Loading [MathJax]/jax/element/mml/optable/MathOperators.js
A journal of IEEE and CAA , publishes high-quality papers in English on original theoretical/experimental research and development in all areas of automation
C. J. Li and X. F. Zong, “Group hybrid coordination control of multi-agent systems with time-delays and additive noises,” IEEE/CAA J. Autom. Sinica, vol. 10, no. 3, pp. 737–748, Mar. 2023. doi: 10.1109/JAS.2022.105917
Citation: C. J. Li and X. F. Zong, “Group hybrid coordination control of multi-agent systems with time-delays and additive noises,” IEEE/CAA J. Autom. Sinica, vol. 10, no. 3, pp. 737–748, Mar. 2023. doi: 10.1109/JAS.2022.105917

Group Hybrid Coordination Control of Multi-Agent Systems With Time-Delays and Additive Noises

doi: 10.1109/JAS.2022.105917
Funds:  This work was supported by the National Natural Science Foundation of China (62073305), and the Fundamental Research Funds for the Central Universities, China University of Geosciences (Wuhan) (CUG170610)
More Information
  • A new kind of group coordination control problem-group hybrid coordination control is investigated in this paper. The group hybrid coordination control means that in a whole multi-agent system (MAS) that consists of two subgroups with communications between them, agents in the two subgroups achieve consensus and containment, respectively. For MASs with both time-delays and additive noises, two group control protocols are proposed to solve this problem for the containment-oriented case and consensus-oriented case, respectively. By developing a new analysis idea, some sufficient conditions and necessary conditions related to the communication intensity between the two subgroups are obtained for the following two types of group hybrid coordination behavior: 1) Agents in one subgroup and in another subgroup achieve weak consensus and containment, respectively; 2) Agents in one subgroup and in another subgroup achieve strong consensus and containment, respectively. It is revealed that the decay of the communication impact between the two subgroups is necessary for the consensus-oriented case. Finally, the validity of the group control results is verified by several simulation examples.

     

  • COORDINATION control of multi-agent systems (MASs) has become an important theme in the control community, including consensus control, containment control and so on. For these problems, fruitful achievements have been obtained. Under time-delays, many consensus results have been obtained [1]–[4]. In general, time-delays have negative impact on consensus, but Yu et al. [4] revealed that appropriate time-delays can play a positive role in delay-induced consensus problems. The effects of additive and multiplicative measurement noises on consensus behaviors were revealed in [5]–[8], [9] and [10], respectively. The consensus problems were studied for MASs with both time-delays and additive noises in [11]–[14].

    As a kind of coordination control problem, the containment problem has also received extensive attention. Under a relatively ideal environment, many interesting results about containment problems have been obtained, as shown in [15]–[18]. When time-delays exist in MASs, the research results for the containment problem can be seen in [19]–[21]. The effects of additive noises on containment behaviors were revealed in [22] and the containment control of MASs with both time-delays and additive noises was studied in [23].

    It can be seen that the above studies all focused on the problem where agents in a whole system achieve global coordination behavior. However, in some practical application scenarios where multiple tasks such as tracking and striking need to be performed simultaneously, global coordination control may not be able to accomplish these tasks. Hence, it is necessary to investigate group coordination control problems such as group consensus and group containment problems. These require agents in different subgroups to achieve different coordination states. In fact, many scholars have made efforts for solving group coordination control problems. Yu and Wang [24] and [25] studied group consensus under the assumption that the effect between the two subgroups is balanced. Then, the assumption was relaxed in [26] and [27], where the group consensus of MASs with generally connected topology were investigated. Some interesting results about group consensus under measurement noises can be seen in [28] and [29]. The results about the bipartite consensus problem can be found in [30]–[32].

    Note that the aforementioned works about group coordination control all focused on the problem that all subgroups achieve consensus behavior. However, agents in different subgroups may be required to achieve different coordination behaviors for some multi-objective tasks such as military tasks [33]. Moreover, for opinion dynamics in social networks, there may appear agreement or fluctuation which need to be characterized by different coordination behaviors in a whole organization [34]. Hence, it is natural to think about the conditions that consensus can be achieved for one subgroup and containment be achieved for another subgroup in a whole MAS. This problem is called group hybrid coordination control problem, which has not been addressed, even in an ideal environment.

    In this paper, we study the group hybrid coordination control problem of MASs with both time-delays and additive noises. We consider a whole MAS that consists of two subgroups, where communications between the two subgroups are unidirectional and only one subgroup can receive information from another subgroup. The case where a subgroup receiving information from another subgroup is required to achieve containment is called the containment-oriented case. Similarly, the case where the subgroup receiving information from another subgroup is required to achieve consensus is called the consensus-oriented case. Since there are communications between the two subgroups and different subgroups are required to achieve different coordination behaviors for a whole MAS, the system dynamics cannot simply be analyzed as a whole. This makes the existing control protocols, methods and conditions, such as those in [14], [22], and [24]–[32], fail to solve the group hybrid coordination control problem. Then, in this paper, our goal is to explore appropriate analysis methods and obtain some conditions for solving the new kind of group coordination control problem, i.e., the group hybrid coordination control problem of MASs with time-delays and additive noises.

    The main contribution of this paper can be concluded as follows.

    1) A new kind of group coordination control problem-group hybrid coordination control is investigated and two new group control protocols are designed to solve this problem. By developing a new analysis idea, some sufficient conditions and necessary conditions are obtained for the group hybrid coordination control problem under time-delays and additive noises. It is found that the conditions for the containment-oriented case are weaker than those for the consensus-oriented case. And for the consensus-oriented case, the conditions for weak consensus are weaker than those for strong consensus.

    2) The influence mechanism of the communication impact between the two subgroups on group hybrid coordination control problem of MASs with both time-delays and additive noises is revealed: the decay of the communication impact between the two subgroups is necessary for the consensus-oriented case. This provides hints in the design of control gain functions for the similar group coordination control problem.

    The structure of this paper is organized like this: In Section II, the system model, some basic concepts, and lemmas are introduced. Sections III and IV investigate the containment-oriented case and the consensus-oriented case, respectively. Some simulation results are given in Section V. Section VI summarizes this work and gives some other interesting questions for future research.

    Notations: The following notations are used in the process of analysis. 1n and ϑn,i are two n-dimensional column vectors, where each element is one for 1n, and the ith element is one and others are zero for ϑn,i. The m-dimensional identity matrix is denoted by Im. For a,bR, ab= min{a,b}. For any complex number λ in complex space C, its real part is denoted by Re(λ). For the matrix or vector P, its transpose and Euclidean norm are denoted by PT and P, respectively. For a given random variable or vector X, its mathematical expectation is denoted by EX.

    Let G=(V,E,A) be a weighted directed graph, where V={1,,i,,n} is the set of nodes with i representing the ith agent, EV×V denotes the set of edges and (i,j)E is a edge of G. A=[aij]Rn×n is the weighted adjacency matrix with real adjacency weighting elements aij=1 or 0 indicating whether or not there is an information flow from agent j to agent i directly. Moreover, we assume aii=0 for all iV. The set of neighbors of node i is denoted by Ni, that is, aij=1 for jNi. The Laplacian matrix of G is denoted by L. A directed path is a sequence of edges with the form (i1,i),(i,i+1),.... A directed graph contains a directed spanning tree if there exists at least one node that has a directed path to all other nodes. In the directed graph G, the node which has no parent node is called a root. A digraph G has a spanning forest if there exists a directed forest which is the spanning subgraph of G. More details about the spanning subgraph, directed forest and spanning forest can be found in [22].

    In this paper, n+m(n1,m>2) first-order continuous-time agents compose a MAS with a directed communication graph G=(V,E,A). We divide G=(V,E,A) into two subgraphs G1=(V1,E1,A1) and G2=(V2,E2,A2), where V1={1,,n}, V2={n+1,,n+m} and E1=V1×V1, E2=V2×V2, that is, the first n agents belong to one subgroup, which is denoted by Ψ1, and the last m agents belong to another subgroup, which is denoted by Ψ2. Moreover, let Ni=N1iN2i, where N1iV1 and N2iV2 are the sets of neighbors belonging to Ψ1 and Ψ2 of the ith agent respectively. Here, each agent has the following dynamics:

    ˙xi(t)=ui(t),i=1,,n+m (1)

    where xi(t)R and ui(t)R denote the state and control input of agent i, respectively. Denote X1(t)=[x1(t),,xn(t)]T, X2(t)=[xn+1(t),,xn+m(t)]T.

    In this paper, we require the agents in subgroup Ψ1 and subgroup Ψ2 to achieve consensus and containment, respectively. Then, we know that there should be several leaders in subgroup Ψ2. We assume that agents n+1,,n+a(1<a<m), who have no neighbor, are stationary leaders and the other agents n+a+1,,n+m are followers. Denote X21(t)= [xn+1(t),,xn+a(t)]T, X22(t)=[xn+a+1(t),,xn+m(t)]T. Then, for the communication topology subgraphs G1, G2 and the communications between the two subgroups, we give the following assumption.

    Assumption 1: G1 contains a spanning tree and G2 has a spanning forest whose roots are exactly the leaders in MASs. Moreover, the communications between the two subgroups are unidirectional and only one subgroup can receive information from another subgroup.

    In the communication environment considered in this paper, the exchange of state information between agents can not be performed accurately. It is often assumed that the agent i can receive information from its neighbors as follows:

    yji(t)=xj(tτ)+σjiηji(t),jNi,t>0

    where yji(t) denotes the measured value of the jth agent’s state by the ith agent, τ0 is the time-delay, ηji(t)R denotes the measurement noise and σji is the corresponding intensity function. Here, let the initial states be constants, X1(t)=Δ1Rn, X2(t)=Δ2Rm, X21(t)=Δ21Ra and X22(t)=Δ22Rma, t[τ,0].

    In this paper, independent Gaussian white noises are used to model the measurement noises.

    Assumption 2: The stochastic process ηji(t) satisfies t0ηji(s)ds=ωji(t),t0,i,j=1,,n+m, where {ωji(t),i,j=1,,n+m} are independent Brownian motions.

    In this paper, the goal of group hybrid coordination control is to achieve consensus for subgroup Ψ1 and containment for subgroup Ψ2. Below, we give the definitions of consensus and containment, respectively.

    Definition 1 [14]: For all distinct agents i,j in a subgroup Ψ, if limtE|xi(t)xj(t)|2=0 (or limt|xi(t)xj(t)|=0 almost surely (a.s.)), then the agents in this subgroup are said to achieve mean square weak consensus (or almost sure weak consensus).

    Definition 2 [14]: For all agents i in a subgroup Ψ, if there is a random variable ˉx such that E|ˉx|< (or P{|ˉx|<}=1) and limtE|xi(t)ˉx|2=0 (or limt|xi(t)ˉx|=0, a.s.), then the agents in this subgroup are said to achieve mean square strong consensus (or almost sure strong consensus).

    Definition 3 [22]: For all followers i in a subgroup Ψ, if there exist deterministic variables ˉxicoL \{x_{(n+1)f},\ldots, x_{(n+a)f}\}= \{x_f|x_f = \sum_{q=1}^{a} k_qx_{(n+q)f}, k_q\ge 0,\sum_{q=1}^{a} k_q = 1\} ( co_L denotes the convex hull spanned by the leaders’ final positions x_{jf}=\lim_{t\to \infty}x_j(t), \;j=n+1,\ldots,n+a) such that \mathbb E|\bar x_i^*| < \infty (or \mathbb P\{|\bar x_i^*| < \infty\} = 1 ) and \lim_{t\to \infty}\mathbb E|x_i(t)-\bar x_i^*|^2 = 0 (or \lim_{t\to \infty}|x_i(t)- \bar x_i^*|= 0, a.s.), i=n+a+1,\ldots,n+m , then the agents in this subgroup are said to achieve mean square containment (or almost sure containment).

    Remark 1: In general, under a stochastic background system, the coordination control problems of MASs are studied in the mean square and almost sure senses. Compared with mean square coordination behaviors, almost sure coordination behaviors are more intuitive and acceptable. This paper will study group hybrid coordination control problem in both senses.

    Based on above definitions of consensus and containment, we make the following statement: the agents in a subgroup Ψ are said to achieve stochastic consensus (or stochastic containment) if the agents in this subgroup achieve consensus (or containment) in both the mean square and almost sure senses. Then, in this paper, the following two types of group hybrid coordination behavior will be investigated for both the containment-oriented case and consensus-oriented case: 1) Agents in subgroup \Psi_1 achieve stochastic weak consensus and agents in subgroup \Psi_2 achieve stochastic containment; 2) Agents in subgroup \Psi_1 achieve stochastic strong consensus and agents in subgroup \Psi_2 achieve stochastic containment. As mentioned above, we will study group hybrid coordination control in both the above two senses. These studies will enrich the group coordination control problem.

    To study group hybrid coordination control under time-delays and additive noises, we need two important lemmas as follows.

    Lemma 1 [14]: Suppose that the directed graph contains a spanning tree. For the Laplacian matrix {\cal{ L}}' , we have the following statements:

    1) There exists a probability measure π such that \pi^T {\cal{ L}}' = 0 .

    2) There exists a nonsingular matrix Q = (({1}/{\sqrt n})\mathbf 1_n,\widetilde{Q}) and

    \begin{equation*} Q^{-1} = \begin{pmatrix} \upsilon^T \\ \overline{Q} \end{pmatrix},\;\; Q^{-1}{\cal{ L}}'Q = \begin{pmatrix} 0 & 0 \\ 0 & \widetilde{{\cal{ L}}}' \end{pmatrix} \end{equation*}

    where n is the number of nodes, \overline{Q}\; \in \;\mathbb{R}^{(n-1)\times n} , \widetilde {{\cal{ L}}}' \in \mathbb{R}^{(n-1)\times (n-1)} . For \upsilon , one has \upsilon^T{\cal{ L}}'=0 and ({1}/{\sqrt n})\upsilon^T\mathbf 1_n=1.

    3) The directed graph contains a spanning tree if and only if each eigenvalue of \widetilde{{\cal{ L}}}' has positive real part. Moreover, if the directed graph contains a spanning tree, then the probability measure π is unique and \upsilon=\sqrt n\pi .

    Lemma 2 [14]: For the linear scalar equation

    \dot {\bar X}(t)=-\lambda c(t)\bar X(t-\tau),\;\; t>0

    \bar X(t)=\xi(t) for t\in[-\tau,0] , where Re(\lambda)>0, \tau \ge 0 and \xi \in C([-\tau, 0],\mathbb{C}), C([-\tau,0],\mathbb{C}) is the space of all continuous \mathbb{C} -valued functions defined on [-\tau,0] . The solution has the form \bar X(t)=\Gamma(t,s) \bar X(s),\;\; \forall t\ge s\ge 0, where \Gamma(t,s) is the differential resolvent function, satisfying \Gamma(t,t)=1 for t>0 , \Gamma(t,s)=0 for t<s and

    \frac{\partial}{\partial t} \Gamma(t,s)=-\lambda c(t)\Gamma(t-\tau,s),\;\; t>s.

    If there is a constant t_0\ge 0 such that \tau \bar c_{t_0}({\vert \lambda\vert^2}/{Re(\lambda)} )< 1, then

    \vert\Gamma(t,s)\vert^2\le b(\lambda)e^{-\varrho (\lambda)\int_{s}^{t}c(u)du},\;\; t>s\ge t_0

    here, \bar c_{t_0}:={\rm{sup}}_{t\ge t_0}c(t), t_0\ge 0. b(\lambda) is a positive constant depending on λ and \varrho(\lambda):=\rho_1(\lambda)\wedge \rho_2(\lambda) , where \rho_1(\lambda) is the unique root of the equation 3\rho \vert\lambda\vert^2 \tau^2 \bar c_{t_0}^2 e^{\rho \bar c_{t_0}\tau}+2\rho- 2(Re(\lambda)- \vert\lambda\vert^2\tau\bar c_{t_0})=0 and \rho_2(\lambda)=({1}/{\bar c_{t_0}\tau}){\rm {log}}({1}/{\vert\lambda\vert \bar c_{t_0}\tau}).

    Remark 2: Lemmas 1 and 2 are powerful for analyzing the group hybrid coordination control problem of MASs with time-delays and additive noises. In fact, after fully mining the properties of \mathcal{L}' , Lemma 1 is naturally obtained. Moreover, with the help of Lemma 2, we can not only analyze the stability of stochastic systems with time-delays, but also reveal the decay rate.

    In this paper, according to the topology structure of {\cal{ G}} , the Laplacian matrix {\cal{ L}} can be written as

    {\cal{ L}} = \begin{pmatrix} {\cal{ L}}_{11} & {\cal{ L}}_{12} \\ {\cal{ L}}_{21} & {\cal{ L}}_{22} \end{pmatrix}

    where {\cal{ L}}=\left[l_{ij}\right] is defined as

    l_{ij}= \begin{cases} -a_{ij}, & j\neq i \\ \displaystyle\sum\limits_{k=1,k\neq i}^{n+m} a_{ik}, & j=i \end{cases}

    and {\cal{ L}}_{11}\in \mathbb R^{n\times n} is the Laplacian matrix of {\cal{ G}}_1 , {\cal{ L}}_{22}\in \mathbb R^{m\times m}, \;{\cal{ L}}_{12}\in \mathbb R^{n\times m} and {\cal{ L}}_{21}\in \mathbb R^{m\times n} . Here, we define

    {\cal{ L}}_{21}=\begin{pmatrix} 0 \\ {\cal{{\widehat{L}}}}_{21} \end{pmatrix}, \;\;{\cal{ L}}_{22} = \begin{pmatrix} 0 & 0 \\ {\cal{\widetilde{L}}}_{22} & {\cal{{\widehat{L}}}}_{22} \end{pmatrix}

    where {\cal{{\widehat{L}}}}_{21}\in \mathbb R^{(m-a)\times n},\; {\cal{\widetilde{L}}}_{22}\in \mathbb R^{(m-a)\times a} and {\cal{{\widehat{L}}}}_{22}\in \mathbb R^{(m-a)\times (m-a)} .

    Then, in next two sections, we will study the group hybrid coordination control problem in detail. By examining the containment-oriented case and consensus-oriented case respectively, we will obtain some useful results and find some interesting phenomena for group hybrid coordination behaviors under time-delays and additive noises.

    For the containment-oriented case, we give the following assumption for the communications between the two subgroups.

    Assumption 3: The followers in subgroup \Psi_2 can measure the state information from the agents in subgroup \Psi_1 .

    Then, under this communication model, we assume that each agent can receive its neighbor information in the following form:

    \begin{cases} \bar y_{ji}(t)=x_j(t-\tau)+\sigma_{ji}\eta_{ji}(t),\;\; i\in {\cal{ V}}_1,\;\;j\in N_i \\ \bar{\bar y}_{ji}(t)=\bar{\bar y}_{ji1}(t)+\bar{\bar y}_{ji2}(t),\;\; i\in {\cal{ V}}_2,\;\;j\in N_i \end{cases}

    where \bar y_{ji}(t) and \bar{\bar y}_{ji}(t) denote the state measurement of agent j by the ith agent in subgroup \Psi_1 and \Psi_2 , respectively. \bar{\bar y}_{ji1}(t)=x_j(t-\tau)+\sigma_{ji}\eta_{ji}(t), \;i\in {\cal{ V}}_2,\;j\in N_{2i}, is the state measurement of agent j belonging to subgroup \Psi_2 by agent i, and \bar{\bar y}_{ji2}(t)=\kappa(t)(x_j(t-\tau)+\sigma_{ji}\eta_{ji}(t)), \;i\in {\cal{ V}}_2,\;j\in N_{1i}, is the state measurement of agent j belonging to subgroup \Psi_1 by agent i. \kappa(t) is a time-varying function, which reflects the communication intensity between subgroup \Psi_1 and subgroup \Psi_2 .

    Remark 3: Note that there are communications between the two subgroups. This means that the MAS is still a whole system. Under the assumption that the in-degrees from the other subgroups are equal at any time for each node in one subgroup, Shang [28] concluded that the group consensus behavior can be reached under appropriate time-delay and multiplicative noise intensity. Under time-delays and additive noises, we determine to study the group hybrid coordination control problem when the assumption mentioned above is released. This is a key problem that will be explored in this paper.

    Based on interactions between any two agents in the environment with both time-delays and additive noises, we propose the control protocol as follows:

    u_i(t) = \begin{cases} c_1(t)\displaystyle\sum\limits_{j=1}^{n} a_{ij}(\bar y_{ji}(t) - x_i(t-\tau)), \;\;\forall i\in {\cal{ V}}_1 \\ c_2(t)\displaystyle\sum\limits_{j=n+1}^{n+m} a_{ij}(\bar{\bar y}_{ji1}(t) - x_i(t-\tau)) \\ \qquad + c_2(t)\displaystyle\sum\limits_{j=1}^{n} a_{ij}\bar{\bar y}_{ji2}(t),\;\; \forall i\in {\cal{ V}}_2 \end{cases} (2)

    where t>0 , c_1(t) , c_2(t)\in C((0,\infty);[0,\infty)) are two control gain functions to be designed.

    Remark 4: Obviously, the communications between agents in the same subgroup and in different subgroups are considered in this protocol. In fact, the group control protocol (2), which considers time-delay, noise and the communication intensity \kappa(t) , is the generalization of the protocol in [24]. Also, the control protocol (2) is different from protocols in [14], [22], [23], and [28]. In fact, these protocols fail to solve the group hybrid coordination control problem. Moreover, we can see that there may be some coupling relationships between the time-varying function \kappa(t) and control gain functions c_1(t), c_2(t) from protocol (2). This may make it more difficult to analyze the group hybrid coordination control problem.

    Under protocol (2), from (1) we can obtain the following equations:

    \begin{equation} \dot X_1(t) = -c_1(t){\cal{ L}}_{11}X_1(t-\tau) + c_1(t)\sum\limits_{i,j=1}^{n} a_{ij}\sigma_{ji}\vartheta_{n,i}\eta_{ji}(t) \end{equation} (3)

    and

    \begin{split} \dot X_2(t) =\; & -c_2(t)\overline {{\cal{ L}}}_{22}X_2(t-\tau) -c_2(t)\kappa(t){\cal{ L}}_{21}X_1(t-\tau) \\ & +c_2(t)\sum\limits_{i,j=n+1}^{n+m} a_{ij}\sigma_{ji}\vartheta_{m,i-n}\eta_{ji}(t) \\ & +c_2(t)\kappa(t)\sum\limits_{i=n+1}^{n+m}\sum\limits_{j=1}^{n} a_{ij}\sigma_{ji}\vartheta_{m,i-n}\eta_{ji}(t) \end{split} (4)

    where \overline {{\cal{ L}}}_{22}=\left[\bar l_{ij}\right] is the Laplacian matrix of {\cal{ G}}_2 , which is defined as

    \bar l_{ij}= \begin{cases} -a_{ij}, & j\neq i,\;\; i\in {\cal{ V}}_2,\;\;j\in N_{2i} \\ \displaystyle\sum\limits_{k=n+1,k\neq i}^{n+m} a_{ik}, & j=i, \;\;i\in {\cal{ V}}_2,\;\;j\in N_{2i}. \end{cases}

    Moreover, we rewrite \overline {{\cal{ L}}}_{22} as

    \overline{{\cal{ L}}}_{22} = \begin{pmatrix} 0 & 0 \\ {\cal{\widetilde{L}}}_{22} & \widehat{{\cal{L}}}_{22}' \end{pmatrix}

    where {\cal{\widetilde{L}}}_{22}\in \mathbb R^{(m-a)\times a},\; \widehat{{\cal{L}}}_{22}'\in \mathbb R^{(m-a)\times (m-a)}.

    From Assumption 1 we know that each eigenvalue of \widehat{{\cal{L}}}_{22}' has positive real part [22]. Let Q_1 = (({1}/{\sqrt n})\mathbf 1_n,\widetilde Q_1) and Q_1^{-1} = \begin{pmatrix} \upsilon_1^T \\ \overline Q_1 \end{pmatrix},\;\;Q_1^{-1} {\cal{ L}}_{11}Q_1 = \begin{pmatrix} 0 & 0 \\ 0 & \mathcal{\widetilde L}_{11} \end{pmatrix} , where \upsilon_1=\sqrt n\pi_1 =[\upsilon_{11},\ldots, \upsilon_{1n}]^T, \pi_1^T{\cal{ L}}_{11}=0 . Let \{\lambda_j\}{}_{j=2}^{n} and \{\lambda_k\}{}_{k=1}^{m-a} be the eigenvalues of \mathcal{\widetilde L}_{11} and \widehat{{\cal{L}}}_{22}' , respectively. Then, we give the following assumption.

    Assumption 4: There exists a constant t_0\ge 0 such that \tau \tilde c_{t_0} max_{2\le j\le n}{\vert \lambda_j\vert^2}/{Re(\lambda_j)} < 1 and \tau \tilde c_{t_0}' max_{1\le k\le m-a}{\vert \lambda_k\vert^2}/{Re(\lambda_k)} < 1, where \tilde c_{t_0}={\rm{sup}}_{t\ge t_0}c_1(t) and \tilde c_{t_0}'={\rm{sup}}_{t\ge t_0}c_2(t).

    According to the interactions between agents in subgroup \Psi_1 , we know that the consensus problem of subgroup \Psi_1 is the general global consensus problem. Then, the following conditions on c_1(t) were explored to solve the global stochastic weak consensus problem [5], [14]:

    (C1) \int_{0}^{\infty} c_1(t)dt = \infty ;

    (C2) \lim_{t\to \infty} c_1(t) = 0 ;

    (C3) \lim_{t\to \infty}c_1(t)\log \int_{0}^{t}c_1(s)ds=0 .

    Remark 5: For MASs with both time-delays and additive noises, conditions (C1)−(C3) have been explored for the weak consensus problem in both the two senses mentioned above [14]. It can be seen that the form of condition (C3) is more complex, but it can reflect the decay rate of the control gain function accurately and help us to design a more suitable control protocol.

    However, for subgroup \Psi_2 , since the anticipant behavior is different from subgroup \Psi_1 and the state information from subgroup \Psi_1 may be available for the followers, we also need to consider the effect of the information from subgroup \Psi_1 . Note that the existing conditions are not sufficient to analyze the containment control problem of subgroup \Psi_2 . Then, we will analyze the effect of communications between the two subgroups and redesign control gain functions such that the MASs can achieve group hybrid coordination behaviors under time-delays and additive noises. We first give some conditions for c_2(t) and introduce a new condition (C7) on c_1(t) and \kappa(t) as follows:

    (C4) \int_{0}^{\infty} c_2(t)dt = \infty ;

    (C5) \lim_{t\to \infty} c_2(t) = 0 ;

    (C6) \lim_{t\to \infty}c_2(t)\log \int_{0}^{t}c_2(s)ds=0 ;

    (C7) \lim_{t\to \infty} \kappa^2(t)\int_{0}^{t}c_1^2(s)ds=0 .

    Remark 6: Conditions (C4)−(C6) are sufficient for both the global stochastic weak consensus and stochastic containment. For solving the group hybrid coordination control problem of MASs with both time-delays and additive noises, condition (C7) related to the communication intensity between the two subgroups is given. It reveals the influence of the communication impact between the two subgroups and gives a hint that we can impose the joint condition on \kappa(t) and c_1(t) to solve the group hybrid coordination control problem for the containment-oriented case.

    Theorem 1: For the MAS (1), suppose that Assumptions 1−4 hold. Then, under protocol (2), the MAS (1) can achieve group hybrid coordination behavior with stochastic weak consensus and stochastic containment if (C1)−(C7) hold.

    Proof: Here, we will examine the two subgroups respectively. First, we consider the agents in subgroup \Psi_1 . Let J_n= ({1}/{\sqrt n})\mathbf 1_n \upsilon_1^T. By the properties of matrix {\cal{ L}}_{11} and Lemma 1, we have {\cal{ L}}_{11}\mathbf 1_n = 0 , \upsilon_1^T{\cal{ L}}_{11} = 0 and (I_n-J_n){\cal{ L}}_{11} = {\cal{ L}}_{11} = {\cal{ L}}_{11}(I_n- J_n). Let \psi(t)=(I_n-J_n)X_1(t)=[\psi_1(t),\psi_2(t),\ldots,\psi_n(t)]^T , where \psi_i(t)\in \mathbb{R},\;i=1,\ldots,n. Define \tilde\psi(t)=Q_1^{-1}\psi(t)=[\tilde\psi_1(t),\tilde\psi_2(t),\ldots, \tilde\psi_n(t)]^T , \bar\psi(t)=[\tilde\psi_2(t),\ldots,\tilde\psi_n(t)]^T , \tilde \psi_i(t)\in \mathbb R, \;i=1,\ldots,n. By the definition of Q_1^{-1} in Lemma 1, we can get \tilde\psi_1(t)=\upsilon_1^T\psi(t)= \upsilon_1^T(I_n-J_n)X_1(t)=0 and

    \begin{equation*} d\bar \psi(t) = -c_1(t)\widetilde {{\cal{ L}}}_{11}\bar \psi(t-\tau)dt + dM(t)\end{equation*}

    where M(t) = \sum_{i,j=1}^{n}a_{ij}\sigma_{ji}\overline Q_1(I_n-J_n)\vartheta_{n,i} \int_{0}^{t}c_1(s)d\omega_{ji}(s) . Then, by the semi-decoupled skill, variation of constants formula, and some important estimation methods, we can get \lim_{t\to \infty} \mathbb E{\Vert \bar \psi(t)\Vert}^2 = 0 and \lim_{t\to \infty} \Vert \bar \psi(t)\Vert = 0 from (C1)−(C3) directly [14]. Then we can come to the conclusion that the agents in subgroup \Psi_1 can achieve stochastic weak consensus. Now we consider the stochastic containment problem of subgroup \Psi_2 . Let \gamma(t) =Q_1^{-1}X_1(t) =[\gamma_1(t),\gamma_2(t),\ldots,\gamma_n(t)]^T , \bar \gamma(t) = [\gamma_2(t),\ldots,\gamma_n(t)]^T, where \gamma_i(t)\in \mathbb{R},\;i=1,\ldots,n. We can obtain d\gamma_1(t)=\upsilon_1^T c_1(t) \sum_{i,j=1}^{n}a_{ij}\sigma_{ji}\vartheta_{n,i}d\omega_{ji}(t) . Then, \gamma_1(t) = \gamma_1^*+ \upsilon_1^T \sum_{i,j=1}^{n}a_{ij}\sigma_{ji}\vartheta_{n,i}\int_{0}^{t} c_1(s)d\omega_{ji}(s) , where \gamma_1^*=\upsilon_1^T \Delta_1 . We also have

    \begin{equation} d\bar\gamma(t)=-c_1(t) \widetilde {{\cal{ L}}}_{11} \bar\gamma(t-\tau)dt+dM'(t) \end{equation} (5)

    where M'(t)=\sum_{i,j=1}^{n}a_{ij}\sigma_{ji}\overline Q_1\vartheta_{n,i}\int_{0}^{t}c_1(s)d\omega_{ji}(s) . From [7], \lim_{t\to \infty}\bar\gamma(t)= [0,\ldots,0]^T can be proved easily. From (4), we have

    \begin{equation} dX_{21}(t)=0 \end{equation} (6)

    and

    \begin{split} dX_{22}(t)= \;& -c_2(t)({\cal{\widetilde{L}}}_{22}X_{21}(t-\tau)+\widehat{{\cal{L}}}_{22}'X_{22}(t-\tau))dt \\ & -c_2(t)\kappa(t){\cal{{\widehat{L}}}}_{21}X_1(t-\tau)dt \\ & +c_2(t)V_1(t) +c_2(t)\kappa(t)V_2(t) \end{split} (7)

    where V_1(t)=\sum_{i=n+a+1}^{n+m}\sum_{j=n+1}^{n+m} a_{ij}\sigma_{ji}\vartheta_{m-a,i-n-a}d{\omega_{ji}} (t) , and V_2(t)=\sum_{i=n+a+1}^{n+m}\sum_{j=1}^{n} a_{ij}\sigma_{ji}\vartheta_{m-a,i-n-a}d\omega_{ji}(t) . Letting \overline X_{22}(t)= X_{22}(t)+\widehat{{\cal{L}}}_{22}'^{-1}{\cal{\widetilde{L}}}_{22}X_{21}(t) , we obtain

    \begin{split} d\overline X_{22}(t)=\; & -c_2(t)\widehat{{\cal{L}}}_{22}'\overline X_{22}(t-\tau)dt \\ & -c_2(t)\kappa(t){\cal{{\widehat{L}}}}_{21}X_1(t-\tau)dt \\ & +c_2(t)V_1(t) +c_2(t)\kappa(t)V_2(t). \end{split} (8)

    From the above analysis, we have

    \begin{split} d\overline X_{22}(t)= \;& -c_2(t)\widehat{{\cal{L}}}_{22}'\overline X_{22}(t-\tau)dt-c_2(t)\kappa(t)P\gamma(t-\tau)dt \\ & +c_2(t)V_1(t) +c_2(t)\kappa(t)V_2(t) \end{split} (9)

    where P={\cal{{\widehat{L}}}}_{21}Q_1 . From the matrix theorem we have R\widehat{{\cal{L}}}_{22}'R^{-1} = J , where R is a complex invertible matrix, and J is the Jordan normal form of \widehat{{\cal{L}}}_{22}' . And we know J = diag(J_{\lambda_1,n_1},\ldots,J_{\lambda_l,n_l}) , \sum_{k=1}^{l} n_k = m-a , where \lambda_1,\ldots,\lambda_l are all the eigenvalues of \widehat{{\cal{L}}}_{22}' and J_{\lambda_k,n_k} is the Jordan block corresponding to eigenvalue \lambda_k , and its dimension is n_k . Letting Y(t)=R\overline X_{22}(t)=[Y_1(t),\ldots,Y_{m-a}(t)]^T with Y_j(t)\in \mathbb{R},\;j=1,\ldots, m-a, then we have dY(t) = -c_2(t)JY(t-\tau)dt -c_2(t)\kappa(t)\times D\gamma(t\;-\;\tau)dt \;+ \; Rd\bar M(t)\;+\; Rd\bar M'(t), where D=RP , \bar M(t)= \sum_{i=n+a+1}^{n+m}\sum_{j=n+1}^{n+m}a_{ij}\sigma_{ji}\vartheta_{m-a,i-n-a}\int_{0}^{t}c_2(s)d\omega_{ji}(s) , and \bar M'(t)= \sum_{i=n+a+1}^{n+m}\sum_{j=1}^{n}a_{ij}\sigma_{ji}\vartheta_{m-a,i-n-a}\int_{0}^{t}c_2(s)\kappa(s)d \omega_{ji}(s). Considering the kth Jordan block, and letting \varphi_k(t)\;=\;[\varphi_{k,1}(t),\ldots, \varphi_{k,n_k}(t)]^T , D(k) = [D_{k,1}^T,\ldots,D_{k,n_k}^T]^T, R(k) = [R_{k,1}^T,\ldots,R_{k,n_k}^T]^T, where \varphi_{k,j}(t)= Y_{k_j}(t) , D_{k,j} and R_{k,j}=R_{k_j} are k_j th row of D and R with k_j=\sum_{l=1}^{k-1} n_l+j , respectively, one has

    \begin{split} d\varphi_k(t)=\;& -c_2(t)J_{\lambda_k,n_k}\varphi_k(t-\tau)dt \\ &-c_2(t)\kappa(t)D(k)\gamma(t-\tau)dt \\ & + R(k)d\bar M(t)+ R(k)d\bar M'(t). \end{split} (10)

    From (10) we obtain the following semi-decoupled equations:

    \begin{split} d\varphi_{k,n_k}(t)= & -c_2(t)\lambda_k\varphi_{k,n_k}(t-\tau)dt \\ &-c_2(t)\kappa(t)D_{k,n_k}\gamma(t-\tau)dt \\ & + d\bar M_{k,n_k}(t)+ d\bar M_{k,n_k}'(t) \end{split} (11)

    and

    \begin{split} d\varphi_{k,j}(t)= \; & -c_2(t)\lambda_k\varphi_{k,j}(t-\tau)dt-c_2(t)\varphi_{k,j+1}(t-\tau)dt \\ & -c_2(t)\kappa(t)D_{k,j}\gamma(t-\tau)dt \\ & +d\bar M_{k,j}(t)+d\bar M_{k,j}'(t),\;\; j=1,\ldots,n_k-1 \end{split} (12)

    where \bar M_{k,j}(t)\;=\;\sum_{i=n+a+1}^{n+m}\;\sum_{q=n+1}^{n+m} r_{k_j,i}\;a_{iq}\sigma_{qi}\;\;\int_{0}^{t}c_2(s)d \omega_{qi}(s), \bar M_{k,j}'(t)\;= \; \sum_{i=n+a+1}^{n+m}\;\sum_{q=1}^{n} r_{k_j,i}a_{iq}\sigma_{qi}\;\int_{0}^{t}c_2(s)\kappa(s) d\omega_{qi}(s), r_{k_j,i}= R_{k_j}\vartheta_{m-a,i-n-a}, j=1,\ldots,n_k . Denote \Gamma_k(t,t_0) as the differential resolvent function defined in Lemma 2 with λ being replaced with \lambda_k . We let D_{k,n_k}=[d_1,\ldots,d_n] . Then, from (11), we have

    \begin{split} \varphi_{k,n_k}(t)= \; & \Gamma_k(t,t_0)\varphi_{k,n_k}(t_0) \\ & -\int_{t_0}^{t} \Gamma_k(t,s)c_2(s)\kappa(s)d_1\gamma_1(s)ds \\ & -\sum_{p=2}^{n}\int_{t_0}^{t} \Gamma_k(t,s)c_2(s)\kappa(s)d_p\gamma_p(s)ds \\ & +Z_{k,n_k}(t)+Z_{k,n_k}'(t) \end{split} (13)

    where Z_{k,\;n_k}(t)\;=\;\int_{t_0}^{t}\Gamma_k(t,\;s)d\bar M_{k,\;n_k}(s)\,, \;Z_{k,n_k}'(t)\;=\;\int_{t_0}^{t}\Gamma_k(t,s) \times d\bar M_{k,n_k}'(s) . Taking the expectation for (13) and using the Jensen inequality tool, we obtain

    \begin{split} \mathbb{E}{| \varphi_{k,n_k}(t)|}^2 \le\; & 5|\Gamma_k(t,t_0)|^2 \mathbb{E}{| \varphi_{k,n_k}(t_0)|}^2 \\ & +5\mathbb{E} \left|\int_{t_0}^{t} \Gamma_k(t,s)c_2(s)\kappa(s)d_1\gamma_1(s)ds\right|^2 \\ & +5\mathbb{E} \sum_{p=2}^{n}\left|\int_{t_0}^{t} \Gamma_k(t,s)c_2(s)\kappa(s)d_p\gamma_p(s)ds\right|^2 \\ & +C_{k,n_k}\int_{t_0}^{t} \Big|\Gamma_k(t,s)\Big|^2c_2^2(s)ds \\ & +C_{k,n_k}' \int_{t_0}^{t} \Big|\Gamma_k(t,s)\Big|^2c_2^2(s)\kappa^2(s)ds \end{split} (14)

    where C_{k,n_k} =\sum_{i=n+a+1}^{n+m}\sum_{q=n+1}^{n+m}|r_{k_{n_k},i}|^2 a^2_{iq}\sigma^2_{qi} , C_{k,n_k}' = \sum_{i=n+a+1}^{n+m}\times \sum_{q=1}^{n}|r_{k_{n_k},i}|^2 a^2_{iq}\sigma^2_{qi} . From Lemma 2, we obtain

    \begin{split} \mathbb{E}{| \varphi_{k,n_k}(t)|}^2 \le\; & 5b(\lambda_k)e^{-\varrho(\lambda_k)\int_{t_0}^{t}c_2(u)du} \mathbb{E}{| \varphi_{k,n_k}(t_0)|}^2 \\ &+5b(\lambda_k)\mathbb{E}\Bigg(\int_{t_0}^{t}e^{-0.5\varrho(\lambda_k)\int_{s}^{t}c_2(u)du}c_2(s) \\ & \times \kappa(s)|d_1\gamma_1(s)|ds\Bigg)^2 \\ &+5b(\lambda_k)\mathbb{E} \sum_{p=2}^{n}\Bigg(\int_{t_0}^{t}e^{-0.5\varrho(\lambda_k)\int_{s}^{t}c_2(u)du}c_2(s) \\ & \times\kappa(s)|d_p\gamma_p(s)|ds\Bigg)^2 \\ & +C_{k,n_k}b(\lambda_k)\int_{t_0}^{t}e^{-\varrho(\lambda_k)\int_{s}^{t}c_2(u)du} c_2^2(s)ds \\ & +C_{k,n_k}'b(\lambda_k) \int_{t_0}^{t} e^{-\varrho(\lambda_k)\int_{s}^{t}c_2(u)du}c_2^2(s)\kappa^2(s)ds. \end{split} (15)

    We now need to prove that the second term, the third term and the last term on the right hand side of (15) vanish in infinite time because other terms go toward zero, which can be obtained from [5] and [14]. We first consider the second term mentioned above. Fix the k and write S_{k,n_k}(t)= \int_{t_0}^{t} e^{-0.5\varrho(\lambda_k) \int_{s}^{t}c_2(u)du}c_2(s)\kappa(s)d_1\gamma_1(s)ds . Then, \mathop {{\rm{lim}}}\limits_{t\to \infty}\mathbb{E}{| S_{k,n_k}(t)|}^2 \le \lim_{t\to\infty}\mathbb{E}(\int_{0}^{t}e^{-0.5\varrho(\lambda_k)\int_{s}^{t}c_2(u)du} c_2(s)\kappa(s)|d_1\gamma_1(s)|ds)^2 . Let U(t)= \int_{0}^{t} e^{-0.5\varrho(\lambda_k) \int_{s}^{t}c_2(u)du} c_2(s)\kappa(s)|d_1\gamma_1(s)|ds . Then, we can obtain \sqrt{\mathbb{E}(U(t))^2}\;\le \; \int_{0}^{t}e^{-0.5\varrho(\lambda_k)\int_{s}^{t}c_2(u)du}c_2(s)\kappa(s)\sqrt{\mathbb{E}|d_1\gamma_1(s)|^2}ds. By means of L’Hôpital’s rule, we can get

    \begin{align*} \lim_{t\to \infty} \sqrt{\mathbb{E}(U(t))^2}\le\; &\lim_{t\to \infty}\frac{\sqrt{\mathbb{E}|d_1\gamma_1(t)|^2}\kappa(t)}{0.5\varrho(\lambda_k)} \\ \le\;&\lim_{t\to \infty} \frac {\sqrt{c_3\int_{0}^{t}c_1^2(s)ds+c_4\mathbb{E}\Vert\Delta_1\Vert^2}\kappa(t)}{0.5\varrho(\lambda_k)} \end{align*}

    where c_3 = d_1^2 \sum_{k=1}^{n} \upsilon_{1k}^2 \sum_{i,j=1}^{n}a_{ij}^2\sigma_{ji}^2 c_4 is a constant. From (C7), we can get \mathop {{\rm{lim}}}\limits_{t\to \infty}{\mathbb{E}(U(t))^2}=0, after which \mathop {{\rm{lim}}}\limits_{t\to \infty}\mathbb{E}{| S_{k,n_k}(t)|}^2= 0 can be obtained directly. Now we consider the remaining terms mentioned above. Note that \lim_{t\to \infty}\sum_{p=2}^{n}d_p\gamma_p(t)=0 . By similar skills used in estimating the second term, we can get \lim_{t\to \infty}\mathbb{E} \sum_{p=2}^{n} (\int_{t_0}^{t}e^{-0.5\varrho(\lambda_k)\int_{s}^{t}c_2(u)du}c_2(s)\kappa(s)|d_p\gamma_p(s)|ds)^2= 0 directly. From (C7), we have \lim_{t\to \infty}\kappa(t)=0 . By L’Hôpital’s rule, combining (C5) and (C7), we can similarly get

    \begin{equation} \begin{split} &\lim_{t\to \infty}\int_{0}^{t} e^{-0.5\varrho(\lambda_k) \int_{s}^{t}c_2(u)du}c_2^2(s)\kappa^2(s)ds \\ &\qquad =\lim_{t\to \infty}\frac{\int_{0}^{t} e^{0.5\varrho(\lambda_k) \int_{0}^{s}c_2(u)du}c_2^2(s)\kappa^2(s)ds}{e^{0.5\varrho(\lambda_k)\int_{0}^{t}c_2(u)du}} \\ & \qquad=\lim_{t\to \infty}\frac{c_2(t)\kappa^2(t)}{0.5\varrho(\lambda_k)}=0 \end{split} \end{equation} (16)

    then \lim_{t\to \infty}\mathbb{E}{| \varphi_{k,n_k}(t)|}^2=0 can be obtained for the fixed k. Similarly, we can obtain \mathbb{E}{| \varphi_{k,j}(t)|}^2=0, \;j=1,\ldots,n_k-1 for the fixed k. Repeating the above process, similar induction yields \lim_{t\to \infty} \mathbb{E}{\Vert \varphi_{k}(t)\Vert}^2=0 for all k=1,\ldots,l . Hence,

    \begin{equation} \lim_{t\to \infty} \mathbb E{\Vert\overline X_{22}(t)\Vert}^2=0. \end{equation} (17)

    We still consider the kth Jordan block. Similar to the definition of \varphi_k(t) , let \delta_k(t)=[\delta_{k,1}(t),\ldots,\delta_{k,n_k}(t)]^T . When \tau=0 , we can get d\delta_{k,n_k}(t)\;=\;-\;c_2(t)\lambda_k\delta_{k,n_k}(t)dt\;- \; c_2(t)\kappa(t)D_{k,n_k}\gamma(t)dt+ d\bar M_{k,n_k}(t)\;+\; d\bar M_{k,n_k}'(t) and d\delta_{k,j}(t)\;=\;-\;c_2(t)\lambda_k\delta_{k,j}(t)dt\;- c_2(t)\delta_{k,j+1}(t)dt- c_2(t)\kappa(t)D_{k,j}\gamma(t)dt+ d\bar M_{k,j}(t)+ d\bar M_{k,j}'(t),\;\,j=1, \ldots,n_k - 1. According to [35] and above analyses, it is easy to obtain \lim_{t\to \infty}| \delta_{k,j}(t)|\;=\;0, \;j\;=\;1,\ldots,n_k. Letting \theta_{k,n_k}(t)= \delta_{k,n_k}(t)- \varphi_{k,n_k}(t) , we have

    \begin{equation} \begin{split} d\theta_{k,n_k}(t)=\;& -c_2(t)\lambda_k\theta_{k,n_k}(t-\tau)dt+c_2(t)g_{k,n_k}(t)dt \\ & +c_2(t)\kappa(t)h_{k,n_k}(t)dt \end{split} \end{equation} (18)

    where g_{k,n_k}(t)=\lambda_k(\delta_{k,n_k}(t-\tau)-\delta_{k,n_k}(t)) and h_{k,n_k}(t)=D_{k,n_k}\times (\gamma(t-\tau)-\gamma(t)) . From the above analyses we can obtain \lim_{t\to \infty}| g_{k,n_k}(t)|=0 and \lim_{t\to \infty}| h_{k,n_k}(t)|=0 easily. Then, we obtain \theta_{k,n_k}(t)\;=\;\Gamma_k(t,t_0)\theta_{k,n_k}(t_0)+\int_{t_0}^{t} \Gamma_k(t,s)c_2(s)g_{k,n_k}(s)ds+\ \int_{t_0}^{t} \Gamma_k(t,s)c_2(s)\kappa(s)h_{k,n_k}(s)ds . By the similar skills used in above analysis, we can obtain \lim_{t\to \infty}| \theta_{k,n_k}(t)|=0 . From this and \lim_{t\to \infty}| \delta_{k,n_k}(t)|=0 , we have \lim_{t\to \infty}| \varphi_{k,n_k}(t)|=0 for the fixed k. We now assume that \lim_{t\to \infty}| \varphi_{k,j+1}(t)|=0 for some fixed j<n_k . Letting \tilde g_{k,j+1}(t)=\delta_{k,j+1}(t)-\varphi_{k,j+1}(t-\tau) , then we have \lim_{t\to \infty} | \tilde g_{k,j+1}(t)|=0 . We also have d\theta_{k,j}(t)=-c_2(t)\lambda_k\theta_{k,j}(t- \tau)dt+c_2(t)g_{k,j}(t)dt+c_2(t)\kappa(t)h_{k,j}(t)dt\;-\;c_2(t)\tilde g_{k,j+1}(t)dt, where g_{k,j}(t)\;=\;\lambda_k(\delta_{k,j}(t-\tau)-\delta_{k,j}(t)) , h_{k,j}(t)\;=\;D_{k,j}(\gamma(t-\tau)-\gamma(t)) , \lim_{t\to \infty} |g_{k,j}(t)|\;=\;0 and \lim_{t\to \infty}|h_{k,j}(t)|\;=\;0. Also, we have \theta_{k,j}(t)=\Gamma_k(t,t_0)\theta_{k,j}(t_0)+\int_{t_0}^{t} \Gamma_k(t,s)c_2(s)g_{k,j}(s)ds\;+\; \int_{t_0}^{t} \Gamma_k(t,s)\;\times c_2(s)k(s)h_{k,j}(s)ds-\int_{t_0}^{t} \Gamma_k(t,s)c_2(s)\tilde g_{k,j+1}(s)ds. Similarly, we can get \lim_{t\to \infty}| \theta_{k,j}(t)|=0 . From this and \lim_{t\to \infty} |\delta_{k,j}(t)|=0 one has \lim_{t\to \infty}| \varphi_{k,j}(t)|=0 for the fixed k. Furthermore, one can obtain \lim_{t\to \infty} {\Vert\varphi_k(t)\Vert}=0 for all k=1,\ldots,l . Then,

    \begin{split} \lim_{t\to \infty} {\Vert\overline X_{22}(t)\Vert}=0. \end{split} (19)

    From (17), (19) and the definition of \overline X_{22}(t) , we have \lim_{t\to \infty} \mathbb{E}\Vert X_{22}(t)-\bar X^*\Vert^2=0 and \lim_{t\to \infty}\Vert X_{22}(t)-\bar X^*\Vert=0 , where \bar X^*=-\widehat{{\cal{L}}}_{22}'^{-1}{\cal{\widetilde{L}}}_{22}\Delta_{21} , and the ith element of -\widehat{{\cal{L}}}_{22}'^{-1}{\cal{\widetilde{L}}}_{22}\Delta_{21} belongs to co_L in Definition 3. Hence, the agents in subgroup \Psi_2 achieve stochastic containment. That is, the MAS (1) achieve group hybrid coordination behavior with stochastic weak consensus and stochastic containment.

    For agents in subgroup \Psi_1 , we know that in addition to conditions (C1)−(C3), the following condition is also needed to explore the global stochastic strong consensus:

    (C8) \int_{0}^{\infty} c_1^2(t)dt < \infty .

    Then, we have the following theorem.

    Theorem 2: For the MAS (1), suppose that Assumptions 1−4 hold. Then, under protocol (2), the MAS (1) can achieve group hybrid coordination behavior with stochastic strong consensus and stochastic containment if (C1)−(C8) hold, and only if (C1) and (C8) hold.

    Proof: We first prove the “if ” part. We now consider the agents in subgroup \Psi_1 . Combining condition (C8) and the proof of Theorem 1, we can get that \int_{0}^{\infty} c_1(s)d\omega_{ji}(s) is well defined [5], \lim_{t\to \infty} \mathbb E{\Vert\bar\gamma(t)\Vert}^2=0 and \lim_{t\to \infty}\Vert\bar\gamma(t)\Vert= 0 . Then, we have \lim_{t\to \infty}\mathbb{E}\Vert X_1(t)-\bar x^*\mathbf{1}_n\Vert^2=0 and \lim_{t\to \infty}\Vert X_1(t)-\bar x^*\mathbf{1}_n\Vert= 0 , where \bar x^*=\gamma_1(\infty) . For the stochastic containment problem of subgroup \Psi_2 , the proof is similar to the proof of Theorem 1, and is omitted here. Also, the proof of the “only if ” part can be found in [14], [35] and we omit it.

    Above, we explored the group hybrid coordination control problem for the containment-oriented case and obtained some interesting results. Then, for the consensus-oriented case, we need to determine if the above conditions sufficient. We will explore this problem in Section IV.

    To examine the consensus-oriented case, we first give the following assumption for the communications between the two subgroups.

    Assumption 5: The agents in subgroup \Psi_1 can measure the state information from the agents in subgroup \Psi_2 .

    Based on the interactions between any two agents under the communication mode mentioned above, we give the control protocol as follows:

    \begin{equation} u_i(t) = \left\{ \begin{aligned} &c_1(t)\sum_{j=1}^{n} a_{ij}(\bar y_{ji1}(t) - x_i(t-\tau)) \\ &\qquad+ c_1(t)\sum_{j=n+1}^{n+m} a_{ij}\bar y_{ji2}(t),\;\; \forall i\in {\cal{ V}}_1 \\ &c_2(t)\sum_{j=n+1}^{n+m} a_{ij}(\bar{\bar y}_{ji}(t) - x_i(t-\tau)),\;\; \forall i\in {\cal{ V}}_2 \end{aligned}\right. \end{equation} (20)

    where t>0 , c_1(t) , c_2(t) \in C((0,\infty);\;[0,\infty)) are two control gain functions to be designed. \bar y_{ji1}(t)=x_j(t-\tau)+\sigma_{ji}\eta_{ji}(t), \; i\in {\cal{ V}}_1, j\in N_{1i}, is the state measurement of agent j belonging to subgroup \Psi_1 by agent i, and \bar y_{ji2}(t)=\kappa(t)(x_j(t-\tau)+\sigma_{ji}\eta_{ji}(t)), i\in {\cal{ V}}_1,\;j\in N_{2i}, is the state measurement of agent j belonging to subgroup \Psi_2 by agent i. \kappa(t) is a time-varying function, which reflects the communication intensity between subgroup \Psi_1 and subgroup \Psi_2 . \bar{\bar y}_{ji}(t)=x_j(t-\tau)+\sigma_{ji}\eta_{ji}(t),\; i\in {\cal{ V}}_2,\;j\in N_i is the state measurement of agent j by agent i belonging to subgroup \Psi_2 .

    Under protocol (20), from (1), we obtain the following equations:

    \begin{split} \dot X_1(t) = \;& -c_1(t)\overline{{\cal{ L}}}_{11}X_1(t-\tau) -c_1(t)\kappa(t){\cal{ L}}_{12}X_2(t-\tau) \\ & + c_1(t)\sum_{i,j=1}^{n} a_{ij}\sigma_{ji}\vartheta_{n,i}\eta_{ji}(t) \\ & +c_1(t)\kappa(t)\sum_{i=1}^{n}\sum_{j=n+1}^{n+m} a_{ij}\sigma_{ji}\vartheta_{n,i}\eta_{ji}(t) \end{split} (21)

    and

    \begin{split} \dot X_2(t) =\;& -c_2(t){\cal{ L}}_{22}X_2(t-\tau) \\ & +c_2(t)\sum_{i,j=n+1}^{n+m} a_{ij}\sigma_{ji}\vartheta_{m,i-n}\eta_{ji}(t) \end{split} (22)

    where \overline {{\cal{ L}}}_{11}=\left[\bar l_{ij}\right] is defined as

    \bar l_{ij}=\left\{ \begin{aligned} & -a_{ij}, & j\neq i,\;\; i\in {\cal{ V}}_1,\;\;j\in N_{1i} \\ &\sum_{k=1,k\neq i}^{n} a_{ik}, & j=i, \;\;i\in {\cal{ V}}_1,\;\;j\in N_{1i} \end{aligned} \right.

    and it is the Laplacian matrix of {\cal{ G}}_1 . {\cal{ L}}_{22} is the Laplacian matrix of {\cal{ G}}_2 .

    Similarly, from Assumption 1 we know that each eigenvalue of \widehat{{\cal{L}}}_{22} has positive real part [22]. Let Q_2 = (({1}/{\sqrt n})\mathbf 1_n,\widetilde Q_2) and Q_2^{-1} = \begin{pmatrix} \upsilon_2^T \\ \overline Q_2 \end{pmatrix} , Q_2^{-1}\overline {{\cal{ L}}}_{11}Q_2 = \begin{pmatrix} 0 & 0 \\ 0 & \widetilde {{\cal{ L}}}'_{11} \end{pmatrix} , where \upsilon_2=\sqrt n\pi_2= [\upsilon_{21},\ldots,\upsilon_{2n}]^T, \pi_2^T\overline{{\cal{ L}}}_{11}=0 . Let \{\lambda_j\}{}_{j=2}^{n} and \{\lambda_k\}{}_{k=1}^{m-a} be the eigenvalues of \widetilde{{\cal{ L}}}'_{11} and \widehat{{\cal{L}}}_{22} , respectively. Then, we give the following assumption.

    Assumption 6: There exists a constant t_0\ge 0 such that \tau \tilde c_{t_0} max_{2\le j\le n}({\vert \lambda_j\vert^2}\;/\;{Re(\lambda_j)}) \;< \; 1 and \tau \tilde c_{t_0}' max_{1\le k\le m-a}\;({\vert \lambda_k\vert^2}/ {Re(\lambda_k)} ) < 1, where \tilde c_{t_0}={\rm{sup}}_{t\ge t_0}c_1(t) and \tilde c_{t_0}'={\rm{sup}}_{t\ge t_0}c_2(t).

    To investigate this kind of group hybrid coordination behavior for the consensus-oriented case, we need the following condition:

    (C9) \lim_{t\to \infty} \kappa^2(t)\int_{0}^{t}c_2^2(s)ds=0 .

    Theorem 3: For the MAS (1), suppose that Assumptions 1, 2, 5 and 6 hold. Then, under protocol (20), the MAS (1) can achieve group hybrid coordination behavior with stochastic weak consensus and stochastic containment if (C1)−(C6) and (C9) hold.

    Proof: We first consider the stochastic containment problem of subgroup \Psi_2 . From (C4)−(C6), we can obtain the conclusion that the agents in subgroup \Psi_2 can achieve stochastic containment directly [14], [22]. Similar to the above analysis, from (22) we can obtain dX_{21}(t)=0 and

    \begin{align*} dX_{22}(t)=\; & -c_2(t)({\cal{\widetilde{L}}}_{22}X_{21}(t-\tau)+\widehat{{\cal{L}}}_{22}X_{22}(t-\tau))dt \\ & +c_2(t)\sum_{i=n+a+1}^{n+m}\sum_{j=n+1}^{n+m} a_{ij}\sigma_{ji}\vartheta_{m-a,i-n-a}d\omega_{ji}(t). \end{align*}

    Letting \overline X_{22}(t)=X_{22}(t)+\widehat{{\cal{L}}}_{22}^{-1}{\cal{\widetilde{L}}}_{22}X_{21}(t) , we have

    \begin{align*} d\overline X_{22}(t)=\; & -c_2(t)\widehat{{\cal{L}}}_{22}\overline X_{22}(t-\tau)dt \\ & +c_2(t)\sum_{i=n+a+1}^{n+m}\sum_{j=n+1}^{n+m} a_{ij}\sigma_{ji}\vartheta_{m-a,i-n-a}d\omega_{ji}(t). \end{align*}

    From the above analysis, we can obtain \lim_{t\to \infty} {\Vert\overline X_{22}(t)\Vert}^2= 0. Now we consider the consensus problem of subgroup \Psi_1 . From the above analysis and (21), we can get

    \begin{split} d X_1(t) = \;& -c_1(t)\overline{{\cal{ L}}}_{11}X_1(t-\tau)dt -c_1(t)\kappa(t){\cal{ L}}_{12}\bar X(t-\tau)dt \\ & + c_1(t)\sum_{i,j=1}^{n} a_{ij}\sigma_{ji}\vartheta_{n,i}d\omega_{ji}(t) \\ & +c_1(t)\kappa(t)\sum_{i=1}^{n}\sum_{j=n+1}^{n+m} a_{ij}\sigma_{ji}\vartheta_{n,i}d\omega_{ji}(t) \end{split} (23)

    where \bar X(t) =[X_{21}^T(t),(\bar X_{22}(t)-\widehat{{\cal{L}}}_{22}^{-1}{\cal{\widetilde{L}}}_{22}X_{21}(t))^T]^T =[\bar X_1(t),\ldots, \bar X_m(t)]^T. Let J_n=({1}/{\sqrt n})\mathbf 1_n \upsilon_2^T. By the properties of matrix \overline{{\cal{ L}}}_{11} and Lemma 1, we have \overline{{\cal{ L}}}_{11}\mathbf 1_n=0 and \upsilon_2^T \overline{{\cal{ L}}}_{11}=0 , then (I_n-J_n)\overline{{\cal{ L}}}_{11}=\overline{{\cal{ L}}}_{11}=\overline{{\cal{ L}}}_{11}(I_n-J_n) . Let \xi(t)=(I_n-J_n)X_1(t)= [\xi_1(t),\xi_2(t),\ldots,\xi_n(t)]^T, where \xi_i(t)\in \mathbb{R},i=1,\ldots,n . Define \tilde\xi(t)=Q_2^{-1}\xi(t) = [\tilde\xi_1(t),\tilde\xi_2(t),\ldots,\tilde\xi_n(t)]^T, \bar\xi(t)=[\tilde\xi_2(t),\ldots, \tilde\xi_n(t)]^T, \tilde \xi_i(t)\in \mathbb R,\;i=1,\ldots,n. By the definition of Q_2^{-1} in Lemma 1, we can get \tilde\xi_1(t)=\upsilon_2^T\xi(t)= \upsilon_2^T(I_n-J_n)X_1(t)=0 and

    \begin{split} d\bar \xi(t) = \;& -c_1(t)\widetilde {{\cal{ L}}}'_{11}\bar \xi(t-\tau)dt-c_2(t)\kappa(t)H\bar X(t-\tau)dt \\ & + dW(t)+d\bar W(t) \end{split} (24)

    where H\;=\;\overline Q_2(I_n\;-\;J_n)\;{\cal{ L}}_{12}, W(t) \;=\; \sum_{i,j=1}^{n}a_{ij}\sigma_{ji}\overline Q_2\;(I_n\;- J_n)\vartheta_{n,i}\;\int_{0}^{t}c_1(s)\;d\omega_{ji}\;(s), \bar W(t)\;=\; \sum_{i=1}^{n}\;\sum_{j=n+1}^{n+m} a_{ij}\sigma_{ji}\overline Q_2 \; (I_n- J_n)\vartheta_{n,i}\int_{0}^{t}c_1(s)\kappa(s)d\omega_{ji}(s). From [14], we now need to prove that \lim_{t\to \infty} \mathbb E {\Vert \bar \xi(t)\Vert}^2 = 0 for any initial state. Then, according to the matrix theorem, we have \bar R\widetilde {{\cal{ L}}}'_{11}\bar R^{-1} = J , where \bar R is a complex invertible matrix, J is the Jordan normal form of \widetilde {{\cal{ L}}}'_{11} . And we know J = {\rm{diag}}(J_{\lambda_1,n_1},\ldots,J_{\lambda_l,n_l}) , \sum_{k=1}^{l} n_k = n-1 , where \lambda_1,\ldots,\lambda_l are all the eigenvalues of \widetilde {{\cal{ L}}}'_{11} and J_{\lambda_k,n_k} is the Jordan block corresponding to eigenvalue \lambda_k , whose dimension is n_k . Letting \bar Y(t)=\bar R\bar \xi(t)=[\bar Y_1(t),\ldots,\bar Y_{n-1}(t)]^T with \bar Y_j(t)\in \mathbb{R},\;j=1,\ldots,n-1 , then we have d\bar Y(t) = -c_1(t)J\bar Y(t-\tau)dt - c_1(t)\kappa(t)\bar D\bar X(t-\tau)dt + \bar RdW(t)+ \bar Rd\bar W(t), where \bar D=\bar RH . Here, we first consider the kth Jordan block. Let \zeta_k(t)=[\zeta_{k,1}(t),\ldots, \zeta_{k,n_k}(t)]^T , \bar D(k) = [\bar D_{k,1}^T,\ldots,\bar D_{k,n_k}^T]^T , \bar R(k) = [\bar R_{k,1}^T,\ldots,\bar R_{k,n_k}^T]^T , where \zeta_{k,j}(t)=\bar Y_{k_j}(t) , \bar D_{k,j} is k_j th row of \bar D with k_j=\sum_{l=1}^{k-1} n_l+j , \bar D_{k,n_k}=[\bar d_1,\ldots,\bar d_m] , \bar R_{k,j}=\bar R_{k_j} is k_j th row of H with k_j= \sum_{l=1}^{k-1} n_l+j . We obtain

    \begin{split} d\zeta_k(t)=\; & -c_1(t)J_{\lambda_k,n_k}\zeta_k(t-\tau)dt \\ & -c_1(t)\kappa(t)\bar D(k)\bar X(t-\tau)dt \\ &+ \bar R(k)dW(t)+ \bar R(k)d\bar W(t). \end{split} (25)

    We further get the following semi-decoupled equations:

    \begin{split} d\zeta_{k,n_k}(t)= \; & -c_1(t)\lambda_k\zeta_{k,n_k}(t-\tau)dt \\ & -c_1(t)\kappa(t)\bar D_{k,n_k}\bar X(t-\tau)dt \\ & + dW_{k,n_k}(t)+ d\bar W_{k,n_k}(t) \end{split} (26)

    and

    \begin{split} d\zeta_{k,j}(t)= \; & -c_1(t)\lambda_k\zeta_{k,j}(t-\tau)dt-c_1(t)\zeta_{k,j+1}(t-\tau)dt \\ & -c_1(t)\kappa(t)\bar D_{k,j}\bar X(t-\tau)dt \\ & +dW_{k,j}(t)+d\bar W_{k,j}(t), \;\;j=1,\ldots,n_k-1 \end{split} (27)

    where W_{k,j}(t)\;=\;\sum_{i,q=1}^{n} r_{k_j,i}a_{iq}\sigma_{qi}\;\int_{0}^{t}c_1(s)\;d\omega_{qi}(s), \;\bar W_{k,j}\;(t) \; = \sum_{i=1}^{n}\sum_{q=n+1}^{n+m} r_{k_j,i}a_{iq}\sigma_{qi}\int_{0}^{t}c_1(s)\kappa(s)d\omega_{qi}(s), r_{k_j,i}\;=\;\bar R_{k_j}\overline{Q}_2(I_n\;- J_n)\vartheta_{n,i}, j=1,\ldots,n_k . From (26), we obtain

    \begin{split} \zeta_{k,n_k}(t)= \; & \Gamma_k(t,t_0)\zeta_{k,n_k}(t_0) \\ & -\sum_{p=1}^{m}\int_{t_0}^{t} \Gamma_k(t,t_0)c_1(s)\kappa(s)\bar d_p\bar X_p(s)ds \\ & +Z_{k,n_k}(t)+Z_{k,n_k}'(t) \end{split} (28)

    where Z_{k,n_k} (t) = \int_{t_0}^{t} e^{-\lambda_k\int_{s}^{t} c_1(u)du}dW_{k,n_k}(s), Z_{k,n_k}' (t) = \int_{t_0}^{t} e^{-\lambda_k\int_{s}^{t} c_1(u)du} d\bar W_{k,n_k}(s) . By taking the absolute value for (28), we have |\zeta_{k,n_k}(t)| \; \le\;|\Gamma_k(t\;,t_0)| |\;\zeta_{k,n_k}\;(t_0)| \; +\;\sum_{p=1}^{m}|\;\int_{t_0}^{t} \Gamma_k\;(t,t_0)\;c_1(s)\;\kappa(s)\;\times \bar d_p\bar X_p(s)ds| +|Z_{k,n_k}(t)|+|Z_{k,n_k}' (t)| \le \sqrt{b_0}e^{-0.5\varrho_0 \int_{t_0}^{t}c_1(u)du} |\zeta_{k,n_k}(t_0)| + \sqrt{b_0} \sum_{p=1}^{m} \int_{t_0}^{t} \;e^{-0.5\varrho_0 \int_{t_0}^{t}c_1(u)du}c_1(s)\kappa(s)|\bar d_p\bar X_p(s)| ds \;+\;|Z_{k,n_k}(t)|\;+ |Z_{k,n_k}'(t)| , where b_0= max _{2\le j \le n}b(\lambda_j) , \varrho_0= min _{2\le j \le n} \varrho(\lambda_j) . Then, we obtain

    \begin{split} \mathbb{E}{|\zeta_{k,n_k}(t)|}^2 \le\; & 4\sqrt{b_0}e^{-0.5\varrho_0\int_{t_0}^{t}c_1(u)du} \mathbb{E}{| \zeta_{k,n_k}(t_0)|}^2 \\ & +4\sqrt{b_0}\mathbb{E}\sum_{p=1}^{m}\Bigg(\int_{t_0}^{t} e^{-0.5\varrho(\lambda_k)\int_{s}^{t}c_1(u)du}c_1(s) \\ & \times \kappa(s)|\bar d_p\bar X_p(s)|ds\Bigg)^2 \\ & +C_{k,n_k}\sqrt{b_0}\int_{t_0}^{t} e^{-0.5\varrho_0\int_{s}^{t}c_1(u)du} c_1^2(s)ds \\ & +C_{k,n_k}'\sqrt{b_0} \int_{t_0}^{t} e^{-0.5\varrho_0\int_{s}^{t}c_1(u)du} c_1^2(s)\kappa^2(s)ds \end{split} (29)

    where C_{k,n_k} = \sum_{i,q=1}^{n} |r_{k_{j},i}|^2 a^2_{iq}\sigma^2_{qi} and C_{k,n_k}' = \sum_{i=1}^{n} \sum_{q=n+1}^{n+m} |r_{k_{j},i}|^2\times a^2_{iq}\sigma^2_{qi}. Here, let k be fixed and write S_{k,n_k}(t) = \int_{t_0}^{t} e^{-0.5\varrho(\lambda_k) \int_{s}^{t}c_1(u)du}c_1(s)\kappa(s)\bar d_1\bar X_1(s)ds. Similar to the proof of Theorem 1, we can get \lim_{t\to \infty} \mathbb E{|\zeta_{k,n_k}(t)|}^2=0 and \lim_{t\to \infty} {|\zeta_{k,n_k}(t)|}=0 for the fixed k. Similarly, we can obtain \lim_{t\to \infty} \mathbb E{\Vert \zeta_{k}(t) \Vert}^2=0 and \lim_{t\to \infty} {\Vert \zeta_{k}(t) \Vert}=0 for all k=1,\ldots,l . Hence,

    \begin{split} \lim_{t\to \infty} \mathbb E{\Vert \bar \xi(t) \Vert}^2=0,\;\; \lim_{t\to \infty} {\Vert \bar \xi(t) \Vert}=0. \end{split} (30)

    Then, we come to the conclusion that MAS (1) can achieve group hybrid coordination behavior with stochastic weak consensus and stochastic containment.

    Remark 7: Note that condition (C9) holds if \lim_{t\to \infty} \kappa(t)=0 . Then, we need to determine if we can obtain necessary conditions about \kappa(t) . Suppose that the MAS (1) can achieve the above group hybrid coordination behavior and then \lim_{t\to \infty} \mathbb E{| \zeta_{k,n_k}(t)|}^2=0 . From (29) and S_{k,n_k}(t) , we can obtain that when \bar d_1 \neq 0 , \lim_{t\to \infty} \kappa(t)=0 . This together with (C7) reveals the decay rate of the communication impact between the two subgroups and tells us that we can impose the joint condition on c_1(t) , c_2(t) and \kappa(t) to solve the group hybrid coordination control problem of MASs with time-delays and additive noises.

    To investigate this problem, the following condition is needed:

    (C10) \int_{0}^{\infty} c_1(t)\kappa(t)dt < \infty .

    Then, we give the following theorem, which is based on the above results.

    Theorem 4: For the MAS (1), suppose that Assumptions 1, 2, 5 and 6 hold, and c_1(t)\kappa(t) is monotonic. Then, under protocol (20), the MAS (1) can achieve group hybrid coordination behavior with stochastic strong consensus and stochastic containment if (C1)−(C6) and (C8)−(C10) hold, and only if (C1) and (C8) hold.

    Proof: We first prove the “if ” parts. From the above analysis, we can get that the agents in subgroup \Psi_2 can achieve stochastic containment directly. Then, we consider the subgroup \Psi_1 . Letting \xi'(t) =Q_2^{-1}X_1(t)=[\xi'_1(t),\xi'_2(t),\ldots,\xi'_n(t)]^T , \bar \xi'(t) = [\xi'_2(t),\ldots,\xi'_n(t)]^T , \xi'_i(t)\in\mathbb{R}, i=1,\ldots,n , then we can get d\xi'_1\;(t) \;= \;-c_1(t)\;\kappa(t)\;\beta(t)dt\;+\;\upsilon_2^T c_1(t)\; \sum_{i,j=1}^{n}\;a_{ij}\;\sigma_{ji}\vartheta_{n,i} d\omega_{ji}(t)\;+ \; \upsilon_2^T\; c_1(t)\;\kappa(t)\sum_{i=1}^{n}\;\sum_{j=n+1}^{n+m}\;a_{ij}\;\sigma_{ji}\;\vartheta_{n,i} \;d\omega_{ji}(t), where \beta(t)= \sum_{q=1}^{m}d_q' \bar X_q(t-\tau) is the first element of Q_2^{-1}{\cal{ L}}_{12}\bar X(t-\tau) , [d_1',\ldots,d_m'] is the first row of Q_2^{-1}{\cal{ L}}_{12} . We further obtain

    \begin{split} \xi'_1(t) = \; &\xi'^*_1 -\int_{0}^{t}c_1(s)\kappa(s)\beta(s)ds \\ & +\upsilon_2^T \sum_{i,j=1}^{n}a_{ij}\sigma_{ji} \vartheta_{n,i}\int_{0}^{t} c_1(s)d\omega_{ji}(s) \\ &+\upsilon_2^T \sum_{i=1}^{n}\sum_{j=n+1}^{n+m}a_{ij}\sigma_{ji}\vartheta_{n,i}\int_{0}^{t}c_1(s)\kappa(s)d\omega_{ji}(s) \end{split} (31)

    where \xi'^*_1\;=\;\upsilon_2^T \Delta_1, \upsilon_2^T \sum_{i,j=1}^{n}a_{ij}\sigma_{ji}{\vartheta_{n,i}}\;\int_{0}^{\infty}c_1(s)\;d\omega_{ji}(s) and \upsilon_2^T \sum_{i=1}^{n}\sum_{j=n+1}^{n+m}a_{ij}\sigma_{ji}\vartheta_{n,i}\int_{0}^{\infty}c_1(s)\kappa(s)d\omega_{ji}(s) are well defined. Then, we have \lim_{t\to \infty}\int_{0}^{t}c_1(s)\kappa(s)\beta(s)ds=\lim_{t\to \infty}\sum_{q=1}^{m} d_q'\times \int_{0}^{t}\bar X_q(s)c_1(s)\kappa(s)ds. Noting that \lim_{t\to \infty} \sum_{q=1}^{m}d_q'\bar X_q(t)<\infty , then from (C10) and Abel test we have \lim_{t\to \infty}\int_{0}^{t}c_1(s)\kappa(s)\times \;\beta(s)ds < \infty. We also have

    \begin{split} \mathbb E |\xi'_1(t)|^2 \le \; & 4|\xi'^*_1|^2+ 4\mathbb E|\int_{0}^{t}c_1(s)\kappa(s)\beta(s)ds|^2 \\ &+a_1\int_{0}^{t}c_1^2(s)ds+a_2\int_{0}^{t}c_1^2(s)\kappa^2(s)ds \end{split} (32)

    where a_1 , a_2 are all constants and a_1=\sum_{k=1}^{n}\upsilon_{2k}^2\sum_{i,j=1}^{n}a^2_{ij} \times \sigma^2_{ji} > 0 and a_2\;=\;\sum_{k=1}^{n} \upsilon_{2k}^2\sum_{i=1}^{n}\;\sum_{j=n+1}^{n+m}a^2_{ij}\;\sigma^2_{ji}\; > \;0. Similar to the proof of \lim_{t\to \infty}\int_{0}^{t}c_1(s)\kappa(s)\beta(s)ds<\infty , we have \lim_{t\to \infty}\mathbb{E}|\int_{0}^{t}\;c_1(s)\kappa(s)\;\beta(s)ds|^2 \;< \;\infty. Then, we obtain

    \begin{equation} \lim_{t\to\infty}{\mathbb{E}|\xi'_1(t)-\xi'^{\infty}_1|}^2=0, \lim_{t\to\infty}\xi'_1(t)=\xi'^{\infty}_1\end{equation} (33)

    where \xi'^{\infty}_1\;=\;\xi'^*_1 -\int_{0}^{\infty}c_1(s)\kappa(s)\beta(s)ds\;+\; \upsilon_2^T\sum_{i,j=1}^{n}a_{ij} \sigma_{ji}\vartheta_{n,i}\times \int_{0}^{\infty} c_1(s)d\omega_{ji}(s) + \upsilon_2^T \sum_{i=1}^{n}\sum_{j=n+1}^{n+m}a_{ij}\sigma_{ji}\vartheta_{n,i}\int_{0}^{\infty} c_1(s)\kappa(s)d\omega_{ji}(s) < \infty. We also have d\bar\xi'(t)=-c_1(t) \widetilde {{\cal{ L}}}'_{11} \xi'(t-\tau)dt -c_1(t)\kappa(t)P'dt+ d\tilde W(t)+d\tilde W'(t), where P'=\overline Q_2{\cal{ L}}_{12}\bar X(t-\tau) , \tilde W(t)=\sum_{i,j=1}^{n}a_{ij}\times \sigma_{ji}\overline Q_2 {\vartheta_{n,i}} \int_{0}^{t}c_1(s) d\omega_{ji}(s) and \tilde W'(t) = \sum_{i=1}^{n} \sum_{j=n+1}^{n+m} a_{ij}\sigma_{ji}\overline Q_2{\vartheta_{n,i}}\times \int_{0}^{t} c_1(s) \kappa(s)d\omega_{ji}(s) . Utilizing similar skills in the above analysis, we can easily get \lim_{t\to \infty}\mathbb{E}{\Vert \bar \xi'(t)\Vert}^2=0 and \lim_{t\to \infty} \Vert \bar \xi'(t)\Vert=0 . This together with (33) gives the conclusion that the “if ” part of Theorem 4 holds. We now prove the “only if ” part. The necessity of (C1) can be found in the proof of Theorem 3. From (33), \lim_{t\to \infty}\mathbb E|\xi'_1(t)|^2<\infty , a_1>0 and a_2>0 , we can get the necessity of (C8).

    Remark 8: From the above analyses, we can see that more conditions are needed to achieve group hybrid coordination control compared to global coordination control. Since different subgroups are required to achieve different coordination behaviors, the system dynamics cannot simply be analyzed as a whole. This results in the existing methods and conditions, such as those in [14], [22], and [24]–[32], unable to solve the group hybrid coordination control problem. Hence, in this paper, we develop a new analysis idea to study this problem. Then, some sufficient conditions and necessary conditions are obtained, and the influence mechanism of the communication impact between the two subgroups is revealed for the group hybrid coordination control problem of MASs with both time-delays and additive noises.

    In this section, several simulation examples are presented to illustrate the validity of the protocols and conditions for the group hybrid coordination control problem.

    For the containment-oriented case, we give a MAS with seven agents and the communication topology is shown in Fig. 1, where the number of subgroups is 2 and the agents 4 and 5 are two stationary leaders. We can see that Assumption 1 and Assumption 3 hold. Here, we require that agents 1, 2, 3 achieve consensus and agents 4, 5, 6, 7 achieve containment. We give the initial states X_1(t)=\Delta_1=\left[5,-1,-3\right]^T , X_2(t)= \Delta_2=\left[-2,-3,7,-6\right]^T, t\in [-\tau,0] , and assume that \sigma_{ji}=1, \;i,j= 1,2,3,4,5,6,7. We first consider \tau=0.4 , the control gain functions c_1(t)=(1+t)^{-0.5} , c_2(t)=(1+t)^{-0.8} and \kappa(t)=(1+t)^{-0.4} . It is easy to get that Assumption 4 and conditions (C1)−(C7) hold. From Theorem 1, we have the conclusion that the given MAS can achieve group hybrid coordination behavior with stochastic weak consensus and stochastic containment, which is shown in Fig. 2. Then, we consider \tau = 0.8 , c_1(t) = 0.6(1+t)^{-1},\; c_2(t)=0.5(1+t)^{-0.8} and \kappa(t)=(1+ t)^{-0.2} . Other conditions are the same as above. It is easy to verify that Assumption 4 and conditions (C1)−(C8) hold. From Theorem 2, we have the conclusion that the given MAS can achieve group hybrid coordination behavior with stochastic strong consensus and stochastic containment, which is shown in Fig. 3.

    Figure  1.  The communication graph: The containment-oriented case.
    Figure  2.  The containment-oriented case: Stochastic weak consensus + Stochastic containment.
    Figure  3.  The containment-oriented case: Stochastic strong consensus + Stochastic containment.

    Then, for the consensus-oriented case, we give another MAS with seven agents and the interactions between agents is shown in Fig. 4, where the number of subgroups is 2 and the agents 4 and 5 are also two stationary leaders. We can also see that Assumptions 1 and 5 hold. We also require agents in the two subgroups to achieve the coordination behavior mentioned above. Now, we give X_1(t)=\Delta_1=\left[-4,1,-2\right]^T , X_2(t)= \Delta_2=\left[1,2,3,-4\right]^T, t\in [-\tau,0] , and assume that \sigma_{ji}=1,\;i,j= 1,2,3,4,5,6,7. We first consider \tau=0.5 , c_1(t)=0.9(1+t)^{-0.4} , c_2(t)=0.7(1+t)^{-1} and \kappa(t)=(1+t)^{-0.7} . It is easy to get that Assumption 6, conditions (C1)−(C6) and (C9) hold. From Theorem 3 we have the conclusion that the given MAS can achieve group hybrid coordination behavior with stochastic weak consensus and stochastic containment, which is shown in Fig. 5. We now consider \tau=1.5 , c_1(t)=0.3(1+t)^{-0.9},\; c_2(t)= 0.3(1+t)^{-1} and \kappa(t)=(1+t)^{-0.6} . Other conditions are the same as above. Then, it is easy to get that Assumption 6, conditions (C1)−(C6) and (C8)−(C10) hold. Theorem 4 tells us that the given MAS can achieve group hybrid coordination behavior with stochastic strong consensus and stochastic containment, which is shown in Fig. 6.

    Figure  4.  The communication graph: The consensus-oriented case.
    Figure  5.  The consensus-oriented case: Stochastic weak consensus + Stochastic containment.
    Figure  6.  The consensus-oriented case: Stochastic strong consensus + Stochastic containment.

    Now, we give a counter-example to show the necessity of the decay of the communication impact between the two subgroups for the consensus-oriented case. We consider \tau= 0.5, c_1(t)=0.8(1+t)^{-0.9}, \;c_2(t)=0.9(1+t)^{-1} and \kappa(t)=1 . Other conditions are the same as above. It is easy to get that Assumption 6, conditions (C1)−(C6), (C8) and (C10) hold but (C9) does not hold. Then, we obtain Fig. 7, which shows that the given MAS fails to achieve group hybrid coordination behavior.

    Figure  7.  A counter-example for the consensus-oriented case.

    This work addressed the group hybrid coordination control problem of MASs with both time-delays and additive noises. The containment-oriented case and the consensus-oriented case were discussed respectively. Using the semi-decoupled skill and some estimation methods, this work provided a new analysis idea to investigate the group hybrid coordination control problem. Then, some sufficient conditions and necessary conditions were obtained and some interesting phenomenons were found. It is concluded that to solve the group hybrid coordination control problem, more conditions are needed than to solve the global coordination control, and the decaying condition of communication impact between the two subgroups is necessary for the consensus-oriented case. Intuitively, in real networks, the communication impact between the two subgroups may gradually weaken with the increase of the distance between agents in different subgroups. This might be beneficial for designing the group control protocol. Furthermore, the following new findings are obtained: 1) The conditions for the containment-oriented case are weaker than for the consensus-oriented case; 2) For the consensus-oriented case, the conditions for weak consensus are weaker than those for strong consensus.

    Recently, many important results on distributed coordination control have been obtained, such as coordination control based on dynamic event-triggered [36] and [37]. Determining how to apply these advanced control strategies to the control problem studied in this paper to improve the system communication efficiency and execution capability is a future direction of investigation.

  • [1]
    R. Olfati-Saber and R.M. Murray, “Consensus problems in networks of agents with switching topology and time-delays,” IEEE Trans. Autom. Control, vol. 49, no. 9, pp. 1520–1533, 2004. doi: 10.1109/TAC.2004.834113
    [2]
    W. Xiao, L. Cao, H. Li, and R. Lu, “Observer-based adaptive consensus control for nonlinear multi-agent systems with time-delay,” Sci. China Inf. Sci., vol. 63, no. 9, p. 132202, 2020.
    [3]
    Lin and Y. Jia, “Average consensus in networks of multi-agents with both switching topology and coupling time-delay,” Physica A, vol. 387, pp. 303–313, 2008. doi: 10.1016/j.physa.2007.08.040
    [4]
    W. Yu, G. Chen, M. Cao, and W. Ren, “Delay-induced consensus and quasi-consensus in multi-agent dynamical systems,” IEEE Trans. Circuits Syst. I-Regul. Pap., vol. 60, pp. 2679–2687, 2013. doi: 10.1109/TCSI.2013.2244357
    [5]
    T. Li and J. Zhang, “Mean square average-consensus under measurement noises and fixed topologies: Necessary and sufficient conditions,” Automatica, vol. 45, pp. 1929–1936, 2009. doi: 10.1016/j.automatica.2009.04.017
    [6]
    L. Cheng, Z. Hou, and M. Tan, “A mean square consensus protocol for linear multi-agent systems with communication noises and fixed topologies,” IEEE Trans. Autom. Control, vol. 59, no. 1, pp. 261–267, 2013.
    [7]
    B. Wang and J. Zhang, “Consensus conditions of multi-agent systems with unbalanced topology and stochastic disturbances,” J. Sys. Sci. Math. Scis., vol. 29, no. 10, pp. 1353–1365, 2009.
    [8]
    Y. Wang, L. Cheng, Z. Hou, M. Tan, H. Liu, and M. Wang, “Necessary and sufficient conditions for solving leader-following problem of multi-agent systems with communication noises,” in Proc. 25th Chinese Control and Decision Conf., 2013, pp. 778–783.
    [9]
    Y. Ni and X. Li, “Consensus seeking in multi-agent systems with multiplicative measurement noises,” Syst. Control Lett., vol. 62, no. 5, pp. 430–437, 2013. doi: 10.1016/j.sysconle.2013.01.011
    [10]
    Y. Zhang, R. Li and G. Chen, “Consensus control of second-order time-delayed multiagent systems in noisy environments using absolute velocity and relative position measurements,” IEEE T. Cybern., vol. 51, no. 11, pp. 5364–5374, 2021. doi: 10.1109/TCYB.2019.2954623
    [11]
    J. Liu, X. Liu, W. Xie, and H. Zhang, “Stochastic consensus seeking with communication delays,” Automatica, vol. 47, pp. 2689–2696, 2011. doi: 10.1016/j.automatica.2011.09.005
    [12]
    T. Li and J. Zhang, “Consensus conditions of multi-agent systems with time-varying topologies and stochastic communication noises,” IEEE Trans. Autom. Control, vol. 55, no. 9, pp. 2043–2057, 2010. doi: 10.1109/TAC.2010.2042982
    [13]
    S. Liu, L. Xie, and H. Zhang, “Distributed consensus for multi-agent systems with delays and noises in transmission channels,” Automatica, vol. 47, pp. 920–934, 2011. doi: 10.1016/j.automatica.2011.02.003
    [14]
    X. Zong, T. Li, and J. Zhang, “Consensus conditions of continuous-time multi-agent systems with time-delays and measurement noises,” Automatica, vol. 99, pp. 412–419, 2019. doi: 10.1016/j.automatica.2018.04.002
    [15]
    H. Liu, G. Xie, and L. Wang, “Necessary and sufficient conditions for containment control of networked multi-agent systems,” Automatica, vol. 48, no. 7, pp. 1415–1422, 2012. doi: 10.1016/j.automatica.2012.05.010
    [16]
    Z. Peng, J. Hu, and B. Ghosh, “Data-driven containment control of discrete-time multi-agent systems via value iteration,” Sci. China Inf. Sci., vol. 63, no. 8, p. 189205, 2020.
    [17]
    Y. Lou and Y. Hong, “Target containment control of multi-agent systems with random switching interconnection topologies,” Automatica, vol. 48, no. 5, pp. 879–885, 2012. doi: 10.1016/j.automatica.2012.02.032
    [18]
    F. Wang, Y. Ni, Z. Liu and Z. Chen, “.Containment control for general second-order multiagent systems with switched dynamics,” IEEE T. Cybern., vol. 50, no. 2, pp. 550–560, 2020. doi: 10.1109/TCYB.2018.2869706
    [19]
    H. Yang, F. Wang, and F. Han, “Containment control of fractional order multi-agent systems with time delays,” IEEE/CAA J. Autom. Sinica, vol. 5, no. 3, pp. 727–732, 2018. doi: 10.1109/JAS.2016.7510211
    [20]
    D. Wang, D. Wang, and W. Wang, “Necessary and sufficient conditions for containment control of multi-agent systems with time delay,” Automatica, vol. 103, pp. 418–423, 2019. doi: 10.1016/j.automatica.2018.12.029
    [21]
    K. Liu, G. Xie, and L. Wang, “Containment control for second-order multi-agent systems with time-varying delays,” Syst. Control Lett., vol. 67, pp. 24–31, 2014. doi: 10.1016/j.sysconle.2013.12.013
    [22]
    Y. Wang, L. Cheng, Z. Hou, M. Tan, and M. Wang, “Containment control of multi-agent systems in a noisy communication environment,” Automatica, vol. 50, no. 7, pp. 1922–1928, 2014. doi: 10.1016/j.automatica.2014.05.018
    [23]
    F. Zhou, Z. Wang, and N. Fan, “Containment consensus with measurement noises and time-varying communication delays,” Chin. Phys. B, vol. 24, no. 2, pp. 23–29, 2015.
    [24]
    J. Yu and L. Wang, “Group consensus of multi-agent systems with directed information exchange,” Int. J. Syst. Sci., vol. 43, pp. 334–348, 2012. doi: 10.1080/00207721.2010.496056
    [25]
    J. Yu and L. Wang, “Group consensus in multi-agent systems with switching topologies and communication delays,” Syst. Control Lett., vol. 59, pp. 340–348, 2010. doi: 10.1016/j.sysconle.2010.03.009
    [26]
    L. Ji, Q. Liu, and X. Liao, “On reaching group consensus for linearly coupled multi-agent networks,” Inf. Sci., vol. 287, pp. 1–12, 2014. doi: 10.1016/j.ins.2014.07.024
    [27]
    B. An, G. Liu, and C. Tan, “Group consensus control for networked multi-agent systems with communication delays,” ISA Trans., vol. 76, pp. 78–87, 2018. doi: 10.1016/j.isatra.2018.03.008
    [28]
    Y. Shang, “Group consensus of multi-agent systems in directed networks with noises and time delays,” Int. J. Syst. Sci., vol. 46, no. 14, pp. 2481–2492, 2015. doi: 10.1080/00207721.2013.862582
    [29]
    M. Song, X. Kong, H. Li, and X. Lu, “Asymptotic consensus and asymptotic group consensus of Boolean networks with additive noise,” in Proc. Int. Conf. Information, Cybernetics, and Computational Social Systems, 2020, pp. 531–536.
    [30]
    W. Li, H. Zhang, Y. Zhou, and Y. Wang, “Bipartite formation tracking for multi-agent systems using fully distributed dynamic edge-event-triggered protocol,” IEEE/CAA J. Autom. Sinica, vol. 9, no. 5, pp. 847–853, 2022. doi: 10.1109/JAS.2021.1004377
    [31]
    C. Ma, W. Zhao, and Y. Zhao, “Bipartite consensus of discrete-time double-integrator multi-agent systems with measurement noise,” J. Syst. Sci. Complex., vol. 31, no. 6, pp. 1525–1540, 2018. doi: 10.1007/s11424-018-7363-x
    [32]
    L. Pan, H. Shao, Y. Xi, and D. Li, “Bipartite consensus problem on matrix-valued weighted directed networks,” Sci. China Inf. Sci., vol. 64, no. 4, p. 149204, 2021.
    [33]
    X. Dong, Q. Li, Q. Zhao, and Z. Ren, “Time-varying group formation analysis and design for general linear multi-agent systems with directed topologies,” Int. J. Robust Nonlinear Control, vol. 27, no. 9, pp. 1640–1652, 2017.
    [34]
    D. Meng, Z. Meng, and Y. Hong, “Disagreement of hierarchical opinion dynamics with changing antagonisms,” SIAM J. Control Optim., vol. 57, no. 1, pp. 718–742, 2019. doi: 10.1137/17M1130344
    [35]
    C. Li and X. Zong, “Group consensus of multi-agent systems with additive noises,” Sci. China Inf. Sci., 2022, DOI: 10.1007/s11432-021-3424-x.
    [36]
    X. Ge, S. Xiao, Q.-L. Han, X.-M. Zhang, and D. Ding, “Dynamic event-triggered scheduling and platooning control co-design for automated vehicles over vehicular ad-hoc networks,” IEEE/CAA J. Autom. Sinica, vol. 9, no. 1, pp. 31–46, 2022. doi: 10.1109/JAS.2021.1004060
    [37]
    X. Ge, Q.-L. Han, L. Ding, Y.-L. Wang, X.-M. Zhang, “Dynamic event-triggered distributed coordination control and its applications: A survey of trends and techniques,” IEEE Trans. Syst. Man Cybern. -Syst., vol. 50, no. 9, pp. 3112–3125, 2020. doi: 10.1109/TSMC.2020.3010825
  • Related Articles

    [1]Meilin Li, Yue Long, Tieshan Li, Hongjing Liang, C. L. Philip Chen. Dynamic Event-Triggered Consensus Control for Input Constrained Multi-Agent Systems With a Designable Minimum Inter-Event Time[J]. IEEE/CAA Journal of Automatica Sinica, 2024, 11(3): 649-660. doi: 10.1109/JAS.2023.123582
    [2]Lihong Feng, Bonan Huang, Jiayue Sun, Qiuye Sun, Xiangpeng Xie. Adaptive Event-Triggered Time-Varying Output Group Formation Containment Control of Heterogeneous Multiagent Systems[J]. IEEE/CAA Journal of Automatica Sinica, 2024, 11(6): 1398-1409. doi: 10.1109/JAS.2024.124260
    [3]Hongru Ren, Hui Ma, Hongyi Li, Zhenyou Wang. Adaptive Fixed-Time Control of Nonlinear MASs With Actuator Faults[J]. IEEE/CAA Journal of Automatica Sinica, 2023, 10(5): 1252-1262. doi: 10.1109/JAS.2023.123558
    [4]Haihua Guo, Min Meng, Gang Feng. Lyapunov-Based Output Containment Control of Heterogeneous Multi-Agent Systems With Markovian Switching Topologies and Distributed Delays[J]. IEEE/CAA Journal of Automatica Sinica, 2023, 10(6): 1421-1433. doi: 10.1109/JAS.2023.123198
    [5]Jingshu Sang, Dazhong Ma, Yu Zhou. Group-Consensus of Hierarchical Containment Control for Linear Multi-Agent Systems[J]. IEEE/CAA Journal of Automatica Sinica, 2023, 10(6): 1462-1474. doi: 10.1109/JAS.2023.123528
    [6]Wenchao Huang, Hailin Liu, Jie Huang. Distributed Robust Containment Control of Linear Heterogeneous Multi-Agent Systems: An Output Regulation Approach[J]. IEEE/CAA Journal of Automatica Sinica, 2022, 9(5): 864-877. doi: 10.1109/JAS.2022.105560
    [7]Yamei Ju, Derui Ding, Xiao He, Qing-Long Han, Guoliang Wei. Consensus Control of Multi-Agent Systems Using Fault-Estimation-in-the-Loop: Dynamic Event-Triggered Case[J]. IEEE/CAA Journal of Automatica Sinica, 2022, 9(8): 1440-1451. doi: 10.1109/JAS.2021.1004386
    [8]Qinglai Wei, Xin Wang, Xiangnan Zhong, Naiqi Wu. Consensus Control of Leader-Following Multi-Agent Systems in Directed Topology With Heterogeneous Disturbances[J]. IEEE/CAA Journal of Automatica Sinica, 2021, 8(2): 423-431. doi: 10.1109/JAS.2021.1003838
    [9]Yanxu Su, Qingling Wang, Changyin Sun. Self-triggered Consensus Control for Linear Multi-agent Systems With Input Saturation[J]. IEEE/CAA Journal of Automatica Sinica, 2020, 7(1): 150-157. doi: 10.1109/JAS.2019.1911837
    [10]Ting Wang, Min Hu, Yanlong Zhao. Consensus Control With a Constant Gain for Discrete-time Binary-valued Multi-agent Systems Based on a Projected Empirical Measure Method[J]. IEEE/CAA Journal of Automatica Sinica, 2019, 6(4): 1052-1059. doi: 10.1109/JAS.2019.1911594
    [11]Long Ma, Haibo Min, Shicheng Wang, Yuan Liu, Zhiguo Liu. Distributed Containment Control of Networked Nonlinear Second-order Systems With Unknown Parameters[J]. IEEE/CAA Journal of Automatica Sinica, 2018, 5(1): 232-239. doi: 10.1109/JAS.2016.7510235
    [12]Hongyong Yang, Fuyong Wang, Fujun Han. Containment Control of Fractional Order Multi-Agent Systems With Time Delays[J]. IEEE/CAA Journal of Automatica Sinica, 2018, 5(3): 727-732. doi: 10.1109/JAS.2016.7510211
    [13]Zhenhua Wang, Juanjuan Xu, Huanshui Zhang. Consensus Seeking for Discrete-time Multi-agent Systems with Communication Delay[J]. IEEE/CAA Journal of Automatica Sinica, 2015, 2(2): 151-157.
    [14]Wen Qin, Zhongxin Liu, Zengqiang Chen. Formation Control for Nonlinear Multi-agent Systems with Linear Extended State Observer[J]. IEEE/CAA Journal of Automatica Sinica, 2014, 1(2): 171-179.
    [15]Airong Wei, Xiaoming Hu, Yuzhen Wang. Tracking Control of Leader-follower Multi-agent Systems Subject to Actuator Saturation[J]. IEEE/CAA Journal of Automatica Sinica, 2014, 1(1): 84-91.
    [16]Jianming Wei, Yunan Hu, Meimei Sun. Adaptive Iterative Learning Control for a Class of Nonlinear Time-varying Systems with Unknown Delays and Input Dead-zone[J]. IEEE/CAA Journal of Automatica Sinica, 2014, 1(3): 302-314.
    [17]Chuanrui Wang, Xinghu Wang, Haibo Ji. A Continuous Leader-following Consensus Control Strategy for a Class of Uncertain Multi-agent Systems[J]. IEEE/CAA Journal of Automatica Sinica, 2014, 1(2): 187-192.
    [18]Hao Zhang, Gang Feng, Huaicheng Yan, Qijun Chen. Distributed Self-triggered Control for Consensus of Multi-agent Systems[J]. IEEE/CAA Journal of Automatica Sinica, 2014, 1(1): 40-45.
    [19]Huiyang Liu, Long Cheng, Min Tan, Zengguang Hou. Containment Control of General Linear Multi-agent Systems with Multiple Dynamic Leaders: a Fast Sliding Mode Based Approach[J]. IEEE/CAA Journal of Automatica Sinica, 2014, 1(2): 134-140.
    [20]Wenhui Liu, Feiqi Deng, Jiarong Liang, Haijun Liu. Distributed Average Consensus in Multi-agent Networks with Limited Bandwidth and Time-delays[J]. IEEE/CAA Journal of Automatica Sinica, 2014, 1(2): 193-203.

Catalog

    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

    Figures(7)

    Article Metrics

    Article views (487) PDF downloads(94) Cited by()

    Highlights

    • A new kind of group coordination control problem——group hybrid coordination control is investigated in this paper. The group hybrid coordination control means that in a whole multi-agent system that consists of two subgroups with communications between them, agents in the two subgroups achieve consensus and containment, respectively
    • Using the semi-decoupled skill and some estimation methods, this work provides a new analysis idea to investigate the group hybrid coordination control problem under time-delays and additive noises. Some sufficient conditions and necessary conditions are obtained for this problem
    • The influence mechanism of the communication impact between the two subgroups on group hybrid coordination control problem of MASs with both time-delays and additive noises is revealed: the decay of the communication impact between the two subgroups is necessary for the consensus-oriented case. This provides hints in the design of control gain functions for the similar group coordination control problem

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return