H2ωij(ˆujˆui)ˆuiiH2(f(ui)1mjH1f(uj))ˆuiiH2bˆv2i+iH2ˆaiˆvi. (4.10)

    Since ˆuj is bounded for jH1, there exists σB>0 such that |ˆuj|σB for jH1. Then, a direct calculation (the details of which we omit) shows that

    iH2jH1ωijˆu2j+14i,jH1ωij(ˆujˆui)2Cωσ2B, (4.11)

    where Cω:=(mmaxjH1(iH2ωij)+m22maxiH2jH1ωij). Hence, for the second term in (4.10), Lemma 4.1 and (4.11) imply that

    ciH2jH1H2ωij(ˆujˆui)ˆuicλH2ω,1iH2ˆu2i+cCωσ2B (4.12)

    For the third term, by the mean-value theorem and the assumption fα, there exists ξBi, iH2, and ζBj, jH1 such that

    iH2(f(ui)1mjH1f(uj))ˆui=iH2(f(ui)f(ˉuB)+f(ˉuB)1mjH1f(uj))ˆui=iH2f(ξBi)(uiˉuB)21miH2jH1f(ζBj)(ˉuBuj)ˆuiαiH2ˆu2i1miH2jH1f(ζBj)(ˉuBuj)ˆui.

    We note that because ζBj is between ˉuB and uj for jH1, it is clear that |ζBj||ˉuB|+|uj|2σB for jH1. Since f is differentiable, there exists KB0 such that f(ζBj)KB. Therefore, by applying Cauchy's inequality with δ, we obtain

    αiH2ˆu2i1miH2jH1f(ζBj)(ˉuBuj)ˆuiαiH2ˆu2i+nmKB4δ(jH1|ujˉuB|)2+δiH2ˆu2i(δα)iH2ˆu2i+nmKB4δσ2B. (4.13)

    Hence, for η(0,1), we obtain

    ˙E(t)min{1ϵ(cλH2ω,1+α1δ),ηb}iH2(ϵˆu2i+ˆv2i)+14b(1η)iH2ˆa2i+nmKB4δσ2Bmin{1ϵ(cλH2ω,1+α1δ),ηb}iH2(ϵˆu2i+ˆv2i)+|H2|σ24b(1η)+nmKB4δσ2BλBηE(t)+ABη,

    in the same way that we obtained (3.5), where λBη=2min{1ϵ(cλH2ω,1+α1δ),ηb}, and ABη=|H2|σ24b(1η)+nmKB4δσ2B. Since η is arbitrarily chosen in (0,1) and cλH2ω,1+α1δ>0, we obtain the desired result by the same argument as in the proof of Theorem 3.1.

    Remark 4.3. In [30], Plotnikov et al. addressed the case that b=0, f(s)=s3/3, and a network in which all but one node are in synchrony for the model (4.1)-(4.4); they obtained the inequality c>1/di as a sufficient condition, where di is the degree of the i-th neuron whose general definition is di:=kH1H2ωik, iH1H2. We note that since they considered a standard weighted graph, in this case, di denotes the number of adjacent nodes. Additionally, they showed that the i-th neuron synchronizes with the other nodes with a precision level equal to 2σ. More precisely, t0>0 such that |uiˉuB|Δ1=O(σ), and |viˉvB|Δ2=O(σ) holds: for t>t0.

    Theorem 4.2 leads to similar results to that obtained in [30]. First, it is clear that the precision level presented in Theorem 4.2 is given by

    |uiˉuB|Δ=O(σ)+O(σB), and |viˉvB|Δ=O(σ)+O(σB).

    Moreover, since all nodes in the first layer synchronize with the precision level by Theorem 3.1, we also have σB=O(σ). Therefore, in the case where σ=0, complete synchronization is achieved. In particular, if all the neurons in H1 are completely synchronized, that is, ˆui=0, iH1, then σB=0, which implies Δ=O(σ). Thus, in this case, we can see that Theorem 4.2 and the result in [30] present the same decay rate for the precision level. Second, a similar condition to [30] has been proposed in Theorem 4.2, which is cλH2ω,1+α1δ>0. We note that the first Dirichlet eigenvalue λH2ω,1 is also related to the degree. By a proof similar to that provided for Lemma 1.9 in [12], it is easy to see that the first (Dirichlet) eigenvalue satisfies

    λH2ω,1minijωij(D+1)|H2|,

    where D is the diameter of H2. In particular, if we assume that each node in H2 is connected to at least one boundary node in H1, as the generalization of the network condition in [30], then it follows from the Rayleigh quotient (2.2) that

    λH2ω,1=i,jH1H2ωij(ϕjϕi)2iH2ϕ2iiH2jH1ωij(ϕjϕi)2ni=1ϕ2i|H1|miniH2,jH1ijωij

    where ϕ is the eigenvector corresponding to λH2ω,1. Hence, we can see that the network structure affects synchronization. Moreover, by increasing the number of boundary nodes, we can derive that the system (4.1)–(4.4) synchronizes with precision levels. It is natural that adding more boundary nodes with the same tendency induces synchronization, even if the attractive force (i.e., the coupling strength c) between the interior nodes is weak. Theorem 4.2 explains this phenomenon.

    We also note that in [30], they did not lead to a mathematical analysis when more than one node was not synchronized with the rest of the network. By contrast, Theorem 4.2 provides a mathematical analysis of the occurrence of synchronization in that case.

    Now, we consider the second case where the neurons in the first layer are not synchronized and affect neurons in the second layer. This situation is modeled as a heterogeneous FHN network with the Robin-type boundary:

    ϵ˙ui=uif(ui)vi+cjH1H2ωij(ujui) (4.14)
    ˙vi=uibvi+ai, (4.15)

    for iH2, which is an index of interior vertices and

    ϵ˙ui=uif(ui)vi+cjH1ωij(ujui) (4.16)
    ˙vi=uibvi+ai, (4.17)

    for iH1, which is an index of boundary vertices.

    We can show that this system synchronizes with precision levels by the same arguments used in the proof of Theorem 4.2. More specifically, for the same fluctuations ˆui, ˆvi, and ˆai as in Theorem 4.2, we have

    ϵ˙ˆui=(uif(ui)vi+cjH1H2ωij(ujui))1mjH1(ujf(uj)vj+ckH1ωkj(ukuj))

    and ˙ˆvi=ˆuibˆvi+ˆai. In particular, since jH1kH1ωkj(ukuj)=0, we obtain

    ϵ˙ˆui=(uif(ui)vi+cjH1H2ωij(ujui))1mjH1(ujf(uj)vj),

    which is the same as (4.11). Hence, by the same arguments with the proof of Theorem 4.2, we obtain the following result.

    Theorem 4.4. Suppose σ>0 such that |aiaj|σ for i,jH1H2 and fα0 in R. If c and δ are sufficiently large to cλH2ω,1+α1δ>0, where λH2ω,1 is the first Dirichlet eigenvalue, then the FHN network (4.14)-(4.15) with the Robin-type boundary condition (4.16)-(4.17) synchronizes with the precision levels.

    In this section, we discuss the main problem addressed in this study (i.e., the heterogeneous FHN network with hierarchical architecture) when applying Theorem 3.1 and Theorem 4.4. First, we address the simple hierarchical architecture with n layers introduced in Section 2.2. Subsequently, we discuss the general hierarchical architecture.

    Since the first layer is not dependent on the other layers, the heterogeneous FHN network is modeled by (3.1)-(3.2) in Section 3. Hence, by Theorem 3.1, the system synchronizes with the precision levels. For the second layer, because the neurons in the second layer are only affected by the neurons in the first layer, we can consider it as the boundary value system (4.14)-(4.17). Therefore, from Theorem 4.4, we can see that the second layer also synchronizes with the precision levels. In particular, in the case where ai0, since it follows from Theorem 3.1 that ˆui0 and ^vi0 for iH1, we have ui(1/n1)iH1ui and vi(1/n1)iH1vi for iI2 by Theorem 4.4, where nk:=|Hk|, k=1,2,3,, that is, ui and vi in the second layer converge to the average of ui and vi in the first layer, respectively.

    We now discuss the system for the third layer in the case of the simple hierarchical architecture introduced in Section 2.2:

    ϵ˙ui=uif(ui)vi+cjH2H3ωij(ujui)˙vi=uibvi+ai

    for iH3, subject to the Robin-type boundary condition

    ϵ˙ui=uif(ui)vi+cjH1H2ωij(ujui)˙vi=uibvi+ai

    for iH2. Hence, from the results in Section 4, if we assume, without loss of generality, that the solutions ui and vi on the first and second layers are bounded, then we have the following result for the third layer.

    Theorem 5.1. Suppose that there exists σ>0 and α0 such that |aiaj|σ for i,jH2H3 and f(s)α for all sR. If c and δ satisfy cλH3ω,1+α1δ>0, then the FHN network with three-hierarchical architectures synchronizes with the precision level. Moreover, there exist C>0 and σB>0 such that the functional E(t):=12iH2(ϵˆu2i(t)+ˆv2i(t)), t0, satisfies

    E(t)[E(0)ABη2bη]exp(2bηt)+ABη2bη

    for t0, where η is a constant in (0,1) satisfying η1bϵ(cλH2ω,1+α1δ), and ABη:=|H3|σ24b(1η)+Cσ2B.

    Proof. First, we define fluctuations ˆui=uiˉuB, ˆvi=viˉvB, and ˆai=aiˉaB for i=H2H3, where

    ˉuB:=1n2iH2ui,ˉvB:=1n2iH2vi,ˉaB:=1n2iH2ai.

    Then, we have

    ϵ˙ˆui=ˆui(f(ui)1n2jH2f(uj))ˆvi+cjH2H3ωji(ˆujˆui)cn2kH2jH1ωjk(ujuk) (5.1)

    and

    ˙ˆvi=ˆuibˆvi+ˆai

    for iH3. Now, we consider the following Lyapunov function:

    E(t)=12iH3(ϵˆu2i+ˆv2i),

    and differentiate it with respect to t. Then, it follows from similar arguments used in the proof of Theorem 4.2 that

    ˙EλH3ω,1iH3ˆu2i+Cωσ2B+(δ2α)iH3ˆu2i+n2n32δσ2B+iH3ˆu2ibiH3ˆv2i+b(1η)iH3ˆv2i+14b(1η)iH3ˆa2icn2iH3kH2jH1ωjk(ujuk)ˆui.

    By applying Cauchy's inequality with δ to the last term on the right-hand side of the inequality above, the last term satisfies

    |c|H2|iH3kH2jH1ωjk(ujuk)ˆui|c2|H3|2δ|H2|2(kH2jH1ωjk|ujuk|)2+δ2iH3ˆu2i2c2|H1||H3|maxjH1,iH2ωijδ|H2|σ2B+δ2iH3ˆu2i.

    Therefore, we obtain

    ˙E(t)min{1ϵ(cλH3ω,1+α1δ),ηb}iH3(ϵˆu2i+ˆv2i)+14b(1η)iH3ˆa2i+Cσ2Bmin{1ϵ(cλH3ω,1+α1δ),ηb}iH3(ϵˆu2i+ˆv2i)+|H3|σ24b(1η)+Cσ2B

    for some C>0. Thus, taking a sufficiently small η, we obtain

    ˙E(t)2ηbE(t)+ABη.

    Finally, the Grönwall inequality leads to the desired result.

    Now, repeating the above procedure up to the m-th layer, we obtain the following result.

    Theorem 5.2. Suppose that there exists σ>0 and α0 such that |aiaj|σ for i,jH1Hm and f(s)α for all sR. If c>0 and δ>0 satisfy cλHkω,1+α1δ>0, for k=1,2,,m, then the FHN network with m-hierarchical architectures synchronizes with the precision level.

    We are now in a position to prove the main result of this paper for heterogeneous FHN networks with a general hierarchical architecture:

    ϵ˙ui=uif(ui)vi+cjHk,lHk,lωij(ujui), (5.2)
    ˙vi=uibvi+ai, (5.3)

    for iHk,l, subject to the Robin-type boundary condition:

    ϵ˙ui=uif(ui)vi+cjHk,l(Hk,l)ωij(ujui), (5.4)
    ˙vi=uibvi+ai, (5.5)

    for iHk,l. The assumptions of Hk,l and Hk,l are described in Section 2.2.

    Since the first layer is connected and has no boundary, from Theorem 3.1, it is clear that the system synchronizes with the precision levels. In the case where a component H2,l, since H2,l=H1 for all l1, we have λH2,lω,1>0 for all l1. Thus, by the same proof, as provided for Theorem 4.4, we obtain synchronization with the precision levels. Now, let us consider the l-th component H3,l in the third layer, and ˆui=uiˉuB, ˆvi=viˉvB, and ˆai=aiˉaB for i=H3,lH3,l, where

    ˉuB:=1n3,liH3,lui,ˉvB:=1n3,liH3,lvi,ˉaB:=1n3,liH3,lai.

    Here, n3,l:=|H3,l|. We note that the boundary H3,l is not an empty set, but it consists of several disjointed components (H3,l)m that are either H1 or a component in the second layer. Hence, we have

    ϵ˙ˆui=ˆui(f(ui)1n2jH3,lf(uj))ˆvi+cjH3,lH3,lωji(ˆujˆui)cn3,lmk(H3,l)mj((H3,l)m)ωjk(ujuk)

    and

    ˙ˆvi=ˆuibˆvi+ˆai

    for iH3,l. We note that since H3,l is a component in the third layer, there exists m such that (H3,l)m is a component in the second layer, which implies that

    mk(H3,l)mj((H3,l)m)ωjk(ujuk)0.

    Moreover, because of the above argument, the solutions ui and vi on the first and second layers are bounded. By applying the same arguments as the proof of Theorem 5.1, we can prove that for each k3 and l1, the system (5.2)–(5.5) synchronizes with precision levels. Hence, by repeating this process for each component Hm,l, we obtain the main result of this paper.

    Theorem 5.3. Suppose that there exists σ>0 and α0 such that |aiaj|σ for i,jH1,1Hm,n and f(s)α for all sR. If c and δ satisfy cλHk,lω,1+α1δ>0, for k=1,2,,m, and l1, then the FHN network with hierarchical architectures synchronizes with the precision level.

    Remark 5.4. In proving Theorem 5.3, it can be seen that the connectivity of the first layer plays a very important role in inducing synchronization. However, in the second layer, the connectivity of each layer is not important. We can see that the most important condition for inducing synchronization is the existence of a path from a node in each layer to a node in the first layer.

    In this section, we present a numerical experience to illustrate the contents of Theorem 5.3. This simulation was run using MATLAB'S ODE solver "ode15s." For the heterogeneous FHN network with hierarchical architecture, we first consider a directed graph with four layers for the hierarchical architecture, as shown in Figure 2. The first layer marked in red is connected, and the second, third, and fourth layers, marked in green, blue, and magenta, have four, two, and three components, respectively. We note that it is a little difficult to distinguish, but the green component in the lower right of Figure 2 is a total of two overlapping components. The index set of nodes of one component is {6,7,12,19,22,24,27}, and the index set of the nodes of the other component is {8,11,14,18,20,25}. For i<j, the k-th component Hi,k in i-th layer and the l-th component Hj,l in the j-th layer are connected by arcs (directed edges) from a node in Hi,k to a node in Hj,l. The weights on the edges and arcs are given randomly in the range [0,1]. For the system (5.2)–(5.5), we set b=0.5, ϵ=0.01, and f(s)=s3/3. The natural frequency ai has a uniform distribution on [0.5,0.7], and the initial configurations of u and v are randomly given in [20,20] and [10,10], respectively. Finally, to determine the coupling strength c, we must know the first (Dirichlet) eigenvalue for each layer. Because it is not easy to find the values of the first (Dirichlet) eigenvalues for the components, we propose its lower bound instead of the first (Dirichlet) eigenvalue. By the proof similar to Lemma 1.9 in [12], it is easy to see that the first (Dirichlet) eigenvalue satisfies

    λHk,lω,1minijωij(Dk,l+1)nk,l,

    where Dk,l is the diameter of Hk,l and nk,l is the number of nodes in Hk,l, k=1,2,3,4. In this simulation, maxk,l(Dk,lnk,l)=35, and the minimum of all the weights given on the edges is 0.0125. Hence, we choose δ=0.5 and c=4000. Then, all assumptions in Theorem 5.3 are fulfilled. Hence, we can see that the heterogeneous FHN network with hierarchical architecture synchronizes with precision levels. as shown in Figure 3.

    Figure 2.  Directed graph structure. Red, green, blue, and magenta nodes indicate units belonging to the 1st, 2nd, 3rd, and 4th layers in that order. Numbers are the index of each node. Line segments represent (undirected) edges in each layer, and arrows represent arcs (directed edges) between layers.
    Figure 3.  Graphs of the functions ui and vi. (a) graphs of ui's; (b) dynamics of ui's; (c) graphs of ui's on a log scale versus time; (d) graphs of vi's; (e) dynamics of vi's; (d) trajectory for (ui(t),vi(t)) for each i. System parameters are given by n=53, δ=0.5, c=4000, f(s)=s3/3, and parameters ai are uniform distribution on [0.5,0.7]. The initial data for ui and vi are randomly given in [-20 20] and [-10, 10], respectively. The graph (b) and (e) are heat maps for ui and vi. In (b) and (e), the y-axis is the index of the nodes, and the values of ui and vi are represented by colors.

    From Figure 3 (a) and (b), we can see that u synchronizes very quickly, and the initial behavior of ui can be confirmed in Figure 3 (c), which are graphs of ui on a log scale versus time. By contrast, the synchronization of vi is relatively slow.

    Finally, in Figure 4 (a), we provide graphs for iˆu2i, iˆv2i, and E. The y-axis is logarithmically scaled and E is the sum of all Lyapunov functions of the layers. In this figure, since each function appears to move in an almost straight line in some time interval, each function has an overall exponential decay on the time interval. We can also see that iˆui has a large change in a very short time, as shown in Figures 4 (b). In particular, by Figure 3 (a) and (d), we can easily predict that E will oscillate periodically. On the other hand, in Figure 4 (c), this phenomena is not very visible, and it seems as if there is no oscillation. However, as shown in Figure 5, we can see that E is oscillating periodically with a very small amplitude.

    Figure 4.  Graphs of energy functions. (a) graphs of iˆu2i, iˆv2i, and E; (b) enlarged graph of (a); (c) graphs of iˆu2i, iˆv2i, and E. The y-axis of the graph (a) and (b) is on a log scale. Note that the x-axis in graph (b) is in units of 103. The graph (c) is a typical graphs for iˆu2i, iˆv2i, and E with no log scale applied.
    Figure 5.  Graph of the Lyapunov function E. This is the graph of the Lyapunov function E in Figure 4 (c) and a zoomed-in portion of it. The oscillations are not visible in the full graph, but by zooming in on a portion of it, we can see that it oscillates with a very small amplitude and has a period.

    In this simulation, although it is not easy to obtain the first (Dirichlet) eigenvalues, the lower bound is proposed. If we directly find the first (Dirichlet) eigenvalues, then we can find a smaller c satisfying the condition cλHk,lω,1+α1δ>0.

    We considered a heterogeneous FHN network with a hierarchical architecture—that is, a neuronal model in which the neuronal axon network is not homogeneous and spans various brain regions, and we demonstrated that synchronization occurs in this model through mathematical analyses of the Lyapunov function. It is well known that networks with heterogeneous units are much more difficult to completely synchronize than networks with identical characteristics. Thus, in this paper, synchronization means that all units have similar trajectories within a small deviation.

    First, we proposed a generalized model for the first layer based on a traditional heterogeneous FHN network. After the second layer, we constructed a Robin-type boundary value problem by reflecting the characteristics of the hierarchical structure of the neurons. This is because the preceding layer affects the next layer, but not vice versa. Moreover, in these models, we can see that one layer either directly or indirectly affects the next and subsequent layers.

    After modeling, we proposed a sufficient condition for synchronization, which is critically dependent on the coupling strength and connectivity of the network topology. In particular, according to the final result of this paper, the importance of the connectivity of the network topology is limited to the first layer and if there exists a path from the first layer to each layer, even if each layer except the first layer is not connected, all layers are synchronized with precision levels. Moreover, we showed that small deviations in synchronization depend on the parameter σ, which is the deviation of the natural frequencies ai. If σ=0, then we have complete synchronization. Finally, we performed simulations to validate the theoretical result of this paper. In fact, there is a limitation in this simulation, that the Dirichlet eigenvalue cannot be directly calculated and its lower bound was used. For this reason, a larger value of the coupling strength c was required. Therefore, computing the Dirichlet eigenvalues is left for future work. In addition, there are different extended models of the present work we are currently addressing: FHN network with delay, a nonlinear dynamical network of Hindmarsh-Rose neurons, and FHN network with nonlinear operators instead of graph Laplacian. Additionally, we expect that the results and mathematical methodology can be applied to study other similar neural models with hierarchical architectures.

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

    This work was supported by a National Research Foundation of Korea(NRF) grant funded by the Korean government(MSIT) (No. 2021R1A2C1093929) and Heuron company.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.



    [1] M. Aldana, C. Huepe, Phase transitions in self-driven many-particle systems and related non-equilibrium models: A network approach, J. Stat. Phys., 112 (2003), 135–153. http://dx.doi.org/10.1023/A:1023675519930 doi: 10.1023/A:1023675519930
    [2] M. S. Ali, M. Hymavathi, B. Priya, S. A. Kauser, G. K. Thakur, Stability analysis of stochastic fractional-order competitive neural networks with leakage delay, AIMS Math., 6 (2021), 3205–3241. http://dx.doi.org/10.3934/math.2021193 doi: 10.3934/math.2021193
    [3] B. Ambrosio, M. Aziz-Alaoui, Synchronization and control of a network of coupled reaction-diffusion systems of generalized FitzHugh-Nagumo type, ESAIM: Proc., 39 (2013), 15–24. http://dx.doi.org/10.1051/proc/201339003 doi: 10.1051/proc/201339003
    [4] J. Baladron, D. Fasoli, O. Faugeras, J. Touboul, Mean-field description and propagation of chaos in networks of Hodgkin-Huxley and FitzHugh-Nagumo neurons, J. Math. Neurosc., 2 (2013), 1–50. http://dx.doi.org/10.1186/2190-8567-2-10 doi: 10.1186/2190-8567-2-10
    [5] A. Blaustein, F. Filbet, Concentration Phenomena in Fitzhugh–Nagumo Equations: A Mesoscopic Approach, SIAM J. Math. Anal., 55 (2023), 367–404. http://dx.doi.org/10.1137/22m1469961 doi: 10.1137/22m1469961
    [6] I. I. Blekhman, Gradient estimates for solutions of nonlinear elliptic and parabolic equations, New York: ASME, 1988.
    [7] M. Bossy, J. Fontbona, H. Olivero, Synchronization of stochastic mean field networks of Hodgkin–Huxley neurons with noisy channels, J. Math. Biol., 78 (2019), 1771–1820. http://dx.doi.org/10.1007/s00285-019-01326-7 doi: 10.1007/s00285-019-01326-7
    [8] Q. Cao, X. Long, New convergence on inertial neural networks with time-varying delays and continuously distributed delays, AIMS Math., 5 (2020), 5955–5968. http://dx.doi.org/10.3934/math.2020381 doi: 10.3934/math.2020381
    [9] T. Carletti, H. Nakao, Turing patterns in a network-reduced FitzHugh-Nagumo model, Phys. Rev. E., 101 (2020), 022203. http://dx.doi.org/10.1103/physreve.101.022203 doi: 10.1103/physreve.101.022203
    [10] D. Centola, M. Macy, Complex contagions and the weakness of long ties, Am. J. Sociol., 113 (2007), 702–734. http://dx.doi.org/10.1086/521848 doi: 10.1086/521848
    [11] F. Cheng, J. Li, Q. Yu, The existence of solitary wave solutions for the neuron model with conductance-resistance symmetry, AIMS Math., 8 (2023), 3322–3337. http://dx.doi.org/10.3934/math.2023171 doi: 10.3934/math.2023171
    [12] F. R. Chung, F. C. Graham, Spectral graph theory, In: American Mathematical Soc., 1997. http://dx.doi.org/10.1201/b16113-54
    [13] P. Degond, S. Motsch, Large scale dynamics of the persistent turning walker model of fish behavior, J. Statist. Phys., 131 (2008), 989–1021. http://dx.doi.org/10.1007/s10955-008-9529-8 doi: 10.1007/s10955-008-9529-8
    [14] J. Fell, N. Axmacher, The role of phase synchronization in memory processes, Nat. Rev. Neurosci., 12 (2011), 105–118. http://dx.doi.org/10.1038/nrn2979 doi: 10.1038/nrn2979
    [15] R. FitzHugh, Impulses and physiological states in theoretical models of nerve membrane, Biophys. J., 1 (1961), 445–466. http://dx.doi.org/10.1016/s0006-3495(61)86902-6 doi: 10.1016/s0006-3495(61)86902-6
    [16] P. Fries, A mechanism for cognitive dynamics: neuronal communication through neuronal coherence, Trends Cogn. Sci., 9 (2005), 474–480. http://dx.doi.org/10.1016/j.tics.2005.08.011 doi: 10.1016/j.tics.2005.08.011
    [17] M. Gerster, R. Berner, J. Sawicki, A. Zakharova, A. Škoch, J. Hlinka, et al., FitzHugh–Nagumo oscillators on complex networks mimic epileptic-seizure-related synchronization phenomena, Chaos, 30 (2020), 123130. http://dx.doi.org/10.1063/5.0021420 doi: 10.1063/5.0021420
    [18] M. M. Ibrahim, S. Iram, M. A. Kamran, M. M. Naeem Mannan, M. U. Ali, I. H. Jung, et al., Synchronization Analysis of Multiple FitzHugh-Nagumo Noisy and Nonnoisy Neurobiological Networks, Math. Probl. Eng., 2022 (2022), 9302758. http://dx.doi.org/10.1155/2022/9302758 doi: 10.1155/2022/9302758
    [19] M. M. Ibrahim, I. H. Jung, Complex synchronization of a ring-structured network of Fitzhugh-Nagumo neurons with single-and dual-state gap junctions under ionic gates and external electrical disturbance, IEEE Access, 7 (2019), 57894–57906. http://dx.doi.org/10.1109/access.2019.2913872 doi: 10.1109/access.2019.2913872
    [20] N. Iqbal, R. Wu, Y. Karaca, R. Shah, W. Weera, Pattern dynamics and Turing instability induced by self-super-cross-diffusive predator-prey model via amplitude equations, AIMS Math., 8 (2023), 2940–2960. http://dx.doi.org/10.3934/math.2023153 doi: 10.3934/math.2023153
    [21] A. Jadbabaie, J. Lin, A. S. Morse, Coordination of groups of mobile autonomous agents using nearest neighbor rules, IEEE Trans. Automat. Contr., 48 (2003), 988–1001. http://dx.doi.org/10.1109/tac.2003.812781 doi: 10.1109/tac.2003.812781
    [22] Y. Jiang, J. Wu, H. Yang, F. Xu, M. Wang, S. Huang, et al., Chimera States mediated by nonlocally attractive-repulsive coupling in FitzHugh–Nagumo neural networks, Chin. J. Phys., 66 (2020), 172–179. http://dx.doi.org/10.1016/j.cjph.2020.03.019 doi: 10.1016/j.cjph.2020.03.019
    [23] A. Kazemy, J. Lam, X. M. Zhang, Event-triggered output feedback synchronization of master–slave neural networks under deception attacks, IEEE T. Neur. Net. Lear., 33 (2020), 952–961. http://dx.doi.org/10.1109/tnnls.2020.3030638 doi: 10.1109/tnnls.2020.3030638
    [24] Y. Kuramoto, Chemical oscillations, waves, and turbulence, Courier Corporation, 2003. http://dx.doi.org/10.1007/978-3-642-69689-3
    [25] J. Liu, H. Wu, J. Cao, Event-triggered synchronization in fixed time for semi-Markov switching dynamical complex networks with multiple weights and discontinuous nonlinearity, Commun. Nonlinear Sci. Numer. Simul., 90 (2020), 105400. http://dx.doi.org/
    [26] F. Min, A. C. J. Luo, Complex Dynamics of Projective Synchronization of Chua Circuits with Different Scrolls, Int. J. Bifurcation Chaos, 25 (2015), 1530016. http://dx.doi.org/10.1142/S0218127415300165 doi: 10.1142/S0218127415300165
    [27] J. Nagumo, S. Arimoto, S. Yoshizawa, An active pulse transmission line simulating nerve axon, Proc. IRE, 50 (1962), 2061–2070. http://dx.doi.org/10.1109/jrproc.1962.288235 doi: 10.1109/jrproc.1962.288235
    [28] Z. T. Njitacke, J. Ramadoss, C. N. Takembo, K. Rajagopal, J. Awrejcewicz, An enhanced FitzHugh–Nagumo neuron circuit, microcontroller-based hardware implementation: Light illumination and magnetic field effects on information patterns, Chaos, Soliton. Fract., 167 (2023), 113014. http://dx.doi.org/10.1016/j.chaos.2022.113014 doi: 10.1016/j.chaos.2022.113014
    [29] Z. T. Njitacke, C. N. Takembo, J. Awrejcewicz, H. P. E. Fouda, J. Kengne, Hamilton energy, complex dynamical analysis and information patterns of a new memristive FitzHugh-Nagumo neural network, Chaos, Soliton. Fract., 160 (2022), 112211. http://dx.doi.org/10.1016/j.chaos.2022.112211 doi: 10.1016/j.chaos.2022.112211
    [30] S. Plotnikov, J. Lehnert, A. Fradkov, E. Schöll, Synchronization in heterogeneous FitzHugh-Nagumo networks with hierarchical architecture, Phys. Rev. E, 94 (2016), 012203. http://dx.doi.org/10.1103/physreve.94.012203 doi: 10.1103/physreve.94.012203
    [31] S. A. Plotnikov, A. L. Fradkov, On synchronization in heterogeneous FitzHugh–Nagumo networks, Chaos, Soliton. Fract., 121 (2019), 85–91. http://dx.doi.org/10.1016/j.chaos.2019.02.006 doi: 10.1016/j.chaos.2019.02.006
    [32] C. Quininao, J. D. Touboul, Clamping and Synchronization in the Strongly Coupled FitzHugh–Nagumo Model, SIAM J. Appl. Dyn. Syst., 19 (2020), 788–827. http://dx.doi.org/10.1137/19m1283884 doi: 10.1137/19m1283884
    [33] K. Rajgopal, A. Karthikeyan, V. R. Varun Raj, Dynamical behavior of pancreatic β cells with memductance flux coupling: Considering nodal properties and wave propagation in the excitable media, Chaos, Soliton. Fract., 165 (2022), 112857. http://dx.doi.org/10.1016/j.chaos.2022.112857 doi: 10.1016/j.chaos.2022.112857
    [34] Y. Shen, X. Liu, Event-based master–slave synchronization of complex-valued neural networks via pinning impulsive control, Neural Networks, 145 (2022), 374–385. http://dx.doi.org/10.1016/j.neunet.2021.10.025 doi: 10.1016/j.neunet.2021.10.025
    [35] L. Skrzypek, Y. You, Dynamics and synchronization of boundary coupled FitzHugh-Nagumo neural networks, Appl. Math. Comput., 388 (2021), 125545. http://dx.doi.org/10.1016/j.amc.2020.125545 doi: 10.1016/j.amc.2020.125545
    [36] D. Somers, N. Kopell, Rapid synchronization through fast threshold modulation, Biol. Cybern., 68 (1993), 393–407. http://dx.doi.org/10.1007/bf00198772 doi: 10.1007/bf00198772
    [37] P. Wang, Q. Zhang, H. Su, Aperiodically intermittent control for synchronization of discrete-time delayed neural networks, J. Franklin Inst., 359 (2022), 4915–4937. http://dx.doi.org/10.1016/j.jfranklin.2022.04.033 doi: 10.1016/j.jfranklin.2022.04.033
    [38] D. Wei, X. Luo, Y. Zou, Firing activity of complex space-clamped FitzHugh-Nagumo neural networks, Eur. Phys. J. B, 63 (2008), 279–282. http://dx.doi.org/10.1140/epjb/e2008-00227-5 doi: 10.1140/epjb/e2008-00227-5
    [39] Y. Xu, Y. Li, W. Li, Adaptive finite-time synchronization control for fractional-order complex-valued dynamical networks with multiple weights, Commun. Nonlinear Sci. Numer. Simul., 85 (2020), 105239. http://dx.doi.org/10.1016/j.cnsns.2020.105239 doi: 10.1016/j.cnsns.2020.105239
    [40] Z. Yong, Z. Su-Hua, Z. Tong-Jun, A. Hai-Long, Z. Zhen-Dong, H. Ying-Rong, et al., The synchronization of FitzHugh–Nagumo neuron network coupled by gap junction, Chin. Phys. B, 17 (2008), 2297–2303. http://dx.doi.org/10.1088/1674-1056/17/6/061 doi: 10.1088/1674-1056/17/6/061
    [41] X. Zhang, F. Min, Y. Dou, Y. Xu, Bifurcation analysis of a modified FitzHugh-Nagumo neuron with electric field, Chaos, Soliton. Fract., 170 (2023), 113415. http://dx.doi.org/10.1016/j.chaos.2023.113415 doi: 10.1016/j.chaos.2023.113415
    [42] X. Zhang, W. Zhou, H. R. Karimi, Y. Sun, Finite- and fixed-time cluster synchronization of nonlinearly coupled delayed neural networks via pinning control, IEEE T. Neur. Net. Lear., 32 (2020), 5222–5231. http://dx.doi.org/10.1109/tnnls.2020.3027312 doi: 10.1109/tnnls.2020.3027312
    [43] B. Zhen, Z. Song, The Study for Synchronization between Two Coupled FitzHugh-Nagumo Neurons Based on the Laplace Transform and the Adomian Decomposition Method, Neural Plast., 2021 (2021), 6657835. http://dx.doi.org/10.1155/2021/6657835 doi: 10.1155/2021/6657835
    [44] S. Zhu, H. Bao, J. Cao, Bipartite synchronization of coupled delayed neural networks with cooperative-competitive interaction via event-triggered control, Phys. A, 600 (2022), 127586. http://dx.doi.org/10.1016/j.physa.2022.127586 doi: 10.1016/j.physa.2022.127586
  • Reader Comments
  • © 2023 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(1519) PDF downloads(81) Cited by(0)

Figures and Tables

Figures(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog