This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Full-State Prescribed Performance-Based Consensus of Double-Integrator Multi-Agent Systems with Jointly Connected Topologies

Yahui Hou and Bin Cheng, Member, IEEE Yahui Hou is with the School of Automation, Central South University, Changsha 410083, China (e-mail: [email protected]). Bin Cheng is with the Department of Control Science and Engineering, College of Electronics and Information Engineering, Tongji University, Shanghai 201804, China (e-mail: [email protected]). Corresponding authors: Yahui Hou and Bin Cheng.
Abstract

This paper addresses the full-state prescribed performance-based consensus problem for double-integrator multi-agent systems with jointly connected topologies. To improve the transient performance, a distributed prescribed performance control protocol consisting of the transformed relative position and the transformed relative velocity is proposed, where the communication topology satisfies the jointly connected assumption. Different from the existing literatures, two independent transient performance specifications imposed on relative positions and relative velocities can be guaranteed simultaneously. A numerical example is ultimately used to validate the effectiveness of proposed protocol.

Index Terms:
Consensus, prescribed performance control, jointly connected topologies, double-integrator dynamics, multi-agent systems.

I Introduction

Recently, multi-agent systems (MASs) have received considerable attention in completing a range of tasks such as rescue, transportation, and exploration, mainly because they are efficient and inexpensive[1]. More and more research topics were conducted on the cooperative control of MASs, in which consensus reaching is a classical topic[2]. The consensus of MASs was achieved for different dynamics, including single-integrator dynamic [3], double-integrator dynamic [4, 5], linear dynamic [6, 7], and nonlinear dynamic [8].

Although the control protocols in literatures [3, 4, 5, 6, 7, 8] achieved asymptotic convergence as expected, the steady and transient-state performance during the convergence process were usually neglected. To address this issue, the prescribed performance control (PPC) method based on performance functions and error transformation functions was first proposed in [9], where the tracking errors were confined within an arbitrarily small residual set, with predefined minimum convergence rate and maximum overshoot. Soon afterwards, the PPC method was applied to the single-integrator and double-integrator MASs to achieve consensus [10, 11], in which the trajectories of relative positions defined as the difference among the connected agents respected the performance bound.

In detail, we then review the application of PPC method in MASs. In [11], Macellari et al. proposed a distributed control protocol consisting of a linear combination term between relative positions and relative velocities as well as an additional absolute velocity term, such that the performance specifications imposed on the above linear combination term can be guaranteed. The performance specifications are imposed on the linear combination term, however, relative positions and relative velocities cannot be confined within their respective performance bounds. In terms of the controller designed in [11], modulating the performance specifications on relative velocities makes the stability analysis more complicated. In [12], the designed controller guaranteed a transient performance for the robot joint position and velocity tracking errors. Recently, more research results on this topic can be found in [13, 14, 15, 16].

Up to this point, no research results were found in which the relative positions and relative velocities among connected agents evolve within their respective predefined performance bounds. In short, these independent performance specifications associated with positions and velocities form the so-called full-state prescribed performance for double-integrator MASs. The above literatures mainly considered fixed topology, while switching topology cases caused by the failures and changes of communication link among neighboring agents are rarely involved [17, 18].

Motivated by the above discussion, the full-state prescribed performance-based consensus control problem for double-integrator MASs with jointly connected topologies is addressed. To be specific, a vital advantage of this paper, is that a novel distributed control protocol is designed to improve the transient performance of both relative positions and relative velocities, which are different from the existing PPC methods based on MASs with single-integrator [10, 13] and double-integrator [11, 15, 16] dynamics. Furthermore, considering the jointly connected topology case, all agents are driven to converge asymptotically by applying prescribed performance-based control protocol, and the design parameters are independent of topology information. It is worth mentioning that topology graphs are no longer limited to tree and connected graph cases [13, 14, 15], and MASs can achieve the accurate consensus rather than practical consensus [12, 14].

II Background and Problem Formulation

II-A Notations and Graph Theory

For convenience, we summarize some notations as follows. N\mathbb{R}^{N}: the set of NN dimension real column vectors; INI_{N}: the N×NN\times N identity matrix; x𝖳x^{{\sf\tiny T}}: the transposition of xx; diag{d1,d2,,dn}{\rm diag}\{d_{1},d_{2},\cdots,d_{n}\}: diagonal matrix with entries d1,d2,,dnd_{1},d_{2},\cdots,d_{n}.

An undirected graph 𝒢σ(t)\mathcal{G}_{\sigma(t)} is composed of the node set 𝒱={1,2,,N}\mathcal{V}=\{1,2,\cdots,N\} and edge set σ(t)𝒱×𝒱\mathcal{E}_{\sigma(t)}\subset\mathcal{V}\times\mathcal{V}, where σ(t):[0,+)𝒫\sigma(t):[0,+\infty)\to\mathcal{P}, 𝒫={1,2,,n0}\mathcal{P}=\{1,2,\cdots,n_{0}\} is a piecewise constant switching signal. Let 𝒜σ(t)=[aij(t)]N×N\mathcal{A}_{\sigma(t)}=[a_{ij}(t)]\in\mathbb{R}^{N\times N} be the adjacency matrix where aij(t)=1a_{ij}(t)=1 if (i,j)σ(t)(i,j)\in\mathcal{E}_{\sigma(t)} and aij(t)=0a_{ij}(t)=0 otherwise. The neighbor set of node ii is 𝒩i(t)={j𝒱:(i,j)σ(t)}\mathcal{N}_{i}(t)=\{j\in\mathcal{V}:(i,j)\in\mathcal{E}_{\sigma(t)}\}. The Laplacian matrix is defined as σ(t)=𝒟σ(t)𝒜σ(t)\mathcal{L}_{\sigma(t)}=\mathcal{D}_{\sigma(t)}-\mathcal{A}_{\sigma(t)}, where degree matrix 𝒟σ(t)=diag{d1(t),d2(t),,dN(t)}\mathcal{D}_{\sigma(t)}={\rm diag}\{d_{1}(t),d_{2}(t),\cdots,d_{N}(t)\} with diagonal entry di(t)=j𝒩i(t)aij(t)d_{i}(t)=\sum_{j\in\mathcal{N}_{i}(t)}a_{ij}(t). The incidence matrix is defined as Bσ(t)=[bil(t)]N×Mσ(t)B_{\sigma(t)}=[b_{il}(t)]\in\mathbb{R}^{N\times M_{\sigma(t)}}, i𝒱i\in\mathcal{V}, lσ(t)l\in\mathcal{H}_{\sigma(t)} where σ(t)={1,,Mσ(t)}\mathcal{H}_{\sigma(t)}=\{1,\cdots,M_{\sigma(t)}\} is a edge set. Besides, edge (i,j)σ(t)(i,j)\in\mathcal{E}_{\sigma(t)}, j𝒩i(t)j\in\mathcal{N}_{i}(t) is described as edge lσ(t)l\in\mathcal{H}_{\sigma(t)}. For the connected graph, the edge Laplacian matrix σ(t)e=Bσ(t)𝖳Bσ(t)\mathcal{L}_{\sigma(t)}^{e}=B_{\sigma(t)}^{{\sf\tiny T}}B_{\sigma(t)} is positive semidefinite [15, 16].

Suppose there exists an infinite, bounded and contiguous time interval [tk,tk+1)[t_{k},t_{k+1}), kk\in\mathbb{N} where t0=0t_{0}=0 and tk+1tkνt_{k+1}-t_{k}\leq\nu for positive constant ν\nu. During each interval [tk,tk+1)[t_{k},t_{k+1}), there is a sequence of nonoverlapping subintervals [tk0,tk1),,[tkq,tkq+1),,[tkmk1,tkmk)[t_{k}^{0},t_{k}^{1}),\cdots,[t_{k}^{q},t_{k}^{q+1}),\cdots,[t_{k}^{m_{k}-1},t_{k}^{m_{k}}), with tk0=tkt_{k}^{0}=t_{k} and tkmk=tk+1t_{k}^{m_{k}}=t_{k+1}, satisfying tkq+1tkqτ>0t_{k}^{q+1}-t_{k}^{q}\geq\tau>0, 0qmk0\leq q\leq m_{k}, and τ\tau is called dwell time. The graph 𝒢σ(t)\mathcal{G}_{\sigma(t)} is fixed for t[tkq,tkq+1)t\in[t_{k}^{q},t_{k}^{q+1}). If the union graph q=0mk𝒢σ(tkq)\cup_{q=0}^{m_{k}}\mathcal{G}_{\sigma(t_{k}^{q})} is connected, then 𝒢σ(t)\mathcal{G}_{\sigma(t)} is jointly connected.

II-B Prescribed Performance Control

Partly referring to [9, 10], the prescribed performance is achieved if element sl(t)s_{l}(t) of tracking errors s(t)Mσ(t)s(t)\in\mathbb{R}^{M_{\sigma(t)}} evolves within the following performance bounds

ρs,l(t)<sl(t)<ρs,l(t),lσ(t)\displaystyle-\rho_{s,l}(t)<s_{l}(t)<\rho_{s,l}(t),~{}\forall l\in\mathcal{H}_{\sigma(t)} (1)

where performance function ρs,l(t)=(ρs,l0ρs,l)e𝔤t+ρs,l\rho_{s,l}(t)=(\rho_{s,l}^{0}-\rho_{s,l}^{\infty})e^{-\mathfrak{g}t}+\rho_{s,l}^{\infty} with positive constants ρs,l0\rho_{s,l}^{0}, ρs,l\rho_{s,l}^{\infty} and 𝔤\mathfrak{g}. Next, we normalize sl(t)s_{l}(t) by s^l(t)=sl(t)ρs,l(t)\hat{s}_{l}(t)=\frac{s_{l}(t)}{{\rho_{s,l}(t)}}, and define the prescribed performance regions Ds^l{s^l(t):s^l(t)(1,1)}D_{\hat{s}_{l}}\triangleq\left\{\hat{s}_{l}(t)\in{\mathbb{R}}:\hat{s}_{l}(t)\in(-1,1)\right\} with Ds^Ds^1×Ds^2××Ds^Mσ(t)D_{\hat{s}}\triangleq D_{\hat{s}_{1}}\times D_{\hat{s}_{2}}\times\cdots\times D_{\hat{s}_{M_{\sigma(t)}}}. Let s^=[s^1,,s^Mσ(t)]𝖳\hat{s}=[\hat{s}_{1},\cdots,\hat{s}_{M_{\sigma(t)}}]^{{\sf\tiny T}} and ρs(t)=diag{ρs,1(t),,ρs,Mσ(t)(t)}\rho_{s}(t)={\rm diag}\{\rho_{s,1}(t),\cdots,\rho_{s,{M_{\sigma(t)}}}(t)\}. Thus s^l(t)Ds^l\hat{s}_{l}(t)\in D_{\hat{s}_{l}}, t0t\geq 0 is equivalent to (1).

The normalized value s^l(t)\hat{s}_{l}(t) is transformed by tracking error transformation functions that define a smooth and strictly increasing mapping εs,l:Ds^l(,)\varepsilon_{s,l}:D_{\hat{s}_{l}}\to(-\infty,\infty), with εs,l(0)=0\varepsilon_{s,l}(0)=0. More specifically, one can select transformation function as

εs,l(s^l)=ln(1+s^l1s^l).\displaystyle\varepsilon_{s,l}(\hat{s}_{l})=\ln\bigg{(}\frac{1+\hat{s}_{l}}{1-\hat{s}_{l}}\bigg{)}. (2)

The time derivative of εs,l(s^l)\varepsilon_{s,l}(\hat{s}_{l}) is

ε˙s,l(s^l)=Js,l(s^l,t)[s˙l+αs,l(t)sl]\displaystyle\dot{\varepsilon}_{s,l}(\hat{s}_{l})=J_{s,l}(\hat{s}_{l},t)\left[{\dot{s}_{l}+\alpha_{s,l}(t)s_{l}}\right] (3)

with Js,l(s^l,t)=21s^l2>0J_{s,l}(\hat{s}_{l},t)=\frac{2}{1-\hat{s}_{l}^{2}}>0 and αs,l(t)=ρ˙s,l(t)ρs,l(t)\alpha_{s,l}(t)=-\frac{\dot{\rho}_{s,l}(t)}{\rho_{s,l}(t)}. Let αs(t)=diag{αs,1(t),,αs,Mσ(t)(t)}\alpha_{s}(t)={\rm diag}\{\alpha_{s,1}(t),\cdots,\alpha_{s,{M_{\sigma(t)}}}(t)\}. By calculating, the following inequality

supt0,lσ(t)[αs,l(t)]α¯s\displaystyle\sup_{t\geq 0,l\in\mathcal{H}_{\sigma(t)}}[\alpha_{s,l}(t)]\leq\bar{\alpha}_{s} (4)

holds for a positive constant α¯s\bar{\alpha}_{s}. In view of the properties of transformation function εs,l(s^l)\varepsilon_{s,l}(\hat{s}_{l}), with constants ζs,1,ζs,2>0\zeta_{s,1},\zeta_{s,2}>0, the following inequality

sJs(s^,t)εs(s^)ζs,1εs2(s^),sJs(s^,t)εs(s^)ζs,2s^2\displaystyle sJ_{s}(\hat{s},t)\varepsilon_{s}(\hat{s})\geq\zeta_{s,1}\varepsilon_{s}^{2}(\hat{s}),~{}sJ_{s}(\hat{s},t)\varepsilon_{s}(\hat{s})\geq\zeta_{s,2}\hat{s}^{2} (5)

holds[12], where s=[s1,,sMσ(t)]𝖳s=[s_{1},\cdots,s_{M_{\sigma(t)}}]^{{\sf\tiny T}}, s^=[s^1,,s^Mσ(t)]𝖳\hat{s}=[\hat{s}_{1},\cdots,\hat{s}_{M_{\sigma(t)}}]^{{\sf\tiny T}}, εs(s^)=[εs,1(s^1),,εs,Mσ(t)(s^Mσ(t))]𝖳\varepsilon_{s}(\hat{s})=[\varepsilon_{s,1}(\hat{s}_{1}),\cdots,\varepsilon_{s,{M_{\sigma(t)}}}(\hat{s}_{M_{\sigma(t)}})]^{{\sf\tiny T}}, and Js(s^,t)=diag{Js,1(s^1,t),,Js,Mσ(t)(s^Mσ(t),t)}J_{s}(\hat{s},t)={\rm diag}\{J_{s,1}(\hat{s}_{1},t),\cdots,J_{s,{M_{\sigma(t)}}}(\hat{s}_{M_{\sigma(t)}},t)\}.

Lemma 1 ([19])

Given any positive constants kk and DD, we have

kεs𝖳(s^)Js(s^,t)s+D0\displaystyle-k\varepsilon_{s}^{{\sf\tiny T}}(\hat{s})J_{s}(\hat{s},t)s+D\leq 0

for all εs(s^)>ε¯s\|\varepsilon_{s}(\hat{s})\|>\bar{\varepsilon}_{s}, with constant ε¯s>0\bar{\varepsilon}_{s}>0.

II-C Problem Formulation

Consider a MAS with NN agents modeled by double-integrator dynamic

x˙i(t)=vi(t),v˙i(t)=ui(t),i𝒱\dot{x}_{i}(t)=v_{i}(t),\ \dot{v}_{i}(t)=u_{i}(t),\ i\in\mathcal{V} (6)

where xi(t),vi(t),ui(t){x}_{i}(t),v_{i}(t),u_{i}(t)\in\mathbb{R} represent the position, velocity and control input of agent ii, respectively.

Assumption 1

The graph 𝒢σ(t)\mathcal{G}_{\sigma(t)} are jointly connected across each interval [tk,tk+1)[t_{k},t_{k+1}), kk\in\mathbb{N}.

Based on the matrix Bσ(t)B_{\sigma(t)}, we define relative position

yl(t)=xi(t)xj(t),\displaystyle y_{l}(t)=x_{i}(t)-x_{j}(t),

and relative velocity

zl(t)=vi(t)vj(t),t[tkq,tkq+1)\displaystyle z_{l}(t)=v_{i}(t)-v_{j}(t),~{}t\in[t_{k}^{q},t_{k}^{q+1})

where edge ll is connected between agents ii and jj. Letting x=[x1,,xN]𝖳x=[x_{1},\cdots,x_{N}]^{{\sf\tiny T}}, v=[v1,,vN]𝖳v=[v_{1},\cdots,v_{N}]^{{\sf\tiny T}}, y=[y1,,yMσ(t)]𝖳y=[y_{1},\cdots,y_{M_{\sigma(t)}}]^{{\sf\tiny T}} and z=[z1,,zMσ(t)]𝖳z=[z_{1},\cdots,z_{M_{\sigma(t)}}]^{{\sf\tiny T}}, we have y=Bσ(t)𝖳xy=B_{\sigma(t)}^{{\sf\tiny T}}x and z=Bσ(t)𝖳vz=B_{\sigma(t)}^{{\sf\tiny T}}v.

In the following, the performance specifications are imposed on tracking errors yl(t)y_{l}(t) and zl(t)z_{l}(t). For convenience, the symbol s{\rm s} of section II-B is replaced with y{\rm y} and q{\rm q} respectively.

Assumption 2 ([11])

The values y^l(tkq)\hat{y}_{l}(t_{k}^{q}) and z^l(tkq)\hat{z}_{l}(t_{k}^{q}) are inside the performance bounds (1), lσ(t)\forall l\in\mathcal{H}_{\sigma(t)}.

Objective: For the MAS (6) with switching topologies 𝒢σ(t)\mathcal{G}_{\sigma(t)} and performance functions ρy,l(t)\rho_{y,l}(t) and ρz,l(t)\rho_{z,l}(t), design a distributed control protocol ui(t)u_{i}(t) to achieve consensus in the sense of

limtyl(t)=0andlimtzl(t)=0,lσ(t)\displaystyle\lim_{t\to\infty}y_{l}(t)=0~{}{\rm and}~{}\lim_{t\to\infty}z_{l}(t)=0,~{}\forall l\in\mathcal{H}_{\sigma(t)}

while transient performance y^(t)Dy^\hat{y}(t)\in D_{\hat{y}} and z^(t)Dz^\hat{z}(t)\in D_{\hat{z}}, t0t\geq 0 are guaranteed.

III Main Results

To achieve the above objective, we propose the prescribed performance control protocol as follows

ui(t)=\displaystyle u_{i}(t)= l=1Mσ(t)bil(t)Jy,l(y^l,t)εy,l(y^l)\displaystyle-\sum_{l=1}^{{M_{\sigma(t)}}}b_{il}(t)J_{y,l}(\hat{y}_{l},t)\varepsilon_{y,l}(\hat{y}_{l})
ϕl=1Mσ(t)bil(t)Jz,l(z^l,t)εz,l(z^l),i𝒱\displaystyle-\phi\sum_{l=1}^{{M_{\sigma(t)}}}b_{il}(t)J_{z,l}(\hat{z}_{l},t)\varepsilon_{z,l}(\hat{z}_{l}),\;i\in\mathcal{V} (7)

where ϕ\phi is a positive constant. Invoking (6) and (III), the closed-loop system is rewritten as

x¨=Bσ(t)Jy(y^,t)εy(y^)ϕBσ(t)Jz(z^,t)εz(z^).\displaystyle\ddot{x}=-B_{\sigma(t)}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})-\phi B_{\sigma(t)}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z}). (8)
Remark 1

Compared to existing literatures [10, 13, 20] based on single-integrator MASs, the velocity information of double-integrator MAS (6) makes controller design and stability analysis more complicated. The control protocol (III) consists of two nonlinear error transformation terms, which is a vital advantage in improving the transient performance of relative positions and relative velocities, instead of single relative positions mentioned in [11].

Partly inspired by [16], the following lemmas are obtained.

Lemma 2 ([16, 21])

Given a matrix Q=[h1INh2INh3INh4IN]Q=\scriptsize{\begin{bmatrix}h_{1}I_{N}&-h_{2}I_{N}\\ h_{3}I_{N}&h_{4}I_{N}\end{bmatrix}} with positive constants h1h_{1}, h2h_{2}, h3h_{3}, h4h_{4} and any vectors ϵ1,ϵ2N\epsilon_{1},\epsilon_{2}\in\mathbb{R}^{N}. If 4h1h4>(h3h2)24h_{1}h_{4}>(h_{3}-h_{2})^{2}, then

[ϵ1ϵ2]𝖳Q[ϵ1ϵ2]>0.\displaystyle\begin{bmatrix}\epsilon_{1}\\ \epsilon_{2}\end{bmatrix}^{{\sf\tiny T}}Q\begin{bmatrix}\epsilon_{1}\\ \epsilon_{2}\end{bmatrix}>0. (9)
Lemma 3 ([16, 21])

Under Assumption 2, consider the MAS (6) and protocol (III). If initial values x(0)x(0) and v(0)v(0) are bounded, we obtain that x(t)x(t) and v(t)v(t) are uniformly continuous and bounded for t[0,tmax)t\in[0,t_{max}).

Proof:

Simplifying the equality (III), we obtain

v˙i(t)=Si(t),i𝒱\displaystyle\dot{v}_{i}(t)=S_{i}(t),\;i\in\mathcal{V} (10)

where Si(t)=l=1Mσ(t)bil(t)Jy,l(y^l,t)εy,l(y^l)ϕl=1Mσ(t)bil(t)Jz,l(z^l,t)εz,l(z^l)S_{i}(t)=-\sum_{l=1}^{M_{\sigma(t)}}b_{il}(t)J_{y,l}(\hat{y}_{l},t)\varepsilon_{y,l}(\hat{y}_{l})-\phi\sum_{l=1}^{M_{\sigma(t)}}b_{il}(t)J_{z,l}(\hat{z}_{l},t)\varepsilon_{z,l}(\hat{z}_{l}), for t[tkq,tkq+1)t\in[t_{k}^{q},t_{k}^{q+1}). By solving differential equation (10), one has, with t[tkq,tkq+1)t\in[t_{k}^{q},t_{k}^{q+1}),

vi(t)=vi(tkq)+tkqtSi(τ)𝑑τ.\displaystyle v_{i}(t)=v_{i}(t_{k}^{q})+\int_{t_{k}^{q}}^{t}S_{i}(\tau)d\tau.

Since y^(0)Dy^\hat{y}(0)\in D_{\hat{y}} and z^(0)Dz^\hat{z}(0)\in D_{\hat{z}}, and x(0)x(0) and v(0)v(0) are bounded, vi(t)v_{i}(t) and xi(t)=xi(0)+0t01vi(τ)𝑑τx_{i}(t)=x_{i}(0)+\int_{0}^{t_{0}^{1}}v_{i}(\tau)d\tau, t[0,t01)t\in[0,t_{0}^{1}) are continuous and bounded. By calculating, Si(t01)S_{i}(t_{0}^{1}) is bounded, and thus vi(t)v_{i}(t) and xi(t)x_{i}(t), t[t01,t02)t\in[t_{0}^{1},t_{0}^{2}) are continuous and bounded. In return, the same conclusion can be reached for x(t)x(t) and v(t)v(t), t[tkq,tkq+1)t\in[t_{k}^{q},t_{k}^{q+1}). Letting ttmaxt\to t_{max}, it follows from [19] that xi(t)x_{i}(t) and vi(t)v_{i}(t), t[0,tmax)t\in[0,t_{max}) are uniformly continuous and bounded. The proof is completed. ∎

Theorem 1

Under Assumptions 1 and 2, consider the MAS (6) and protocol (III) with performance functions ρy,l(t)\rho_{y,l}(t) and ρz,l(t)\rho_{z,l}(t). If positive constants h1,h2,h3,h4,h5,h6,ϕh_{1},h_{2},h_{3},h_{4},h_{5},h_{6},\phi are selected such that

4h1h4>(h3h2)2,h3h22h5α¯y>0,\displaystyle 4h_{1}h_{4}>(h_{3}-h_{2})^{2},\;h_{3}-h_{2}-2h_{5}\bar{\alpha}_{y}>0,
h4ϕh6α¯z>0, 2h6ϕa2ϕ(h3h2)2h6a4>0,\displaystyle h_{4}\phi-h_{6}\bar{\alpha}_{z}>0,\;2h_{6}\phi-a_{2}\phi(h_{3}-h_{2})-2h_{6}a_{4}>0,

then tracking errors yl(t)y_{l}(t) and zl(t)z_{l}(t) evolve within performance bound (1) respectively, and converge to zero as tt\to\infty.

Proof:

The proof is divided into four steps. To begin with, there exists a unique maximum solution w(t)=[y^(t),z^(t)]𝖳w(t)=[\hat{y}(t),\hat{z}(t)]^{{\sf\tiny T}} over the set D=Dy^×Dz^D=D_{\hat{y}}\times D_{\hat{z}}, i.e., w(t)Dw(t)\in D, t[0,τmax)\forall t\in[0,\tau_{\max}). Nextly, it is verified that the protocol (III)\eqref{u1} ensures, y^(t)\hat{y}(t) and z^(t)\hat{z}(t) for t[0,τmax)t\in[0,\tau_{\max}) are bounded and strictly within the compact subset of Dy^D_{\hat{y}} and Dz^D_{\hat{z}} respectively. Then proof by contradiction leads to τmax=\tau_{\max}=\infty. The consensus is finally proven by Barbalat’s Lemma.

Phase I. Since y^(0)Dy^\hat{y}(0)\in D_{\hat{y}} and z^(0)Dz^\hat{z}(0)\in D_{\hat{z}}, one has w(0)Dw(0)\in D. By calculating the derivative of y^(t)\hat{y}(t) and z^(t)\hat{z}(t), it is confirmed that w˙(t)\dot{w}(t) is continuous and locally Lipschitz on ww. In view of Theorem 54 of [22], there exists the maximal solution w(t)w(t), and hence w(t)Dw(t)\in D is guaranteed for t[0,τmax)t\in[0,\tau_{\max}).

Phase II. It follows from Phase I that yl(t)y_{l}(t) and zl(t)z_{l}(t), t[0,τmax)\forall t\in[0,\tau_{\max}) satisfy bound (1) respectively. Consider a potential function as follows

V(t)=12ξ𝖳Qξ+h52εy𝖳(y^)εy(y^)+h62εz𝖳(z^)εz(z^)\displaystyle V(t)=\frac{1}{2}\xi^{{\sf\tiny T}}Q\xi+\frac{h_{5}}{2}\varepsilon_{y}^{{\sf\tiny T}}(\hat{y})\varepsilon_{y}(\hat{y})+\frac{h_{6}}{2}\varepsilon_{z}^{{\sf\tiny T}}(\hat{z})\varepsilon_{z}(\hat{z}) (11)

where ξ=[x𝖳,v𝖳]𝖳\xi=[x^{{\sf\tiny T}},v^{{\sf\tiny T}}]^{{\sf\tiny T}}, Q=[h1INh2INh3INh4IN]>0Q=\scriptsize{\begin{bmatrix}h_{1}I_{N}&-h_{2}I_{N}\\ h_{3}I_{N}&h_{4}I_{N}\end{bmatrix}>0}, h5>h4h_{5}>h_{4}, and h6>0h_{6}>0. Computing V˙(t)\dot{V}(t), we obtain

V˙(t)=\displaystyle\dot{V}(t)= 12(h3h2)y𝖳Jy(y^,t)εy(y^)\displaystyle-\frac{1}{2}(h_{3}-h_{2})y^{{\sf\tiny T}}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})
+h5ε𝖳(y^)Jy(y^,t)αy(t)yh6ϕBσ(t)Jz(z^,t)εz(z^)2\displaystyle+h_{5}\varepsilon^{{\sf\tiny T}}(\hat{y})J_{y}(\hat{y},t)\alpha_{y}(t)y-h_{6}\phi\|B_{\sigma(t)}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})\|^{2}
h4ϕz𝖳Jz(z^,t)εz(z^)+h6εz𝖳(z^)Jz(z^,t)αz(t)z\displaystyle-h_{4}\phi z^{{\sf\tiny T}}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})+h_{6}\varepsilon_{z}^{{\sf\tiny T}}(\hat{z})J_{z}(\hat{z},t)\alpha_{z}(t)z
+h1v𝖳xϕ2(h3h2)y𝖳Jz(z^,t)εz(z^)\displaystyle+h_{1}v^{{\sf\tiny T}}x-\frac{\phi}{2}(h_{3}-h_{2})y^{{\sf\tiny T}}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})
+(h5h4)z𝖳Jy(y^,t)εy(y^)+12(h3h2)v𝖳v\displaystyle+(h_{5}-h_{4})z^{{\sf\tiny T}}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})+\frac{1}{2}(h_{3}-h_{2})v^{{\sf\tiny T}}v
h6εz𝖳(z^)Jz(z^,t)Bσ(t)𝖳Bσ(t)Jy(y^,t)εy(y^).\displaystyle-h_{6}\varepsilon_{z}^{{\sf\tiny T}}(\hat{z})J_{z}(\hat{z},t)B_{\sigma(t)}^{{\sf\tiny T}}B_{\sigma(t)}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y}).

By the Young’s inequality [14], with positive constants a1,a2,a3,a4a_{1},a_{2},a_{3},a_{4} and a5a_{5}, one has

v𝖳xa1x𝖳x+14a1v𝖳v,\displaystyle v^{{\sf\tiny T}}x\leq a_{1}x^{{\sf\tiny T}}x+\frac{1}{4a_{1}}v^{{\sf\tiny T}}v,
y𝖳Jz(z^,t)εz(z^)a2Bσ(t)Jz(z^,t)εz(z^)2+14a2x𝖳x,\displaystyle y^{{\sf\tiny T}}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})\leq a_{2}\|B_{\sigma(t)}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})\|^{2}+\frac{1}{4a_{2}}x^{{\sf\tiny T}}x,
z𝖳Jy(y^,t)εy(y^)a3Bσ(t)Jy(y^,t)εy(y^)2+14a3v𝖳v,\displaystyle z^{{\sf\tiny T}}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})\leq a_{3}\|B_{\sigma(t)}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})\|^{2}+\frac{1}{4a_{3}}v^{{\sf\tiny T}}v,
εz𝖳(z^)Jz(z^,t)Bσ(t)𝖳Bσ(t)Jy(y^,t)εy(y^)\displaystyle\varepsilon_{z}^{{\sf\tiny T}}(\hat{z})J_{z}(\hat{z},t)B_{\sigma(t)}^{{\sf\tiny T}}B_{\sigma(t)}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})
a4Bσ(t)Jz(z^,t)εz𝖳(z^)2+14a4Bσ(t)Jy(y^,t)εy(y^)2.\displaystyle\quad\leq a_{4}\|B_{\sigma(t)}J_{z}(\hat{z},t)\varepsilon_{z}^{{\sf\tiny T}}(\hat{z})\|^{2}+\frac{1}{4a_{4}}\|B_{\sigma(t)}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})\|^{2}.

Combining the above inequalities and inequality (4) gives

V˙(t)\displaystyle\dot{V}(t)\leq (h3h22h5α¯y)y𝖳Jy(y^,t)εy(y^)\displaystyle-(\frac{h_{3}-h_{2}}{2}-h_{5}\bar{\alpha}_{y})y^{{\sf\tiny T}}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})
(h6ϕa2ϕ(h3h2)2h6a4)Bσ(t)Jz(z^,t)εz(z^)2\displaystyle-(h_{6}\phi-\frac{a_{2}\phi(h_{3}-h_{2})}{2}-h_{6}a_{4})\|B_{\sigma(t)}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})\|^{2}
(h4ϕh6α¯z)z𝖳Jz(z^,t)εz(z^)\displaystyle-(h_{4}\phi-h_{6}\bar{\alpha}_{z})z^{{\sf\tiny T}}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})
+(h1a1+ϕ(h3h2)8a2)x𝖳x\displaystyle+(h_{1}a_{1}+\frac{\phi(h_{3}-h_{2})}{8a_{2}})x^{{\sf\tiny T}}x
+(h14a1+(h5h4)4a3+(h3h2)2)v𝖳v\displaystyle+(\frac{h_{1}}{4a_{1}}+\frac{(h_{5}-h_{4})}{4a_{3}}+\frac{(h_{3}-h_{2})}{2})v^{{\sf\tiny T}}v
+((h5h4)a3+h64a4)Bσ(t)Jy(y^,t)εy(y^)2.\displaystyle+((h_{5}-h_{4})a_{3}+\frac{h_{6}}{4a_{4}})\|B_{\sigma(t)}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})\|^{2}.

By feat of Lemma 3, x𝖳xx^{{\sf\tiny T}}x and v𝖳vv^{{\sf\tiny T}}v are bounded and the upper bound are supposed as D1D_{1} and D2D_{2} respectively. On account of the boundedness of y(t)y(t) for t[0,τmax)t\in[0,\tau_{\max}), Bσ(t)Jy(y^,t)εy(y^)2\|B_{\sigma(t)}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})\|^{2} is bounded and its upper bound is recorded as D3D_{3}. With h6ϕa2ϕ2(h3h2)h6a4>0h_{6}\phi-\frac{a_{2}\phi}{2}(h_{3}-h_{2})-h_{6}a_{4}>0, one has

V˙(t)\displaystyle\dot{V}(t)\leq κ1y𝖳Jy(y^,t)εy(y^)+κ3D1+κ4D2\displaystyle-\kappa_{1}y^{{\sf\tiny T}}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})+\kappa_{3}D_{1}+\kappa_{4}D_{2}
κ2z𝖳Jz(z^,t)εz(z^)+κ5D3\displaystyle-\kappa_{2}z^{{\sf\tiny T}}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})+\kappa_{5}D_{3}

where κ1=12(h3h2)h5α¯y\kappa_{1}=\frac{1}{2}(h_{3}-h_{2})-h_{5}\bar{\alpha}_{y}, κ2=h4ϕh6α¯z\kappa_{2}=h_{4}\phi-h_{6}\bar{\alpha}_{z}, κ3=h1a1+ϕ(h3h2)8a2\kappa_{3}=h_{1}a_{1}+\frac{\phi(h_{3}-h_{2})}{8a_{2}}, κ4=h14a1+(h5h4)4a3+(h3h2)2\kappa_{4}=\frac{h_{1}}{4a_{1}}+\frac{(h_{5}-h_{4})}{4a_{3}}+\frac{(h_{3}-h_{2})}{2}, and κ5=(h5h4)a3+h64a4\kappa_{5}=(h_{5}-h_{4})a_{3}+\frac{h_{6}}{4a_{4}}. By Lemma 1, the following holds

κ12y𝖳Jy(y^,t)εy(y^)+κ3D1+κ4D2\displaystyle-\frac{\kappa_{1}}{2}y^{{\sf\tiny T}}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})+\kappa_{3}D_{1}+\kappa_{4}D_{2} 0,\displaystyle\leq 0,
κ22z𝖳Jz(z^,t)εz(z^)+κ5D3\displaystyle-\frac{\kappa_{2}}{2}z^{{\sf\tiny T}}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})+\kappa_{5}D_{3} 0,\displaystyle\leq 0,

for all εy(y^)>ε¯y\|\varepsilon_{y}(\hat{y})\|>\bar{\varepsilon}_{y} and εz(z^)>ε¯z\|\varepsilon_{z}(\hat{z})\|>\bar{\varepsilon}_{z}. By applying the inequality (5), we have

V˙(t)\displaystyle\dot{V}(t)\leq κ12y𝖳Jy(y^,t)εy(y^)κ22z𝖳Jz(z^,t)εz(z^)\displaystyle-\frac{\kappa_{1}}{2}y^{{\sf\tiny T}}J_{y}(\hat{y},t)\varepsilon_{y}(\hat{y})-\frac{\kappa_{2}}{2}z^{{\sf\tiny T}}J_{z}(\hat{z},t)\varepsilon_{z}(\hat{z})
\displaystyle\leq κ1ζy,12εy(y^)2κ2ζz,12εz(z^)2\displaystyle-\frac{\kappa_{1}\zeta_{y,1}}{2}\|\varepsilon_{y}(\hat{y})\|^{2}-\frac{\kappa_{2}\zeta_{z,1}}{2}\|\varepsilon_{z}(\hat{z})\|^{2}
\displaystyle\leq 0.\displaystyle 0. (12)

It can be seen that V˙(t)0\dot{V}(t)\leq 0 is guaranteed by

|εy,l(y^)|\displaystyle|\varepsilon_{y,l}(\hat{y})| εy=max{εy(y^(0)),ε¯y},\displaystyle\leq\varepsilon_{y}^{*}=\max\{\varepsilon_{y}(\hat{y}(0)),\bar{\varepsilon}_{y}\},
|εz,l(y^)|\displaystyle|\varepsilon_{z,l}(\hat{y})| εz=max{εz(z^(0)),ε¯z},lσ(t)\displaystyle\leq\varepsilon_{z}^{*}=\max\{\varepsilon_{z}(\hat{z}(0)),\bar{\varepsilon}_{z}\},\forall l\in\mathcal{H}_{\sigma(t)}

within t[0,tmax)t\in[0,t_{max}). By using the inverse operation of the transformation function εs,l(s^l)\varepsilon_{s,l}(\hat{s}_{l}), we obtain, with t[0,tmax)t\in[0,t_{max}),

y^l(t)\displaystyle\hat{y}_{l}(t) [δ¯y,l,δ¯y,l][εy,l1(εy),εy,l1(εy)],\displaystyle\in[\underline{\delta}_{y,l},\bar{\delta}_{y,l}]\triangleq[-\varepsilon^{-1}_{y,l}(\varepsilon_{y}^{*}),\varepsilon^{-1}_{y,l}(\varepsilon_{y}^{*})],
z^l(t)\displaystyle\hat{z}_{l}(t) [δ¯z,l,δ¯z,l][εz,l1(εz),εz,l1(εz)].\displaystyle\in[\underline{\delta}_{z,l},\bar{\delta}_{z,l}]\triangleq[-\varepsilon^{-1}_{z,l}(\varepsilon_{z}^{*}),\varepsilon^{-1}_{z,l}(\varepsilon_{z}^{*})].

Phase III. Up to this point, we prove that τmax\tau_{\max} can be extended to \infty. It is obtained from (III) that

w(t)D=Dy^×Dz^,t[0,τmax)\displaystyle w(t)\in D^{{}^{\prime}}=D_{\hat{y}}^{{}^{\prime}}\times D_{\hat{z}}^{{}^{\prime}},~{}\forall t\in[0,\tau_{\max})

where Dy^=[δ¯y1,δ¯y1]××[δ¯yM,δ¯yM]D_{\hat{y}}^{{}^{\prime}}=[\underline{\delta}_{y_{1}},\bar{\delta}_{y_{1}}]\times\cdots\times[\underline{\delta}_{y_{M}},\bar{\delta}_{y_{M}}] and Dz^=[δ¯z1,δ¯z1]××[δ¯zM,δ¯zM]D_{\hat{z}}^{{}^{\prime}}=[\underline{\delta}_{z_{1}},\bar{\delta}_{z_{1}}]\times\cdots\times[\underline{\delta}_{z_{M}},\bar{\delta}_{z_{M}}]. DD^{{}^{\prime}} is obviously a nonempty subset of DD. Suppose τmax<\tau_{\max}<\infty, it is inferred from Proposition C.3.6 of [22] that the inequality w(t)Dw(t^{{}^{\prime}})\notin D^{{}^{\prime}} holds for a instant t[0,τmax)t^{{}^{\prime}}\in[0,\tau_{\max}), which is a clear contradiction. From Theorem 11 of [19], one obtain τmax=\tau_{\max}=\infty and w(t)DDw(t)\in D^{{}^{\prime}}\subset D, thus V(t)V(t) is finite and bounded, t>0\forall t>0.

Phase IV. We finally prove that y(t)y(t) and z(t)z(t) converge to zero as tt\to\infty. Taking a part of calculation in (III) gives

V˙(t)κ¯εy(y^)2κ¯εz(z^)2\displaystyle\dot{V}(t)\leq-\underline{\kappa}\|\varepsilon_{y}(\hat{y})\|^{2}-\underline{\kappa}\|\varepsilon_{z}(\hat{z})\|^{2} (13)

where κ¯=min{κ1ζ12,κ2ζ32}\underline{\kappa}={\min}\{\frac{\kappa_{1}\zeta_{1}}{2},\frac{\kappa_{2}\zeta_{3}}{2}\}. Since V˙(t)0\dot{V}(t)\leq 0 and V(t)>0V(t)>0, V(t)V(t) is bounded and limtV(t)\lim_{t\to\infty}V(t) exists. Given an infinite sequences V(tk)V(t_{k}), kk\in\mathbb{N}, by utilizing Cauchy’s Convergence Criterion, one has, for ϵ>0\forall\epsilon>0, Mϵ𝒵+\exists M_{\epsilon}\in\mathcal{Z}_{+}, such that kMϵ\forall k\geq M_{\epsilon},

|V(tk+1)V(tk)|<ϵor|tktk+1V˙(s)𝑑s|<ϵ.\displaystyle|V(t_{k+1})-V(t_{k})|<\epsilon~{}{\rm or}~{}\Big{|}\int_{t_{k}}^{t_{k+1}}\dot{V}(s)ds\Big{|}<\epsilon. (14)

It follows that

tk0tk1V˙(s)ds++tkmk1tkmkV˙(s)ds<ϵ.\displaystyle\int_{t_{k}^{0}}^{t_{k}^{1}}-\dot{V}(s)ds+\cdots+\int_{t_{k}^{m_{k}-1}}^{t_{k}^{m_{k}}}-\dot{V}(s)ds<\epsilon. (15)

For each subinterval [tkq,tkq+1)[{t_{k}^{q}},{t_{k}^{q+1}}), q=0,1,,mk1q=0,1,\cdots,m_{k}-1, one has

tkqtkq+1V˙(s)ds\displaystyle\int_{t_{k}^{q}}^{t_{k}^{q+1}}-\dot{V}(s)ds κ¯tkqtkjq+1εy(y^(s))2+εz(z^(s))2ds\displaystyle\geq\underline{\kappa}\int_{t_{k}^{q}}^{t_{k}^{jq+1}}\|\varepsilon_{y}(\hat{y}(s))\|^{2}+\|\varepsilon_{z}(\hat{z}(s))\|^{2}ds
κ¯tkqtkq+τεy(y^(s))2+εz(z^(s))2ds.\displaystyle\geq\underline{\kappa}\int_{t_{k}^{q}}^{t_{k}^{q}+\tau}\|\varepsilon_{y}(\hat{y}(s))\|^{2}+\|\varepsilon_{z}(\hat{z}(s))\|^{2}ds.

Combining the above two inequalities gives

ϵ>\displaystyle\epsilon> κ¯q=0mk1tkqtkq+τεy(y^(s))2+εz(z^(s))2ds.\displaystyle\underline{\kappa}\sum_{q=0}^{m_{k}-1}\int_{t_{k}^{q}}^{t_{k}^{q}+\tau}\|\varepsilon_{y}(\hat{y}(s))\|^{2}+\|\varepsilon_{z}(\hat{z}(s))\|^{2}ds.

Since finite switches take place during [tk,tk+1)[t_{k},t_{k+1}), the number mkm_{k} is finite for kk\in\mathbb{N}. It implies that, with a bounded constant Mk>0M_{k}>0,

limttt+τMk(εy(y^(s))2+εz(z^(s))2)𝑑s=0.\displaystyle\lim_{t\to\infty}\int_{t}^{t+\tau}M_{k}(\|\varepsilon_{y}(\hat{y}(s))\|^{2}+\|\varepsilon_{z}(\hat{z}(s))\|^{2})ds=0.

Since εy(y^(t))\varepsilon_{y}(\hat{y}(t)) and εz(z^(t))\varepsilon_{z}(\hat{z}(t)) are bounded, V˙(t)\dot{V}(t) and V¨(t)\ddot{V}(t) are bounded, t>0\forall t>0. Utilizing the Barbalat’s Lemma [16] gives limtεy(y^(t))=0\lim_{t\to\infty}\varepsilon_{y}(\hat{y}(t))=0 and limtεz(z^(t))=0\lim_{t\to\infty}\varepsilon_{z}(\hat{z}(t))=0, and consequently, limty(t)=0\lim_{t\to\infty}y(t)=0 and limtz(t)=0\lim_{t\to\infty}z(t)=0. The proof is completed. ∎

Remark 2

Generally speaking, the controller ui(t)u_{i}(t) is updated based on the connected agents’ state feedback. As the graph switches, the neighbor set of agent will change. By applying the control protocol (III), the consensus control problem caused by switching topologies is solved. If the graph 𝒢σ(t)\mathcal{G}_{\sigma(t)} is a tree, the edge Laplacian matrix Bσ(t)𝖳Bσ(t)B_{\sigma(t)}^{{\sf\tiny T}}B_{\sigma(t)} is positive definite. However, this positive definiteness condition is not required in the proof, such that protocol (III) is applicable to generally connected graphs containing cycles.

Remark 3

From Theorem 1, it follows that asymptotic consensus is achieved with transient performance, which is more accurate than practical consensus in [12]. Besides, asymptotic convergence can restrain the external disturbances and bound fluctuations caused by performance bounds [14, 20].

Refer to caption
Figure 1: Topology graphs. (a) 𝒢¯1\bar{\mathcal{G}}_{1}. (b) 𝒢¯2\bar{\mathcal{G}}_{2}. (c) 𝒢¯3\bar{\mathcal{G}}_{3}.

IV Simulation

Consider a MAS composed of N=5N=5 agents with the dynamic model (6) and (III). Suppose the topology graph described in Fig.1 switches as 𝒢¯1𝒢¯2𝒢¯3𝒢¯1\bar{\mathcal{G}}_{1}\to\bar{\mathcal{G}}_{2}\to\bar{\mathcal{G}}_{3}\to\bar{\mathcal{G}}_{1}\to\cdots, with the dwelling time τ=0.1\tau=0.1 s. We choose performance functions ρy,l(t)=(50.1)e1.5t+0.1\rho_{y,l}(t)=(5-0.1)e^{-1.5t}+0.1 and ρz,l(t)=(50.1)e0.8t+0.1\rho_{z,l}(t)=(5-0.1)e^{-0.8t}+0.1 for all edges. Then we have αy,l(t)=ρ˙y,l(t)ρy,l(t)=1.5×ρy,l(t)0.1ρy,l(t)1.5\alpha_{y,l}(t)=-\frac{\dot{\rho}_{y,l}(t)}{\rho_{y,l}(t)}=1.5\times\frac{\rho_{y,l}(t)-0.1}{\rho_{y,l}(t)}\leq 1.5 and αz,l(t)0.8\alpha_{z,l}(t)\leq 0.8, and hence α¯y=1.5\bar{\alpha}_{y}=1.5 and α¯z=0.8\bar{\alpha}_{z}=0.8. For all edges, the same transformation functions are selected as

εy,l(y^l)=ln(1+y^l1y^l)andεz,l(z^l)=ln(1+z^l1z^l).\displaystyle\varepsilon_{y,l}(\hat{y}_{l})=\ln\left(\frac{1+\hat{y}_{l}}{1-\hat{y}_{l}}\right)~{}{\rm and}~{}\varepsilon_{z,l}(\hat{z}_{l})=\ln\left(\frac{1+\hat{z}_{l}}{1-\hat{z}_{l}}\right).

The initial positions of agents are x(0)=[0.5, 1, 2.5, 1.5, 2]𝖳x(0)=[-0.5,\,1,\,2.5,\,1.5,\,2]^{{\sf\tiny T}}, and initial velocities are v(0)=[1.5,0.5,2.5,3,2]𝖳v(0)=[1.5,\,-0.5,\,-2.5,\,-3,\,-2]^{{\sf\tiny T}}. For Theorem 1, we choose h1=10,h2=1,h3=6,h4=1.5,h5=1.6,h6=1.5,a2=0.1,a3=0.5,a4=0.1h_{1}=10,h_{2}=1,h_{3}=6,h_{4}=1.5,h_{5}=1.6,h_{6}=1.5,a_{2}=0.1,a_{3}=0.5,a_{4}=0.1 and ϕ=1\phi=1. The state trajectories of all agents for tt from 0 to 55s, are depicted in Figs. 3 and 3. It follows from Fig. 3 that the relative positions and relative velocities are confined within the performance bounds as expected, and the state consensus of all agents is eventually achieved as shown in Fig. 3.

Refer to caption
Figure 2: Convergence trajectories with prescribed performance. Top: relative positions; bottom: relative velocities.
Refer to caption
Figure 3: State trajectories of all agents. Top: position trajectories; bottom: velocity trajectories.

V Conclusion

This paper addresses the full-state prescribed performance-based consensus problem for double-integrator MASs with jointly connected topologies. A distributed prescribed performance control protocol is proposed such that the relative positions and relative velocities are constrained within the respective performance bounds. The proposed protocol drives all agents to achieve consensus with jointly connected topologies. Future research topics focus on more general agent dynamics.

References

  • [1] X. M. Zhang, Q. L. Han, X. Ge, D. Ding, L. Ding, D. Yue, and C. Peng, “Networked control systems:a survey of trends and techniques,” IEEE/CAA Journal of Automatica Sinica, vol. 007, no. 001, pp. 1–17, 2020.
  • [2] J. Qin, Q. Ma, Y. Shi, and L. Wang, “Recent advances in consensus of multi-agent systems: A brief survey,” IEEE Transactions on Industrial Electronics, vol. 64, no. 6, pp. 4972–4983, 2017.
  • [3] R. Olfati-Saber and R. M. Murray, “Consensus problems in networks of agents with switching topology and time-delays,” IEEE Transactions on Automatic Control, vol. 49, no. 9, pp. 1520–1533, 2004.
  • [4] W. Yu, L. Zhou, X. Yu, J. Lü, and R. Lu, “Consensus in multi-agent systems with second-order dynamics and sampled data,” IEEE Transactions on Industrial Informatics, vol. 9, no. 4, pp. 2137–2146, 2013.
  • [5] J. Zeng, P. Wang, H. Su, and C. Xu, “Second-order consensus for multi-agent systems with various intelligent levels via intermittent sampled-data control,” IEEE Transactions on Circuits and Systems II: Express Briefs, vol. 69, no. 12, pp. 4899–4903, 2022.
  • [6] Y. Ren, Q. Wang, and Z. Duan, “Optimal distributed leader-following consensus of linear multi-agent systems: A dynamic average consensus-based approach,” IEEE Transactions on Circuits and Systems II: Express Briefs, vol. 69, no. 3, pp. 1208–1212, 2021.
  • [7] Y. Sun and S. Peng, “Consensus of discrete-time leader-following linear multi-agent systems under lyapunov-function-based event-triggered mechanism,” IEEE Transactions on Circuits and Systems II: Express Briefs, vol. 70, no. 12, pp. 4409 – 4413, 2023.
  • [8] X. Tan, M. Cao, and J. Cao, “Distributed dynamic event-based control for nonlinear multi-agent systems,” IEEE Transactions on Circuits and Systems II: Express Briefs, vol. 68, no. 2, pp. 687–691, 2020.
  • [9] C. P. Bechlioulis and G. A. Rovithakis, “Robust adaptive control of feedback linearizable MIMO nonlinear systems with prescribed performance,” IEEE Transactions on Automatic Control, vol. 53, no. 9, pp. 2090–2099, 2008.
  • [10] Y. Karayiannidis, D. V. Dimarogonas, and D. Kragic, “Multi-agent average consensus control with prescribed performance guarantees,” in IEEE Conference on Decision & Control, 2012, pp. 2219–2225.
  • [11] L. Macellari, Y. Karayiannidis, and D. V. Dimarogonas, “Multi-agent second order average consensus with prescribed transient behavior,” IEEE Transactions on Automatic Control, vol. 62, no. 10, pp. 5282–5288, 2017.
  • [12] Y. Karayiannidis, D. Papageorgiou, and Z. Doulgeri, “A model-free controller for guaranteed prescribed performance tracking of both robot joint positions and velocities,” IEEE Robotics & Automation Letters, vol. 1, no. 1, pp. 267–273, 2016.
  • [13] F. Chen and D. V. Dimarogonas, “Consensus control for leader-follower multi-agent systems under prescribed performance guarantees,” in 2019 IEEE 58th Conference on Decision and Control.   IEEE, 2019, pp. 4785–4790.
  • [14] ——, “Second order consensus for leader-follower multi-agent systems with prescribed performance,” in IFAC-PapersOnLine, vol. 52, no. 20, 2019, pp. 103–108.
  • [15] Y. Hou and W. Hu, “Event-triggered consensus of second-order multi-agent systems with prescribed performance,” in 2021 China Automation Congress (CAC), 2021, pp. 1146–1150.
  • [16] Y. Hou and B. Cheng, “Event-based HH_{\infty} consensus of double-integrator multi-agent systems: A prescribed performance control approach,” IEEE Transactions on Circuits and Systems II: Express Briefs, early access, November. 15, 2023, doi: 10.1109/TCSII.2023.3332746.
  • [17] B. Cheng, X. Wang, and Z. Li, “Event-triggered consensus of homogeneous and heterogeneous multiagent systems with jointly connected switching topologies,” IEEE Transactions on Cybernetics, vol. 49, no. 12, pp. 4421–4430, 2018.
  • [18] X. Wan, Y. Guo, and X. Wu, “Differentially private consensus for multi-agent systems under switching topology,” IEEE Transactions on Circuits and Systems II: Express Briefs, vol. 70, no. 9, pp. 3499–3503, 2023.
  • [19] C. J. Stamouli, C. P. Bechlioulis, and K. J. Kyriakopoulos, “Robust dynamic average consensus with prescribed transient and steady state performance,” Automatica, vol. 144, p. 110503, 2022.
  • [20] E. Restrepo and D. V. Dimarogonas, “On asymptotic stability of leader–follower multiagent systems under transient constraints,” IEEE Control Systems Letters, vol. 6, pp. 3164–3169, 2022.
  • [21] Y. Hou, W. Hu, J. Li, and T. Huang, “Prescribed performance control for double-integrator multi-agent systems: A unified event-triggered consensus framework,” IEEE Transactions on Circuits and Systems I: Regular Papers, 2023, under review.
  • [22] E. D. Sontag, Mathematical control theory.   London, U. K.: Springer, 1998.