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

Consensus of A Class of Nonlinear Systems with Varying Topology: A Hilbert Metric Approach

Dongjun Wu This work was supported in part by the European Research Council (ERC) through the European Union’s Horizon 2020 Research and Innovation Program under Grant 834142 (ScalableControl). The author is with Department of Automatic Control, Lund University, 22100 Lund, Sweden (email: dongjun.wu@control.lth.se).
Abstract

In this technical note, we introduce a novel approach to studying consensus of continuous-time nonlinear systems with varying topology based on Hilbert metric. We demonstrate that this metric offers significant flexibility in analyzing consensus properties, while effectively handling nonlinearities and time dependencies. Notably, our approach relaxes key technical assumptions from some standard results while yielding stronger conclusions with shorter proofs. This framework provides new insights into nonlinear consensus under varying topology.

Index Terms:
Hilbert metric, nonlinear consensus, varying topology

I Introduction

Consensus under varying topology is a fundamental research question in multi-agent and network dynamical systems [1]. This topic has been extensively studied over the last three decades within the control community. In the seminal work [2], Vicsek et al. first observed that consensus (or coordination, as referred to in their study ) is closely related to the topology of multi-agent systems. Following this observation, lots of efforts have been made to establish the theoretical foundations for consensus with varying topology. Notable contributions include the works of Jadbabaie et. al. [3], Moreau [1], and Ren and Beard [4], among others. These early 21st century studies sparked a surge in research on consensus with varying topology in subsequent years. The theoretical developments in this area have found applications across various domains, including power grids [5], social networks [6], and formation control [7]. More recent advancements can be found in works such as [8], [9], and [10].

Most of the aforementioned works have primarily focused on linear systems. However, nonlinearities are pervasive in multi-agent systems, such as coupled oscillators [11], mobile robots [12] and social networks [6]. For these systems, the tools for analyzing linear consensus are not directly applicable, necessitating the use of nonlinear techniques. Various tools have been introduced to address the problem, including dissipative analysis [13, 14], monotone system theory [15, 16], and non-smooth techniques [17, 18]. For further reviews on this topic, the readers are referred to [19, 20, 21] and the references therein.

As far as we know, the result obtained in [17] stands among the most general ones regarding nonliner consensus with varying topology. Prior to this, Lin et al. [22] studied consensus of a class of nonlinear systems with fixed topology using non-smooth analysis techniques, with invariance principle being key of the proof. They later extended their result to nonlinear systems with switching topology [17], requiring the switching signal to be sufficiently regular (the definition will be provided later). When switching is present, invariance principle can no longer be used. To handle this, a technical result from [23] was employed in [17]. It is worth mentioning that, even though the system in consideration might possibly be smooth, non-smooth techniques had to be used [22, 17]. The result in [17] can also be seen as a nonlinear extension of the work [1], as both works relied on the quasi-strongly connectedness of multi-agent system as a key assumption.

The aim of this note is to relax some of the technical assumptions made in [17] and to provide new understandings for the problem. To achieve this goal, we adopt a novel approach, different from standard practice of utilizing Lyapunov or non-smooth analysis, by leveraging the so-called Hilbert metric to analyze the system dynamics, which proves to be quite successful.

Hilbert metric has already been used to study consensus. For instance, in [24] Hilbert metric was used to further relax and extend the results obtained in [25]; in [26], it was used for consensus in non-commutative spaces. However, to the best of our knowledge, it has not been introduced to study nonlinear consensus with varying topology. This might be related to the fact that the Hilbert metric was considered more suitable for linear systems; e.g., one of the most widely used results – Birkhoff’s Theorem [27] – is applicable for linear systems (or more generally, homogeneous systems). In this note, however, we show that Hilbert metric can also serve as a systematic tool for studying nonlinear consensus. In particular, it performs well in handling nonlinearities and time dependencies.

Before outlining the contributions of this note, it is necessary to briefly recall the main result obtained in [17]. Consider the following multi-agent system with switching topology controlled by the switching signal σ:0{1,,N}\sigma:\mathbb{R}_{\geq 0}\to\{1,\cdots,N\}:

{x˙1=fσ(t)1(x1,,xn)x˙n=fσ(t)n(x1,,xn)\begin{cases}\dot{x}_{1}=f_{\sigma(t)}^{1}(x_{1},\cdots,x_{n})\\ \vdots\\ \dot{x}_{n}=f_{\sigma(t)}^{n}(x_{1},\cdots,x_{n})\end{cases} (1)

where ximx_{i}\in\mathbb{R}^{m}. For each p{1,,N}p\in\{1,\cdots,N\}, we associate a digraph 𝒢p\mathcal{G}_{p} with the system defined by vector fields {fpi}i=1n\{f_{p}^{i}\}_{i=1}^{n}. 𝒢p\mathcal{G}_{p} has nn vertices denoted as {1,,n}\{1,\cdots,n\}, and a link (i,j)(i,j) is in 𝒢p\mathcal{G}_{p} if fpif_{p}^{i} depends explicitly on xjx_{j}. The digraph is called quasi strongly connected (QSC) if there exists a node kk such that for each node jkj\neq k, there is a directed path joining node kk to node jj. Such a node is called a center. In other words, a digraph is QSC if there is a directed spanning tree, and the root of the tree is a center. For a switching graph corresponding to the system (1), we say the graph is uniformly quasi strongly connected (UQSC) if there exists a constant T>0T>0 such that the union of the graph on the interval [t,t+T][t,t+T] for any tt is QSC. Let 𝒞i\mathcal{C}_{i} be the polytope formed by xix_{i} and its neighboring agents, and Txi𝒞iT_{x_{i}}\mathcal{C}_{i} the tangent cone of 𝒞i\mathcal{C}_{i} at xix_{i} in m\mathbb{R}^{m}, see [17] for more detailed explanations of these terminologies.

One of the main results in [17] can be restated as:

Theorem 1 (Lin et al. [17]).

Consider the system (1). Assume that for each p{1,,N}p\in\{1,\cdots,{N}\}: 1) fpif_{p}^{i} is locally Lipschitz and fpif_{p}^{i} is in the relative interior of the cone Txi𝒞iT_{x_{i}}\mathcal{C}_{i}, i.e., fpiri(Txi𝒞i)f_{p}^{i}\in{\rm ri}(T_{x_{i}}\mathcal{C}_{i}); 2) the switching signal is piece-wise constant with minimum dwell time τD\tau_{D}; then the system achieves global consensus if and only if the system (1) is UQSC.

A switching signal satisfying the assumptions (having a minimum dwell time τD\tau_{D}) in Theorem 1 is said to be regular.

Note that under the assumption of Theorem 1, the vector field fpif_{p}^{i} can be written as fip=aijp(x)(xjxi)f_{i}^{p}=\sum a_{ij}^{p}(x)(x_{j}-x_{i}) for some non-negative scalar functions aijpa_{ij}^{p}. Equivalently, this means that the system can be written as x˙=Ap(x)x\dot{x}=A_{p}(x)x with (Ap(x))ij=aijp(x)(A_{p}(x))_{ij}=a_{ij}^{p}(x), in which the matrix Ap(x)A_{p}(x) is Metzler, that is, Aij0A_{ij}\geq 0 for all iji\neq j, and each row sums to zero. System of the form x˙=A(x)x\dot{x}=A(x)x with A(x)A(x) being Metzler has been recently considered by Kawano and Cao [28], where such systems were referred to as “virtually positive”. Due to the special structure x˙=Aσ(t)(x)x\dot{x}=A_{\sigma(t)}(x)x, consensus of such systems shares a lot in common with linear time varying multi-agent systems. A remarkable result concerning the latter was obtained by Moreau in [25]:

Theorem 2 (Moreau [25]).

Consider the LTV system x˙=A(t)x\dot{x}=A(t)x. Assume that A()A(\cdot) is uniformly bounded and piecewise continuous. Assume that, for every time tt, A(t)=(aij(t))n×nA(t)=(a_{ij}(t))\in\mathbb{R}^{n\times n} is Metzler with zero row sums. If there exists an index k{1,,n}k\in\{1,\cdots,n\}, a threshold value δ>0\delta>0 and an interval length T>0T>0 such that for all t0t\geq 0,

tT+taik(s)𝑑sδ,i{1,,n}\{k},\int_{t}^{T+t}a_{ik}(s)ds\geq\delta,\quad\forall i\in\{1,\cdots,n\}\backslash\{k\}, (2)

then the system achieves exponential consensus.

The proof of Theorem 2 provided in [25] was based on Lyapunov analysis, see also [3, 1] for discrete time versions. In particular, separable Lyapunov functions were used. This is a technique commonly used in monotone systems, see for example [29, 30, 31, 32].

Theorem 1 and Theorem 2 and share the same spirit. Nevertheless, their proof techniques were quite different: the former relies on on non-smooth analysis while the latter on Lyapunov analysis. It is then a question whether the two results can be understood in a unifying framework. This has remained an open question. In this note, we provide an affirmative answer to it.

We address the problem by a non-Lyapunov-based method, i.e., through the analysis of the evolution of the system dynamics under the Hilbert metric. The contributions of the note are twofold:

  1. 1.

    Utilizing Hilbert metric to analyze nonlinear consensus is new. It does not rely on Lyapunov [25] or non-smooth analysis [17]. Our proof based on Hilbert metric for consensus is largely simplified compared to existing works. This has led to new understandings and insights of nonlinear consensus.

  2. 2.

    Instead of regular switching topology considered in [17], we are able to analyze more general time varying topologies, requiring only measurability of the switching. This includes cases such as piece-wise continuous switching studied in [25]. Consequently, our results extend classical findings, including Theorem 1 and 2, by relaxing key technical assumptions. Furthermore, we obtain stronger results; for instance, while [17] establishes only asymptotic consensus, our analysis demonstrates exponential consensus under mild additional assumptions.

Organization of the paper: In Section II, we provide the problem setting and prove some technical results. In particular, we give some explicit formulas and new definitions which will be crucial for the next section. Section III contains the main results of this note and Section IV demonstrate a simulation result.

Notations: |||\cdot| stands for Euclidean 22-norm. For a dynamical system, use ϕ(t,t0,x0)\phi(t,t_{0},x_{0}) to represent the solution at tt from initial state (t0,x0)(t_{0},x_{0}). The interior of a set SS is denoted IntS{\rm Int}S. Being XX, YY some topological spaces, denote C(X,Y)C(X,Y) the space of continuous maps from XX to YY. Given a set SXS\subseteq X, the indicator function 1S:X{0,1}1_{S}:X\to\{0,1\} is defined to be 1S(x)=11_{S}(x)=1 if xSx\in S and 0 otherwise. For a Metzler matrix AA with row sum zero, 𝒢A\mathcal{G}^{A} represents the graph associated with AA. Given x,yx,y, d(x,y)d(x,y) stands for the Hilbert metric between xx and yy. +\mathbb{N}_{+} is the set of positive natural numbers. 𝟙n\mathds{1}_{n} a column vector of dimension nn with all ones. A continuous function β:[0,a)×[0,)[0,)\beta:[0,a)\times[0,\infty)\to[0,\infty) is called a class 𝒦\mathcal{KL} function if 1) rβ(r,s)r\mapsto\beta(r,s) is strictly increasing and β(0,s)0\beta(0,s)\equiv 0 for all s0s\geq 0; 2) sβ(r,s)s\mapsto\beta(r,s) is decreases to 0 as ss\to\infty.

II Preliminary Results

II-A Problem setting

We consider continuous time multi-agent nonlinear systems of the form

x˙i=j=1naij(t,x)(xjxi),\dot{x}_{i}=\sum_{j=1}^{n}a_{ij}(t,x)(x_{j}-x_{i}), (3)

for i=1,,ni=1,\cdots,n, where the state of each agent is in \mathbb{R} and taij(t,)t\mapsto a_{ij}(t,\cdot) is measurable; in addition, aij(t,x)0a_{ij}(t,x)\geq 0 for all t0t\geq 0, xnx\in\mathbb{R}^{n} and i,j{1,,n}i,j\in\{1,\cdots,n\}, iji\neq j.

Remark 1.

We can also consider more general systems having the following form

x˙i=a(t)xi+b(t)+j=1naij(t,x)(xjxi)\dot{x}_{i}=a(t)x_{i}+b(t)+\sum_{j=1}^{n}a_{ij}(t,x)(x_{j}-x_{i}) (4)

where a()a(\cdot) and b()b(\cdot) are bounded on +\mathbb{R}_{+}. Indeed, define y(t)=e0ta(s)dsx(0te0τa(s)dsb(τ)dτ)𝟙y(t)=e^{\int_{0}^{t}a(s){\rm d}s}x-\left(\int_{0}^{t}e^{\int_{0}^{\tau}a(s){\rm d}s}b(\tau){\rm d}\tau\right)\mathds{1} we have y˙=A~(t,y)y\dot{y}=\tilde{A}(t,y)y where A~\tilde{A} is Metzler and satisfies A~𝟙=0\tilde{A}\mathds{1}=0.

We mention a few practical examples that can be written as (3).

  • Kuramoto oscillators: This is a widely studied nonlinear consensus model with applications in various engineering domains. Consider a network of Kuramoto oscillators with time-varying and state-dependent coupling:

    θ˙i=ωi(t)+jkij(t,θ)sin(θiθj)\dot{\theta}_{i}=\omega_{i}(t)+\sum_{j}k_{ij}(t,\theta)\sin(\theta_{i}-\theta_{j})

    where ωi(t)\omega_{i}(t) are the natural frequencies, and kij(t,θ)k_{ij}(t,\theta) are non-negative coupling functions. Phase synchronization occurs when |θi(t)θj(t)|0|\theta_{i}(t)-\theta_{j}(t)|\to 0 as tt\to\infty. This can only be achieved if all natural frequencies are identical. In this case, the model can be transformed into the standard form (3) by defining aij(t,θ):=kij(t,θ)sin(θiθj)θiθja_{ij}(t,\theta):=k_{ij}(t,\theta)\frac{\sin(\theta_{i}-\theta_{j})}{\theta_{i}-\theta_{j}}. Generalizations of the Kuramoto oscillator that can be expressed in the form of (3) can be found in [21].

  • Cucker-Smale model: This model is commonly used to describe flocking behavior [33]:

    x˙i\displaystyle\dot{x}_{i} =vi\displaystyle=v_{i}
    v˙i\displaystyle\dot{v}_{i} =λNj=1Nψij(x,v)(vjvi)\displaystyle=\frac{\lambda}{N}\sum_{j=1}^{N}\psi_{ij}(x,v)(v_{j}-v_{i})

    where NN is the number of agents, ψij(x,v)0\psi_{ij}(x,v)\geq 0 represents the interaction force, and λ>0\lambda>0 is a constant. The system achieves flocking if vi(t)vj(t)0v_{i}(t)-v_{j}(t)\to 0 for all i,ji,j. The second equation can clearly be written in the form of (3) and thus analyzed using the proposed methods.

  • Hegselmann–Krause model: This model is widely used for opinion dynamics [6]. A time-varying version can be written as

    x˙i=j=1Nϕij(t,xi,xj)(xixj)\dot{x}_{i}=\sum_{j=1}^{N}\phi_{ij}(t,x_{i},x_{j})(x_{i}-x_{j})

    where

    ϕij(t,xi,xj)={1,|xixj|ϵ(t)0,otherwise\phi_{ij}(t,x_{i},x_{j})=\begin{cases}1,&|x_{i}-x_{j}|\leq\epsilon(t)\\ 0,&\text{otherwise}\end{cases}

    and ϵ:+[0,1]\epsilon:\mathbb{R}_{+}\to[0,1] is a measurable function.

  • Animal group models: The following is widely used to simulate animal group behavior (see, e.g., [34]):

    x˙i=jAi(t)ϕa(xi,xj)|xixj|(xjxi)+jRi(t)ϕr(xi,xj)|xixj|(xixj)\dot{x}_{i}=\sum_{j\in A_{i}(t)}\frac{\phi_{a}(x_{i},x_{j})}{|x_{i}-x_{j}|}(x_{j}-x_{i})+\sum_{j\in R_{i}(t)}\frac{\phi_{r}(x_{i},x_{j})}{|x_{i}-x_{j}|}(x_{i}-x_{j})

    where AiA_{i} and ϕa\phi_{a} represent attraction, and RiR_{i} and ϕr\phi_{r} represent repulsion. Both ϕa\phi_{a} and ϕr\phi_{r} are non-negative. This model also has the form (3).

The system (3) can also be written in matrix form as

x˙=A(t,x)x\dot{x}=A(t,x)x (5)

in which A(t,x)ij=aij(t,x)A(t,x)_{ij}=a_{ij}(t,x) for iji\neq j and A(t,x)ii=jiaij(t,x)A(t,x)_{ii}=-\sum_{j\neq i}a_{ij}(t,x). Note that A(t,x)A(t,x) is Metzler and has the property A(t,x)𝟙=0A(t,x)\mathds{1}=0.

Associated with the system (5) is a varying digraph 𝒢A(t,x)\mathcal{G}^{A(t,x)} understood in the following sense: for iji\neq j, if aij(t,x)>0a_{ij}(t,x)>0, then (i,j)(i,j) is a directed link from node ii to jj and the weight on this link is aij(t,x)a_{ij}(t,x); if aij(t,x)=0a_{ij}(t,x)=0, then there is no link from ii to jj. Therefore, for any Metzler matrix AA, there is an associated graph. Note that, we do not consider self loop, i.e., the graph 𝒢A\mathcal{G}^{A} is characterized only by the off-diagonal elements of AA. The following definition collects a few important notions that we will use frequently in the paper.

Definition 1.

Let A,B,C(z)n×n,zZA,B,C(z)\in\mathbb{R}^{n\times n},\;z\in Z be some Metzler matrices with row sum zero, ZZ some index set,

  1. 1.

    We denote 𝒢A\mathcal{G}^{A} the graph associated with the matrix AA.

  2. 2.

    We say that 𝒢A𝒢B\mathcal{G}^{A}\geq\mathcal{G}^{B} if AijBijA_{ij}\geq B_{ij} for all iji\neq j.

  3. 3.

    We say that the graph valued function z𝒢C(z)z\mapsto\mathcal{G}^{C(z)} is continuous if zC(z)z\mapsto C(z) is continuous.

  4. 4.

    The graph 𝒢A\mathcal{G}^{A} is said to be quasi-strongly connected (QSC) if k+\exists k\in\mathbb{N}_{+}, such that Aik>0A_{ik}>0 for all iki\neq k; it is called δ\delta-connected for some δ>0\delta>0, if k+\exists k\in\mathbb{N}_{+}, such that AikδA_{ik}\geq\delta for all iki\neq k.

In this note, consensus is understood in the following sense:

Definition 2.

Given a forward invariant set DnD\subseteq\mathbb{R}^{n}, the system (1) is said to achieve

  • asymptotic consensus on DD if there exists a class 𝒦\mathcal{KL} function β\beta, such that |xi(t)xj(t)|β(|xi(0)xj(0)|,t)|x_{i}(t)-x_{j}(t)|\leq\beta(|x_{i}(0)-x_{j}(0)|,t) for all i,j{1,,n}i,j\in\{1,\cdots,n\}, t0t\geq 0 and x(0)Dx(0)\in D;

  • exponential consensus on DD if there exist some constants k,λ>0k,\lambda>0 such that |xi(t)xj(t)|keλt|xi(0)xj(0)||x_{i}(t)-x_{j}(t)|\leq ke^{-\lambda t}|x_{i}(0)-x_{j}(0)| for all i,j{1,,n}i,j\in\{1,\cdots,n\}, t0t\geq 0 and x(0)Dx(0)\in D.

Remark 2.

For any a<ba<b, [a,b]n[a,b]^{n} is invariant under the system flow (3). Thus we may assume DD in Definition 2 is compact.

The following regularity condition will be imposed on A(t,x)A(t,x) throughout the paper unless otherwise stated.

Assumption A1.

Assume, for any compact set DnD\subseteq\mathbb{R}^{n},

  1. 1.

    the mapping tA(t,x)t\mapsto A(t,x) is measurable for every fixed xDx\in D;

  2. 2.

    there is a locally integrable function kD(t)k_{D}(t) such that

    |A(t,x)xA(t,y)y|kD(t)|xy||A(t,x)x-A(t,y)y|\leq k_{D}(t)|x-y|

    for all x,yD,t0x,y\in D,\;t\geq 0.

  3. 3.

    there exists a constant CD>0C_{D}>0, such that |A(t,x)|CD|A(t,x)|\leq C_{D} for every xD,t0x\in D,\;t\geq 0.

Under Assumption A1, the solution to system (5) exists and is unique through every point (t0,x0)+×D(t_{0},x_{0})\in\mathbb{R}_{+}\times D [35, Theorem 5.1, Theorem 5.3].

II-B The Hilbert metric

In this paper, we use Hilbert metric to analyze consensus problems. Hilbert metric is a metric defined on cones. More precisely, in our setting, a cone is some closed subset KnK\subseteq\mathbb{R}^{n} satisfying the following four properties 1) The interior of KK is non-empty; 2) For v,wKv,w\in K, v+wv+w is also in KK. 3) For all λ0\lambda\geq 0, and vKv\in K, λv\lambda v is also in KK. 4) KK={0}K\cap-K=\{0\}, where K:={x:xK}-K:=\{-x:x\in K\}. Given a cone, we can define a partial ordering as xyx\leq y if yxKy-x\in K and x<yx<y if yxIntKy-x\in{\rm Int}\;K. For x,yIntKx,y\in{\rm Int}K, define two numbers M(x/y)=inf{λ:xλy}M(x/y)=\inf\{\lambda:x\leq\lambda y\} or \infty if the set is empty, and m(x/y)=sup{μ:μyx}m(x/y)=\sup\{\mu:\mu y\leq x\}. Then the Hilbert metric between xx, yy is defined as d(x,y)=lnM(x,y)m(x,y)d(x,y)=\ln\frac{M(x,y)}{m(x,y)}. Define the diameter of a set SKS\subseteq K as diam(S)=supx,ySd(x,y).{\rm diam}(S)=\sup_{x,y\in S}d(x,y). For any cone SInt+n{0}S\subseteq{\rm Int}\mathbb{R}_{+}^{n}\cup\{0\}, we have diamS<+{\rm diam}S<+\infty. In fact, d(x,y)=lnmaxi(xi/yi)mini(xi/yi)d(x,y)=\ln\frac{\max_{i}(x_{i}/y_{i})}{\min_{i}(x_{i}/y_{i})}, where x=(x1,,xn)x=(x_{1},\cdots,x_{n}), y=(y1,,yn)y=(y_{1},\cdots,y_{n}), which is bounded on SS. A mapping A:KKA:K\to K on a cone is called non-negative. If in addition, AA maps the interior of KK into its interior, we call AA a positive mapping. For example, when K=+nK=\mathbb{R}_{+}^{n}, then a non-negative matrix AA represents a non-negative mapping while a positive matrix represents a positive mapping.

In the literature, positive mapping is often studied in the positive orthant +n\mathbb{R}_{+}^{n}. The following simple example shows that the positive orthant may not be the right cone for studying consensus. Consider the system x˙1=0\dot{x}_{1}=0, x˙2=x1x2\dot{x}_{2}=x_{1}-x_{2}. It is then obvious that the system achieves exponential consensus. The vector plot of the system is shown in Fig. 1. We can see that the system does not contract the positive orthant into its interior since x1=0x_{1}=0 is invariant. However, the system does contract the smaller cone painted in gray. This turns out to be a key observation we need.

Refer to caption


Figure 1: The red arrows represent the vector fields of a system. The x2x_{2}-axis is invariant and hence cannot be mapped into the interior of the positive orthant. However, the system contracts the smaller cone painted in gray.

To proceed, we need to justify that contraction in Hilbert metric is equivalent to consensus that we defined earlier in Definition 2.

Lemma 1.

Suppose that 𝒦\mathcal{K} is a proper cone in +n\mathbb{R}_{+}^{n} satisfying 𝒦Int+n{0}\mathcal{K}\subseteq\text{Int}\ \mathbb{R}_{+}^{n}\cup\{0\}. Then a system achieves asymptotic (resp. exponential) consensus on 𝒦\mathcal{K} if and only if there exists class 𝒦\mathcal{KL} function (resp. positive constants K,λK,\lambda) such that

d(x(t),𝟙)\displaystyle d(x{(t)},\mathds{1}) β(d(x(0),𝟙),t) (asymptotic)\displaystyle\leq\beta(d(x{(0)},\mathds{1}),t)\text{ (asymptotic)}
d(x(t),𝟙)\displaystyle d(x{(t)},\mathds{1}) Keλtd(x(0),𝟙) (exponential)\displaystyle\leq Ke^{-\lambda t}d(x{(0)},\mathds{1})\text{ (exponential)}

where d(x,y)d(x,y) stands for the Hilbert metric between xx and yy.

Proof.

We prove the asymptotic case – the exponential case is similar. Suppose that the system achieves asymptotic consensus on 𝒦\mathcal{K}, i.e., there exists a class 𝒦\mathcal{KL} function β\beta such that |xi(t)xj(t)|β(|xi(0)xj(0)|,t)|x_{i}(t)-x_{j}(t)|\leq\beta(|x_{i}(0)-x_{j}(0)|,t) for all tt and x(0)𝒦x(0)\in\mathcal{K} and x(t)𝒦x(t)\in\mathcal{K} for all t0t\geq 0. Let

An(x)=|x|x|n𝟙|2,Bn(x)=i,j=1n(xixj)2;A_{n}(x)=\left|x-\frac{|x|}{\sqrt{n}}\mathds{1}\right|^{2},\quad B_{n}(x)=\sum_{i,j=1}^{n}(x_{i}-x_{j})^{2};

we estimate:

d(x(t),𝟙)2\displaystyle d(x(t),\mathds{1})^{2} =d(x(t),|x(t)|n𝟙)21c1An(x(t)) (Lemma 5)\displaystyle=d\left(x(t),\frac{|x(t)|}{\sqrt{n}}\mathds{1}\right)^{2}\leq\frac{1}{c_{1}}A_{n}(x(t))\text{ (Lemma \ref{lem:2norm-Hnorm}})
1c1nBn(x(t)) (Lemma 4)\displaystyle\leq\frac{1}{c_{1}n}B_{n}(x(t))\text{ (Lemma \ref{lem:An-Bn})}
=1c1ni,j=1n(xi(t)xj(t))2\displaystyle=\frac{1}{c_{1}n}\sum_{i,j=1}^{n}(x_{i}(t)-x_{j}(t))^{2}
1c1ni,j=1nβ(t,(xi(0)xj(0))2)\displaystyle\leq\frac{1}{c_{1}n}\sum_{i,j=1}^{n}\beta(t,(x_{i}(0)-x_{j}(0))^{2})
nc1β(t,Bn(x(0)))\displaystyle\leq\frac{n}{c_{1}}\beta(t,B_{n}(x(0)))
nc1β~(t,d(x(0),𝟙)) (Lemma 4,5)\displaystyle\leq\frac{n}{c_{1}}\tilde{\beta}(t,d(x(0),\mathds{1}))\text{ (Lemma \ref{lem:An-Bn},\ref{lem:2norm-Hnorm})}

in which β~\tilde{\beta} is some class 𝒦\mathcal{KL} function. The converse proceeds in a similar fashion. ∎

Remark 3.

Although in Lemma 1, the initial condition is restricted to a cone 𝒦\mathcal{K}, in practice this is sufficient for consensus on any compact set DnD\subseteq\mathbb{R}^{n}. Indeed, let y=x+α𝟙y=x+\alpha\mathds{1}. Then y˙=A(t,yα𝟙)y\dot{y}=A(t,y-\alpha\mathds{1})y and consensus of yy is equivalent to consensus of xx. By choosing α>0\alpha>0 sufficiently large, we may assume DD is in the interior of +n\mathbb{R}_{+}^{n}.

We propose to study the following type of small cones. For γ[0,1n)\gamma\in[0,\frac{1}{\sqrt{n}}), define a family of cones

𝒦(γ):={xn:xi|x|1nγ,i=1,n}.\mathcal{K}(\gamma):=\left\{x\in\mathbb{R}^{n}:\frac{x_{i}}{|x|}\geq\frac{1}{\sqrt{n}}-\gamma,\;\forall i=1,\cdots n\right\}. (6)

See Fig. 2 for an illustration of such cones. The following lemma summarizes some properties of the diameter (in Hilbert metric) of these cones.

Refer to caption

Figure 2: Cone 𝒦(γ)\mathcal{K}(\gamma) when n=2n=2.
Lemma 2.

Given γ[0,1n)\gamma\in[0,\frac{1}{\sqrt{n}}), the diameter of 𝒦(γ)\mathcal{K}(\gamma) is

diam𝒦(γ)=log(1n+n(1nγ)2){\rm diam}\mathcal{K}(\gamma)=\log\left(1-n+\frac{n}{(1-\sqrt{n}\gamma)^{2}}\right)

Denote α(γ)=diam𝒦(γ)\alpha(\gamma)={\rm diam}\mathcal{K}(\gamma), then

  • α\alpha is smooth, strictly increasing on [0,1/n)[0,1/\sqrt{n}), and α(0)=0\alpha(0)=0, α(t)\alpha(t)\to\infty as t1nt\to\frac{1}{\sqrt{n}}. The derivative of α\alpha is lower bounded away from zero.

  • For any ϵ0[0,1n)\epsilon_{0}\in[0,\frac{1}{\sqrt{n}}) and C(0,1)C\in(0,1), there exist some positive constants k1,k2k_{1},k_{2} and k(0,1)k\in(0,1), such that k1γα(γ)k2γk_{1}\gamma\leq\alpha(\gamma)\leq k_{2}\gamma, and

    α(Cγ)kα(γ)\alpha(C\gamma)\leq k\alpha(\gamma) (7)

    for all γ[0,ϵ0]\gamma\in[0,\epsilon_{0}].

Proof.

We show (7). Notice that

α(r)α(Cr)\displaystyle\alpha(r)-\alpha(Cr) =α(ξ)(1C)rcrck2Cα(Cr)\displaystyle=\alpha^{\prime}(\xi)(1-C)r\geq cr\geq\frac{c}{k_{2}C}\alpha(Cr)

from which it follows that

α(Cr)11+c/k2Cα(r).\alpha(Cr)\leq\frac{1}{1+c/k_{2}C}\alpha(r).

The rest is straightforward. ∎

II-C Accumulated graph

We have noted that in both [25] and [17], consensus is related to the accumulation of the graphs over time, either the union of the switching graph in [17], or the integration of the system matrix in [25]. This motivates us to define the accumulated graph for a time-varying graph 𝒢(t)\mathcal{G}(t), t0\forall t\geq 0.

Definition 3 (Accumulated graph).

Let 𝒢()=(aij())\mathcal{G}(\cdot)=(a_{ij}(\cdot)) be a measurable time-varying graph, i.e., taij(t)t\mapsto a_{ij}(t) is a measurable function for all iji\neq j. The accumulating graph of 𝒢()\mathcal{G}(\cdot) over the interval [t1,t2][t_{1},t_{2}] is the graph 𝒢|t1t2\mathcal{G}|_{t_{1}}^{t_{2}} defined by the Lebesgue integral

𝒢|t1t2=t1t2𝒢(t)dt.\mathcal{G}|_{t_{1}}^{t_{2}}=\int_{t_{1}}^{t_{2}}\mathcal{G}(t){\rm d}t.

in the sense that (𝒢|t1t2)ij=t1t2aij(t)dt(\mathcal{G}|_{t_{1}}^{t_{2}})_{ij}=\int_{t_{1}}^{t_{2}}a_{ij}(t){\rm d}t.

Example 1.

For a system (1) with switching topology, the graph associated with it can be written as 𝒢(t)=i=1N1Ai(t)𝒢i\mathcal{G}(t)=\sum_{i=1}^{N}1_{A_{i}}(t)\mathcal{G}_{i} where 1Ai1_{A_{i}} is the indicator function of some measurable sets AiA_{i} and 𝒢i\mathcal{G}_{i} the graph corresponding to p=ip=i. Now the union graph on [t,t+T][t,t+T] used in [17] was nothing but the accumulated graph 𝒢|tt+T\mathcal{G}|_{t}^{t+T}. This is quite similar to the construction of the Lebesgue integration – define first for simple functions and then extend to larger class of functions.

With these technical preparations, we are now ready to present the main results of this paper.

III Main results

We start with a lemma which explains how connectedness of the graph is related to the contraction property under the Hilbert metric. The proof can be found in the Appendix.

Lemma 3.

Let An×nA\in\mathbb{R}^{n\times n} be a non-negative matrix and there exist a constant δ(0,1)\delta\in(0,1), and an integer kk such that

  • A𝟙=𝟙A\mathds{1}=\mathds{1};

  • aik>δa_{ik}>\delta for all i=1,,ni=1,\cdots,n;

then for any 0<ϵ<1n0<\epsilon<\frac{1}{\sqrt{n}}, there exists a constant C(0,1)C\in(0,1) such that

A𝒦(ϵ)𝒦(Cϵ)A\mathcal{K}(\epsilon)\subseteq\mathcal{K}(C\epsilon) (8)

where 𝒦(ϵ)\mathcal{K}(\epsilon) is defined as in (6). Moreover, CC can be taken as 1δ1nϵδ\frac{1-\delta}{1-\sqrt{n}\epsilon\delta}. As a result, there exist a positive constant c(0,1)c\in(0,1) such that

diam(Am𝒦(ϵ))cmdiam(𝒦(ϵ)),m1.{\rm diam}(A^{m}\mathcal{K}(\epsilon))\leq c^{m}{\rm diam}(\mathcal{K}(\epsilon)),\quad\forall m\geq 1. (9)

where kk can be taken as c=11+ηCc=\frac{1}{1+\eta C} and η>0\eta>0 depends on ϵ\epsilon, and nn.

Our first main result is the following theorem.

Theorem 3.

Let DD be a compact set in n\mathbb{R}^{n}. Consider the multi-agent system (5) satisfying Assumption A1. Let {tk}1\{t_{k}\}_{1}^{\infty} be an increasing sequence with tkkt_{k}\xrightarrow{k\to\infty}\infty and supk|tk+1tk|<\sup_{k}|t_{k+1}-t_{k}|<\infty. If there exists a graph 𝒢B(x)\mathcal{G}^{B(x)} such that B(x)B(x) is quasi-strongly connected and

tktk+1𝒢A(t,ϕ(t,tk,x))dt𝒢B(x)\int_{t_{k}}^{t_{k+1}}\mathcal{G}^{A(t,\phi(t,t_{k},x))}{\rm d}t\geq\mathcal{G}^{B(x)} (10)

for all xDx\in D, k1k\geq 1, then the system achieves asymptotic consensus on DD. If in addition, the graph 𝒢B(x)\mathcal{G}^{B(x)} is continuous, the system achieves exponential consensus on DD.

Proof.

We prove the second part first. As remarked earlier, by defining the coordinate transform y=x+α𝟙y=x+\alpha\mathds{1} for α>0\alpha>0 large, we may assume that DD lies in a sufficiently small cone in +n\mathbb{R}_{+}^{n}. First, assume tA(t,x)t\mapsto A(t,x) is piece-wise continuous. This assumption will be removed later. Consider the Euler approximation scheme on the interval [tk,tk+1][t_{k},t_{k+1}]:

xi+1=xi+hA(tk+ih,xi)xi,i=0,,N1x_{i+1}=x_{i}+hA(t_{k}+ih,x_{i})x_{i},\quad i=0,\cdots,N-1

with h=tk+1tkNh=\frac{t_{k+1}-t_{k}}{N}, N+N\in\mathbb{Z}_{+} large and x0=xx_{0}=x. Choose λ\lambda large enough such that A¯(t,x):=A(t,x)+λI0\bar{A}(t,x):=A(t,x)+\lambda I\geq 0 for all t0t\geq 0 and xDx\in D. The following calculation is in order

x0\displaystyle x_{0} =x\displaystyle=x
x1\displaystyle x_{1} =(1hλ)x+hA¯(tk,x)x\displaystyle=(1-h\lambda)x+h\bar{A}(t_{k},x)x
x2\displaystyle x_{2} =[(1hλ)2I+h(1hλ)(A¯(tk,x)+A¯(tk+h,x1))+]x\displaystyle=[(1-h\lambda)^{2}I+h(1-h\lambda)(\bar{A}(t_{k},x)+\bar{A}(t_{k}+h,x_{1}))+*]x
\displaystyle\vdots
xN\displaystyle x_{N} =[(1hλ)NI+(1hλ)N1i=0N1hA¯(tk+ih,xi)+]x\displaystyle=[(1-h\lambda)^{N}I+(1-h\lambda)^{N-1}\sum_{i=0}^{N-1}h\bar{A}(t_{k}+ih,x_{i})+*]x (11)

where * stands for non-negative terms. Now let PkN(x)P_{k}^{N}(x) be the matrix in the bracket on the right hand side of (11) and define Pk(x)=limNPkN(x)P_{k}(x)=\lim_{{N}\to\infty}P_{k}^{N}(x). Note that

ϕ(tk+1,tk,x)=Pk(x)x.\phi(t_{k+1},t_{k},x)=P_{k}(x)x. (12)

We claim that Pk(x)P_{k}(x) has the following properties: 1) Pk(x)P_{k}(x) is non-negative and Pk(x)𝟙=𝟙P_{k}(x)\mathds{1}=\mathds{1}; 2) There exists a non-negative matrix SS defining a QSC graph 111Here we are slightly abusing the concept of a QSC graph. It was previously defined as the graph associated with a continuous-time system 5. Here, it should be understood in the sense that there exists some k+k\in\mathbb{N}_{+}, such that Sik>0S_{ik}>0 for all iki\neq k. , independent of k,xk,x such that Pk(x)SP_{k}(x)\geq S. Item 1) is obvious, we verify 2). From (11) we see

Pk(x)eλ(tk+1tk)(I+tktk+1A¯(t,ϕ(t,tk,x))dt)SP_{k}(x)\geq e^{-\lambda(t_{k+1}-t_{k})}\left(I+\int_{t_{k}}^{t_{k+1}}\bar{A}(t,\phi(t,t_{k},x)){\rm d}t\right)\geq S

for some non-negative SS such that 𝒢S\mathcal{G}^{S} is QSC thanks to (10), the continuity of 𝒢B(x)\mathcal{G}^{B(x)} and compactness of DD. Without loss of generality, we may assume that 𝒢S\mathcal{G}^{S} is δ\delta-connected (since otherwise, we can consider the matrix Pk+nPk+n1PkP_{k+n}P_{k+n-1}\cdots P_{k}, which will be lower-bounded by SnS^{n} that is δ\delta-connected). Invoking Lemma 3, we conclude that the system achieves exponential consensus on a cone of the form (6) which includes DD in its interior.

To remove the assumption on piece-wise continuity of tA(t,x)t\mapsto A(t,x) on [tk,tk+1][t_{k},t_{k+1}], it suffices to replace Riemann integration by Lebesgue integration as follows. View φ:tA(t,)\varphi:t\mapsto A(t,\cdot) as a mapping from [tk,tk+1][t_{k},t_{k+1}] to C(D;n×n)C(D;\mathbb{R}^{n\times n}) equipped with norm g=supxDg(x)||g||_{*}=\sup_{x\in D}||g(x)||. Then φL([tk,tk+1];C(D;n×n))\varphi\in L^{\infty}([t_{k},t_{k+1}];C(D;\mathbb{R}^{n\times n})), which can be approximated by simple functions. Let η>0\eta>0 be an arbitrarily small constant, and 𝒜(t,x)=i=0N11[si,si+1)(t)Ai(x)\mathscr{A}(t,x)=\sum_{i=0}^{N-1}1_{[s_{i},s_{i+1})}(t)A_{i}(x) the simple function such that φ𝒜<η\|\varphi-\mathscr{A}\|_{\infty}<\eta. Assume that the partition is uniform (otherwise we can always refine the partition to make it close to uniform and then use approximation arguments), i.e., si+1si=tk+1tkNs_{i+1}-s_{i}=\frac{t_{k+1}-t_{k}}{N} for all ii. Let A¯i(x)=Ai(x)+λI0\bar{A}_{i}(x)=A_{i}(x)+\lambda I\geq 0, 𝒜¯(t,x)=𝒜(t,x)+λI\bar{\mathscr{A}}(t,x)=\mathscr{A}(t,x)+\lambda I. As before, the Euler approximation scheme gives (11). Now

i=0N1hA¯(tk\displaystyle\sum_{i=0}^{N-1}h\bar{A}(t_{k} +ih,xi)=i=0N1hA¯i(xi)\displaystyle+ih,x_{i})=\sum_{i=0}^{N-1}h\bar{A}_{i}(x_{i})
=tktk+1𝒜¯(t,ϕ(t,tk,x))dt\displaystyle=\int_{t_{k}}^{t_{k+1}}\bar{\mathscr{A}}(t,\phi(t,t_{k},x)){\rm d}t
tktk+1A¯(t,ϕ(t,tk,x))dtη(tk+1tk)𝟙n×n.\displaystyle\geq\int_{t_{k}}^{t_{k+1}}\bar{A}(t,\phi(t,t_{k},x)){\rm d}t-\eta(t_{k+1}-t_{k})\mathds{1}_{n\times n}.

Since |tk+1tk||t_{k+1}-t_{k}| is uniformly bounded, we can choose η\eta sufficiently small such that PkN(x)P_{k}^{N}(x) is lower-bounded by a QSC graph which is independent of kk. Thus the system achieves exponential consensus on DD.

It remains to prove asymptotic consensus when 𝒢B(x)\mathcal{G}^{B(x)} is not necessarily continuous. As before, we can assume that tA(t,x)t\mapsto A(t,x) is piecewise continuous. The Euler approximation still converges due to local Lipschitz continuity of the system vector fields. Now, instead of having a uniform lower bound on Pk(x)P_{k}(x) (see (12)), we only know that it is bounded by some QSC graph. But still, we know that for small ϵ0>0\epsilon_{0}>0, Pk(x)𝒦(ϵ0)𝒦(ϵ1)P_{k}(x)\mathcal{K}(\epsilon_{0})\subseteq\mathcal{K}(\epsilon_{1}). Since DD is compact, we may assume that D𝒦(ϵ0)D\subseteq\mathcal{K}(\epsilon_{0}). Define D1=ϕ(t1,t0,D)D_{1}=\phi(t_{1},t_{0},D), Dk+1=ϕ(tk+1,tk,Dk)D_{k+1}=\phi(t_{k+1},t_{k},D_{k}) for k1k\geq 1 and ηk=diamDk\eta_{k}={\rm diam}D_{k} which is a non-negative decreasing sequence. Set a decreasing sequence {ϵk}1\{\epsilon_{k}\}_{1}^{\infty}, such that Dk𝒦(ϵk)D_{k}\subseteq\mathcal{K}(\epsilon_{k}). Clearly, ηk\eta_{k} is strictly decreasing and is bounded from below by 0. Therefore ηk\eta_{k} converges to a limit c0c\geq 0.

Refer to caption

Figure 3: Asymptotic contraction.

Suppose c>0c>0, and choose ηm\eta_{m} sufficiently close to cc. There must exist y𝒦(ϵm)\𝒦(c)y\in\mathcal{K}(\epsilon_{m})\backslash\mathcal{K}(c), such that ϕ(tm+1,tm,y)𝒦(ϵm)\𝒦(c)\phi(t_{m+1},t_{m},y)\in\mathcal{K}(\epsilon_{m})\backslash\mathcal{K}(c). Choose z𝒦(c)z\in\mathcal{K}(c) close to yy, then due to the continuity of yϕ(t,t0,y)|ϕ(t,t0,y)|y\mapsto\frac{\phi(t,t_{0},y)}{|\phi(t,t_{0},y)|}, the error between ϕ(tm+1,tm,y)|ϕ(tm+1,tm,y)|\frac{\phi(t_{m+1},t_{m},y)}{|\phi(t_{m+1},t_{m},y)|} and ϕ(tm+1,tm,z)|ϕ(tm+1,tm,z)|\frac{\phi(t_{m+1},t_{m},z)}{|\phi(t_{m+1},t_{m},z)|} is of order O(|ϵmc|)O(|\epsilon_{m}-c|) which can be made sufficiently small by choosing mm large enough. But

minw𝒦(c)Dϕ(tm+1,tm,w))i|ϕ(tm+1,tm,w)|=1nc>1nc\min_{w\in\mathcal{K}(c)\cap D}\frac{\phi(t_{m+1},t_{m},w))_{i}}{|\phi(t_{m+1},t_{m},w)|}=\frac{1}{\sqrt{n}}-c^{\prime}>\frac{1}{\sqrt{n}}-c

(since ηm+1<ηm\eta_{m+1}<\eta_{m}) implying that ϕ(tm+1,tm,z)i|ϕ(tm+1,tm,z)|\frac{\phi(t_{m+1},t_{m},z)_{i}}{|\phi(t_{m+1},t_{m},z)|} is away from 1nc\frac{1}{\sqrt{n}}-c for any mm. This is a contradiction since ϕ(tm+1,tm,y)|ϕ(tm+1,tm,y)|𝒦(ϵm)\𝒦(c)\frac{\phi(t_{m+1},t_{m},y)}{|\phi(t_{m+1},t_{m},y)|}\in\mathcal{K}(\epsilon_{m})\backslash\mathcal{K}(c) and that ϕ(tm+1,tm,y)|ϕ(tm+1,tm,y)|\frac{\phi(t_{m+1},t_{m},y)}{|\phi(t_{m+1},t_{m},y)|} and ϕ(tm+1,tm,z)|ϕ(tm+1,tm,z)|\frac{\phi(t_{m+1},t_{m},z)}{|\phi(t_{m+1},t_{m},z)|} are sufficiently close. Thus we conclude that c=0c=0 and asymptotic consensus is achieved. The proof strategy is shown in Fig. 3.

The proof is now complete. ∎

Theorem 3 needs the computation of the integral tktk+1A(t,ϕ(t,tk,x))dt\int_{t_{k}}^{t_{k+1}}A(t,\phi(t,t_{k},x)){\rm d}t, which is impossible in most cases – except that A(t,x)A(t,x) does not dependent on xx. The following corollary is more convenient for practical use.

Corollary 1.

Consider the system (3) under Assumption A1. Let DD be a compact invariant set. Suppose that there exists a (continuous) QSC graph 𝒢B(x)\mathcal{G}^{B(x)}, continuous on DD, and an increasing sequence {tk}1\{t_{k}\}_{1}^{\infty} with tkkt_{k}\xrightarrow{k\to\infty}\infty and sup|tk+1tk|<\sup|t_{k+1}-t_{k}|<\infty, such that

tktk+1𝒢A(t,x)dt𝒢B(x),xD,k1\int_{t_{k}}^{t_{k+1}}\mathcal{G}^{A(t,x)}{\rm d}t\geq\mathcal{G}^{B(x)},\quad\forall x\in D,\;k\geq 1

then the system achieves (asymptotic) exponential consensus on DD.

Proof.

Assume B(x)B(x) is continuous. We utilize Euler approximation as before. Note that for fixed rs{1,,n}r\neq s\in\{1,\cdots,n\}, there exists xrsDx_{rs}\in D such that

i=1N1hars(tk+hi,ϕ(tk+hi,tk,x))i=1N1hars(tk+hi,xrs)\sum_{i=1}^{N-1}ha_{rs}(t_{k}+hi,\phi(t_{k}+hi,t_{k},x))\geq\sum_{i=1}^{N-1}ha_{rs}(t_{k}+hi,x_{rs})

since for every kk and ii, xars(tk+hi,ϕ(tk+hi,tk,x))x\mapsto a_{rs}(t_{k}+hi,\phi(t_{k}+hi,t_{k},x)) is continuous. But the right hand side is an approximation of tktk+1ars(t,xrs)dt\int_{t_{k}}^{t_{k+1}}a_{rs}(t,x_{rs}){\rm d}t, which is lower bounded by Brs(xrs)B_{rs}(x_{rs}). Define a matrix B~:=(Brs(xrs))\tilde{B}:=(B_{rs}(x_{rs})). Due to the continuity of xB(x)x\mapsto B(x), B(x)B(x) and hence B~\tilde{B}, are lower bounded by some matrix SS such that 𝒢S\mathcal{G}^{S} is QSC. The conclusion follows invoking Theorem 3. ∎

Example 2.

1) Theorem 2 is now a corollary of Theorem 3. Our result is slightly stronger: the mapping tA(t)t\mapsto A(t) is only required to be bounded measurable while in Theorem 2, this mapping is assumed to be piecewise continuous. Note that the proof strategy is rather different for the two theorems.

2) For Theorem 1, the system (1) can be written as

x˙=[k=1p1k(σ(t))Ak(x)]x.\dot{x}=\left[\sum_{k=1}^{p}1_{k}(\sigma(t))A_{k}(x)\right]x.

Fix an interval [t0,t0+T][t_{0},t_{0}+T], then on [t0τD,t0+T+τD][t_{0}-\tau_{D},t_{0}+T+\tau_{D}], we integrate

t0τDt0+T+τDk=1p1k(σ(t))Ak(x)dtτDk=1pAk(x).\int_{t_{0}-\tau_{D}}^{t_{0}+T+\tau_{D}}\sum_{k=1}^{p}1_{k}(\sigma(t))A_{k}(x){\rm d}t\geq\tau_{D}\sum_{k=1}^{p}A_{k}(x).

By assumption, k=1pAk(x)\sum_{k=1}^{p}A_{k}(x) is QSC and hence the system achieves asymptotic consensus.

If we assume further that Ak(x)A_{k}(x) is continuous, then we get exponential consensus from Theorem 3. But as far as we know, it is not clear how to prove this using the techniques in [17]. In addition, Theorem 1 was proven only for switching multi-agent system with regular switching signal while Theorem 3 only requires the “switchings” to be measurable which is always satisfied in practice.

Example 3.

Consider a Kuramoto model with identical frequency

θ˙i=ω+j𝒩t(i)aij(t)sin(θjθi),i=1,n\dot{\theta}_{i}=\omega+\sum_{j\in\mathcal{N}_{t}(i)}a_{ij}(t)\sin(\theta_{j}-\theta_{i}),\quad i=1,\cdots n (13)

where 𝒩t(i)\mathcal{N}_{t}(i) stands for the neighboring node of ii at time tt and aij()L(0,>0)a_{ij}(\cdot)\in L^{\infty}(\mathbb{R}_{\geq 0},\mathbb{R}_{>0}). Let a,ba,b be real numbers such that 0ba<π0\leq b-a<\pi and the graph 𝒢A(t)\mathcal{G}^{A(t)} satisfies the assumption of Theorem 3 where the lower bound of the accumulated graph is now state-independent.

We claim that the system (13) achieves exponential consensus on [a,b]n[a,b]^{n}. In particular, if θ˙i=ω+j𝒩(i)sin(θjθi)\dot{\theta}_{i}=\omega+\sum_{j\in\mathcal{N}(i)}\sin(\theta_{j}-\theta_{i}) and the graph associated with the system is QSC, then exponential consensus is achieved on [a,b]n[a,b]^{n}. To see the claim, let xi=θiωx_{i}=\theta_{i}-\omega, the above model can be written as

x˙i=j𝒩t(i)aij(t)sin(xjxi)\dot{x}_{i}=\sum_{j\in\mathcal{N}_{t}(i)}a_{ij}(t)\sin(x_{j}-x_{i})

or in matrix form x˙=A(t,x)x\dot{x}=A(t,x)x where (A(t,x))ij=sin(xjxi)xjxiaij(t)(A(t,x))_{ij}=\frac{\sin(x_{j}-x_{i})}{x_{j}-x_{i}}a_{ij}(t) for iji\neq j and j𝒩t(i)j\in\mathcal{N}_{t}(i). Then on [a,b]n[a,b]^{n}, (A(t,x))ijcaij(t)(A(t,x))_{ij}\geq ca_{ij}(t) for some positive constant cc. Thus consensus is determined by the graph 𝒢A(t)\mathcal{G}^{A(t)}.

IV Simulation Result

We simulate Example 3 for ω=0\omega=0. Consider a chain structure as in Fig. 4 of N=10N=10 oscillators. The weights on the link (xixi+1)(x_{i}\to x_{i+1}) is ai,i+1(t)a_{i,i+1}(t). The weights are generated in the following manner. First, generate some random intervals [tk,tk+1][t_{k},t_{k+1}] for k=1,,Mk=1,\cdots,M for some large MM. Then, divide each [tk,tk+1][t_{k},t_{k+1}] into NN smaller pieces [sik,si+1k][s_{i}^{k},s^{k}_{i+1}] randomly. After that, Euler scheme will run on each interval [sik,si+1k][s_{i}^{k},s^{k}_{i+1}]. In each step of the Euler scheme, we choose three random numbers p,q,rp,q,r from {1,,N1}\{1,\cdots,N-1\} and generate three random positives numbers ap,p+1a_{p,p+1}, aq,q+1a_{q,q+1}, ar,r+1a_{r,r+1} lower bounded by a threshold δ>0\delta>0 and the rest of ai,i+1a_{i,i+1} are set to zero.

By construction, the weights ai,i+1a_{i,i+1} are zero for most of the time and since the these weights are generated randomly, they are quite irregular. But still, we can see from Fig. 5 that the system achieves consensus.

Refer to caption

Figure 4: A chain of coupled oscillators.

Refer to caption Refer to caption

Figure 5: Simulation result. Left: the signal a1,1+1(t)a_{1,1+1}(t). Right: the state components.

V Conclusion

In this technical note, we have shown that the Hilbert metric can serve as alternative tool to study consensus properties. It is advantageous in dealing with nonlinearities and time dependencies, and requires very weak regularity assumptions. The results obtained in this note are somewhat preliminary and open the door for future research.

VI Appendix

Proof of Lemma 3 .

Consider the set

(ϵ)={e:1nϵei1n,i=1,,n,|e|1}.\mathcal{E}(\epsilon)=\left\{e:\frac{1}{\sqrt{n}}-\epsilon\leq e_{i}\leq\frac{1}{\sqrt{n}},\;\forall i=1,\cdots,n,\;|e|\leq 1\right\}.

Then 𝒦(ϵ)={αe:α>0,e(ϵ)}\mathcal{K}(\epsilon)=\{\alpha e:\forall\alpha\in\mathbb{R}_{>0},\;e\in\mathcal{E}(\epsilon)\}. Thus it is sufficient to prove

(Ae)i|Ae|1nCϵ,i1,e(ϵ).\frac{(Ae)_{i}}{|Ae|}\geq\frac{1}{\sqrt{n}}-C\epsilon,\quad\forall i\geq 1,\;e\in\mathcal{E}(\epsilon). (14)

Write e=1ne~e=\frac{1}{\sqrt{n}}-\tilde{e}, then 0e~iϵ0\leq\tilde{e}_{i}\leq\epsilon and (Ae)i=1n(Ae~)i(Ae)_{i}=\frac{1}{\sqrt{n}}-(A\tilde{e})_{i}. On the one hand, (Ae~)i=jknaije~j+aike~kδe~k(A\tilde{e})_{i}=\sum_{j\neq k}^{n}a_{ij}\tilde{e}_{j}+a_{ik}\tilde{e}_{k}\geq\delta\tilde{e}_{k}. On the other hand,

jknaije~j+aike~k\displaystyle\sum_{j\neq k}^{n}a_{ij}\tilde{e}_{j}+a_{ik}\tilde{e}_{k} ϵjknaij+aike~k\displaystyle\leq\epsilon\sum_{j\neq k}^{n}a_{ij}+a_{ik}\tilde{e}_{k}
=ϵ(1aik)+aike~k\displaystyle=\epsilon(1-a_{ik})+a_{ik}\tilde{e}_{k}
δ(e~kϵ)+ϵ.\displaystyle\leq\delta(\tilde{e}_{k}-\epsilon)+\epsilon.

Thus we obtain the inequality

1nϵ(1δ)δe~k(Ae)i1nδe~k.\frac{1}{\sqrt{n}}-\epsilon(1-\delta)-\delta\tilde{e}_{k}\leq(Ae)_{i}\leq\frac{1}{\sqrt{n}}-\delta\tilde{e}_{k}.

As a result |Ae|1nδe~k|Ae|\leq 1-\sqrt{n}\delta\tilde{e}_{k} and

(Ae)i|Ae|1nδe~kϵ(1δ)1nδe~k=1n1δ1ne~kδϵ\frac{(Ae)_{i}}{|Ae|}\geq\frac{\frac{1}{\sqrt{n}}-\delta\tilde{e}_{k}-\epsilon(1-\delta)}{1-\sqrt{n}\delta\tilde{e}_{k}}=\frac{1}{\sqrt{n}}-\frac{1-\delta}{1-\sqrt{n}\tilde{e}_{k}\delta}\epsilon

Recall that e~kϵ<1n\tilde{e}_{k}\leq\epsilon<\frac{1}{\sqrt{n}}, we get 1δ1ne~kδ1δ1nϵδ:=C(0,1)\frac{1-\delta}{1-\sqrt{n}\tilde{e}_{k}\delta}\leq\frac{1-\delta}{1-\sqrt{n}\epsilon\delta}:=C\in(0,1) and (14) follows. In other words, we have shown A𝒦(ϵ)𝒦(Cϵ)A\mathcal{K}(\epsilon)\subseteq\mathcal{K}(C\epsilon).

To show (9), following the same procedure, we can prove A𝒦(Cϵ)𝒦(CCϵ)A\mathcal{K}(C\epsilon)\subseteq\mathcal{K}(C^{\prime}C\epsilon), where C=1δ1nCϵδ<CC^{\prime}=\frac{1-\delta}{1-\sqrt{n}C\epsilon\delta}<C. Thus A2𝒦(ϵ)A𝒦(Cϵ)𝒦(C2ϵ)A^{2}\mathcal{K}(\epsilon)\subseteq A\mathcal{K}(C\epsilon)\subseteq\mathcal{K}(C^{2}\epsilon). By induction, we can show Am𝒦(ϵ)𝒦(Cmϵ)A^{m}\mathcal{K}(\epsilon)\subseteq\mathcal{K}(C^{m}\epsilon). Thus the inequality (9). ∎

Lemma 4.

For x0x\geq 0, define two quantities as

An(x)=|x|x|n𝟙|2,Bn(x)=i,j=1n(xixj)2A_{n}(x)=\left|x-\frac{|x|}{\sqrt{n}}\mathds{1}\right|^{2},\quad B_{n}(x)=\sum_{i,j=1}^{n}(x_{i}-x_{j})^{2}

then nAnBn2nAnnA_{n}\leq B_{n}\leq 2nA_{n}.

Proof.

It suffices to prove for |x|=1|x|=1. Define X=i=1nxiX=\sum_{i=1}^{n}x_{i}, then

An=22nX,Bn=2n2X2.A_{n}=2-\frac{2}{\sqrt{n}}X,\quad B_{n}=2n-2X^{2}.

Since |x|=1|x|=1, we have 0Xn0\leq X\leq\sqrt{n}. Thus Bn2n2nX=nAnB_{n}\geq 2n-2\sqrt{n}X=nA_{n}. On the other hand, 2nAnBn=2(Xn)202nA_{n}-B_{n}=2(X-\sqrt{n})^{2}\geq 0. ∎

Lemma 5.

Consider the cone +n\mathbb{R}_{+}^{n}. Suppose that 𝒦\mathcal{K} is a proper cone such that 𝒦Int+n{0}\mathcal{K}\subseteq\text{Int}\ \mathbb{R}_{+}^{n}\cup\{0\}. Then there exist some constants C2>C1>0C_{2}>C_{1}>0 such that C1d(x,w)|xw|C2d(x,w)C_{1}d(x,w)\leq|x-w|\leq C_{2}d(x,w) for all x+n,x\in\mathbb{R}_{+}^{n}, w𝒦w\in\mathcal{K} with |w|=|x|=1|w|=|x|=1.

Proof.

Since |w|=|x|=1|w|=|x|=1, we can find a constant C>0C>0, such that

Ctanh(12d(x,w))|xw|exp(d(x,w))1C\tanh\left(\frac{1}{2}d(x,w)\right)\leq|x-w|\leq\exp\left(d(x,w)\right)-1

invoking [36, Theorem 4.1], see [36, (4.1) and (4.3)]. The conclusion follows by noticing that d(x,w)d(x,w) is bounded for all x+nx\in\mathbb{R}_{+}^{n}, w𝒦w\in\mathcal{K} with |w|=|x|=1|w|=|x|=1. ∎

References

  • [1] L. Moreau, “Stability of multiagent systems with time-dependent communication links,” IEEE Transactions on automatic control, vol. 50, no. 2, pp. 169–182, 2005.
  • [2] T. Vicsek, A. Czirók, E. Ben-Jacob, I. Cohen, and O. Shochet, “Novel type of phase transition in a system of self-driven particles,” Physical review letters, vol. 75, no. 6, p. 1226, 1995.
  • [3] A. Jadbabaie, J. Lin, and A. S. Morse, “Coordination of groups of mobile autonomous agents using nearest neighbor rules,” IEEE Transactions on automatic control, vol. 48, no. 6, pp. 988–1001, 2003.
  • [4] W. Ren and R. W. Beard, “Consensus seeking in multiagent systems under dynamically changing interaction topologies,” IEEE Transactions on automatic control, vol. 50, no. 5, pp. 655–661, 2005.
  • [5] F. Dorfler and F. Bullo, “Synchronization and transient stability in power networks and nonuniform kuramoto oscillators,” SIAM Journal on Control and Optimization, vol. 50, no. 3, pp. 1616–1642, 2012.
  • [6] H. Rainer and U. Krause, “Opinion dynamics and bounded confidence: models, analysis and simulation,” Journal of Artifical Societies and Social Simulation, 2002.
  • [7] K.-K. Oh, M.-C. Park, and H.-S. Ahn, “A survey of multi-agent formation control,” Automatica, vol. 53, pp. 424–440, 2015.
  • [8] G. Shi and K. H. Johansson, “The role of persistent graphs in the agreement seeking of social networks,” IEEE Journal on Selected Areas in Communications, vol. 31, no. 9, pp. 595–606, 2013.
  • [9] B. D. Anderson, G. Shi, and J. Trumpf, “Convergence and state reconstruction of time-varying multi-agent systems from complete observability theory,” IEEE Transactions on Automatic Control, vol. 62, no. 5, pp. 2519–2523, 2016.
  • [10] N. Barabanov and R. Ortega, “Global consensus of time-varying multiagent systems without persistent excitation assumptions,” IEEE Transactions on Automatic Control, vol. 63@bookhale2009ordinary, title=Ordinary differential equations, author=Hale, Jack K, year=2009, publisher=Courier Corporation , no. 11, pp. 3935–3939, 2018.
  • [11] J. A. Acebrón, L. L. Bonilla, C. J. Pérez Vicente, F. Ritort, and R. Spigler, “The kuramoto model: A simple paradigm for synchronization phenomena,” Reviews of modern physics, vol. 77, no. 1, pp. 137–185, 2005.
  • [12] D. V. Dimarogonas and K. J. Kyriakopoulos, “On the rendezvous problem for multiple nonholonomic agents,” IEEE Transactions on automatic control, vol. 52, no. 5, pp. 916–922, 2007.
  • [13] G.-B. Stan and R. Sepulchre, “Analysis of interconnected oscillators by dissipativity theory,” IEEE Transactions on Automatic Control, vol. 52, no. 2, pp. 256–270, 2007.
  • [14] J. Yao, Z.-H. Guan, and D. J. Hill, “Passivity-based control and synchronization of general complex dynamical networks,” Automatica, vol. 45, no. 9, pp. 2107–2113, 2009.
  • [15] W. Yu, G. Chen, and M. Cao, “Consensus in directed networks of agents with nonlinear dynamics,” IEEE Transactions on Automatic Control, vol. 56, no. 6, pp. 1436–1441, 2011.
  • [16] C. Altafini, “Consensus problems on networks with antagonistic interactions,” IEEE transactions on automatic control, vol. 58, no. 4, pp. 935–946, 2012.
  • [17] Z. Lin, B. Francis, and M. Maggiore, “State agreement for continuous-time coupled nonlinear systems,” SIAM Journal on Control and Optimization, vol. 46, no. 1, pp. 288–307, 2007.
  • [18] S.-J. Chung and J.-J. E. Slotine, “Cooperative robot control and concurrent synchronization of lagrangian systems,” IEEE transactions on Robotics, vol. 25, no. 3, pp. 686–700, 2009.
  • [19] Y. Cao, W. Yu, W. Ren, and G. Chen, “An overview of recent progress in the study of distributed multi-agent coordination,” IEEE Transactions on Industrial informatics, vol. 9, no. 1, pp. 427–438, 2012.
  • [20] W. Ren and R. W. Beard, Distributed consensus in multi-vehicle cooperative control.   Springer, 2008, vol. 27, no. 2.
  • [21] F. Dörfler and F. Bullo, “Synchronization in complex networks of phase oscillators: A survey,” Automatica, vol. 50, no. 6, pp. 1539–1564, 2014.
  • [22] Z. Lin, B. Francis, and M. Maggiore, “On the state agreement problem for multiple nonlinear dynamical systems,” in IFAC Proceedings Volumes, vol. 38, no. 1.   Elsevier, 2005, pp. 82–87.
  • [23] K. S. Narendra and A. M. Annaswamy, “Persistent excitation in adaptive systems,” International Journal of Control, vol. 45, no. 1, pp. 127–160, 1987.
  • [24] J. Cortés, “Distributed algorithms for reaching consensus on general functions,” Automatica, vol. 44, no. 3, pp. 726–737, 2008.
  • [25] L. Moreau, “Stability of continuous-time distributed consensus algorithms,” in 2004 43rd IEEE conference on decision and control (CDC), vol. 4.   IEEE, 2004, pp. 3998–4003.
  • [26] R. Sepulchre, A. Sarlette, and P. Rouchon, “Consensus in non-commutative spaces,” in 49th IEEE conference on decision and control (CDC).   IEEE, 2010, pp. 6596–6601.
  • [27] G. Birkhoff, “Extensions of Jentzsch's theorem,” Transactions of the American Mathematical Society, vol. 85, no. 1, pp. 219–227, 1957.
  • [28] Y. Kawano and M. Cao, “Contraction analysis of virtually positive systems,” Systems & Control Letters, vol. 168, p. 105358, 2022.
  • [29] G. Dirr, H. Ito, A. Rantzer, and B. Rüffer, “Separable lyapunov functions for monotone systems: Constructions and limitations,” Discrete Contin. Dyn. Syst. Ser. B, vol. 20, no. 8, pp. 2497–2526, 2015.
  • [30] A. Rantzer, “Scalable control of positive systems,” European Journal of Control, vol. 24, pp. 72–80, 2015.
  • [31] H. R. Feyzmahdavian, B. Besselink, and M. Johansson, “Stability analysis of monotone systems via max-separable lyapunov functions,” IEEE Transactions on Automatic Control, vol. 63, no. 3, pp. 643–656, 2017.
  • [32] J. Tsitsiklis, D. Bertsekas, and M. Athans, “Distributed asynchronous deterministic and stochastic gradient optimization algorithms,” IEEE transactions on automatic control, vol. 31, no. 9, pp. 803–812, 1986.
  • [33] F. Cucker and S. Smale, “Emergent behavior in flocks,” IEEE Transactions on automatic control, vol. 52, no. 5, pp. 852–862, 2007.
  • [34] E. Cristiani, P. Frasca, and B. Piccoli, “Effects of anisotropic interactions on the structure of animal groups,” Journal of mathematical biology, vol. 62, pp. 569–588, 2011.
  • [35] J. K. Hale, Ordinary differential equations.   Krieger Publishing Company, 1980.
  • [36] P. J. Bushell, “Hilbert’s metric and positive contraction mappings in a banach space,” Archive for Rational Mechanics and Analysis, vol. 52, pp. 330–338, 1973.