Loading [MathJax]/jax/output/SVG/jax.js

Emergence of synchronization in Kuramoto model with frustration under general network topology

  • Received: 01 August 2021 Revised: 01 December 2021 Published: 22 February 2022
  • Primary: 34D06, 34C15; Secondary: 92B25, 70F99

  • In this paper, we will study the emergent behavior of Kuramoto model with frustration on a general digraph containing a spanning tree. We provide a sufficient condition for the emergence of asymptotical synchronization if the initial data are confined in half circle. As lack of uniform coercivity in general digraph, we apply the node decomposition criteria in [25] to capture a clear hierarchical structure, which successfully yields the dissipation mechanism of phase diameter and an invariant set confined in quarter circle after some finite time. Then the dissipation of frequency diameter will be clear, which eventually leads to the synchronization.

    Citation: Tingting Zhu. Emergence of synchronization in Kuramoto model with frustration under general network topology[J]. Networks and Heterogeneous Media, 2022, 17(2): 255-291. doi: 10.3934/nhm.2022005

    Related Papers:

    [1] Tingting Zhu . Synchronization of the generalized Kuramoto model with time delay and frustration. Networks and Heterogeneous Media, 2023, 18(4): 1772-1798. doi: 10.3934/nhm.2023077
    [2] Xiaoxue Zhao, Zhuchun Li . Synchronization of a Kuramoto-like model for power grids with frustration. Networks and Heterogeneous Media, 2020, 15(3): 543-553. doi: 10.3934/nhm.2020030
    [3] Seung-Yeal Ha, Yongduck Kim, Zhuchun Li . Asymptotic synchronous behavior of Kuramoto type models with frustrations. Networks and Heterogeneous Media, 2014, 9(1): 33-64. doi: 10.3934/nhm.2014.9.33
    [4] Seung-Yeal Ha, Jaeseung Lee, Zhuchun Li . Emergence of local synchronization in an ensemble of heterogeneous Kuramoto oscillators. Networks and Heterogeneous Media, 2017, 12(1): 1-24. doi: 10.3934/nhm.2017001
    [5] Seung-Yeal Ha, Se Eun Noh, Jinyeong Park . Practical synchronization of generalized Kuramoto systems with an intrinsic dynamics. Networks and Heterogeneous Media, 2015, 10(4): 787-807. doi: 10.3934/nhm.2015.10.787
    [6] Seung-Yeal Ha, Jeongho Kim, Jinyeong Park, Xiongtao Zhang . Uniform stability and mean-field limit for the augmented Kuramoto model. Networks and Heterogeneous Media, 2018, 13(2): 297-322. doi: 10.3934/nhm.2018013
    [7] Seung-Yeal Ha, Hansol Park, Yinglong Zhang . Nonlinear stability of stationary solutions to the Kuramoto-Sakaguchi equation with frustration. Networks and Heterogeneous Media, 2020, 15(3): 427-461. doi: 10.3934/nhm.2020026
    [8] Young-Pil Choi, Seung-Yeal Ha, Seok-Bae Yun . Global existence and asymptotic behavior of measure valued solutions to the kinetic Kuramoto--Daido model with inertia. Networks and Heterogeneous Media, 2013, 8(4): 943-968. doi: 10.3934/nhm.2013.8.943
    [9] Vladimir Jaćimović, Aladin Crnkić . The General Non-Abelian Kuramoto Model on the 3-sphere. Networks and Heterogeneous Media, 2020, 15(1): 111-124. doi: 10.3934/nhm.2020005
    [10] Wenlian Lu, Fatihcan M. Atay, Jürgen Jost . Consensus and synchronization in discrete-time networks of multi-agents with stochastically switching topologies and time delays. Networks and Heterogeneous Media, 2011, 6(2): 329-349. doi: 10.3934/nhm.2011.6.329
  • In this paper, we will study the emergent behavior of Kuramoto model with frustration on a general digraph containing a spanning tree. We provide a sufficient condition for the emergence of asymptotical synchronization if the initial data are confined in half circle. As lack of uniform coercivity in general digraph, we apply the node decomposition criteria in [25] to capture a clear hierarchical structure, which successfully yields the dissipation mechanism of phase diameter and an invariant set confined in quarter circle after some finite time. Then the dissipation of frequency diameter will be clear, which eventually leads to the synchronization.



    Synchronized behavior in complex systems is ubiquitous and has been extensively investigated in various academic communities such as physics, biology, engineering [2,7,29,36,37,39,40,42], etc. Recently, sychronization mechanism has been applied in control of robot systems and power systems [12,13,34]. The rigorous mathematical treatment of synchronization phenomena was started by two pioneers Winfree [43] and Kuramoto [27,28] several decades ago, who introduced different types of first-order systems of ordinary differential equations to describe the synchronous behaviors. These models contain rich emergent behaviors such as synchronization, partially phase-lcoking and nonlinear stability, etc., and have been extensively studied in both theoretical and numerical level [3,5,11,14,15,16,17,19,26,32,39].

    In this paper, we address the synchronous problem of Kuramoto model on a general graph under the effect of frustration. To fix the idea, we consider a digraph G=(V,E) consisting of a finite set V={1,,N} of vertices and a set EV×V of directed arcs. We assume that Kuramoto oscillators are located at vertices and interact with each other via the underlying network topology. For each vertex i, we denote the set of its neighbors by Ni, which is the set of vertices that directly influence vertex i. Now, let θi=θi(t) be the phase of the Kuramoto oscillator at vertex i, and define the (0,1)-adjacency matrix (χij) as follows:

    χij={1if thejth oscillator influences theith oscillator,0otherwise.

    Then, the set of neighbors of i-th oscillator is actually Ni:={j:χij>0}. In this setting, the Kuramoto model with frustration on a general network is governed by the following ODE system:

    {˙θi(t)=Ωi+KjNisin(θj(t)θi(t)+α),t>0,iV,θi(0)=θi0, (1)

    where Ωi,K,N and α(0,π2) are the natural frequency of the ith oscillator, coupling strength, the number of oscillators and the uniform frustration between oscillators, respectively. For the case of nonpositive frustration, we can reformulate such a system into (1) form by taking ˆθi=θi for i=1,2,,N. Note that the well-posedness of system (1) is guaranteed by the standard Cauchy-Lipschitz theory since the vector field on the R.H.S of (1) is analytic.

    Comparing to the original Kuramoto model, there are two additional structures, i.e., frustration and general digraph. The frustration was introduced by Sakaguchi and Kuramoto [38], due to the observation that a pair of strongly coupled oscillators eventually oscillate with a common frequency that deviates from the average of their natural frequencies. On the other hand, the original all-to-all symmetric network is an ideal setting, thus it is natural to further consider general digraph case. Therefore, the frustration model with general digraph is more realistic in some sense. Moreover, these two structures also lead to richer phenomenon. For instance, the author in [6] observed that the frustration is common in disordered interactions, and the author in [44] found that frustration can induce the desynchronization through varying the value of α in numerical simulations. For more information, please refer to [4,10,21,23,24,25,30,33,35,41].

    However, mathematically, for the Kuramoto model, the frustration and general digraph structures generate a lot of difficulties in rigorous analysis. For instance, the conservation law and gradient flow structure are lost, and thus the asymptotic states and dissipation mechanism become non-trivial. For all-to-all and symmetric case with frustration, in [20], the authors provided sufficient frameworks leading to complete synchronization under the effect of uniform frustration. In their work, they required initial configuration to be confined in half circle. Furthermore, the authors in [31] dealt with the stability and uniqueness of emergent phase-locked states. In particular, the authors in [22] exploited order parameter approach to study the identical Kuramoto oscillators with frustration. They showed that an initial configuration whose order parameter is bounded below will evolve to the complete phase synchronization or the bipolar state exponentially fast. On the other hand, for non-all-to-all case without frustration, the authors in [9] lifted the Kuramoto model to second-order system such that the second-order formulation enjoys several similar mathematical structures to that of Cucker-Smale flocking model [8]. But this method only works when the size of initial phases is less than a quarter circle, as we know the cosine function becomes negative if π2<θ<π. In [45], for the Kuramoto model on general network whose initial configuration is distributed in half circle, the authors exploited the idea of node decomposition in [25] and showed that frequency synchronization emerges under sufficient frameworks. To the best knowledge of the authors, there is few work on the Kuramoto model over general digraph with frustration. The authors in [18] studied the Kuramoto model with frustrations on a complete graph which is a small perturbation of all-to-all network, and provided synchronization estimates in half circle.

    Our interest in this paper is studying the system (1) with uniform frustration on a general digraph. As far as the authors know, when the ensemble is distributed in half circle, the dissipation structure of the Kuramoto type model with general digragh is still unclear. The main difficulty comes from the loss of uniform coercive inequality, which is due to the non-all-to-all and non-symmetric interactions. Thus we cannot expect to capture the dissipation from Gronwall-type inequality of phase diameter. For example, the time derivative of the phase diameter may be zero at some time for general digraph case. To this end, we switch to apply the idea of node decomposition in [25] to gain the dissipation through hypo-coercivity. Due to the lack of monotonicity of sine function in half circle, we follow the delicate constructions and estimates of the convex combinations in [45]. Eventually, we have the following main theorem.

    Theorem 1.1. Suppose the network topology (χij) in system (1) contains a spanning tree, D is a given positive constant such that D<π2, and all oscillators are initially confined in half circle, i.e.,

    D(θ(0))<π.

    Then for sufficient large coupling strength K and small frustration α, there exists a finite time t>0 such that

    D(θ(t))D,t[t,),

    and

    D(ω(t))C1eC2(tt),tt,

    where C1 and C2 are positive constants.

    Remark 1. The first part of Theorem 1.1 claims that all oscillators confined initially in half circle will enter into a region less than quarter cirlce after some finite time. It is natural to ask how large K and how small α we need to guarantee the first part of Theorem 1.1. In fact, according to the proof in later sections, we have the following explicit constraints on K and α,

    tanα<1(1+(d+1)ζζD(θ(0)))2Ncβd+1D[4(2N+1)c]d,D+α<π2,1>(1+(d+1)ζζD(θ(0)))c[4(2N+1)c]dβd+1D(D(Ω)Kcosα+2Nsinαcosα), (2)

    where d is the number of general nodes which is smaller than N (see Lemma 2.7), D(Ω) is the diameter of natural frequency, and the other parameters ζ, γ, η, β and c are positive constants which satisfy the following properties,

    D(θ(0))<ζ<γ<π,η>max{1sinγ,21ζγ},β=12η,c=(N1j=1ηjA(2N,j)+1)γsinγ, (3)

    where A(2N,j) denotes the number of permutations. It's obvious that we can find admissible parameters satisfying (3) since D(θ(0))<π. Once the parameters are fixed, we immediately conclude (2) holds for small α and large K.

    Remark 2. For t[0,t), the phase diameter D(θ(t)) satisfies D(θ(t))<ζ which follows from t<ˉt in Lemma 4.5 and Lemma 2.8. After t, all oscillators are confined in a region less than quarter circle, and Kuramoto model (1) will be equivalent to Cucker-Smale type model with frustration (see (53)). Therefore, we can directly apply the methods and results in [9] to conclude the emergence of frequency synchronization for large coupling and small frustration (see Lemma 4.6). Therefore, to guarantee the emergence of synchronization, it suffices to show the detailed proof of the first part of Theorem 1.1.

    The rest of the paper is organized as follows. In Section 2, we recall some concepts on the network topology and provide a priori local-in-time estimate on the phase diameter of whole ensemble with frustration. In Section 3, we consider a strongly connected ensemble with frustration for which the initial phases are distributed in a half circle. We show that for large coupling strength and small frustration, the phase diameter is uniformly bounded by a value less than π2 after some finite time. In Section 4, we study the general network with a spanning tree structure under the effect of uniform frustration. We use the inductive argument and show that Kuramoto oscillators will concentrate into a region less than quarter circle in finite time, which eventually leads to the emergence of synchronization exponentially fast. In Section 5, we present some simulations to illustrate the main results in Theorem 1.1. Section 6 is devoted to a brief summary.

    In this section, we first introduce some fundamental concepts such as synchronization, spanning tree and node decomposition of a general network (1). Then, we will provide some necessary notations and a priori estimate that will be frequently used in later sections.

    For the Kuramoto-type model, we recall the definition of synchronization as follows.

    Definition 2.1. Let θ(t):=(θ1(t),θ2(t),,θN(t)) be a time-dependent Kuramoto phase vecor. The configuration θ(t) exhibits (asymptotic) complete frequency synchronization if and only if the relative frequencies tend to zero asymptotically:

    limt|˙θi(t)˙θj(t)|=0, ij.

    Let the network topology be registered by the neighbor set Ni which consists of all neighbors of the ith oscillator. Then, for a given set of {Ni}Ni=1 in system (1), we have the following definition.

    Definition 2.2. (1)The Kuramoto digraph G=(V,E) associated to system (1) consists of a finite set V={1,2,,N} of vertices, and a set EV×V of arcs with ordered pair (j,i)E if jNi.

    (2)A path in G from i1 to ik is a sequence i1,i2,,ik such that

    isNis+1for 1sk1.

    If there exists a path from j to i, then vertex i is said to be reachable from vertex j.

    (3)The Kuramoto digraph contains a spanning tree if we can find a vertex such that any other vertex of G is reachable from it.

    In order to guarantee the emergence of synchronization, we will always assume the existence of a spanning tree throughout the paper. Now we recall the concepts of root and general root introduced in [25]. Let l,kN with 1lkN, and let Cl,k=(cl,cl+1,,ck) be a vector in Rkl+1 such that

    ci0,likandki=lci=1.

    For an ensembel of N-oscillators with phases {θi}Ni=1, we set Lkl(Cl,k) to be a convex combination of {θi}ki=l with the coefficient Cl,k:

    Lkl(Cl,k):=ki=lciθi.

    Note that each θi is a convex combination of itself, and particularly θN=LNN(1) and θ1=L11(1).

    Definition 2.3. (Root and general root)

    1. We say θk is a root if it is not affected by the rest oscillators, i.e., jNk for any j{1,2,,N}{k}.

    2. We say Lkl(Cl,k) is a general root if Lkl(Cl,k) is not affected by the rest oscillators, i.e., for any i{l,l+1,,k} and j{1,2,,N}{l,l+1,,k}, we have jNi.

    Lemma 2.4.[25] The following assertions hold.

    1. If the network contains a spanning tree, then there is at most one root.

    2. Assume the network contains a spanning tree. If LNk(Ck,N) is a general root, then Ll1(C1,l) is not a general root for each l{1,2,,k1}.

    In this part, we will recall the concept of maximum node introduced in [25]. Then, we can follow node decomposition introduced in [25] to represent the whole graph G (or say vertex set V) as a disjoint union of a sequence of nodes. The key point is that the node decomposition shows a hierarchical structure, then we can exploit this advantage to apply the induction principle. Let G=(V,E),V1V, and a subgraph G1=(V1,E1) is the digraph with vertex set V1 and arc set E1 which consists of the arcs in G connecting agents in V1. For a given digraph G=(V,E), we will identify a subgraph G1=(V1,E1) with its vertex set V1 for convenience. Now we first present the definition of nodes below.

    Definition 2.5. [25] (Node) Let G be a digraph. A subset G1 of vertices is called a node if it is strongly connected, i.e., for any subset G2 of G1, G2 is affected by G1G2. Moreover, if G1 is not affected by GG1, we say G1 is a maximum node.

    Intuitively, a node can be understood through a way that a set of oscillators can be viewed as a "large" oscillator. The concept of node can be exploited to simplify the structure of the digraph, which indeed helps us to catch the attraction effect more clearly in the network topology.

    Lemma 2.6.[25] Any digraph G contains at least one maximum node. A digraph G contains a unique maximum node if and only if G has a spanning tree.

    Lemma 2.7.[25] (Node decomposition)Let G be any digraph. Then we can decompose G to be a union as G=di=0(kij=1Gji) such that

    1. Gj0 are the maximum nodes of G, where 1jk0.

    2. For any p,q where 1pd and 1qkp, Gqp are the maximum nodes of G(p1i=0(kij=1Gji)).

    Proof. As G is assumed to be any digraph, from Lemma 2.2, we see that G contains at least one maximum node. Therefore, we can find all maximum nodes of G and denote them by Gj0 with 1jk0, where k0 is the number of maximum nodes. Next, we get rid of k0j=1Gj0 and collect all maximum nodes in the remaining digraph G(k0j=1Gj0). Denote all maximum nodes of the remainder G(k0j=1Gj0) by Gj1 with 1jk1, provided that there are k1 maximum nodes for G(k0j=1Gj0). Then we can repeat the same process and find the maximum nodes Gqp of G(p1i=0(kij=1Gji)) with 1qkp. After d steps, we can construct G=(di=0(kij=1Gji)) as G consists of finite N oscillators.

    Remark 3. Lemma 2.7 shows a clear hierarchical structure on a general digraph. For the convenience of later analysis, we make some comments on important notations and properties that are used throughout the paper.

    1. From the definition of maximum node, for 1qqkp, we see that Gqp and Gqp do not affect each other. Actually, Gqp will only be affected by G0 and Gji, where 1ip1, 1jki. Thus in the proof of our main theorem (see Theorem 1.1), without loss of generality, we may assume ki=1 for all 1id. Hence, the decomposition can be further simplified and expressed by

    G=di=0Gi,

    where Gp is a maximum node of G(p1i=0Gi).

    2.For a given oscillator θk+1iGk+1, we denote by k+1j=0Nk+1i(j) the set of neighbors of θk+1i, where Nk+1i(j) represents the neighbors of θk+1i in Gj. Note that the node decomposition and spanning tree structure in G guarantee that kj=0Nk+1i(j).

    In this part, for notational simplicity, we introduce some notations, such as the extreme phase, phase diameter of G and the first k+1 nodes, natural frequency diameter, and cardinality of subdigraph:

    θ=(θ1,θ2,,θN),ω=(ω1,ω2,,ωN),Ω=(Ω1,Ω2,,ΩN),θM=max1kN{θk}=max0idmax1jNi{θij},θm=min1kN{θk}=min0idmin1jNi{θij},D(θ)=θMθm,Dk(θ)=max0ikmax1jNi{θij}min0ikmin1jNi{θij},ΩM=max0idmax1jNi{Ωij},Ωm=min0idmin1jNi{Ωij},D(Ω)=ΩMΩm,Ni=|Gi|,Sk=ki=0Ni,0kd,di=0Ni=N.

    Finally, we provide a priori local-in-time estimate on the phase diameter to finish this section, which states that all oscillators can be confined in half circle in short time.

    Lemma 2.8. Let θi be a solution to system (1) and suppose the initial phase diameter satisfies

    D(θ(0))<ζ<γ<π,

    where ζ and γ are positive constants less than π.Then there exists a finite time ˉt>0 such that the phase diameter of whole ensemble remains less than ζ before ˉt, i.e.,

    D(θ(t))<ζ,t[0,ˉt)whereˉt=ζD(θ(0))D(Ω)+2NKsinα.

    Proof. From system (1), we see that the dynamics of extreme phases is given by the following equations

    ˙θM(t)=ΩM+KjNMsin(θj(t)θM(t)+α),˙θm(t)=Ωm+KjNmsin(θj(t)θm(t)+α),

    where θM and θm denote the maximum and minimum phases. We combine the above equations to estimate the dynamics of phase diameter

    ˙D(θ(t))=˙θM(t)˙θm(t)=ΩMΩm+KjNMsin(θjθM+α)KjNmsin(θjθm+α)D(Ω)+KjNM[sin(θjθM)cosα+cos(θjθM)sinα]KjNm[sin(θjθm)cosα+cos(θjθm)sinα]=D(Ω)+Kcosα(jNMsin(θjθM)jNmsin(θjθm))+Ksinα(jNMcos(θjθM)jNmcos(θjθm)), (4)

    where ΩM and Ωm are the natural frequencies of the extreme phases θM and θm respectively, and we use the formula

    sin(x+y)=sinxcosy+cosxsiny

    When the phase diameter satisfies D(θ(t))ζ<π, it is obvious that

    sin(θjθM)0, jNMandsin(θjθm)0, jNm.

    Then we see from (4) that

    ˙D(θ)D(Ω)+2NKsinα. (5)

    That is to say, when D(θ(t))ζ<π, the growth of phase diameter is no greater than the linear growth with positive slope D(Ω)+2NKsinα. Now we integrate on both sides of (5) from 0 to t to have

    D(θ(t))D(θ(0))+(D(Ω)+2NKsinα)t.

    Therefore, it yields that there exists a finite time ˉt>0 such that

    D(θ(t))<ζ, t[0,ˉt),

    where ˉt is given as below

    ˉt=ζD(θ(0))D(Ω)+2NKsinα.

    We will first study the special case, i.e., the network is strongly connected. Without loss of generality, we denote by G0 the strongly connected digraph. According to Definition 2.5, Lemma 2.6 and Lemma 2.7, the strongly connected network consists of only one maximum node. Then in the present special case, according to Remark 2, we only need to show that the phase diameter is uniformly bounded by a value less than π2 after some finite time, which ultimately yields the emergence of frequency synchronization. We now introduce an algorithm to construct a proper convex combination of oscillators, which can involve the dissipation from interaction of general network. More precisely, the algorithm for G0 consists of the following three steps:

    Step 1. For any given time t, we reorder the oscillator indexes to make the oscillator phases from minimum to maximum. More specifically, by relabeling the agents at time t, we set

    θ01(t)θ02(t)θ0N0(t). (6)

    In order to introduce the following steps, we first provide the process of iterations for ˉLN0k(ˉCk,N0) and L_l1(C_1,l) as follows:

    (A1): If ˉLN0k(ˉCk,N0) is not a general root, then we construct

    ˉLN0k1(ˉCk1,N0)=ˉak1ˉLN0k(ˉCk,N0)+θ0k1ˉak1+1.

    (A2): If L_l1(C_1,l) is not a general root, then we construct

    L_l+11(C_1,l+1)=a_l+1L_l1(C_1,l)+θ0l+1a_l+1+1

    Step 2. According to the strong connectivity of G0, we immediately know that ˉLN01(ˉC1,N0) is a general root, and ˉLN0k(ˉCk,N0) is not a general root for k>1. Therefore, we may start from θ0N0 and follow the process A1 to construct ˉLN0k(ˉCk,N0) until k=1.

    Step 3. Similarly, we know that L_N01(C_1,N0) is a general root and L_l1(C_1,l) is not a general root for l<N0. Therefore, we may start from θ01 and follow the process A2 until l=N0.

    It is worth emphasizing that the order of the oscillators may change along time t, but at each time t, the above algorithm does work. For convenience, the algorithm from Step 1 to Step 3 will be referred to as Algorithm A. Then, based on Algorithm A, we will provide a priori estimates on a monotone property about the sine function, which will be crucially used later.

    Lemma 3.1. Let {θ0i} be a solution to system (1) with strongly connected network G0. Moreover at time t, we also assume that the oscillators are well-ordered as (6), and the phase diameter and quantity η satisfiy the following conditions:

    D0(θ(t))<γ<π,η>max{1sinγ,21ζγ},

    where ζ,γ are given in the condition (3).Then at time t, the following assertions hold

    {N0i=n(ηinminjN0i(0)jisin(θ0jθ0i))sin(θ0ˉknθ0N0), ˉkn=minjN0i=nN0i(0)j, 1nN0,ni=1(ηnimaxjN0i(0)jisin(θ0jθ0i))sin(θ0k_nθ01), k_n=maxjni=1N0i(0)j, 1nN0.

    Proof. For the proof of this lemma, please see [45] for details.

    Recall the strongly connected ensemble G0, and denote by θ0i (i=1,2,,N0) the members in G0. For the oscillators in G0, based on a priori estimates in Lemma 3.1, we will design proper coefficients of convex combination which helps us to capture the dissipation structure. Now we assume that at time t, the oscillators in G0 are well-ordered as follows,

    θ01(t)θ02(t)θ0N0(t).

    Then we apply the process A1 from θ0N0 to θ01 and the process A2 from θ01 to θ0N0 to respectively construct

    ˉLN0k1(ˉCk1,N0) with ˉa0N0=0, ˉa0k1=η(2N0k+2)(ˉa0k+1),2kN0,L_k+11(C_1,k+1) with a_01=0, a_0k+1=η(k+1+N0)(a_0k+1),1kN01, (7)

    where N0 is the cardinality of G0 and η is given in the condition (3). By induction criteria, we can deduce explict expressions about the constructed coefficients:

    ˉa0k1=N0k+1j=1ηjA(2N0k+2,j),2kN0,a_0k+1=kj=1ηjA(k+1+N0,j),1kN01.

    Note that ˉa0N0+1i=a_0i, i=1,2,N0. And we set

    ˉθ0k:=ˉLN0k(ˉCk,N0),θ_0k:=L_k1(C_1,k),1kN0. (8)

    We define a non-negative quantity Q0=ˉθ0θ_0 where ˉθ0=ˉθ01 and θ_0=θ_0N0. Note that Q0(t) is Lipschitz continuous with respect to t. We then establish the comparison relation between Q0 and the phase diameter D0(θ) of G0 in the following lemma.

    Lemma 3.2. Let {θ0i} be a solution to system (1) with strongly connected digraph G0. Assume that for the group G0, the coefficients ˉa0k's and a_0k's satisfy the scheme (7). Then at each time t, we have the following relation

    βD0(θ(t))Q0(t)D0(θ(t)),β=12η,

    where η satisfies the condition (3).

    Proof. As we choose the same design for coefficients of convex combination as that in [45], the proof of this lemma is same as that in [45], please see [45] for details.

    From Lemma 3.2, we see that the quantity Q0 can control the phase diameter D0(θ), which play a key role in analysing the bound of phase diameter. Based on Algorithm A and Lemma 3.1, we first study the dynamics of the constructed quantity Q0.

    Lemma 3.3. Let {θ0i} be the solution to system (1) with strongly connected digraph G0. Moreover, for a given positive constant D<min{π2,ζ}, assume the following conditions hold,

    D0(θ(0))<ζ<γ<π,η>max{1sinγ,21ζγ},tanα<1(1+ζζD(θ(0)))2N0cβD,D+α<π2,K>(1+ζζD0(θ(0))(D(Ω)+2N0Ksinα)ccosα1βD, (9)

    where ζ,γ are constants, β is given in Lemma 3.2 andc=(N01j=1ηjA(2N0,j)+1)γsinγ.Then the phase diameter of the graph G0 is uniformly bounded by γ:

    D0(θ(t))<γ,t[0,+),

    and the dynamics of Q0(t) is controlled by the following differential inequality

    ˙Q0(t)D(Ω)+2N0KsinαKcosαcQ0(t),t[0,+).

    Proof. The proof is similar to [45] under the assumption that the frustration α is sufficiently small. However, due to the presence of frustration, there are some slight differences in the process of analysis, thus we put the detailed proof in the Appendix A.

    Lemma 3.3 states that the phase diameter of the digraph G0 remains less than π and provides the dynamics of Q0. We next exploit the dynamics of Q0 and find some finite time such that all oscillators in G0 will be trapped into a region of quarter circle after the time.

    Lemma 3.4. Let {θ0i} be a solution to system (1) with strongly connected digraph G0, and suppose the assumptions in Lemma 3.3 hold. Then there exists time t00 such that

    D0(θ(t))D,for t[t0,+),

    where t0 can be estimated as below and bounded by ˉt given in Lemma 2.8

    t0<ζKcosαcβD(D(Ω)+2N0Ksinα)<ˉt. (10)

    Remark 4. According to Lemma 2.8, we see that D0(θ(t))<ζ for t[0,t0) since t0<ˉt.

    Proof. From Lemma 3.3, we see that the dynamics of quantity Q0(t) is governed by the following inequality

    ˙Q0(t)D(Ω)+2N0KsinαKcosαcQ0(t),t[0,+). (11)

    We next show that there exists some time t0 such that the quantity Q0 in (11) is uniformly bounded after t0. There are two cases we consider separately.

    Case 1. We first consider the case that Q0(0)>βD. When Q0(t)[βD,Q0(0)], from (66), we have

    ˙Q0(t)D(Ω)+2N0KsinαKcosαcQ0(t)D(Ω)+2N0KsinαKcosαcβD<0. (12)

    That is to say, when Q0(t) is located in the interval [βD,Q0(0)], Q0(t) will keep decreasing with a rate bounded by a uniform slope. Then we can define a stopping time t0 as follows,

    t0=inf{t0 | Q0(t)βD}.

    And based on the definition of t0, we see that Q0 will decrease before t0 and has the following property at t0,

    Q0(t0)=βD. (13)

    Moerover, from (12), it is easy to see that the stopping time t0 satisfies the following upper bound estimate,

    t0Q0(0)βDKcosαcβD(D(Ω)+2N0Ksinα). (14)

    Now we study the upper bound of Q0 on [t0,+). In fact, we can apply (12), (13) and the same arguments in (64) to derive

    Q0(t)βD, t[t0,+). (15)

    Case 2. For another case that Q0(0)βD. We can apply the similar analysis in (64) to obtain

    Q0(t)βD,t[0,+). (16)

    Then in this case, we directly set t0=0.

    Therefore, from (15), (16), and Lemma 3.2, we derive the upper bound of D0(θ) on [t0,+) as below

    D0(θ(t))Q0(t)βD,for t[t0,+). (17)

    On the other hand, in order to verify (10), we further study t0 in (17). Combining (14) in Case 1 and t0=0 in Case 2, we see that

    t0<ζKcosαcβD(D(Ω)+2N0Ksinα). (18)

    Here, we use the truth that Q0(0)D0(θ(0))<ζ. Then from the assumption about K in (9), i.e.,

    K>(1+ζζD0(θ(0))(D(Ω)+2N0Ksinα)ccosα1βD,

    it yields the following estimate about t0,

    t0<ζ(1+ζζD0(θ(0)))(D(Ω)+2N0Ksinα)(D(Ω)+2N0Ksinα)=ζD0(θ(0))D(Ω)+2N0Ksinα=ˉt, (19)

    where in this special strongly connected case, it's clear that N0=N and D0(θ)=D(θ) in Lemma 2.8.

    Thus, combining (17), (18) and (19), we derive the desired results.

    In this section, we investigate the general network with a spanning tree structure, and prove our main result Theorem 1.1, which state that synchronization will emerge for Kuramoto model with frustration. According to Definition 2.5 and Lemma 2.6, we see that the digraph G associated to system (1) has a unique maximum node if it contains a spanning tree structure. And From Remark 3, without loss of generality, we assume G is decomposed into a union as G=di=0Gi, where Gp is a maximum node of G(p1i=0Gi).

    We have studied the situation d=0 in Section 3, and we showed that the phase diameter of the digraph G0 is uniformly bounded by a value less than π2 after some finite time, i.e., the oscillators of G0 will concentrate into a region of quarter circle. However, for the case that d>0, Gk's are not maximum nodes in G for k1. Hence, the methods in Lemma 3.3 and Lemma 3.4 can not be directly exploited for the situation d>0. More precisely, the oscillators in Gi with i<k perform as an attraction source and affect the agents in Gk. Thus when we study the behavior of agents in Gk, the information from Gi with i<k can not be ignored.

    From Remark 3 and node decomposition, the graph G can be represented as

    G=dk=0Gk,|Gk|=Nk,

    and we denote the oscillators in Gk by θki with 1iNk. Then we assume that at time t, the oscillators in each Gk are well-ordered as below:

    θk1(t)θk2(t)θkNk(t),0kd.

    For each subdigraph Gk with k0 which is strongly connected, we follow the process in Algorithm A1 and A2 to construct ˉLNkl1(ˉCl1,Nk) and L_l+11(C_1,l+1) by redesigning the coefficients ˉakl and a_kl of convex combination as below:

    {ˉLNkl1(ˉCl1,Nk) with ˉakNk=0, ˉakl1=η(2Nl+2)(ˉakl+1),2lNk,L_l+11(C_1,l+1) with a_k1=0, a_kl+1=η(l+1+2NNk)(a_kl+1),1lNk1. (20)

    By induction principle, we deduce that

    {ˉakl1=Nkl+1j=1ηjA(2Nl+2,j),2lNk,a_kl+1=lj=1ηjA(l+1+2NNk,j),1lNk1.

    Note that ˉakNk+1i=a_ki, i=1,2,Nk. By simple calculation, we have

    ˉak1=Nk1j=1(ηjA(2N,j)),ˉak1N1j=1(ηjA(2N,j)),0kd. (21)

    And we further introduce the following notations,

    ˉθkl:=ˉLNkl(ˉCl,Nk),θ_kl:=L_l1(C_1,l),1lNk,0kd, (22)
    ˉθk:=ˉLNk1(ˉC1,Nk),θ_k:=L_Nk1(C_1,Nk),0kd, (23)
    Qk(t):=max0ik{ˉθi}min0ik{θ_i},0kd. (24)

    Due to the analyticity of the solution, Qk(t) is Lipschitz continuous. Similar to Section 3, we will first establish the comparison between the quantity Qk(t) and phase diameter Dk(θ(t)) of ki=0Gi, which plays a crucial role in later analysis.

    Lemma 4.1. Let θ={θi} be a solution to system (1), and assume that the network contains a spanning tree and for each subdigraph Gk, the coefficients ˉakl and a_kl of convex combination in Algorithm A satisfy the scheme (20). Then at each time t, we have the following relation

    βDk(θ(t))Qk(t)Dk(θ(t)),0kd,β=12η,

    where Dk(θ)=max0ikmax1jNi{θij}min0ikmin1jNi{θij} and η satisfies the condition (3).

    Proof. As we adopt the same construction of coefficients of convex combination in [45] which deals with the Kuramoto model without frustration on a general network, thus for the detailed proof of this lemma, please see [45].

    Now we are ready to prove our main Theorem 1.1. To this end, we will follow similar arguments in Section 3 to complete the proof. Actually, we will investigate the dynamics of the constructed quantity Qk(t) that involves the influences from Gi with i<k, which yields the hypo-coercivity of the phase diameter. Applying similar arguments in Lemma 3.3 and Lemma 3.4, we have the following estimates for G0.

    Lemma 4.2. Suppose that the network topology contains a spanning tree, and let θ={θi} be a solution to (1). Moreover, assume that the initial data and the quantity η satisfy

    D(θ(0))<ζ<γ<π,η>max{1sinγ,21ζγ}, (25)

    where ζ,γ are positive constants. And for a given sufficiently small D<min{π2,ζ}, assume the frustration α and coupling strength κ satisfy

    tanα<1(1+(d+1)ζζD(θ(0)))2Ncβd+1D[4(2N+1)c]d,D+α<π2,K>(1+(d+1)ζζD(θ(0)))(D(Ω)+2NKsinα)ccosα[4(2N+1)c]dβd+1D, (26)

    where d is the number of general nodes andc=(N1j=1ηjA(2N,j)+1)γsinγ.Then the following two assertions hold for the maximum node G0:

    1. The dynamics of Q0(t) is governed by the following equation

    ˙Q0(t)D(Ω)+2NKsinαKcosαcQ0(t),t[0,+),

    2. there exists time t00 such that

    D0(θ(t))βdD[4(2N+1)c]d,for t[t0,+),

    where t0 can be estimated as below and bounded by ˉt given in Lemma 2.8

    t0<ζKcosαcβd+1D[4(2N+1)c]d(D(Ω)+2NKsinα)<ˉt.

    Next, inspiring from Lemma 4.2, we make the following reasonable ansatz for Qk(t) with 0kd.

    Ansatz:

    1. The dynamics of quantity Qk(t) in time interval [0,) is governed by the following differential inequality,

    ˙Qk(t)D(Ω)+2NKsinα+(2N+1)KcosαDk1(θ(t))KcosαcQk(t), (27)

    where we assume D1(θ(t))=0.

    2. There exists a finite time tk0 such that, the phase diameter Dk(θ(t)) of ki=0Gi is uniformly bounded after tk, i.e.,

    Dk(θ(t))βdkD[4(2N+1)c]dk, t[tk,+), (28)

    where tk subjects to the following estimate,

    tk<(k+1)ζKcosαcβd+1D[4(2N+1)c]d(D(Ω)+2NKsinα)<ˉt=ζD(θ(0))D(Ω)+2NKsinα. (29)

    In the subsequence, we will split the proof of the ansatz into two lemmas by induction criteria. More precisely, based on the results in Lemma 4.2 as the initial step, we suppose the ansatz holds for Qk and Dk(θ) with 0kd1, and then prove that the ansatz also holds for Qk+1 and Dk+1(θ).

    Lemma 4.3. Suppose the assumptions in Lemma 4.2 are fulfilled, and the ansatz in (27), (28) and (29) holds for some k with 0kd1. Then the ansatz (27) holds for k+1.

    Proof. We will use proof by contradiction criteria to verify the ansatz for Qk+1. To this end, define a set

    Bk+1={T>0 : Dk+1(θ(t))<γ,  t[0,T)}.

    From Lemma 2.8, we see that

    Dk+1(θ(t))D(θ(t))<ζ<γ, t[0,ˉt).

    It is clear that ˉtBk+1. Thus the set Bk+1 is not empty. Define T=supBk+1. We will prove by contradiction that T=+. Suppose not, i.e., T<+. It is obvious that

    ˉtT,Dk+1(θ(t))<γ,  t[0,T),Dk+1(θ(T))=γ. (30)

    As the solution to system (1) is analytic, in the finite time interval [0,T), ˉθi and ˉθj either collide finite times or always stay together. Similar to the analysis in Lemma 3.3, without loss of generality, we only consider the situation that there is no pair of ˉθi and ˉθj staying together through all period [0,T). That means the order of {ˉθi}k+1i=0 will only exchange finite times in [0,T), so does {θ_i}k+1i=0. Thus, we divide the time interval [0,T) into a finite union as below

    [0,T)=rl=1Jl,Jl=[tl1,tl).

    such that in each interval Jl, the orders of both {ˉθi}k+1i=0 and {θ_i}k+1i=0 are preseved, and the order of oscillators in each subdigraph Gi with 0ik+1 does not change. In the following, we will show the contradiction via two steps.

    Step 1. In this step, we first verify the Ansatz (27) holds for Qk+1 on [0,T), i.e.,

    ˙Qk+1(t)D(Ω)+2NKsinα+(2N+1)KcosαDk(θ(t))KcosαcQk+1(t). (31)

    As the proof is slightly different from that in [45] and rather lengthy, we put the detailed proof in Appendix B.

    Step 2. In this step, we will study the upper bound of Qk+1 in (31) in time interval [tk,T), where tk is given in Ansatz (28) for Dk(θ). For the purposes of discussion, we rewrite the equation (31) in the interval [0,T) as below

    ˙Qk+1(t)Kcosαc(Qk+1(t)(2N+1)cDk(θ(t))(D(Ω)+2NKsinα)cKcosα), (32)

    where c is expressed by the following equation

    c=(N1j=1ηjA(2N,j)+1)γsinγ. (33)

    For the term Dk(θ) in (32), under the assumption of induction criteria, the Ansatz (28) holds for Dk(θ), i.e., there exists time tk such that

    Dk(θ(t))βdkD[4(2N+1)c]dk, t[tk,+),tk<ˉt. (34)

    And from the condition (26), it is obvious that

    K>(1+(d+1)ζζD(θ(0)))(D(Ω)+2NKsinα)ccosα[4(2N+1)c]dβd+1D>(D(Ω)+2NKsinα)ccosα[4(2N+1)c]dβd+1D.

    This directly yields that

    (D(Ω)+2NKsinα)cKcosα<βd+1D[4(2N+1)c]d<βdkD4dk[(2N+1)c]dk1, (35)

    where 0kd1,β<1,c>1. Then for the purposes of analysing the last two terms in the bracket of (32), we add the esimates in (34) and (35) to get

    (2N+1)cDk(θ(t))+(D(Ω)+2NKsinα)cKcosα(2N+1)cβdkD[4(2N+1)c]dk+βdkD4dk[(2N+1)c]dk1βdkD2[4(2N+1)c]dk1<βdkD[4(2N+1)c]dk1,t[tk,+). (36)

    From Lemma 2.8, we have tk<ˉtT, thus it makes sense when we consider the time interval [tk,T). Now based on the above estiamte (36), we apply the differential equation (32) and study the upper bound of Qk+1 on [tk,T). We claim that

    Qk+1(t)max{Qk+1(tk),βdkD[4(2N+1)c]dk1}:=Mk+1,t[tk,T). (37)

    Suppose not, then there exists some ˜t(tk,T) such that Qk+1(˜t)>Mk+1. We construct a set

    Ck+1:={tkt<˜t:Qk+1(t)Mk+1}.

    Since Qk+1(tk)Mk+1, the set Ck+1 is not empty. Define t=supCk+1. Then it is easy to see that

    t<˜t,Qk+1(t)=Mk+1,Qk+1(t)>Mk+1for t(t,˜t]. (38)

    From the construction of Mk+1, (36) and (38), it is clear that for t(t,˜t]

    Kcosαc(Qk+1(t)(2N+1)cDk(θ(t))(D(Ω)+2NKsinα)cKcosα)<Kcosαc(Mk+1βdkD[4(2N+1)c]dk1)0.

    Wen apply the above inequality and integrate on both sides of (32) from t to ˜t to get

    Qk+1(˜t)Mk+1=Qk+1(˜t)Qk+1(t)˜ttKcosαc(Qk+1(t)(2N+1)cDk(θ(t))(D(Ω)+2NKsinα)cKcosα)dt<0,

    which contradicts to the truth Qk+1(˜t)Mk+1>0. Thus we complete the proof of (37).

    Step 3. In this step, we will construct a contradiction to (30). From (37), Lemma 2.8 and the fact that

    βdkD[4(2N+1)c]dk1<D,tk<ˉt,Qk+1(tk)Dk+1(θ(tk))D(θ(tk))<ζ,

    we directly obtain

    Qk+1(t)max{Qk+1(tk),βdkD[4(2N+1)c]dk1}<max{ζ,D}=ζ,t[tk,T).

    From Lemma 4.1 and the condition (25), it yields that

    Dk+1(θ(t))Qk+1(t)β<ζβ<γ,t[tk,T).

    Since Dk+1(θ(t)) is continuous, we have

    Dk+1(θ(T))=limt(T)Dk+1(θ(t))ζβ<γ,

    which obviously contradicts to the assumption Dk+1(θ(T))=γ in (30).

    Thus, we combine all above analysis to conclude that T=+, that is to say,

    Dk+1(θ(t))<γ, t[0,+). (39)

    Then for any finite time T>0, we apply (39) and repeat the analysis in Step 1 to obtain that the differential inequality (27) holds for Qk+1 on [0,T). Thus we obtain the dynamics of Qk+1 in whole time interval [0,+) as below:

    ˙Qk+1(t)D(Ω)+2NKsinα+(2N+1)KcosαDk(θ(t))KcosαcQk+1(t). (40)

    Therefore, we complete the proof of the Ansatz (27) for Qk+1.

    Lemma 4.4. Suppose the conditions in Lemma 4.2 are fulfilled, and the ansatz in (27), (28) and (29) holds for some k with 0kd1. Then the ansatz (28) and (29) holds for k+1.

    Proof. From Lemma 4.3, we know the dynamic of Qk+1 is governed by (40). For the purposes of discussion, we rewrite the differential equation (40) as below and discuss it on [tk,+), i.e.,

    ˙Qk+1(t)Kcosαc(Qk+1(t)(2N+1)cDk(θ(t))(D(Ω)+2NKsinα)cKcosα), (41)

    where c is given in (33). In the subsequence, we will apply (41) to find a finite time tk+1 such that the quantity Qk+1 in (41) is uniformly bounded by some value less than π2 after tk+1. We split into two cases to discuss.

    Case 1. We first consider the case that Qk+1(tk)>βdkD[4(2N+1)c]dk1. In this case, When Qk+1(t)[βdkD[4(2N+1)c]dk1,Qk+1(tk)], we combine (36) and (41) to have

    ˙Qk+1(t)Kcosαc(βdkD[4(2N+1)c]dk1βdkD2[4(2N+1)c]dk1)=KcosαcβdkD2[4(2N+1)c]dk1<0. (42)

    That is to say, when Qk+1(t) is located in the interval [βdkD[4(2N+1)c]dk1,Qk+1(tk)], Qk+1(t) will keep decreasing with a rate bounded by a uniform slope. Therefore, we can define a stopping time tk+1 as follows,

    tk+1=inf{ttk | Qk+1(t)βdkD[4(2N+1)c]dk1}.

    Then, based on (42) and the definition of tk+1, we see that Qk+1 will decrease before tk+1 and has the following property at tk+1,

    Qk+1(tk+1)=βdkD[4(2N+1)c]dk1. (43)

    Moreover, from (42), it yields that the stopping time tk+1 satisfies the following upper bound estimate,

    tk+1Qk+1(tk)βdkD[4(2N+1)c]dk1KcosαcβdkD2[4(2N+1)c]dk1+tk. (44)

    Now we study the upper bound of Qk+1 on [tk+1,+). In fact, we can apply (42), (43) and the same arguments in (37) to derive

    Qk+1(t)βdkD[4(2N+1)c]dk1,t[tk+1,+). (45)

    On the other hand, in order to verify (29), we further study tk+1 in (44). For the first part on the right-hand side of (44), from Lemma 2.8 and the fact that

    Qk+1(tk)Dk+1(θ(tk))D(θ(tk))<ζ,βdkD2[4(2N+1)c]dk1>βd+1D[4(2N+1)c]d,

    we have the following estimates

    Qk+1(tk)βdkD[4(2N+1)c]dk1KcosαcβdkD2[4(2N+1)c]dk1<ζKcosαcβd+1D[4(2N+1)c]d(D(Ω)+2NKsinα), (46)

    where the denominator on the right-hand side of above inequality is positive from the conditions about K and α in (26). For the term tk in (44), based on the assumption (29) for tk, we have

    tk<(k+1)ζKcosαcβd+1D[4(2N+1)c]d(D(Ω)+2NKsinα)<ˉt=ζD(θ(0))D(Ω)+2NKsinα. (47)

    Thus it yields from (44), (46) and (47) that the time tk+1 satisfies

    tk+1<(k+2)ζKcosαcβd+1D[4(2N+1)c]d(D(Ω)+2NKsinα). (48)

    Moreover, from (26), it is easy to see that the coupling strength K satisfies the following inequality

    K>(1+(d+1)ζζD(θ(0)))(D(Ω)+2NKsinα)ccosα[4(2N+1)c]dβd+1D(1+(k+2)ζζD(θ(0)))(D(Ω)+2NKsinα)ccosα[4(2N+1)c]dβd+1D,0kd1. (49)

    Thus we combine (48) and (49) to verify the Ansatz (29) for k+1 in the first case, i.e., the time tk+1 subjects to the following estimate,

    tk+1<ˉt=ζD(θ(0))D(Ω)+2NKsinα. (50)

    Case 2. For another case that Qk+1(tk)βdkD[4(2N+1)c]dk1. Similar to the analysis in (37), we apply (42) to conclude that

    Qk+1(t)βdkD[4(2N+1)c]dk1,t[tk,+). (51)

    In this case, we directly set tk+1=tk. Then, from (47), it yields that the inequalities (48) and (50) also hold, which finish the verification of the Ansatz (29) in the second case.

    Finally, we are ready to verify the ansatz (28) for k+1. Actually, we can apply (45), (51) and Lemma 4.1 to have the upper bound of Dk+1(θ) on [tk+1,+) as below

    Dk+1(θ(t))Qk+1(t)ββdk1D[4(2N+1)c]dk1,t[tk+1,+). (52)

    Then we combine (48), (50) and (52) in Case 1 and similar analysis in Case 2 to conclude that the Ansatz (28) and (29) is true for Dk+1(θ).

    Now, we are ready to prove our main result.

    Lemma 4.5. Let θ=(θ1,θ2,,θN) be a solution to system (1), and suppose that the network contains a spanning tree and the assumptions in Lemma 4.2 are fulfilled. Then there exists a finite time t0 such that

    D(θ(t))D,for t[t,+),

    where t<ˉt and ˉt is given in Lemma 2.8.

    Proof. Combining Lemma 4.2, Lemma 4.3 and Lemma 4.4, we apply inductive criteria to conclude that the Ansatz (27) –(29) hold for all 0kd. Then, it yields from (28) and (29) that there exists a finite time 0td<ˉt such that

    D(θ(t))=Dd(θ(t))D,for t[td,+).

    Thus we set t=td and derive the desired result.

    Remark 5. For the Kuramoto model with frustration, in Lemma 4.5, we show the phase diameter of whole ensemble will be uniformly bounded by a value D less than π2 after some finite time. Under the assumption that α is sufficiently small such that D+α<π2, the interaction function cosx in the dynamics of frequency is positive after the finite time. Thus, we can lift system (1) to the second-order formulation, which enjoys the similar form to Cucker-Smale model with the interaction function cosx.

    More precisely, we can introduce phase velocity or frequency ωi(t):=˙θi(t) for each oscillator, and directly differentiate (1) with respect to time t to derive the equivalent second-order Cucker-Smale type model as below

    {˙θi(t)=ωi(t),t>0,i=1,2,,N,˙ωi(t)=KjNicos(θj(t)θi(t)+α)(ωj(t)ωi(t)),(θi(0),ωi(0))=(θi(0),˙θi(0)). (53)

    Now for the second-order system (53), we apply the results in [9] for Kuramoto model without frustration on a general digraph and present the frequency synchronization for Kuramoto model with frustrations.

    Lemma 4.6. Let (θ(t),ω(t)) be a solution to system (53), and suppose the network contains a spanning tree and the assumptions in Lemma 4.2 are fulfilled.Then there exist positive constants C1 and C2 such that

    D(ω(t))C1eC2(tt),tt,

    where D(ω(t))=max1iN{ωi(t)}min1iN{ωi(t)} is the diameter of frequency.

    Proof. According to Lemma 4.5 and the condition D+α<π2 in (26), we see that there exists time t0 such that

    D(θ(t))+αD+α<π2,t[t,+).

    Therefore, we can apply the methods and results in the work of Dong et al. [9] for Kuramoto model without frustration to yield the emergence of exponentially fast synchronization. As the proof is almost the same as that in [9], we omit its details.

    Proof of Theorem 1.1: Combining Lemma 4.5 and Lemma 4.6, we ultimately derive the desired result in Theorem 1.1.

    In this section, we present several numerical simulations to illustrate the main results in Theorem 1.1, which state that in sufficiently large coupling strength and small frustration regimes, synchronous behavior will emerge for the Kuramoto model with frustration in half circle case.

    For the simulation, we use the fourth-order Runge-Kutta method and employ the parameter N=10. The natural frequencies Ωi are randomly chosen from the interval (1,1), and the initial configuration θi0 are randomly chosen from the interval (0,5π6) which are confined in half circle. The interaction network we exploit are presented in Figure 1, which contains a spanning tree structure. And the corresponding (0,1)-adjacency matrix (χij) is given as below:

    (χij)=(1010000000110000000001100000000101001000000110000000001100000001011000001000010100000001100000000011).
    Figure 1. 

    The interaction network

    .

    For the fixed digraph in Figure 1, we choose large K=20 and small α=0.01 in Figure 2. Figure 2a shows that the phase diameter is uniformly bounded by a value less than π2 after some finite time, and we observe that the frequency diameter D(ω(t)) decays to zero exponentially fast in Figure 2b. This is consistent with the analytical results in Theorem 1.1.

    Figure 2. 

    Frequency synchronization with K=20,α=0.01

    .

    In Figure 3, we fix the coupling strength K=1, and respectively employ α=0,0.01,0.1,0.3 to observe the effects of frustration. Figure 3a3d shows the dynamics of frequencies and we observe that in small frustration regime (i.e., α=0,0.01), the asymptotic behavior evolves to frequency synchronization, while the state of desynchronization occurs for large frustration value (i.e., α=0.1,0.3). Figure 3e displays the dynamics of frequency diameter for two cases α=0.01,0.3.

    Figure 3. 

    Asymptotic behavior for K=1 when α=0,0.01,0.1,0.3

    .

    In Figure 4, we fix the frustration α=0.1 and respectively choose K=0.8,1,1.2,1.4 for simulation. Figure 4a4d presents that the state of desynchronization occurs when K is small (i.e., K=0.8,1), while frequency synchronization asymptotically emerges for large K (i.e., K=1.2,1.4). For comparison, the dynamics of frequency diameter when K=0.8,1.4 is displayed in Figure 4e.

    Figure 4. 

    Asymptotic behavior for α=0.1 when K=0.8,1,1.2,1.4

    .

    Moreover, for K=1, we observe in Figure 3a that frequency synchronization emerges asymptotically with zero frustration (i.e., α=0), whereas desynchronization occurs when α=0.1 in Figure 3c. This implies that the frustration hinders synchronization. And in Figure 4c4d, synchronization emerges when K=1.2,1.4 for α=0.1. This indicates that a larger coupling strength K is needed to guarantee the emergence of synchronization than that in zero frustration case.

    In this paper, under the effect of frustration, we provide sufficient frameworks leading to the complete synchronization for the Kuramoto model with general network containing a spanning tree. To this end, we follow a node decomposition introduced in [25] and construct hypo-coercive inequalities through which we can study the upper bounds of phase diameters. When the initial configuration is confined in a half circle, for sufficiently small frustration and sufficiently large coupling strength, we show that the relative differences of Kuramoto oscillators adding a phase shift will be confined into a region of quarter circle in finite time, thus we can directly apply the methods and results in [9] to prove that the complete frequency synchronization emerges exponentially fast. And we provide some numerical simulations to illustrate the main results.

    We really appreciate the editors and reviewers for their thorough reviews and insightful suggestions.

    We will split the proof into six steps. In the first step, we suppose by contrary that the phase diameter of G0 is bounded by γ in a finite time interval. In the second, third and forth steps, we use induction criteria to construct the differential inequality of Q0(t) in the finite time interval. In the last two steps, we exploit the derived differential inequality of Q0(t) to conclude that phase diameter of G0 is bounded by γ on [0,+), and thus the differential inequality of Q0(t) obtained in the forth step also holds on [0,+).

    Step 1. Define a set

    B0:={T>0: D0(θ(t))<γ,  t[0,T)}.

    From Lemma 2.8 where N=N0 in the present section, the set B0 is non-empty since

    D0(θ(t))=D(θ(t))<ζ<γ, t[0,ˉt),

    which directly yields that ˉtB0. Define T=supB0. And we claim that T=+. Suppose not, i.e., T<+, then we apply the continuity of D0(θ(t)) to have

    D0(θ(t))<γ, t[0,T),D0(θ(T))=γ. (54)

    In particular, we have ˉtT. The analyticity of the solution to system (1) is guaranteed by the standard Cauchy-Lipschitz theory. Therefore, in the finite time interval [0,T), any two oscillators either collide finite times or always stay together. If there are some θi and θj always staying together in [0,T], we can view them as one oscillator and thus the total number of oscillators that we need to study can be reduced. This is a more simpler situation, and we can similarly deal with it. Therefore, we only consider the case that there is no pair of oscillators staying together in [0,T). For this case, there are only finite many collisions occurring through [0,T). Thus, we divide the time interval [0,T) into a finite union as below

    [0,T)=rl=1Jl,Jl=[tl1,tl),

    where the end point tl denotes the collision instant. It is easy to see that there is no collision in the interior of Jl. Now we pick out any time interval Jl and assume that

    θ01(t)θ02(t)θ0N0(t),tJl.

    Step 2. According to the notations in (8), we follow the process A1 and A2 to construct ˉθ0n and θ_0n, 1nN0, respecively. We first study the dynamics of ˉθ0N0=θ0N0, i.e.,

    ˙θ0N0(t)=Ω0N0+KjN0N0(0)sin(θ0jθ0N0+α)ΩM+KjN0N0(0)[sin(θ0jθ0N0)cosα+cos(θ0jθ0N0)sinα]ΩM+N0Ksinα+KcosαminjN0N0(0)sin(θ0jθ0N0). (55)

    For the dynamics of ˉθ0N01, according to the process A1 and ˉa0N01=η(N0+2) in (7), we apply (55) and estimate the derivative of ˉθ0N01 as follows,

    ˙ˉθ0N01=ddt(ˉa0N01θ0N0+θ0N01ˉa0N01+1)=ˉa0N01ˉa0N01+1˙θ0N0+1ˉa0N01+1˙θ0N01ˉa0N01ˉa0N01+1(ΩM+N0Ksinα+KcosαminjN0N0(0)sin(θ0jθ0N0))+1ˉa0N01+1(Ω0N01+KjN0N01(0)sin(θ0jθ0N01+α))ΩM+Kcosα1ˉa0N01+1η(N0+2)minjN0N0(0)sin(θ0jθ0N0)+ˉa0N01ˉa0N01+1N0Ksinα+Kcosα1ˉa0N01+1jN0N01(0)sin(θ0jθ0N01)+Ksinα1ˉa0N01+1jN0N01(0)cos(θ0jθ0N01)ΩM+Kcosα1ˉa0N01+12ηminjN0N0(0)sin(θ0jθ0N0)+ˉa0N01ˉa0N01+1N0Ksinα+Kcosα1ˉa0N01+1(jN0N01(0)jN01sin(θ0jθ0N01)+sin(θ0N0θ0N01))+1ˉa0N01+1N0KsinαΩM+Kcosα1ˉa0N01+1ηminjN0N0(0)sin(θ0jθ0N0)+Kcosα1ˉa0N01+1minjN0N01(0)jN01sin(θ0jθ0N01)+Kcosα1ˉa0N01+1(ηminjN0N0(0)sin(θ0jθ0N0)+sin(θ0N0θ0N01))I1+N0Ksinα, (56)

    where we use

    |jN0N01(0)cos(θ0jθ0N01)|N0, Kcosα1ˉa0N01+1ηN0minjN0N0(0)sin(θ0jθ0N0)0,jN0N01(0)jN01sin(θ0jθ0N01)minjN0N01(0)jN01sin(θ0jθ0N01).

    Next we show the term I1 is non-positive. We only consider the situation γ>π2, and the case γπ2 can be similarly dealt with. It is clear that

    minjN0N0(0)sin(θ0jθ0N0)sin(θ0ˉkN0θ0N0)where ˉkN0=minjN0N0(0)j.

    Note that ˉkN0N01 since ˉLN0N0(ˉCN0,N0) is not a general root. Therefore, if 0θ0N0(t)θ0ˉkN0(t)π2, we immediately obtain that

    0θ0N0(t)θ0N01(t)θ0N0(t)θ0ˉkN0(t)π2,

    which implies that

    I1ηsin(θ0ˉkN0θ0N0)+sin(θ0N0θ0N01)sin(θ0ˉkN0θ0N0)+sin(θ0N0θ0N01)0.

    On the other hand, if π2<θ0N0(t)θ0ˉkN0(t)<γ, we use the fact

    η>1sinγandsin(θ0N0(t)θ0ˉkN0(t))>sinγ,

    to conclude that ηsin(θ0ˉkN0θ0N0)1. Hence, in this case, we still obtain that

    I1ηsin(θ0ˉkN0θ0N0)+sin(θ0N0θ0N01)1+10.

    Thus, for tJl, we combine above analysis to conclude that

    I1=ηminjN0N0(0)sin(θ0jθ0N0)+sin(θ0N0θ0N01)0. (57)

    Then combining (56) and (57), we derive that

    ˙ˉθ0N01ΩM+N0Ksinα+Kcosα1ˉa0N01+1(ηminjN0N0(0)sin(θ0jθ0N0)+minjN0N01(0)jN01sin(θ0jθ0N01)). (58)

     

    Step 3. Now we apply the induction principle to cope with ˉθ0n in (8), which is construced in the iteration process A1. We will prove for 1nN0 that,

    ˙ˉθ0n(t)ΩM+N0Ksinα+Kcosα1ˉa0n+1N0i=nηinminjN0i(0)jisin(θ0j(t)θ0i(t)). (59)

    In fact, it is known that (59) already holds for n=N0,N01 from (55) and (58). Then by induction criteria, suppose (59) holds for n. Next we verify that (59) still holds for n1. According to the Algorithm A1 and similar calculations in (56), the dynamics of the quantity ˉθ0n1(t) subjects to the following estimates

    ˙ˉθ0n1ΩM+N0Ksinα+Kcosα1ˉa0n1+1ηN0i=nηinminjN0i(0)jisin(θ0jθ0i)+Kcosα1ˉa0n1+1minjN0n1(0)jn1sin(θ0jθ0n1)+Kcosα1ˉa0n1+1×(η(N0n+1)N0i=nηinminjN0i(0)jisin(θ0jθ0i)+jN0n1(0)j>n1sin(θ0jθ0n1)I2). (60)

    Moreover, we can prove the term I2 is non-positive. As the proof is very similar as that in the previous step, we omit the details and directly claim that I20, which together with (60) verifies (59).

    Step 4. Now, we set n=1 in (59) and apply Lemma 3.1 to have

    ˙ˉθ01ΩM+N0Ksinα+Kcosα1ˉa01+1N0i=1ηi1minjN0i(0)jisin(θ0jθ0i)ΩM+N0Ksinα+Kcosα1ˉa01+1sin(θ0ˉk1θ0N0)=ΩM+N0Ksinα+Kcosα1ˉa01+1sin(θ01θ0N0), (61)

    where ˉk1=minjN0i=1N0i(0)j=1 due to the strong connectivity of G0. Similarly, we can follow the process A2 to construct θ_0k in (8) until k=N0. Then, we can apply the similar argument in (59) to obtain that,

    ddtθ_0N0(t)ΩmN0Ksinα+Kcosα1ˉa01+1sin(θ0N0θ01). (62)

    Then we recall the notations ˉθ0=ˉθ01 and θ_0=θ_0N0, and combine (61) and (62) to obtain that

    ˙Q0(t)=ddt(ˉθ0θ_0)D(Ω)+2N0KsinαKcosα2ˉa01+1sin(θ0N0θ01)D(Ω)+2N0KsinαKcosα1N01j=1ηjA(2N0,j)+1sin(θ0N0θ01),

    where we use the property

    ˉa01=N01j=1ηjA(2N0,j).

    As the function sinxx is monotonically decreasing in (0,π], we apply (54) to obtain that

    sin(θ0N0θ01)sinγγ(θ0N0θ01).

    Moreover, due to the fact Q0(t)θ0N0(t)θ01(t), we have

    ˙Q0(t)D(Ω)+2N0KsinαKcosα1N01j=1ηjA(2N0,j)+1sinγγ(θ0N0θ01)D(Ω)+2N0KsinαKcosα1N01j=1ηjA(2N0,j)+1sinγγQ0(t),tJl. (63)

    Note that the constructed quantity Q0(t)=ˉθ0(t)θ_0(t) is Lipschitz continuous on [0,T). Moreover, the above analysis does not depend on the time interval Jl, l=1,2,,r, thus the differential inequality (63) holds almost everywhere on [0,T).

    Step 5. Next we study the upper bound of Q0(t) in the period [0,T). Define

    M0=max{Q0(0),βD}.

    We claim that

    Q0(t)M0for all t[0,T). (64)

    Suppose not, then there exists some ˜t[0,T) such that Q0(˜t)>M0. We construct a set

    C0:={t<˜t | Q0(t)M0}.

    Since 0C0, the set C0 is not empty. Then we denote t=supC0. It is easy to see that

    t<˜t,Q0(t)=M0,Q0(t)>M0for t(t,˜t]. (65)

    For the given constant D<min{π2,ζ}, based on the assumptions about the frustration and the coupling strength in (9), it is clear that

    K>(1+ζζD(θ(0))(D(Ω)+2N0Ksinα)ccosα1βD>(D(Ω)+2N0Ksinα)ccosα1βD, (66)

    where c=(N01j=1ηjA(2N0,j)+1)γsinγ. Thus combing the construction of M0, (65) and (66), we obtain that for t(t,˜t], the following estimate holds,

    D(Ω)+2N0KsinαKcosα1N01j=1ηjA(2N0,j)+1sinγγQ0(t)<D(Ω)+2N0KsinαKcosα1N01j=1ηjA(2N0,j)+1sinγγβD<0.

    Then, we apply the above inequality and integrate on the both sides of (63) from t to ˜t to get

    Q0(˜t)M0=Q0(˜t)Q0(t)˜tt(D(Ω)+2N0KsinαKcosα1N01j=1ηjA(2N0,j)+1sinγγQ0(t))dt<0,

    which obviously contradicts to the fact Q0(˜t)M0>0, and verifies (64).

    Step 6. Now we are ready to show the contradiction to (54), which implies that T=+. In fact, from the fact that β<1,D<ζ and Q0(0)D0(θ(0))<ζ, we see

    Q0(t)M0=max{Q0(0),βD}<ζ,t[0,T).

    Then we apply the relation βD0(θ(t))Q0(t) given in Lemma 3.2 and the assumption η>21ζγ in (9) to obtain that

    D0(θ(t))Q0(t)β<ζβ<γ,t[0,T)where β=12η.

    As D0(θ(t)) is continuous, we have

    D0(θ(T))=limt(T)D0(θ(t))ζβ<γ,

    which contradicts to the situation that D0(θ(T))=γ in (54). Therefore, we conclude that T=+, which implies that

    D0(θ(t))<γ,for all t[0,+). (67)

    Then for any finite time T>0, we apply (67) and repeat the same argument in the second, third, forth steps to obtain the dynamics of Q0(t) in (63) holds on [0,T). Thus we obtain the following differential inequality of Q0 on the whole time interval:

    ˙Q0(t)D(Ω)+2N0KsinαKcosα1N01j=1ηjA(2N0,j)+1sinγγQ0(t), t[0,+).

    We will show the detailed proof of Step 1 in Lemma 4.3. Now we pick out any interval Jl with 1lr, where the orders of both {ˉθi}k+1i=0 and {θ_i}k+1i=0 are preseved and the order of oscillators in each subdigraph Gi with 0ik+1 will not change in each time interval. Then, we consider four cases depending on the possibility of relative position between ki=0G and Gk+1.

    Figure 5 shows the four possible relations between ki=0G and Gk+1 at any time t. Case 1 and Case 4 are similar and relative simple, while the analysis on Case 2 and Case 3 are similar but much more complicated. Therefore, we will only show the detailed proof of Case 2 for simplicity. In this case, we have from Figure 5 that

    max0ik+1{ˉθi}=ˉθk+1,min0ik+1{θ_i}=θ_k+1fortJl.
    Figure 5. 

    The four cases

    .

    Without loss of generality, we assume that

    θk+11θk+12θk+1Nk+1,for tJl.

    Step 1. Similar to (59), we claim that for 1nNk+1, the following inequalities hold

    ddtˉθk+1n(t)ΩM+Sk+1Ksinα+SkKcosαDk(θ(t))+Kcosα1ˉak+1n+1Nk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1j(t)θk+1i(t))), (68)

    where Sk=ki=0Ni. In the following, we will prove the claim (68) via induction principle.

    Step 1.1. As an initial step, we first verify that (68) holds for n=Nk+1. In fact, we have

    ddtˉθk+1Nk+1=Ωk+1Nk+1+KcosαjNk+1Nk+1(k+1)sin(θk+1jθk+1Nk+1)+Kcosαkl=0jNk+1Nk+1(l)sin(θljθk+1Nk+1)+Ksinαk+1l=0jNk+1Nk+1(l)cos(θljθk+1Nk+1)ΩM+Sk+1Ksinα+KcosαjNk+1Nk+1(k+1)sin(θk+1jθk+1Nk+1)I11+Kcosαkl=0jNk+1Nk+1(l)sin(θljθk+1Nk+1)I12, (69)

    where we use

    |k+1l=0jNk+1Nk+1(l)cos(θljθk+1Nk+1)|k+1l=0Nl=Sk+1.

    Estimates on I11 in (69). We know that θk+1Nk+1 is the largest phase among Gk+1, and all the oscillators in k+1i=0Gi are confined in half circle before T. Therefore, it is clear that

    sin(θk+1jθk+1Nk+1)0,for jNk+1Nk+1(k+1).

    Then we immediately have

    I11=jNk+1Nk+1(k+1)sin(θk+1jθk+1Nk+1)minjNk+1Nk+1(k+1)sin(θk+1jθk+1Nk+1). (70)

    Estimates on I12 in (69). For θlj which is the neighbor of θk+1Nk+1 in Gl with 0lk, i.e., jNk+1Nk+1(l), we consider two possible orderings between θlj and θk+1Nk+1:

    If θljθk+1Nk+1, we immediately have

    sin(θljθk+1Nk+1)0.

    If θlj>θk+1Nk+1, from the fact that

    θiNiˉθiθ_iθi1,0id,

    we immediately obtain

    θk+1Nk+1ˉθk+1=max0ik+1{ˉθi}max0ik{ˉθi}min0ik{θ_i}min0ikmin1jNi{θij}. (71)

    Thus we use the property of sinxx, x0 and (71) to get

    sin(θljθk+1Nk+1)θljθk+1Nk+1θljmin0ikmin1jNi{θij}Dk(θ(t)).

    Therefore, combining the above discussion, we have

    I12=kl=0jNk+1Nk+1(l)sin(θljθk+1Nk+1)SkDk(θ(t)). (72)

    From (69), (70) and (72), it yields that (68) holds for n=Nk+1.

    Step 1.2. Next, we assume that (68) holds for n with 2nNk+1, and we will show that (68) holds for n1. Following the process A1 and similar analysis in (69), we have

    ˙ˉθk+1n1ΩM+ˉak+1n1ˉak+1n1+1Sk+1Ksinα+ˉak+1n1ˉak+1n1+1SkKcosαDk(θ(t))+Ksinα1ˉak+1n1+1Sk+1+Kcosα1ˉak+1n1+1η(Nk+1n+2+Sk)Nk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1jθk+1i))+Kcosα1ˉak+1n1+1minjNk+1n1(k+1)jn1sin(θk+1jθk+1n1)+Kcosα1ˉak+1n1+1jNk+1n1(k+1)j>n1sin(θk+1jθk+1n1)I21+Kcosα1ˉak+1n1+1kl=0jNk+1n1(l)sin(θljθk+1n1)I22. (73)

    Next we do some estimates about the terms I21 and I22 in (73) seperately.

    Estimates on I21 in (73). Without loss of generality, we only deal with I21 under the situation γ>π2. We first apply the strong connectivity of Gk+1 and Lemma 3.1 to obtain that

    Nk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1j(t)θk+1i(t)))sin(θk+1ˉknθk+1Nk+1), (74)

    where ˉkn=minjNk+1i=nNk+1i(k+1)jn1. According to (74), we consider two cases depending on comparison between θk+1Nk+1θk+1ˉkn and π2.

    (i) For the first case that 0θk+1Nk+1θk+1ˉknπ2, we immediately obtain that for jNk+1n1(k+1), j>n1,

    0θk+1j(t)θk+1n1(t)θk+1Nk+1(t)θk+1n1(t)θk+1Nk+1(t)θk+1ˉkn(t)π2. (75)

    Then it yieldst from (74), (75) and η>2 that

    η(Nk+1n+1)Nk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1j(t)θk+1i(t)))+I21η(Nk+1n+1)sin(θk+1ˉknθk+1Nk+1)+jNk+1n1(k+1)j>n1sin(θk+1jθk+1n1)(Nk+1n+1)sin(θk+1ˉknθk+1Nk+1)+(Nk+1n+1)sin(θk+1Nk+1θk+1n1)0.

    (ii) For the second case that π2<θk+1Nk+1θk+1ˉkn<γ, it is known that

    η>1sinγandsin(θk+1Nk+1θk+1ˉkn)>sinγ, (76)

    which yields ηsin(θk+1ˉknθk+1Nk+1)1. Thus we immediately derive that

    η(Nk+1n+1)Nk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1j(t)θk+1i(t)))+I21η(Nk+1n+1)sin(θk+1ˉknθk+1Nk+1)+jNk+1n1(k+1)j>n1sin(θk+1jθk+1n1)(Nk+1n+1)+(Nk+1n+1)=0.

    Then, we combine the above arguments in (i) and (ii) to obtain

    η(Nk+1n+1)Nk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1jθk+1i))+I210. (77)

    Estimates on I22 in (73). For the term I22, there are three possible relations between θk+1n1 and θlj with 0lk:

    (i) If θljθk+1n1, we immediately have sin(θljθk+1n1)0.

    (ii) If θk+1n1<θljθk+1Nk+1, we consider two cases separately:

    (a) For the case that 0θk+1Nk+1θk+1ˉknπ2, it is clear that

    0θljθk+1n1θk+1Nk+1θk+1n1θk+1Nk+1θk+1ˉknπ2.

    Thus from the above inequality and (74), we have

    ηNk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1j(t)θk+1i(t)))+sin(θljθk+1n1)ηsin(θk+1ˉknθk+1Nk+1)+sin(θljθk+1n1)sin(θk+1ˉknθk+1Nk+1)+sin(θk+1Nk+1θk+1ˉkn)=0.

    (b) For another case that π2<θk+1Nk+1θk+1ˉkn<γ, it is yields from (76) that

    ηNk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1j(t)θk+1i(t)))+sin(θljθk+1n1)ηsin(θk+1ˉknθk+1Nk+1)+sin(θljθk+1n1)1+1=0.

    Hence, combining the above arguments in (a) and (b), we obtain that

    ηNk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1j(t)θk+1i(t)))+sin(θljθk+1n1)0.

     

    (iii) If θlj>θk+1Nk+1, we exploit the concave property of sine function in [0,π] to get

    sin(θljθk+1n1)sin(θljθk+1Nk+1)+sin(θk+1Nk+1θk+1n1). (78)

    For the second part on the right-hand side of above inequality (78), we apply the same analysis in (ii) to obtain

    ηNk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1jθk+1i))+sin(θk+1Nk+1θk+1n1)0.

    For the first part on the right-hand side of (78), the calculation is the same as (72), thus we have

    sin(θljθk+1Nk+1)θljθk+1Nk+1θljmin0ikmin1jNi{θij}Dk(θ(t)).

    Therefore, we combine the above estimates to obtain

    ηSkNk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1j(t)θk+1i(t)))+I22ηSksin(θk+1ˉknθk+1Nk+1)+kl=0jNk+1n1(l)sin(θljθk+1n1)SkDk(θ(t)). (79)

    Then from (73), (77), and (79), it yields that

    ddtˉθk+1n1ΩM+Sk+1Ksinα+ˉak+1n1ˉak+1n1+1SkKcosαDk(θ(t))++Kcosα1ˉak+1n1+1SkDk(θ(t))+Kcosα1ˉak+1n1+1ηNk+1i=n(ηinminjNk+1i(k+1)jisin(θk+1jθk+1i))+Kcosα1ˉak+1n1+1minjNk+1n1(k+1)jn1sin(θk+1jθk+1n1)ΩM+Sk+1Ksinα+SkKcosαDk(θ(t))+Kcosα1ˉak+1n1+1Nk+1i=n1(ηi(n1)minjNk+1i(k+1)jisin(θk+1jθk+1i)).

    This means that the claim (68) does hold for n1. Therefore, we apply the inductive criteria to complete the proof of the claim (68).

    Step 2. Now we are ready to prove (27) on Jl for Case 2. In fact, we apply Lemma 3.1 and the strong connectivity of Gk+1 to have

    Nk+1i=1(ηi1minjNk+1i(k+1)jisin(θk+1jθk+1i))sin(θk+11θk+1Nk+1).

    From the notations in (22) and (23), it is known that

    ˉθk+11=ˉθk+1,θ_k+1Nk+1=θ_k+1.

    Thus, we exploit the above inequality and set n=1 in (68) to obtain

    ddtˉθk+1=ddtˉθk+11ΩM+Sk+1Ksinα+SkKcosαDk(θ(t))+Kcosα1ˉak+11+1Nk+1i=1(ηi1minjNk+1i(k+1)jisin(θk+1jθk+1i))ΩM+Sk+1Ksinα+SkKcosαDk(θ(t))+Kcosα1ˉak+11+1sin(θk+11θk+1Nk+1). (80)

    We further apply the similar arguments in (80) to derive the differential inequality of θ_k+1 as below

    ddtθ_k+1ΩmSk+1KsinαSkKcosαDk(θ(t))+Kcosα1ˉak+11+1sin(θk+1Nk+1θk+11). (81)

    Due to the monotone decreasing property of sinxx in (0,π] and from (30), it is obvious that

    sin(θk+1Nk+1θk+11)sinγγ(θk+1Nk+1θk+11).

    Then we combine the above inequality, (80), (81) and (21) to get

    ˙Qk+1(t)=ddt(ˉθk+1θ_k+1)D(Ω)+2Sk+1Ksinα+2SkKcosαDk(θ(t))Kcosα2ˉak+11+1sin(θk+1Nk+1θk+11)D(Ω)+2Sk+1Ksinα+2SkKcosαDk(θ(t))Kcosα1ˉak+11+1sinγγ(θk+1Nk+1θk+11)D(Ω)+2NKsinα+(2N+1)KcosαDk(θ(t))Kcosα1N1j=1ηjA(2N,j)+1sinγγQk+1(t),tJl,

    where we use the fact that Qk+1(t)θk+1Nk+1(t)θk+11(t) and (21). Eventually, for Case 2, we obtain the dynamics for Qk+1(t) in (27) on Jl, i.e.,

    ˙Qk+1(t)D(Ω)+2NKsinα+(2N+1)KcosαDk(θ(t))KcosαcQk+1(t). (82)

    As our analysis does not depend on the choice of Jl, the differential inequality (82) holds on the interval [0,T).



    [1] A shocking display of synchrony. Phys. D (2000) 143: 21-55.
    [2] Biology of synchronous flashing of fireflies. Nature (1996) 211: 562-564.
    [3] Asymptotic formation and orbital stability of phase-locked states for the Kuramoto model. Phys. D (2012) 241: 735-754.
    [4] Synchronization of nonuniform Kuramoto oscillators for power grids with general connectivity and dampings. Nonlinearity (2019) 32: 559-583.
    [5] On exponential synchronization of Kuramoto oscillators. IEEE Trans. Automat. Control (2009) 54: 353-357.
    [6] Quasientrainment and slow relaxation in a population of oscillators with random and frustrated interactions. Phys. Rev. Lett. (1992) 68: 1073-1076.
    [7] Large-scale dynamics of the persistent turning walker model of fish behavior. J. Stat. Phys. (2008) 131: 989-1021.
    [8] Interplay of time-delay and velocity alignment in the Cucker-Smale model on a general digraph. Discrete Contin. Dyn. Syst.-Ser. B (2019) 24: 5569-5596.
    [9] Emergent Behavior of the Kuramoto model with a time-delay on a general digraph. SIAM J. Appl. Dyn. Syst. (2020) 19: 304-328.
    [10] Synchronization analysis of Kuramoto oscillators. Commun. Math. Sci. (2013) 11: 465-480.
    [11] On the critical coupling for Kuramoto oscillators. SIAM J. Appl. Dyn. Syst. (2011) 10: 1070-1099.
    [12] Synchronization and transient stability in power networks and nonuniform Kuramoto oscillators. SIAM J. Control Optim. (2012) 50: 1616-1642.
    [13] Synchronization in complex oscillator networks and smart grids. Proc. Natl. Acad. Sci. (2013) 110: 2005-2010.
    [14] On the complete synchronization of the Kuramoto phase model. Phys. D (2010) 239: 1692-1700.
    [15]

    S.-Y. Ha, D. Kim, J. Kim and X. Zhang, Asymptotic behavior of discrete Kuramoto model and uniform-in-time transition from discrete to continuous dynamics, J. Math. Phys., 60 (2019), 051508, 21 pp.

    [16]

    S.-Y. Ha, D. Kim, J. Lee and S. E. Noh, Synchronization conditions of a mixed Kuramoto ensemble in attractive and repulsive couplings, J. Nonlinear Sci., 31 (2021), Paper No. 39, 34 pp.

    [17] Remarks on the complete synchronization of Kuramoto oscillators. Nonlinearity (2015) 28: 1441-1462.
    [18] Remarks on the complete synchronization for the Kuramoto model with frustrations. Anal. Appl. (2018) 16: 525-563.
    [19] Emergence of phase-locked states for the Kuramoto model in a large coupling regime. Commun. Math. Sci. (2016) 14: 1073-1091.
    [20] Asymptotic synchronous behavior of Kuramoto type models with frustrations. Netw. Heterog. Media (2014) 9: 33-64.
    [21] Large-time dynamics of Kuramoto oscillators under the effects of inertia and frustration. SIAM J. Appl. Dyn. Syst. (2014) 13: 466-492.
    [22] Emergence of phase-locking in the Kuramoto model for identical oscillators with frustration. SIAM J. Appl. Dyn. Syst. (2018) 17: 581-625.
    [23] Complete synchronization of Kuramoto oscillators with hierarchical leadership. Commun. Math. Sci. (2014) 12: 485-508.
    [24] Formation of phase-locked states in a population of locally interacting Kuramoto oscillators. J. Differ. Equ. (2013) 255: 3053-3070.
    [25] On the critical exponent of the one-dimensional Cucker Smale model on a general graph. Math. Models Meth. Appl. Sci. (2020) 30: 1653-1703.
    [26] Asymptotic phase-Locking dynamics and critical coupling strength for the Kuramoto model. Commun. Math. Phys. (2020) 377: 811-857.
    [27]

    Y. Kuramoto, Self-entrainment of a population of coupled non-linear oscillators, In International Symposium on Mathematical Problems in Theoretical Physics, (ed. H. Araki), Springer Berlin Heidelberg, (1975), 420–411.

    [28]

    Y. Kuramoto, Chemical Oscillations, Waves and Turbulence, Springer-Verlag, Berlin, 1984.

    [29] Collective motion, sensor networks, and ocean sampling. Proc. IEEE (2007) 95: 48-74.
    [30] Emergent multistability and frustration in phase-repulsive networks of oscillators. Phys. Rev. E (2011) 84: 016231.
    [31] Uniqueness and well-ordering of emergent phase-locked states for the Kuramoto model with frustration and inertia. Math. Models Methods Appl. Sci. (2016) 26: 357-382.
    [32] The spectrum of the partially locked state for the Kuramoto model. J. Nonlinear Sci. (2007) 17: 309-347.
    [33] Modular synchronization in complex networks with a gauge Kuramoto model. EPL (2008) 83: 68003.
    [34] Oscillator models and collective motion: Spatial patterns in the dynamics of engineered and biological networks. IEEE Control Sys. (2007) 27: 89-105.
    [35] Glass synchronization in the network of oscillators with random phase shift. Phys. Rev. E (1998) 57: 5030-5035.
    [36] Extension of the Cucker-Smale control law to space flight formations. J. Guid. Control Dynam. (2009) 32: 527-537.
    [37] (2001) Synchronization: A Universal Concept in Nonlinear Sciences.Cambridge University Press.
    [38] A soluble active rotator model showing phase transitions via mutual entrainment. Prog. Theor. Phys. (1986) 76: 576-581.
    [39] From Kuramoto to Crawford: Exploring the onset of synchronization in populations of coupled oscillators. Phys. D (2000) 143: 1-20.
    [40] Coupled oscillators and biological synchronization. Sci. Amer. (1993) 269: 101-109.
    [41] Dynamics in co-evolving networks of active elements. Forma (2009) 24: 17-22.
    [42] Flocks, herds, and schools: A quantitative theory of flocking. Phys. Rev. E (1998) 58: 4828-4858.
    [43] Biological rhythms and behavior of populations of coupled oscillators. J. Theor. Biol. (1967) 16: 15-42.
    [44] Frustration effect on synchronization and chaos in coupled oscillators. Chin. Phys. Soc. (2011) 10: 703-707.
    [45]

    X. Zhang and T. Zhu, Emergence of synchronization in Kuramoto model with general digraph, preprint, arXiv: 2107.06487.

  • This article has been cited by:

    1. Conor Carty, Young-Pil Choi, Chiara Cicolani, Cristina Pignotti, Asymptotic synchronization of Kuramoto oscillators with time delay and non-universal interaction, 2025, 76, 0044-2275, 10.1007/s00033-024-02406-5
  • 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(1082) PDF downloads(170) Cited by(1)

Figures and Tables

Figures(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog