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

A Hybrid Observer for Estimating the State of a Distributed Linear System

Lili Wang lili.wang.zj@gmail.com    Ji Liu ji.liu@stonybrook.edu    A. Stephen Morse as.morse@yale.edu Samueli School of Engineering, University of California, Irvine Department of Electrical and Computer Engineering, Stony Brook University Department of Electrical Engineering, Yale University
Abstract

A hybrid observer is described for estimating the state of an m>0m>0 channel, nn-dimensional, continuous-time, linear system of the form x˙=Ax,yi=Cix,i{1,2,,m}\dot{x}=Ax,\;y_{i}=C_{i}x,\;i\in\{1,2,\ldots,m\}. The system’s state xx is simultaneously estimated by mm agents assuming each agent ii senses yiy_{i} and receives appropriately defined data from each of its current neighbors. Neighbor relations are characterized by a time-varying directed graph (t)\mathbb{N}(t) whose vertices correspond to agents and whose arcs depict neighbor relations. Agent ii updates its estimate xix_{i} of xx at “event times” ti1,ti2,ti3,t_{i1},t_{i2},t_{i3},\ldots using a local continuous-time linear observer and a local parameter estimator which iterates qq times during each event time interval [ti(s1),tis),s1[t_{i(s-1)},t_{is}),\;s\geq 1 to obtain an estimate of x(tis)x(t_{is}). Subject to the assumptions that none of the CiC_{i}’s are zero, the neighbor graph (t)\mathbb{N}(t) is strongly connected for all time, and the system whose state is to be estimated is jointly observable, it is shown that for any number λ>0\lambda>0, it is possible to choose qq and the local observer gains so that each estimate xix_{i} converges to xx at least as fast as eλte^{-\lambda t} does. This result holds whether or not agents communicate synchronously, although in the asynchronous case it is necessary to assume that (t)\mathbb{N}(t) changes in a suitably defined sense. Exponential convergence is also assured if the event time sequences of the mm agents are slightly different than each other, although in this case only if the system being observed is exponentially stable; this limitation however, is primarily a robustness issue shared by all state estimators, centralized or not, which are operating in “open loop” in the face of small modeling errors. The result also holds facing abrupt changes in the number of vertices and arcs in the inter-agent communication graph upon which the algorithm depends.

keywords:
Hybrid Systems; Distributed Observer; Robustness; Resilience.
thanks: This paper was not presented at any IFAC meeting. Portions of this paper were presented, in abbreviated form and without proofs, at the 2017 IEEE Conference on Decision and Control [1]. Corresponding author: L. Wang.

, ,

1 Introduction

In [2] a distributed observer is described for estimating the state of an m>0m>0 channel, nn-dimensional, continuous-time, jointly observable linear system of the form x˙=Ax,yi=Cix,i{1,2,,m}\dot{x}=Ax,\;y_{i}=C_{i}x,\;i\in\{1,2,\ldots,m\}. The state xIRnx\in{\rm I\!R}^{n} is simultaneously estimated by mm agents assuming that each agent ii senses yiy_{i} and receives the state of each of its neighbors’ estimates. An attractive feature of the observer described in [2] is that it is able to generate an asymptotically correct estimate of xx exponentially fast at a pre-assigned rate, if each agent’s set of neighbors do not change with time and the neighbor graph characterizing neighbor relations is strongly connected. However, a shortcoming of the observer in [2] is that it is unable to function correctly if the network changes with time. Changing neighbor graphs will typically occur if the agents are mobile. A second shortcoming of the observer described in [2] is that it is “fragile” by which we mean that the observer is not able to cope with the situation when there is an arbitrary abrupt change in the topology of the neighbor graph such as the loss or addition of a vertex or an arc. For example, if because of a component failure, a loss of battery power, or some other reasons, an agent drops out of the network, what remains of the overall observer will typically not be able to perform correctly and may become unstable, even if joint observability is not lost and what remains of the neighbor graph is still strongly connected.

This paper breaks new ground by introducing a hybrid distributed observer which overcomes the aforementioned difficulties without making restrictive assumptions. To the best of our knowledge, this observer is the first provably correct distributed algorithm capable of generating an asymptotically correct estimate of a jointly observable linear system’s state in the presence of a neighbor graph which changes with time under reasonably general assumptions. Although the observer is developed for continuous-time systems, it can very easily be modified in the obvious way to deal with discrete-time systems.

Notation: Given a collection of n×nn\times n matrices, A1,A2,,AmA_{1},\;A_{2},\;\ldots,\;A_{m}, let diagonal{A1,A2,,Am}\text{diagonal}\{A_{1},A_{2},\ldots,A_{m}\} be the block diagonal matrix with AkA_{k} as its kkth diagonal block. Given a collection of n×1n\times 1 vectors, v1,v2;,vmv_{1},\;v_{2};\;\ldots,\;v_{m}, let column{v1,v2,,v}\text{column}\{v_{1},v_{2},\ldots,v_{\}} be the stacked vector with vkv_{k} as its vvth sub-vector. For an n×nn\times n matrix AA, we let imageA\text{image}A denote the linear subspace spanned by matrix AA. For two n×nn\times n matrix A1A_{1}, and A2A_{2}, we let imageA2imageA2\text{image}A_{2}\cap\text{image}A_{2} denote the intersection of the two images.

1.1 The Problem

We are interested in a network of m>0m>0 autonomous agents labeled 1,2,,m1,2,\ldots,m which are able to receive information from their “neighbors” where by the neighbor of agent ii is meant any agent who is in agent ii’s reception range. We write 𝒩i(t)\mathcal{N}_{i}(t) for the set of labels of agent ii’s neighbors at real {continuous} time tt and always take agent ii to be a neighbor of itself. Neighbor relations at time tt are characterized by a directed graph (t)\mathbb{N}(t) with mm vertices and a set of arcs defined so that there is an arc from vertex jj to vertex ii whenever agent jj is a neighbor of agent ii. Since each agent ii is always a neighbor of itself, (t)\mathbb{N}(t) has a self-arc at each of its vertices. Each agent ii can sense a continuous-time signal yiIRsi,i𝐦=Δ{1,2,,m}y_{i}\in{\rm I\!R}^{s_{i}},\;i\in\mathbf{m}\stackrel{{\scriptstyle\Delta}}{{=}}\{1,2,\ldots,m\}, where

yi\displaystyle y_{i} =\displaystyle= Cix,i𝐦\displaystyle C_{i}x,\;\;\;i\in\mathbf{m} (1)
x˙\displaystyle\dot{x} =\displaystyle= Ax\displaystyle Ax (2)

and xIRnx\in{\rm I\!R}^{n}. It is assumed throughout that the system defined by (1) and (2) is jointly observable; i.e., with C=[C1C2Cm]C=[C_{1}^{\prime}\;C_{2}^{\prime}\;\cdots\;C_{m}^{\prime}]^{\prime}, the matrix pair (C,A)(C,A) is observable. For simplicity, it is further assumed that Ci0,i𝐦C_{i}\neq 0,\;i\in\mathbf{m}; generalization to deal with the case when this assumption does not hold is straight forward. The problem of interest is to develop “private estimators”, one for each agent, which, under ideal conditions without modeling or synchronization errors, enable each agent to obtain an estimate of xx which converges to xx exponentially fast at a pre-assigned rate .

1.2 Background

The distributed state estimation problem has been under study in one form or another for years. The problem has been widely studied as a distributed Kalman filter problem [3, 4, 5, 6, 7, 8, 9, 10]. A form of distributed Kalman filtering is introduced in [3] for discrete-time linear systems; the underlying idea is to switch back and forth between conventional state estimation and a data fusion computation. This approach is extended to continuous-time systems in [4]. There are two key limitations of the ideas presented in [3, 4]. First, it is implicitly assumed in each paper that data fusion {i.e., consensus} can be attained in finite time. Second, it is also implicitly assumed that each pair (Ci,A)(C_{i},A) is observable; this restrictive assumption is needed in order to guarantee that each local error covariance matrix Riccati equation has a solution. Both papers also include assumptions about graph connectivity and information exchange which are more restrictive than they need be.

Discrete-time distributed observers have recently appeared in [11, 12, 13, 14, 15, 16, 17, 18, 19, 20]. None of these estimators admit continuous-time extensions. The algorithm in [11] works for fixed graphs with a relatively complicated topology design by studying the roles of each agent in the network. The distributed observer proposed in [12] can track the system only if the so-called Scalar Tracking Capacity condition is satisfied. Noteworthy among these is the paper [15] which described a discrete-time linear system which solves the estimation problem for jointly observable, discrete-time systems with fixed neighbor graphs assuming only that the neighbor graph is directed and strongly connected. This is done by recasting the estimation problem as a classical decentralized control problem [21, 22]. Although these observers are limited to to discrete-time systems, it has proved possible to make use of the ideas in [15] to obtain a distributed observer for continuous-time systems [2]. In particular, [2] explains how to construct a distributed observer for a continuous-time system with a strongly connected neighbor graph, which is capable of estimating state exponentially fast at a pre-assigned rate. It is straightforward to modify this observer to deal with discrete-time systems.

An interesting idea, suggested in [23], seeks to simplify the structure of a distributed estimator for a continuous-time system at the expense of some design flexibility. This is done, in essence, by exploiting the AA-invariance of the unobservable spaces of the pairs (Ci,A)(C_{i},A); this in turn enables one to “split” the estimators into two parts, one based on conventional spectrum assignment techniques and the other based on consensus [23, 24, 25, 26, 27]. Reference [23] addresses the problem in continuous time for undirected, connected neighbor graphs. The work of [24, 25] extends the result of [23] to the case when the neighbor graph is directed and strongly connected. Establishing correctness requires one to choose gains to ensure that certain LMIs hold. In [27], motivated by the distributed least squares solver problem, a modified algorithm which can deal with measurement noise is proposed . In [26] a simplified version of the ideas in [24] is presented. Because the “high gain” constructions used in [24] and [26] don’t apply in discrete-time, significant modifications are required to exploit these ideas in a discrete-time context [28].

Despite the preceding advances, until the appearance of [1], which first outlines the idea presented in this paper, there were almost no results for doing state estimation with time varying neighbor graphs for either discrete-time or continuous-time linear systems. For sure, there were a few partial results. For example, [17] suggests a distributed observer using a consensus filter for the state estimation of discrete-time linear distributed system for specially structured, undirected neighbor graphs. Another example, in [18], an HH_{\infty} based observer is described which is intended to function in the face of a time-varying graph with a Markovian randomly varying topology. It is also worth mentioning [29] which tackles the challenging problem of trying to define a distributed observer which can function correctly in the face of intermittent disruptions in available information. Although the problem addressed in [29] is different than the problem to which this paper is addressed, resilience in the face of intermittent disruptions is to some extent similar to the notion of resilience addressed in this paper.

The first paper to provide a definitive solution to the distributed state estimation problem for time varying neighbor graphs under reasonably relaxed assumptions was presented, in abbreviated form at the 2017 IEEE Conference on Decision and Control [1]. The central contribution of [1] and this paper is to describe a distributed observer for a jointly observable, continuous-time linear system with a time-varying neighbor graph \mathbb{N} which is capable of estimating the system’s state exponentially fast at any prescribed rate. Assuming “synchronous operation”, the only requirement on the graph is that it be strongly connected for all time.

Since the appearance of [1], several other distributed observers have been suggested which are capable of doing state estimation in the face of changing neighbor graphs. For example, expanding on earlier work in [16], [30] provides a procedure for constructing such an observer which exploits in some detail the structure of \mathbb{N} and its relation to the structure of the data matrices defining the system. The resulting algorithm, which is tailored exclusively to discrete-time systems, deals with state estimation under assumptions which are weaker than strong connectivity. Recently we have learned that the split spectrum observer idea first proposed in [23] and later simplified in [24] and [26] can be modified to deal with strongly-connected time-varying neighbor graphs, although only for continuous time systems. See [31] for an unpublished report on the subject.

1.3 Organization

The remainder of the paper is organized as follows. The hybrid observer itself is described in §2 subject to the assumption that all mm agents share the same event time sequence. Two cases are considered, one in which the interchanges of information between agents are performed synchronously and the other case being when it is not. The synchronous case is the one most comparable to the versions of the distributed observer problem treated in [3] - [18]. The main result for this case is Theorem 22 which asserts that so long as the neighbor graph is strongly connected for all time, exponential convergence to zero at a prescribed convergence rate of all mm state estimation errors is achieved. This is a new result which has no counterpart in any of the previously cited references. The same result is achieved in the asynchronous case {cf. Theorem 2}, but to reach this conclusion it is necessary to assume that the neighbor graph changes in a suitably defined sense111It is worth noting at this point that many of the subtleties of asynchronous operation are obscured or at least difficult to recognize in a discrete-time setting where there is invariably a single underlying discrete-time clock shared by all mm agents. These two theorems are the main contributions of this paper. Their proofs can be found in §3.

The aim of §4 is to explain what happens if the assumption that all mm agents share the same event time sequence is not made. For simplicity, this is only done for the case when differing event time sequences are the only cause of asynchronism. As will be seen, the consequence of event-time sequence mismatches turns out to be more of a robustness issue than an issue due to unsynchronized operation. In particular, it will become apparent that if different agents use slightly different event time sequences then asymptotically correct state estimates will not be possible unless AA is a stability matrix, i.e., all the eigenvalues of matrix AA have strictly negative parts. While at first glance this may appear to be a limitation of the distributed observer under consideration, it is in fact a limitation of virtually all state estimators, distributed or not, which are not used in feedback-loops. Since this easily established observation is apparently not widely appreciated, an explanation is given at the end of the section.

By a (passively) resilient algorithm for a distributed process is meant an algorithm which, by exploiting built-in network and data redundancies, is able to continue to function correctly in the face of abrupt changes in the number of vertices and arcs in the inter-agent communication graph upon which the algorithm depends. In §5, it is briefly explained how to configure things so that that the proposed estimator can cope with the situation when there is an arbitrary abrupt change in the topology of the neighbor graph such as the loss or addition of an arc or a vertex provided connectivity is not lost in an appropriately defined sense. Dealing with a loss or addition of an arc proves to to be easy to accomplish because of the ability of the estimator to deal with time-varying graphs. Dealing with the loss or addition of a vertex is much more challenging and for this reason only preliminary results are presented. Finally in §6 simulation results are provided to illustrate the observer’s performance.

2 Hybrid Observer

The overall hybrid observer to be considered consists of mm private estimators, one for each agent. Agent ii’s private estimator, whose function is to generates an estimate xix_{i} of xx, is a hybrid dynamical system consisting of a “local observer” and a “local parameter estimator.” The purpose of local observer ii is to generate an asymptotically correct estimate of LixL_{i}x where LiL_{i} is any pre-specified, full-rank matrix whose kernel equals the kernel of the observability matrix of the pair (Ci,A)(C_{i},A); roughly speaking, LixL_{i}x can be thought of as that “part of xx” which is observable to agent ii. Agent ii’s local observer is then an nin_{i}-dimensional continuous-time, linear system of the form

w˙i=(A¯i+KiC¯i)wiKiyi\dot{w}_{i}=(\bar{A}_{i}+K_{i}\bar{C}_{i})w_{i}-K_{i}y_{i} (3)

where ni=rankLin_{i}={\rm rank\;}L_{i}, KiK_{i} is a gain matrix to be specified, and C¯i\bar{C}_{i} and A¯i\bar{A}_{i} are unique solutions to the equations Ci=C¯iLiC_{i}=\bar{C}_{i}L_{i} and LiA=A¯iLiL_{i}A=\bar{A}_{i}L_{i}, respectively. As is well known, the pair (C¯i,A¯i)(\bar{C}_{i},\bar{A}_{i}) is observable and the local observer estimation error e¯i=ΔwiLix\bar{e}_{i}\stackrel{{\scriptstyle\Delta}}{{=}}w_{i}-L_{i}x satisfies

e¯i(t)=e(A¯i+KiC¯i)te¯i(0),t[0,)\bar{e}_{i}(t)=e^{(\bar{A}_{i}+K_{i}\bar{C}_{i})t}\bar{e}_{i}(0),\;\;\;t\in[0,\infty)

Since (C¯i,A¯i)(\bar{C}_{i},\bar{A}_{i}) is an observable pair, KiK_{i} can be selected so that e¯i(t)\bar{e}_{i}(t) converges to 0 exponentially fast at any pre-assigned rate. We assume that each KiK_{i} is so chosen. Since

wi(t)=Lix(t)+e¯i(t),i𝐦,t[0,)w_{i}(t)=L_{i}x(t)+\bar{e}_{i}(t),\;\;\;i\in\mathbf{m},\;\;\;t\in[0,\infty) (4)

wiw_{i} can be viewed as a signal which approximates LixL_{i}x in the face of exponentially decaying additive noise, namely e¯i\bar{e}_{i}.

The other sub-system comprising agent ii’s private estimator, is a “local parameter estimator” whose function is to generate estimates of xx at each of agent ii’s preselected event times ti1,ti2,t_{i1},t_{i2},\ldots. Here ti0,ti1,ti2,t_{i0},t_{i1},t_{i2},\ldots is an ascending sequence of event times with a fixed spacing of T>0T>0 time units between any two successive event times. In this section it is assume that ti0=0,i𝐦t_{i0}=0,\;i\in\mathbf{m}, and consequently that all event time sequences are the same222It is easy to generalize the results in this section to the case when event times are not evenly spaced provided that the spacings between successive pairs of event times remains positive and bounded.. Thus tis=sT,s0,i𝐦.t_{is}=sT,\;s\geq 0,\;\;\;i\in\mathbf{m}. Between event times, each xix_{i} is generated using the equation

x˙i=Axi\dot{x}_{i}=Ax_{i} (5)

Motivation for the development of the local parameter estimator whose purpose is to enable agent ii to estimate x(ti(s1))x(t_{i(s-1)}) over the event time interval [ti(s1),tis)[t_{i(s-1)},\;t_{is}), stems from the fact that the equations

wj(ti(s1))=Ljp+e¯j(ti(s1)),j𝐦w_{j}(t_{i(s-1)})=L_{j}p+\bar{e}_{j}(t_{i(s-1)}),\;\;\;\;j\in\mathbf{m}

admit a unique solution, namely p=x(ti(s1))p=x(t_{i(s-1)}). Existence follows from (4) whereas uniqueness is a consequence of the assumption of joint observability.

The existence and uniqueness of pp suggest that an approximate value of x(ti(s1))x(t_{i(s-1)}) can be obtained after a finite number of iterations - say qq - using the linear equation solver discussed in[32]. Having obtained such an approximate value of x(ti(s1))x(t_{i(s-1)}), denoted below by zis(q)z_{is}(q), the desired estimate of x(tis)x(t_{is}) can be taken as

xi(tis)=ΔeATzis(q)x_{i}(t_{is})\stackrel{{\scriptstyle\Delta}}{{=}}e^{AT}z_{is}(q) (6)

This is the architecture which will be considered.
The computations needed to update agent ii’s estimate of x(ti(s1))x(t_{i(s-1)}) are carried out by agent ii during the event time interval [ti(s1),tis)[t_{i(s-1)},t_{is}). This is done using a local parameter estimator which generates a sequence of qq auxiliary states zis(1),zis(2),,zis(q)z_{is}(1),z_{is}(2),\ldots,z_{is}(q) where q>0q>0 is a positive integer to be specified below. The sequence is initialized by setting

zis(0)=xi(ti(s1)),z_{is}(0)=x_{i}(t_{i(s-1)}), (7)

and is recursively updated by agent ii at local iteration times τis(k),k𝐪=Δ{1,2,,q},\tau_{is}(k),k\in\mathbf{q}\stackrel{{\scriptstyle\Delta}}{{=}}\{1,2,\ldots,q\}, known only to agent ii. It is assumed that the τis(k)\tau_{is}(k) together with the initialization τis(0)\tau_{is}(0) are of the form

τis(k)=ti(s1)+kΔ+δis(k),k{0,1,,q}\tau_{is}(k)=t_{i(s-1)}+k\Delta+\delta_{is}(k),\;\;\;k\in\{0,1,\ldots,q\} (8)

where δis(0),δis(1),δis(2),,δis(q)\delta_{is}(0),\delta_{is}(1),\delta_{is}(2),\ldots,\delta_{is}(q) is a sequence of small deviations which satisfy

δis(k)[ϵi,ϵi],k{0,1,,q},\delta_{is}(k)\in[-\epsilon_{i},\epsilon_{i}],\;\;\;k\in\{0,1,\ldots,q\}, (9)

Here ϵi\epsilon_{i} is a small nonnegative number whose constraints will be described below and Δ\Delta is a positive number satisfying

Δq+maxi{ϵi,i𝐦}T\Delta q+\max_{i}\{\epsilon_{i},\;i\in\mathbf{m}\}\leq T

The signal zis(q)z_{is}(q) is agent ii’s updated estimate of x(ti(s1))x(t_{i(s-1)}) and is used to define xi(tis)x_{i}(t_{is}) as in (6).
The transfer of information between agents which is needed to generate the zis(k)z_{is}(k), is carried out asynchronously as follows. For k𝐪k\in\mathbf{q} and j𝐦j\in\mathbf{m}, agent jj broadcasts zjs(k1)z_{js}(k-1) at time τjs(k1)+β\tau_{js}(k-1)+\beta where β\beta is any prescribed nonnegative number chosen smaller than Δ\Delta. It is assumed that the bounds ϵi,i𝐦\epsilon_{i},\;i\in\mathbf{m}, appearing in (9) are small enough so that there exist β\beta and Δ\Delta satisfying

ϵi+ϵjβ,andϵi+ϵj+β<Δ,i,j𝐦\epsilon_{i}+\epsilon_{j}\leq\beta,\;\;\;\;\;\text{and}\;\;\;\;\;\epsilon_{i}+\epsilon_{j}+\beta<\Delta,\;\;i,j\in\mathbf{m} (10)

These inequalities ensure that for k𝐪k\in\mathbf{q}, a broadcast by any agent jj at time τjs(k1)+β\tau_{js}(k-1)+\beta will occur within the reception interval [τis(k1),τis(k))[\tau_{is}(k-1),\;\;\tau_{is}(k)) of agent ii. Fig. 1 provides an example of the update and communication times of two different different agents ii and jj.

Refer to caption

Figure 1: A broadcast by any agent jj at time τjs(k1)+β\tau_{js}(k-1)+\beta will occur within the reception interval [τis(k1),τis(k))[\tau_{is}(k-1),\;\;\tau_{is}(k)) of agent ii.

Accordingly, agent jj is a data source or just source for agent ii on [τis(k1),τis(k))[\tau_{is}(k-1),\;\;\tau_{is}(k)) if agent jj is in the reception range of agent ii at time τjs(k1)+β\tau_{js}(k-1)+\beta. Let 𝒮is(k)\mathcal{S}_{is}(k) denote the set of labels of such agents; that is

𝒮is(k)={j:j𝒩i(τjs(k1)+β)}\mathcal{S}_{is}(k)=\{j:j\in\mathcal{N}_{i}(\tau_{js}(k-1)+\beta)\} (11)

Note that i𝒮is(k)i\in\mathcal{S}_{is}(k), for all i𝐦i\in\mathbf{m} so 𝒮is(k)\mathcal{S}_{is}(k) is never empty. Clearly agent ii can use the signals zjs(k1),j𝒮is(k)z_{js}(k-1),\;j\in\mathcal{S}_{is}(k), to compute zis(k)z_{is}(k).

Prompted by [32], the update equation used to recursively generate the zis(k)z_{is}(k) during agent ii’s ssth event time interval [ti(s1),tis)[t_{i(s-1)},\;t_{is}) is given by

zis(k)=z¯is(k1)Qi(Liz¯is(k1)wi(ti(s1))),k𝐪z_{is}(k)=\bar{z}_{is}(k-1)\\ -Q_{i}(L_{i}\bar{z}_{is}(k-1)-w_{i}(t_{i(s-1)})),\;\;\;\;k\in\mathbf{q}\;\; (12)

where Qi=Li(LiLi)1Q_{i}=L_{i}^{\prime}(L_{i}L_{i}^{\prime})^{-1}, z¯is(k1)\bar{z}_{is}(k-1) is an averaged state

z¯is(k1)=1mis(k)j𝒮is(k)zjs(k1),\bar{z}_{is}(k-1)=\frac{1}{m_{is}(k)}\sum_{j\in\mathcal{S}_{is}(k)}z_{js}(k-1), (13)

and mis(k)m_{is}(k) is the number of labels in 𝒮is(k)\mathcal{S}_{is}(k). The overall private estimator for agent ii is thus described by the equations (3), (5) - (8) and (11) - (13). In summary, initialize xi(ti0)x_{i}(t_{i0}), wi(0)w_{i}(0) randomly. For t[ti0,ti1)t\in[t_{i0},t_{i1}), x˙i=Axi\dot{x}_{i}=Ax_{i}. Then for s=1,2,s=1,2,\ldots, the algorithm of the hybrid estimator for anget ii is shown in Algorithm 1.

Algorithm 1 The hybrid estimator of agent ii
1:Initialize xi(ti0),wi(0),Ki,(C¯i,Ai),Li,qx_{i}(t_{i0}),\;w_{i}(0),\;K_{i},\;(\bar{C}_{i},A_{i}),\;L_{i},\;q
2:w˙i=(A¯i+KiC¯i)wiKiyi\dot{w}_{i}=(\bar{A}_{i}+K_{i}\bar{C}_{i})w_{i}-K_{i}y_{i}
3:for t[ti0,ti1)t\in[t_{i0},t_{i1}) do
4:     x˙i=Axi\dot{x}_{i}=Ax_{i} with xi(ti0)x_{i}(t_{i0})
5:end for
6:for s=1,2,s=1,2,\ldots do
7:     zis(0)=xi(ti(s1))z_{is}(0)=x_{i}(t_{i(s-1)})
8:     for  k=1:qk=1:q do
9:         Agent ii gets the sampled value wi(ti(s1))w_{i}(t_{i(s-1)}) from its own estimator, and receives zjs(k1)z_{js}(k-1) from its neighbor jj.
zis(k)=z¯is(k1)Qi(Liz¯is(k1)wi(ti(s1)))z_{is}(k)=\bar{z}_{is}(k-1)-Q_{i}(L_{i}\bar{z}_{is}(k-1)-w_{i}(t_{i(s-1)}))
where z¯is\bar{z}_{is} is as defined in Eq. (13)
10:     end for
11:     for t[tis,ti(s+1))t\in[t_{is},t_{i(s+1)}) do
12:         x˙i=Axi\dot{x}_{i}=Ax_{i} with xi(tis)=eATzis(q)x_{i}(t_{is})=e^{AT}z_{is}(q)
13:     end for
14:end for
15:Output: xix_{i}

In order to complete the definition of the hybrid observer, it is necessary to specify values of the KiK_{i} and qq. Towards this end, suppose that as a design goal it is desired to pick the KiK_{i} and qq so that all mm state estimation errors

ei=Δxix,i𝐦e_{i}\stackrel{{\scriptstyle\Delta}}{{=}}x_{i}-x,\;i\in\mathbf{m} (14)

converge to zero as fast as eλte^{-\lambda t} does where λ>0\lambda>0 is some desired convergence rate. The KiK_{i} would then have to be chosen using spectrum assignment or some other technique so that the matrix exponentials e(A¯i+KiC¯i)te^{(\bar{A}_{i}+K_{i}\bar{C}_{i})t} all converge to zero at least as fast as eλte^{-\lambda t} does. This of course can be accomplished because each matrix pair (C¯i,A¯i)(\bar{C}_{i},\bar{A}_{i}) is observable. In the sequel it will be assumed that for some preselected positive number λ¯>λ\bar{\lambda}>\lambda, the KiK_{i} have been chosen so that for i𝐦i\in\mathbf{m} the local observer estimation errors satisfy

e¯i(t)cieλ¯(tμ)e¯i(μ),tμ0||\bar{e}_{i}(t)||\leq c_{i}e^{-\bar{\lambda}(t-\mu)}||\bar{e}_{i}(\mu)||,\;\;\;t\geq\mu\geq 0 (15)

where the ci,i𝐦c_{i},\;i\in\mathbf{m} are nonnegative constants and ||||||\cdot|| denotes the two-norm.

To describe how to define an appropriate value of qq to attain the desired convergence rate for the state estimation errors ei,i𝐦e_{i},\;i\in\mathbf{m}, it is necessary to take some preliminary steps. First, for each i𝐦i\in\mathbf{m}, let PiP_{i} denote the orthogonal projection on the unobservable space of (Ci,Ai)(C_{i},A_{i}). It is easy to see that Pi=ILi(LiLi)1Li,i𝐦P_{i}=I-L_{i}^{\prime}(L_{i}L_{i}^{\prime})^{-1}L_{i},\;\;i\in\mathbf{m}. Moreover, because of the assumption of joint observability,

i𝐦imagePi={0}\bigcap_{i\in\mathbf{m}}\text{image}\;P_{i}=\{0\} (16)

Next, let 𝒞\mathcal{C} denote the set of all products of the form Pi1Pi2Pi(m1)2+1P_{i_{1}}P_{i_{2}}\cdots P_{i_{(m-1)^{2}+1}} where each projection matrix in {Pi:i𝐦}\{P_{i}:i\in\mathbf{m}\} occurs in each of such product at least once. Note that 𝒞\mathcal{C} is a closed subset of IRn×n{\rm I\!R}^{n\times n}. Since each projection matrix Pi,i𝐦P_{i},\;i\in\mathbf{m} has a two-norm which is no greater than 11, each matrix M𝒞M\in\mathcal{C} has a two-norm less than or equal to 11. Thus 𝒞\mathcal{C} is also a bounded and thus compact subset. In fact, each product in 𝒞\mathcal{C} actually has two-norm strictly less than 11. This is a consequence of (16) and the requirement that each matrix in {Pi:i𝐦}\{P_{i}:i\in\mathbf{m}\} must occur in each product in 𝒞\mathcal{C} at least once {Lemma 2, [33]}. These observations imply that maximum of the two-norms of the matrices in 𝒞\mathcal{C}, namely

ρ=Δmax{M:M𝒞},\rho\stackrel{{\scriptstyle\Delta}}{{=}}\max\{||M||:M\in\mathcal{C}\}, (17)

exists and is a real non-negative number strictly less than 11 333It is worth noting that although the matrices LiL_{i} used in defining the PiP_{i} are not uniquely determined by the unobservable spaces of the pairs (Ci,A)(C_{i},A), the orthogonal projection matrices PiP_{i} nonetheless are. Thus the set 𝒞\mathcal{C} used in the definition of ρ\rho in (17) ultimately depends only on the family of unobservable spaces of the pairs (Ci,A),i𝐦(C_{i},A),\;i\in\mathbf{m} and not on the particular manner in which the LiL_{i} are chosen. Just how to explicitly characterize this dependence is a topic for future research.. This in turn implies that the attenuation constant

α=Δ1(m1)(1ρ)m(m1)2\alpha\stackrel{{\scriptstyle\Delta}}{{=}}1-\frac{(m-1)(1-\rho)}{m^{(m-1)^{2}}} (18)

is also a real non-negative number strictly less than 11. As will become evident below {cf. (40) and Lemma 5}, in the idealized case when all ϵi\epsilon_{i} and e¯i\bar{e}_{i} are zero, for any integer p>0p>0 and any given value of qq satisfying

qp((m1)2+1),q\geq p((m-1)^{2}+1), (19)

the value of the signal

max{zis(k)x(ti(s1)):i𝐦}\max\{||z_{is}(k)-x(t_{i(s-1)})||:i\in\mathbf{m}\}

is attenuated by at least a factor αp\alpha^{p} after qq iterations during each event - time interval [ti(s1),tis)[t_{i(s-1)},\;t_{is}); i.e., for s1s\geq 1,

max{zis(q)x(ti(s1)):i𝐦}\displaystyle\max\{||z_{is}(q)-x(t_{i(s-1)})||:i\in\mathbf{m}\}\leq
αpmax{zis(0)x(ti(s1)):i𝐦}\displaystyle\alpha^{p}\max\{||z_{is}(0)-x(t_{i(s-1)})||:i\in\mathbf{m}\}

It will soon be apparent, if it is not already from (6), (7) and (14), that over each event- time interval [ti(s1),tis)[t_{i(s-1)},\;t_{is}),

max{ei(tis):i𝐦}\displaystyle\max\{||e_{i}(t_{is})||:i\in\mathbf{m}\}\leq (20)
eATαpmax{ei(ti(s1)):i𝐦}\displaystyle e^{||A||T}\alpha^{p}\max\{||e_{i}(t_{i(s-1)})||:i\in\mathbf{m}\}

Since each event - time interval is of length TT, to achieve an exponential convergence rate of λ\lambda in the idealized case, it is necessary to pick qq so that (19) holds where pp is any integer satisfying eATαp<eλTe^{||A||T}\alpha^{p}<e^{-\lambda T}. In other words, the requirement on qq is that (19) hold where

p>(λ+A)Tln(1α),p>\left\lceil\frac{(\lambda+||A||)T}{\ln(\frac{1}{\alpha})}\right\rceil, (21)

with r\lceil r\rceil here denoting, for any nonnegative number rr, the smallest integer krk\geq r. The following theorem, which applies to the synchronous case when all of the ϵi\epsilon_{i} are zero, {but not necessarily the e¯i\bar{e}_{i}} summarizes these observations.

Theorem 1.

Synchronous case: Suppose ϵi=0,i𝐦\epsilon_{i}=0,\;\;i\in\mathbf{m}, and that the neighbor graph (t)\mathbb{N}(t) is strongly connected for all tt. Let ρ\rho and α\alpha be defined by (17) and (18) respectively. Then each state estimation ei=xix,i𝐦e_{i}=x_{i}-x,\;i\in\mathbf{m}, of the hybrid observer defined by (3), (5) - (8) and (11) - (13), tends to zero as fast as eλte^{-\lambda t} does provided qq satisfies

q>((m1)2+1)(λ+A)Tln(1α)q>((m-1)^{2}+1)\left\lceil\frac{(\lambda+||A||)T}{\ln(\frac{1}{\alpha})}\right\rceil (22)

This theorem will be proved in the next section. Several comments are in order. First, the attenuation of max{ei:i𝐦}\max\{||e_{i}||:i\in\mathbf{m}\} by αp\alpha^{p} over an event time interval is not likely to be tight and a larger attenuation constant can almost certainly be expected. This is important because the larger the attenuation constant the smaller the required value of qq needed to achieve a given convergence rate. Second, the hypothesis that (t)\mathbb{N}(t) strongly connected is almost certainly stronger than is necessary, the notion of a repeatedly jointly strongly connected sequence of graphs [33] being a likely less stringent alternative.

To deal with the asynchronous case when at least some of the ϵi\epsilon_{i} are nonzero, it is necessary to assume that (t)\mathbb{N}(t) is constant on each of the time intervals

s(k)=[ϵ+sT+(k1)Δ+β,ϵ+sT+(k1)Δ+β],k𝐪,s1\mathcal{I}_{s}(k)=[-\epsilon+sT+(k-1)\Delta+\beta,\epsilon+sT\\ +(k-1)\Delta+\beta],\;k\in\mathbf{q},\;s\geq 1 (23)

where

ϵ=max{ϵi:i𝐦}\epsilon=\max\{\epsilon_{i}:i\in\mathbf{m}\} (24)

For this assumption to make sense, these intervals cannot overlap. The following lemma establishes that this is in fact the case.

Lemma 1.

Suppose that q2q\geq 2 and that the ϵi\epsilon_{i} are fixed nonnegative numbers satisfying the constraints in (10). Then for each s0s\geq 0, the qq time intervals defined by (23) are non-overlapping and each is a subinterval of [sT,(s+1)T)[sT,(s+1)T).

Proof of Lemma 1: Fix 2kq2\leq k\leq q. Note that 2ϵ<β2\epsilon<\beta because of (10). This implies that ϵ+sT+(k2)Δ+β<ϵ+sT+(k1)Δ+β\epsilon+sT+(k-2)\Delta+\beta<-\epsilon+sT+(k-1)\Delta+\beta and thus that s(k)\mathcal{I}_{s}(k) and s(k1)\mathcal{I}_{s}(k-1) are disjoint. Since this holds for all kk satisfying 2kq2\leq k\leq q, all s(k),k𝐪\mathcal{I}_{s}(k),\;k\in\mathbf{q} are disjoint.

From (10), ϵβ\epsilon\leq\beta and ϵ+β<Δ\epsilon+\beta<\Delta. These inequalities imply that ϵ+sT+βsT-\epsilon+sT+\beta\geq sT and ϵ+sT+(q1)Δ+β<(s+1)T\epsilon+sT+(q-1)\Delta+\beta<(s+1)T respectively. From this it follows that s(1)[sT,(s+1)T)\mathcal{I}_{s}(1)\subset[sT,(s+1)T), that s(q)[sT,(s+1)T)\mathcal{I}_{s}(q)\subset[sT,(s+1)T) and thus that s(k)[sT,(s+1)T),k𝐪\mathcal{I}_{s}(k)\subset[sT,(s+1)T),\;\;k\in\mathbf{q}.  

We are led to the asynchronous version of Theorem 22.

Theorem 2.

Asynchronous case: Suppose the ϵi,i𝐦\epsilon_{i},\;i\in\mathbf{m}, satisfy (10) and that the neighbor graph (t)\mathbb{N}(t) is constant on each interval s(k),k𝐪,s1\mathcal{I}_{s}(k),\;k\in\mathbf{q},\;s\geq 1 and strongly connected for all tt. Let ρ\rho and α\alpha be defined by (17) and (18) respectively and suppose that qq satisfies (22). Then as in Theorem 22, each state estimation ei=xix,i𝐦e_{i}=x_{i}-x,\;i\in\mathbf{m}, of the hybrid observer defined by(3), (5) - (8) and (11) - (13), tends to zero as fast as eλte^{-\lambda t} does.

The proof of this theorem will be given in the next section. Notice that the asynchronous case here can not be recognized in a discrete-time setting with a discrete-time clock shared by all mm agents considering delays[34].

2.1 A special case

It is possible to relax somewhat the lower bound (22) for qq to achieve exponential convergence in the special case when the neighbor graph (t)\mathbb{N}(t) is symmetric and strongly connected for all tt. This can be accomplished by replacing the straight averaging rule defined by (13), with the convex combination rule

z¯is(k1)=(1mis(k)m+1)zis(k1)+1m+1j𝒮is(k)zjs(k1)\bar{z}_{is}(k-1)=(1-\frac{m_{is}(k)}{m+1})z_{is}(k-1)+\frac{1}{m+1}\!\!\!\!\sum_{j\in\mathcal{S}_{is}(k)}\!\!\!\!z_{js}(k-1) (25)

where mis(k)m_{is}(k) is the number of labels in 𝒮is(k)\mathcal{S}_{is}(k).

To proceed, let 𝒢\mathcal{G} denote the set of all symmetric and strongly connected, graphs on mm vertices. Each graph 𝔾𝒢\mathbb{G}\in\mathcal{G} uniquely determines a matrix M𝔾=I1m+1L𝔾M_{\mathbb{G}}=I-\frac{1}{m+1}L_{\mathbb{G}} where L𝔾L_{\mathbb{G}} is the Laplacian of the simple, weakly connected {undirected} graph determined by 𝔾\mathbb{G}. It is easy to see that M𝔾M_{\mathbb{G}} is a symmetric, doubly stochastic matrix with positive diagonals and that 𝔾\mathbb{G} is its graph. The connection between these matrices and the update rule defined by (25) will become apparent later when assumptions are made which enable us to identify the subsets 𝒮is(k)\mathcal{S}_{is}(k) appearing in (25) with the neighbor sets of the neighbor graph (((s1)T+(k1)δ+β)\mathbb{N}(((s-1)T+(k-1)\delta+\beta) {c.f. Lemmas 36 and 3}. Later in this paper it will also be shown that P(M𝔾I)P(M_{\mathbb{G}}\otimes I) is a contraction in the two norm {Lemma 6}. This means that

σ=max𝔾𝒢P(M𝔾I)\sigma=\max_{\mathbb{G}\in\mathcal{G}}||P(M_{\mathbb{G}}\otimes I)|| (26)

is a nonnegative number less than one.

As will become clear, to achieve a convergence rate of λ\lambda, it is sufficient to pick qq large enough to that eATσq<eλTe^{||A||T}\sigma^{q}<e^{-\lambda T}. In other words, in the special case when (t)\mathbb{N}(t) is symmetric and strongly connected for all time, instead of choosing qq to satisfy (22), to achieve an exponential convergence rate of λ\lambda it is enough to choose qq to satisfy the less demanding constraint

q>((λ+A)Tln(1σ))q>\left\lceil\left(\frac{(\lambda+||A||)T}{\ln(\frac{1}{\sigma})}\right)\right\rceil (27)

Justification for this claim is given in §3. Choosing qq in this way is easier that choosing qq according to (22) because the computation of σ\sigma is less demanding than the computation of ρ\rho and consequently α\alpha. On the other hand, this special approach only applies when the neighbor graph is symmetric.

3 Analysis

The aim of this section is to analyze the behavior of the hybrid observer defined in the last section. To do this, use will be made of the notion of a “mixed matrix norm” which will now be defined. For any positive integers kk, mm, nn and pp, let \mathcal{M} denote the real kmnpkmnp- dimensional vector space of block partitioned matrices 𝐌=[Mij]k×m\mathbf{M}=[M_{ij}]_{k\times m} where each block MijM_{ij} is a n×pn\times p matrix. By the mixed matrix norm of 𝐌,\mathbf{M}\in\mathcal{M}, written |𝐌||\mathbf{M}|, is meant the infinity norm of the matrix [Mij]k×m[||M_{ij}||]_{k\times m} where Mij||M_{ij}|| is the two-norm of MijM_{ij}. For example, with ee denoting the “stacked” state estimation error e=Δcolumn{e1,e2,,em}e\stackrel{{\scriptstyle\Delta}}{{=}}\text{column}\{e_{1},e_{2},\ldots,e_{m}\} the quantity max{ei:i𝐦}\max\{||e_{i}||:i\in\mathbf{m}\} mentioned in the last section, is |e||e|, the mixed matrix norm of ee. It is straight forward to verify that |||\cdot| is in fact a norm and that this norm is sub-multiplicative [33].

Recall that the purpose of agent ii’s local parameter estimator defined by (7), (12), and (13) is to estimate x(ti(s1))x(t_{i(s-1)}) after executing qq iterations during the ssth event time interval of agent ii. In view of this, we define the parameter error vectors for i𝐦i\in\mathbf{m},

πis(k)=zis(k)x(ti(s1))k=0,1,,q\pi_{is}(k)=z_{is}(k)-x(t_{i(s-1)})\;\;\;k=0,1,\ldots,q (28)

for all s1s\geq 1. This, (7), and the definition of eie_{i} in (14) imply that

πis(0)=ei(ti(s1)),s1,i𝐦\pi_{is}(0)=e_{i}(t_{i(s-1)}),\;\;s\geq 1,\;\;i\in\mathbf{m} (29)

In addition, from (5), (6) and (14) it is clear that that

ei(tis)=eATπis(q),s1,i𝐦e_{i}(t_{is})=e^{AT}\pi_{is}(q),\;s\geq 1,\;\;i\in\mathbf{m} (30)

To derive the update equation for πis(k)\pi_{is}(k) as kk ranges from 11 to qq, we first note from (13) that

z¯is(k1)x(ti(s1))=1mis(k)j𝒮is(k)πjs(k1)\bar{z}_{is}(k-1)-x(t_{i(s-1)})=\frac{1}{m_{is}(k)}\!\!\!\!\sum_{j\in\mathcal{S}_{is}(k)}\!\!\!\!\pi_{js}(k-1)\; (31)

Next note that because of (4) and (12)

πis(k)=z¯is(k1)x(ti(s1))Qi(Liz¯is(k1)x(ti(s1))e¯i(ti(s1)))\pi_{is}(k)=\bar{z}_{is}(k-1)-x(t_{i(s-1)})\\ -Q_{i}(L_{i}\bar{z}_{is}(k-1)-x(t_{i(s-1)})-\bar{e}_{i}(t_{i(s-1)}))

From this and (31) it follows that for k𝐪,i𝐦,s1k\in\mathbf{q},\;i\in\mathbf{m},\;s\geq 1,

πis(k)=1mi(s)(k)Pij𝒮is(k)πjs(k1)+Qie¯i(ti(s1))\pi_{is}(k)\!\!=\!\!\frac{1}{m_{i(s)}(k)}P_{i}\!\!\!\!\!\sum_{j\in\mathcal{S}_{is}(k)}\!\!\!\!\!\pi_{js}(k-1)+Q_{i}\bar{e}_{i}(t_{i(s-1)}) (32)

where as before, Pi=IQiLiP_{i}=I-Q_{i}L_{i}. These are the local parameter error equations for the hybrid observer.

The next step in the analysis of the system is studied the evolution of the all-agent parameter error vector

πs(k)=column{π1s(k),π2s(k),,πms(k)}\pi_{s}(k)=\text{column}\{\pi_{1s}(k),\pi_{2s}(k),\ldots,\pi_{ms}(k)\}

Note first that because of (29) and (30)

πs(0)=e(s1),s1\pi_{s}(0)=e(s-1),\;\;s\geq 1 (33)

and

e(s)=eA~Tπs(q),s1e(s)=e^{\tilde{A}T}\pi_{s}(q),\;s\geq 1 (34)

where e(s)e(s) is the all-agent state estimation error vector

e(s)=column{e1(tis),,em(tms)},s0e(s)=\text{column}\{e_{1}(t_{is}),\ldots,e_{m}(t_{ms})\},\;\;s\geq 0 (35)

and A~=diagonal{A,A,,A}\tilde{A}=\text{diagonal}\{A,A,\ldots,A\}.

In order to develop an update equation for πs(k)\pi_{s}(k) as kk ranges from 11 to qq, it is necessary to combine the mm update equations in (32) into a single equation and to do this requires a succinct description of the graph determined by the sets 𝒮is,i𝐦\mathcal{S}_{is},\;i\in\mathbf{m} defined in (11). There are two cases to consider: the synchronous case which is when all of the ϵi=0\epsilon_{i}=0 and the asynchronous case when some or all of the ϵi\epsilon_{i} may be non-zero. The following lemmas cover both cases.

Lemma 2.

Synchronous Case: Suppose ϵi=0,i𝐦\epsilon_{i}=0,\;i\in\mathbf{m}. Then for any fixed value of β\beta satisfying (10), including β=0\beta=0,

𝒮is(k)=𝒩i((s1)T+(k1)Δ+β),i𝐦,k𝐪\mathcal{S}_{is}(k)\!=\!\mathcal{N}_{i}((\!s\!-\!1\!)T\!+\!(k-1)\Delta\!+\!\beta),i\in\mathbf{m},k\in\mathbf{q} (36)

Proof of Lemma 36: By hypothesis all ϵi=0\epsilon_{i}=0. Clearly (10) can be satisfied with β=0\beta=0. Moreover from (8) and (9) and the assumption that ti(s1)=(s1)Tt_{i(s-1)}=(s-1)T, it follows that τjs(k1)=(s1)T+(k1)Δ,j𝐦\tau_{js}(k-1)=(s-1)T+(k-1)\Delta,\;j\in\mathbf{m}, so 𝒩i(τjs(k1)+β)=𝒩i((s1)T+(k1)Δ+β)\mathcal{N}_{i}(\tau_{js}(k-1)+\beta)=\mathcal{N}_{i}((s-1)T+(k-1)\Delta+\beta). From this and (11) it follows that (36) is true.  

The following lemma asserts that (36) still holds in the asynchronous case when some of the ϵi\epsilon_{i} are nonzero, provided (t)\mathbb{N}(t) is constant on each interval s(k),k𝐪,s1\mathcal{I}_{s}(k),\;k\in\mathbf{q},\;s\geq 1.

Lemma 3.

If the ϵi\epsilon_{i} satisfy the constraints in (10) and (t)\mathbb{N}(t) is constant on each interval s(k),k𝐪,s1\mathcal{I}_{s}(k),\;k\in\mathbf{q},\;s\geq 1 then (36) is true.

Proof of Lemma 3: Fix i𝐦i\in\mathbf{m}, s1s\geq 1 and k𝐪k\in\mathbf{q}. In light of (8), (9) and the assumption that ti(s1)=(s1)Tt_{i(s-1)}=(s-1)T, it is clear that for any j𝐦j\in\mathbf{m},

τjs(k1)+β[ϵj+(s1)T+(k1)Δ+β,ϵj+(s1)T+(k1)Δ+β](s1)(k)\tau_{js}(k-1)+\beta\in[-\epsilon_{j}+(s-1)T+(k-1)\Delta+\beta,\\ \epsilon_{j}+(s-1)T+(k-1)\Delta+\beta]\subset\mathcal{I}_{(s-1)}(k)

Moreover, (s1)T+(k1)Δ+β(s1)(k)(s-1)T+(k-1)\Delta+\beta\in\mathcal{I}_{(s-1)}(k). But by assumption, (t)\mathbb{N}(t) is constant on (s1)(k)\mathcal{I}_{(s-1)}(k) which means that 𝒩i(t)\mathcal{N}_{i}(t) is constant on (s1)(k)\mathcal{I}_{(s-1)}(k).Therefore 𝒩i(τjs(k1)+β)=𝒩i((s1)T+(k1)Δ+β)\mathcal{N}_{i}(\tau_{js}(k-1)+\beta)=\mathcal{N}_{i}((s-1)T+(k-1)\Delta+\beta). From this and the definition of 𝒮is(k)\mathcal{S}_{is}(k) in (11), it follows that (36) is true.  

In summary, Lemmas 36 and 3 assert that (36) holds in the synchronous case when all ϵi=0\epsilon_{i}=0, or alternatively in the asynchronous case when the neighbor graph (t)\mathbb{N}(t) is constant on each interval s(k),k𝐪,s1.\mathcal{I}_{s}(k),\;k\in\mathbf{q},\;s\geq 1. Because of this, the following steps to obtain an update equation for e(s)e(s) apply to both cases.

Equation (36) implies that the graphs determined by the 𝒮is(k),k𝐪,i𝐦,s1\mathcal{S}_{is}(k),\;\;k\in\mathbf{q},\;i\in\mathbf{m},\;s\geq 1 are the neighbor graphs ((s1)T+(k1)Δ+β),k𝐪,s1\mathbb{N}((s-1)T+(k-1)\Delta+\beta),\;k\in\mathbf{q},\;s\geq 1. Since (t)\mathbb{N}(t) is assumed to be strongly connected for all t0t\geq 0, each of these neighbor graphs is strongly connected. These graphs are used as follows.

Let 𝒢\mathcal{G} denote the set of all directed graphs on mm vertices which have self-arcs at all vertices. Note that 𝒢\mathcal{G} is a finite set and that (t)𝒢,t0\mathbb{N}(t)\in\mathcal{G},\;t\geq 0. Each graph 𝔾𝒢\mathbb{G}\in\mathcal{G} uniquely determines a so-called “flocking-matrix” which is an m×mm\times m stochastic matrix of the form D𝔾1A𝔾D^{-1}_{\mathbb{G}}A^{\prime}_{\mathbb{G}}, where A𝔾A_{\mathbb{G}} and D𝔾D_{\mathbb{G}} are respectively the adjacency matrix and diagonal in-degree matrix of of 𝔾\mathbb{G}; D𝔾D_{\mathbb{G}} is nonsingular because each graph in 𝒢\mathcal{G} has self-arcs at all vertices.

For k𝐪k\in\mathbf{q} and s1s\geq 1, let Fs(k)F_{s}(k) denote the flocking matrix determined by ((s1)T+(k1)Δ+β)\mathbb{N}((s-1)T+(k-1)\Delta+\beta). Then (32) implies for s1s\geq 1 that

πs(k)=P(Fs(k)I)πs(k1)+Qe¯(s1),k𝐪\pi_{s}(k)=P(F_{s}(k)\otimes I)\pi_{s}(k-1)+Q\bar{e}(s-1),\;\;k\in\mathbf{q} (37)

where

e¯(s)=column{e¯1(t1s),,e¯m(tms)},s0\bar{e}(s)=\text{column}\{\bar{e}_{1}(t_{1s}),\ldots,\bar{e}_{m}(t_{ms})\},\;\;s\geq 0 (38)

P=diagonal{P1,,Pm},Q=diagonal{Q1,,Qm}P=\text{diagonal}\{P_{1},\ldots,P_{m}\},Q=\text{diagonal}\{Q_{1},\ldots,Q_{m}\}, and II is the n×nn\times n identity. Thus for s1s\geq 1,

πs(q)=Φs(0)πs(0)+(k=1qΦs(k))Qe¯(s1)\pi_{s}(q)=\Phi_{s}(0)\pi_{s}(0)+\left(\sum_{k=1}^{q}\Phi_{s}(k)\right)Q\bar{e}(s-1) (39)

where Φs(k)\Phi_{s}(k) is the state transition matrix defined by

Φs(k)=P(Fs(q)I)P(Fs(k+1)I)\Phi_{s}(k)=P(F_{s}(q)\otimes I)\cdots P(F_{s}(k+1)\otimes I) (40)

for 0k<q0\leq k<q and by Φs(q)=I\Phi_{s}(q)=I for k=qk=q. From this, (33) and (34) it follows that for s1s\geq 1, the all-agent state estimation error e(s)e(s) satisfies

e(s)=A(s)e(s1)+B(s)e¯(s1),s1e(s)=A(s)e(s-1)+B(s)\bar{e}(s-1),\;\;s\geq 1 (41)

where

A(s)\displaystyle A(s) =\displaystyle= eA~TΦs(0),s1\displaystyle e^{\tilde{A}T}\Phi_{s}(0),\;\;s\geq 1 (42)
B(s)\displaystyle B(s) =\displaystyle= eA~Tk=1qΦs(k)Q,s1\displaystyle e^{\tilde{A}T}\sum_{k=1}^{q}\Phi_{s}(k)Q,\;\;s\geq 1 (43)

To determine the convergence properties of e(s)e(s) as ss\rightarrow\infty use will be made of the following lemma which gives bounds on the norms of the coefficient matrices A(s)A(s) and B(s)B(s) appearing in (41).

Lemma 4.

Suppose that qq satisfies the inequality given in Theorem 22. Then

|A(s)|\displaystyle|A(s)| \displaystyle\leq eλT,s1\displaystyle e^{-\lambda T},\;\;s\geq 1 (44)
|B(s)|\displaystyle|B(s)| \displaystyle\leq qeAT|Q|,s1\displaystyle qe^{||A||T}|Q|,\;\;s\geq 1 (45)

In order to justify the bound on the norm of A(s)A(s) given in (44), use will be made of the following lemma. which is a simple variation on a result in [33].

Lemma 5.

Let \mathcal{F} denote the set of all flocking matrices determined by those graphs in 𝒢\mathcal{G} which are strongly connected. For any set of μ(m1)2+1\mu\geq(m-1)^{2}+1 flocking matrices S1,S2,,SμS_{1},S_{2},\ldots,S_{\mu} in \mathcal{F}

|P(SμI)P(Sμ1I)P(S1I)|α|P(S_{\mu}\otimes I)P(S_{\mu-1}\otimes I)\cdots P(S_{1}\otimes I)|\leq\alpha (46)

where α\alpha is the attenuation constant

α=1(m1)(1ρ)m(m1)2\alpha=1-\frac{(m-1)(1-\rho)}{m^{(m-1)^{2}}}

Proof of Lemma 5: Fix μ(m1)2+1\mu\geq(m-1)^{2}+1, set k=(m1)2k=(m-1)^{2} and let and S1,S2,,SμS_{1},S_{2},\ldots,S_{\mu} be flocking matrices in \mathcal{F}. Then

P(SμI)P(Sμ1I)P(S2I)P(S1I)={P(SμI)P(Sk+2I)}{P(Sk+1I)P(S2I)P}{S1I}P(S_{\mu}\otimes I)P(S_{\mu-1}\otimes I)\cdots P(S_{2}\otimes I)P(S_{1}\otimes I)\\ =\{P(S_{\mu}\otimes I)\cdots P(S_{k+2}\otimes I)\}\\ \{P(S_{k+1}\otimes I)\cdots P(S_{2}\otimes I)P\}\{S_{1}\otimes I\}

But for any flocking matrix SS\in\mathcal{F}, |SI|=S=1|S\otimes I|=||S||_{\infty}=1 where ||||||\cdot||_{\infty} is the infinity norm. From this, the sub-multiplicative property of the mixed matrix norm, and the fact that |P|1|P|\leq 1, it follows that

|P(SμI)P(Sμ1I)P(S2I)P(S1I)||P(Sk+1I)P(S2I)P||P(S_{\mu}\otimes I)P(S_{\mu-1}\otimes I)\cdots P(S_{2}\otimes I)P(S_{1}\otimes I)|\\ \leq|P(S_{k+1}\otimes I)\cdots P(S_{2}\otimes I)P|

In view of equation (26) of [33],

|P(Sk+1I)P(S2I)P|1(m1)(1ρ)m(m1)2|P(S_{k+1}\otimes I)\cdots P(S_{2}\otimes I)P|\leq 1-\frac{(m-1)(1-\rho)}{m^{(m-1)^{2}}}

Therefore (46) is true.  

Proof of Lemma 4: Lemma 5 implies that if for a given integer p>0p>0, if qp((m1)2+1)q\geq p((m-1)^{2}+1) then for any s1s\geq 1,

|(P(Fs(q)I)P(Fs(1)I)|αp|(P(F_{s}(q)\otimes I)\cdots P(F_{s}(1)\otimes I)|\leq\alpha^{p} (47)

Therefore by (40) and (42), if qq is so chosen, then |A(s)|eATαp|A(s)|\leq e^{||A||T}\alpha^{p}. Thus by picking pp so large that

eATαp<eλTe^{||A||T}\alpha^{p}<e^{-\lambda T} (48)

and then setting q=p((m1)2+1)q=p((m-1)^{2}+1) one gets (44). The requirement on pp determined by (48) is equivalent to the requirement on pp determined by (21). It follows that (44) will hold provided qq satisfies the inequality given in Theorem 22.

Recall that |P|1|P|\leq 1 and that |SI|=1|S\otimes I|=1 for any m×mm\times m stochastic matrix SS. From this and the sub-multiplicative property of the mixed matrix norm it follows that the matrix Φs(k)\Phi_{s}(k) defined by (40) satisfies

|Φs(k)|1,k𝐪,s1|\Phi_{s}(k)|\leq 1,\;\;\;k\in\mathbf{q},\;\;s\geq 1 (49)

This and the definition of B(s)B(s) in (43) imply that for all s1s\geq 1, |B(s)|qeAT|Q||B(s)|\leq qe^{||A||T}|Q|. Thus (45) is true.  

It is obvious at this point that because (36) holds in both the synchronous and asynchronous cases, the same arguments can be used to prove both Theorem 22 and Theorem 2.

Proof of Theorems 22 and 2: In view of (41) and Lemma 4 it is possible to write

|e(s)|<eλT|e(s1)|+b|e¯(s1)|,s1|e(s)|<e^{-\lambda T}|e(s-1)|+b|\bar{e}(s-1)|,\;s\geq 1

where b=qeAT|Q|b=qe^{||A||T}|Q|. Therefore

|e(s)|<eλsT|e(0)|+bk=1seλ(sk)T|e¯(k1)||e(s)|<e^{-\lambda sT}|e(0)|+b\sum_{k=1}^{s}e^{-\lambda(s-k)T}|\bar{e}(k-1)| (50)

To deal with the term involving e¯\bar{e} in (50), we proceed as follows. Note first from (15) that

e¯i(tis)cieλ¯Te¯i(ti(s1)),i𝐦,s1||\bar{e}_{i}(t_{is})||\leq c_{i}e^{-\bar{\lambda}T}||\bar{e}_{i}(t_{i(s-1)})||,\;\;i\in\mathbf{m},\;\;\;s\geq 1

Thus e¯i(tis)cieλ¯sTe¯i(ti0)||\bar{e}_{i}(t_{is})||\leq c_{i}e^{-\bar{\lambda}sT}||\bar{e}_{i}(t_{i0})|| for i𝐦s1i\in\mathbf{m}\;s\geq 1. It follows from this and the definition of e¯(s)\bar{e}(s) in (38) that

|e¯(s)|ceλ¯sT|e¯(0)|,s1|\bar{e}(s)|\leq ce^{-\bar{\lambda}sT}|\bar{e}(0)|,\;\;s\geq 1

where c=max{ci,i𝐦}c=\max\{c_{i},\;i\in\mathbf{m}\}. Thus for s1s\geq 1

bk=1seλ(sk)T|e¯(k1)|cbk=1seλ(sk)Teλ¯(k1)T|e¯(0)|=cbe(λsλ¯)Tk=1se(λ¯λ)kT)|e¯(0)|cbe(λsλ¯)Tk=1e(λ¯λ)kT)|e¯(0)|=cbe(λsλ¯)Te(λ¯λ)T1e(λ¯λ)T|e¯(0)|=cbeλsTeλT1e(λ¯λ)T|e¯(0)|b\sum_{k=1}^{s}e^{-\lambda(s-k)T}|\bar{e}(k-1)|\\ \leq cb\sum_{k=1}^{s}e^{-\lambda(s-k)T}e^{-\bar{\lambda}(k-1)T}|\bar{e}(0)|\\ =cbe^{-(\lambda s-\bar{\lambda})T}\sum_{k=1}^{s}e^{-(\bar{\lambda}-\lambda)kT)}|\bar{e}(0)|\\ \leq cbe^{-(\lambda s-\bar{\lambda})T}\sum_{k=1}^{\infty}e^{-(\bar{\lambda}-\lambda)kT)}|\bar{e}(0)|\\ =cbe^{-(\lambda s-\bar{\lambda})T}\frac{e^{-(\bar{\lambda}-\lambda)T}}{1-e^{-(\bar{\lambda}-\lambda)T}}|\bar{e}(0)|\\ =cbe^{-\lambda sT}\frac{e^{\lambda T}}{1-e^{-(\bar{\lambda}-\lambda)T}}|\bar{e}(0)|\hskip 59.98422pt

Using (50) there follows

|e(s)|eλsT(|e(0)|+d|e¯(0)|),s1|e(s)|\leq e^{-\lambda sT}(|e(0)|+d|\bar{e}(0)|),\;s\geq 1 (51)

where

d=cbeλT1e(λ¯λ)Td=cb\frac{e^{\lambda T}}{1-e^{-(\bar{\lambda}-\lambda)T}}

Fix i𝐦i\in\mathbf{m}. In view of (51) and the definition of e(s)e(s) in (35),

ei(ti(s1))eλ(s1)T(|e(0)|+d|e¯(0)|)i𝐦,s1||e_{i}(t_{i(s-1)})||\leq e^{-\lambda(s-1)T}(|e(0)|+d|\bar{e}(0)|)\;\;i\in\mathbf{m},\;\;s\geq 1

But for t(ti(s1),tis)t\in(t_{i(s-1)},\;t_{is}), x˙i=Axi\dot{x}_{i}=Ax_{i}; consequently e˙i=Aei\dot{e}_{i}=Ae_{i} for the same values of tt. Therefore

ei(t)=eA(t(s1)T)ei(ti(s1)),t[ti(s1),tis),s1e_{i}(t)=e^{A(t-(s-1)T)}e_{i}(t_{i(s-1)}),\;\;t\in[t_{i(s-1)},\;t_{is}),\;\;s\geq 1

so

||ei(t)||eAT||ei(ti(s1)||,t[ti(s1),tis),s1||e_{i}(t)||\leq e^{||A||T}||e_{i}(t_{i(s-1)}||,\;\;t\in[t_{i(s-1)},\;t_{is}),\;\;s\geq 1

Therefore for t[ti(s1),tis)t\in[t_{i(s-1)},t_{is}) and s1s\geq 1

ei(t)e(ATλ(s1)T)(|e(0)|+d|e¯(0)|)||e_{i}(t)||\leq e^{(||A||T-\lambda(s-1)T)}(|e(0)|+d|\bar{e}(0)|)

Now for i𝐦i\in\mathbf{m},

eλsTeλt,t[ti(s1),tis)e^{-\lambda sT}\leq e^{-\lambda t},\;\;t\in[t_{i(s-1)},\;t_{is})

so

||ei(t)||e(ATλt)(|e(0)|+d|e¯(0)|),t[(ti(s1),tis)||e_{i}(t)||\leq e^{(||A||T-\lambda t)}(|e(0)|+d|\bar{e}(0)|),\;\;t\in[(t_{i(s-1)},\;t_{is})

Since this holds for all s1s\geq 1

ei(t)e(ATλt)(|e(0)|+d|e¯(0)|),t0||e_{i}(t)||\leq e^{(||A||T-\lambda t)}(|e(0)|+d|\bar{e}(0)|),\;\;t\geq 0

which proves that the state estimation errors ei,i𝐦e_{i},\;i\in\mathbf{m}, all converge to zero as fast as eλte^{-\lambda t} does.  

3.1 Special case

We now turn to the special case mentioned in §2.1. In this case the definition of the state-transition matrix Φ\Phi appearing in (39) changes from (40) to

Φs(k)=P(Ws(q)I)P(Ws(k+1)I)\Phi_{s}(k)=P(W_{s}(q)\otimes I)\cdots P(W_{s}(k+1)\otimes I) (52)

for 0k<q0\leq k<q, and Ws(k)=M((s1)T+(k1)Δ+β)W_{s}(k)\;\stackrel{{\scriptstyle\triangle}}{{=}}\;M_{\mathbb{N}((s-1)T+(k-1)\Delta+\beta)} with graph ((s1)T+(k1)Δ+β)\mathbb{N}((s-1)T+(k-1)\Delta+\beta).

Although the formula for e(s)e(s), namely (41), and the definitions of A(s)A(s) and B(s)B(s) in (42) and (43) are as before, the bounds for A(s)A(s) and and B(s)B(s) given by (44) and (45) no longer apply. To proceed, use will be made of the following lemma.

Lemma 6.

Let FF be an m×mm\times m doubly stochastic matrix with positive diagonals and a strongly connected graph. Suppose that Pi,i𝐦P_{i},\;\;i\in\mathbf{m}, is a set of n×nn\times n orthogonal projection matrices such that

i=1mimagePi=0\bigcap_{i=1}^{m}\text{image}\;P_{i}=0 (53)

Then the matrix P(FI)P(F\otimes I) is a contraction in the 22-norm where P=diag {P2,P2,,Pm}P=\text{\rm diag\;}\{P_{2},P_{2},\ldots,P_{m}\}.

Proof: Write SS for FIF\otimes I and note that SS is doubly stochastic with positive diagonals and a strongly connected graph. Since P1||P||\leq 1, it must be true that that PSS||PS||\leq||S||. Moreover S1||S||\leq 1 because SSS^{\prime}S is stochastic; thus PS1||PS||\leq 1. Hence it is enough to prove that PS1||PS||\neq 1 or equivalently that SP1||S^{\prime}P||\neq 1
Suppose that SP=1||S^{\prime}P||=1 or equivalently that PSSPx=xPSS^{\prime}Px=x for some nonzero vector xx. Clearly PSS’Px = Px which implies that Px=xPx=x and thus that xSSx=xx.x^{\prime}SS^{\prime}x=x^{\prime}x. Therefore Sx=x||S^{\prime}x||=||x||. From this and Lemma 1 of [35] it follows that SSx=xSS^{\prime}x=x. Now SSSS^{\prime} is stochastic. Moreover its graph is strongly connected because SS has a strongly connected graph and positive diagonals, as does SS^{\prime}. Thus by the Perron Frobenius theorem, SSSS^{\prime} has exactly one eigenvalue at 11 and all the rest must be inside the unit circle; in addition the eigenspace for the eigenvalue 11 must be spanned by the one-vector 𝟏nm\mathbf{1}_{nm}. Therefore x=μ𝟏nmx=\mu\mathbf{1}_{nm} for some nonzero scalar μ\mu. Therefore P𝟏nm=𝟏nmP\mathbf{1}_{nm}=\mathbf{1}_{nm} which implies that 𝟏n=Pi𝟏n,i𝐦\mathbf{1}_{n}=P_{i}\mathbf{1}_{n},\;\;i\in\mathbf{m}, But this is impossible because of (3).  

The following lemma gives the bounds on A(s)A(s) and B(s)B(s) for the special case under consideration.

Lemma 7.

Suppose that qq satisfies (27). Then

A(s)eλT,s1\displaystyle\|A(s)\|\leq e^{-\lambda T},\;\;s\geq 1 (54)
B(s)qeATQ,s1\displaystyle\|B(s)\|\leq qe^{\|A\|T}\|Q\|,\;\;s\geq 1 (55)

Proof: Lemma 6 implies that for each 0k<q0\leq k<q, P(Ws(k)I)<1\|P(W_{s}(k)\otimes I)\|<1. Moreover, P(Ws(k)I)σ\|P(W_{s}(k)\otimes I)\|\leq\sigma where σ\sigma is chosen according to (26). From this and the sub-multiplicative property of the two norm it follows that

P(Ws(q)I)P(Ws(1)I)σq\|P(W_{s}(q)\otimes I)\cdots P(W_{s}(1)\otimes I)\|\leq\sigma^{q}

Therefore by (42) and (52), if qq is so chosen to satisfy (27), then

A(s)eATσq<eλT\|A(s)\|\leq e^{\|A\|T}\sigma^{q}<e^{-\lambda T}

Thus (54) is true. Recall that P(Ws(k)I)<1\|P(W_{s}(k)\otimes I)\|<1 and Ws(k)I1\|W_{s}(k)\otimes I\|\leq 1 for 0k<q0\leq k<q. From this and the sub-multiplicative property of the two norm, the matrix Φs(k)\Phi_{s}(k) defined by (52) satisfies

Φs(k)1,k𝐪,s1\|\Phi_{s}(k)\|\leq 1,\;\;k\in\mathbf{q},\;\;s\geq 1

This and (43) imply (55).  

Other than the modifications in the bounds on A(s)A(s) and B(s)B(s) given in the above lemma, everything else is the same for both the synchronous and asynchronous versions of the problem. So what one gains in this special case is exponential convergence at a prescribed rate with a smaller value of qq.

4 Event-time Mismatch - A Robustness Issue

In the preceding section it was shown that the hybrid observer under discussion will function correctly if local iterations are performed synchronously across the network no matter how fast the associated neighbor graph changes, just so long as it is always strongly connected. Correct performance is also assured in the face of asynchronously executed local iterations across the network during each event time interval, provided the neighbor graph changes in a suitably defined sense. Implicitly assumed in these two cases is that the event time sequences of all mm agents are the same. The aim of this section is to explain what happens if this assumption is not made. For simplicity, this will only be done for the case when differing event time sequences are the only cause of asynchronism. As will be seen, the consequence of event-time sequence mismatches turns out to be more of a robustness issue than an issue due to unsynchronized operation. In particular, it will become apparent that if different agents use slightly different event time sequences then asymptotically correct state estimates will not be possible unless AA is a stability matrix. While at first glance this may appear to be a limitation of the distributed observer under consideration, it is in fact a limitation of virtually all state estimators, distributed or not, which are not used in feedback-loops. Since this easily explained observation is apparently not widely appreciated, an explanation of this simple fact will be given at the end of this section.

There are two differences between the setup to be considered here and the setup considered in the last section. First it will now be assumed that the local deviation times δis(k)\delta_{is}(k) appearing in (8) are all zero. Thus in place of (8) the local iteration times for agent ii on [ti(s1),tis)[t_{i(s-1)},t_{is})

τis(k)=ti(s1)+kΔ,k{0,1,,m}\tau_{is}(k)=t_{i(s-1)}+k\Delta,\;k\in\{0,1,\ldots,m\} (56)

Second instead of assuming that the initializations ti0t_{i0} of the mm agents’ event time sequences are all zero, it will be assumed instead that each ti0t_{i0} is a small number known only to agent ii which lies in the interval [ϵi,ϵi][-\epsilon_{i},\epsilon_{i}] where, as before, ϵi\epsilon_{i} is a small nonnegative number. This means that even though the event time sequences of all mm agents are still periodic with period TT, the sequences are not synchronized with each other. As before it is assumed that within event time interval [tj(s1),tjs)[t_{j(s-1)},t_{js}), agent jj broadcasts iterate zjs(k1)z_{js}(k-1) at time τjs(k1)+β\tau_{js}(k-1)+\beta. To ensure that this time falls within the reception interval [ti(s1),tis)[t_{i(s-1)},t_{is}) of each agent ii, it will continue to be assumed that (10) holds. Apart from these modifications the setup to be considered here is the same as the one considered previously. As a consequence, many of the steps in the analysis of the hybrid observers performance are the same as they were for the previously considered case.

Our first objective is to develop the relevant equations for the local parameter error vector πis(k)\pi_{is}(k) defined by (28). Although (29) and (30) continue to hold without change, (32) requires modification. To understand what needs to be changed, it is necessary to first derive a relationship between x(ti(s1))x(t_{i(s-1)}) and x(tj(s1))x(t_{j(s-1)}). Towards this end, note that

x(ti(s1))=x(tj(s1))+x((s1)T+ti0)x((s1)T+tj0)x(t_{i(s-1)})\!=\!x(t_{j(s-1)})+x((s-1)T+t_{i0})-x((s-1)T+t_{j0})

because tk(s1)=tk0+(s1)Tt_{k(s-1)}=t_{k0}+(s-1)T for all k𝐦k\in\mathbf{m}. From this and (5) it follows that

x(ti(s1))=x(tj(s1))+(eAti0eAtj0)x((s1)T)x(t_{i(s-1)})=x(t_{j(s-1)})+\left(e^{At_{i0}}-e^{At_{j0}}\right)x((s-1)T)

Hence (13) can now be used to obtain

z¯is(k1)x(ti(s1))=1mis(k)j𝒮is(k)πj(s1)(k1)+Γis(k)x((s1)T)\bar{z}_{is}(k-1)-x(t_{i(s-1)})=\frac{1}{m_{is}(k)}\sum_{j\in\mathcal{S}_{is}(k)}\pi_{j(s-1)}(k-1)\\ +\Gamma_{is}(k)x((s-1)T)\;\;\;\;\;\;\; (57)

where

Γis(k)=1mis(k)j𝒮is(k)(eAti0eAtj0)\Gamma_{is}(k)=\frac{1}{m_{is}(k)}\sum_{j\in\mathcal{S}_{is}(k)}\left(e^{At_{i0}}-e^{At_{j0}}\right) (58)

Next note that because of (4) and (12)

πis(k)=z¯is(k1)x(ti(s1))Qi(Liz¯is(k1)x(ti(s1))e¯i(ti(s1)))\pi_{is}(k)=\bar{z}_{is}(k-1)-x(t_{i(s-1)})\\ -Q_{i}(L_{i}\bar{z}_{is}(k-1)-x(t_{i(s-1)})-\bar{e}_{i}(t_{i(s-1)}))

From this and (57) it follows that

πis(k)=1mi(s)(k)Pij𝒮is(k)πjs(k1)+Qie¯i(ti(s1))+PiΓis(k)x((s1)T),k𝐪,i𝐦,s1\pi_{is}(k)=\frac{1}{m_{i(s)}(k)}P_{i}\sum_{j\in\mathcal{S}_{is}(k)}\pi_{js}(k-1)+Q_{i}\bar{e}_{i}(t_{i(s-1)})\\ +P_{i}\Gamma_{is}(k)x((s-1)T),\;\;k\in\mathbf{q},\;i\in\mathbf{m},\;s\geq 1 (59)

which is the modified version of (32) needed to proceed. The difference between (32) and (59) is thus the inclusion in (59) of the term PiΓis(k)x((s1)T)P_{i}\Gamma_{is}(k)x((s-1)T).

The assumption that the event time sequences of the agents may start at a different time requires us to make the same assumption as before about the neighbor graph (t)\mathbb{N}(t), namely that it is constant on each interval s(k),k𝐪,s1\mathcal{I}_{s}(k),\;k\in\mathbf{q},\;s\geq 1. The assumption makes sense in the present context for the same reason as before, specifically because the interval (k)\mathcal{I}(k) defined by (23) do not overlap. This, in turn, is because the bounds ϵi\epsilon_{i} have been assumed to satisfy (10) which guarantees that Lemma 1 continues to hold.

The next step in the analysis of the hybrid observer is to study the evolution of the all-agent parameter error vector

πs(k)=column{π1s(k),π2s(k),,πms(k)}\pi_{s}(k)=\text{column}\{\pi_{1s}(k),\pi_{2s}(k),\ldots,\pi_{ms}(k)\}

As before, (33) and (34) continue to hold where e(s)e(s) is the all - agent state estimation error defined by (35). A simple modification of the proof of Lemma 3 can be used to establish the lemma’s validity in the present context. Consequently a proof will not be given. The lemma enables us to combine the individual update equations in (59), thereby obtaining the update equation

πs(k)=P(Fs(k)I)πs(k1)+Qe¯(s1)+PΓs(k)x((s1)T),k𝐪\pi_{s}(k)=P(F_{s}(k)\otimes I)\pi_{s}(k-1)+Q\bar{e}(s-1)\\ +P\Gamma_{s}(k)x((s-1)T),\;\;k\in\mathbf{q}

where

Γs(k)=column{Γ1s(k),,Γms(k)}\Gamma_{s}(k)=\text{column}\{\Gamma_{1s}(k),\ldots,\Gamma_{ms}(k)\} (60)

The steps involved in doing this are essentially the same as the steps involved in deriving (37). Not surprisingly, the only difference between (37) and (60) is the inclusion in the latter of the term PΓs(k)x((s1)T)P\Gamma_{s}(k)x((s-1)T).

From (33), (34), and (60) it follows at once that the all-agent state estimation error vector satisfies

e(s)=A(s)e(s1)+B(s)e¯(s1)+G(s)x((s1)T),s1e(s)=A(s)e(s-1)+B(s)\bar{e}(s-1)\\ +G(s)x((s-1)T),\;\;s\geq 1 (61)

where A(s)A(s) and B(s)B(s) are as defined in (42) and (43) respectively, and

G(s)=eA~Tk=1qΦs(k)PΓs(k)G(s)=e^{\tilde{A}T}\sum_{k=1}^{q}\Phi_{s}(k)P\Gamma_{s}(k)

The following lemma gives a bound on the mixed matrix norm of G(s)G(s).

Lemma 8.

Suppose that qq satisfies the inequality given in Theorem 22. Then

|G(s)|2mqϵAeA(T+β)|G(s)|\leq 2mq\epsilon||A||e^{||A||(T+\beta)} (62)

Note that this bound is small when ϵ\epsilon is small. This means that small deviations of the agent’s event time sequences from the nominal event time sequence 0,T,2T,0,T,2T,\ldots produce small effects on the error dynamics in (61), provided of course xx is well behaved; i.e., AA is a stability matrix! More will be said about this point below.

Proof of Lemma 62: From (58),

Γis(k)j𝐦eAti0eAtj0||\Gamma_{is}(k)||\leq\sum_{j\in\mathbf{m}}||e^{At_{i0}}-e^{At_{j0}}||

In general, for any real square matrix MM, and real numbers tt and τ\tau

eMteMτM(tτ)eMt||e^{Mt}-e^{M\tau}||\leq||M(t-\tau)||e^{||Mt||}

so

||Γis(k)j𝐦||A(ti0tj0)||eAti0||\Gamma_{is}(k)\|\leq\sum_{j\in\mathbf{m}}||A(t_{i0}-t_{j0})||e^{||At_{i0}||}

By assumption |ti0|ϵi|t_{i0}|\leq\epsilon_{i} and |ti0tj0|ϵi+ϵj|t_{i0}-t_{j0}|\leq\epsilon_{i}+\epsilon_{j}. But ϵiβ\epsilon_{i}\leq\beta and ϵ=max{ϵi,i𝐦}\epsilon=\max\{\epsilon_{i},\;i\in\mathbf{m}\}. Thus |ti0|β|t_{i0}|\leq\beta and |ti0tj0|2ϵ|t_{i0}-t_{j0}|\leq 2\epsilon. Therefore

Γis(k)2mϵAeAβ||\Gamma_{is}(k)||\leq 2m\epsilon||A||e^{||A||\beta}

so

|Γs(k)|2mϵAeAβ|\Gamma_{s}(k)|\leq 2m\epsilon||A||e^{||A||\beta}

In view of (49) and the definition of G(s)G(s), |G(s)|qeATΓs(k)|G(s)|\leq qe^{||A||T}||\Gamma_{s}(k)||. It follows that (62) holds.  

Taking the construction leading to (50) as a guide, it is not difficult to derive from (61) the inequality

|e(s)|eλsT(|e(0)|+d|e¯(0)|)+ϵgk=1seλ(sk)Tx((k1)T),s1|e(s)|\leq e^{-\lambda sT}(|e(0)|+d|\bar{e}(0)|)\\ +\epsilon g\sum_{k=1}^{s}e^{-\lambda(s-k)T}||x((k-1)T)||,\;s\geq 1 (63)

where dd is as defined just below (51) and g=2mqAeA(T+β)g=2mq||A||e^{||A||(T+\beta)}. Comparing (51) to (63), we see that the effect of the change in assumptions leads to the inclusion in (63) of the term involving xx.

At this point there are two distinct cases to consider - either eAte^{At} converges to zero or it does not. Consider first the case when eAte^{At} converges to zero. Then there must be positive constants cac_{a} and λa\lambda_{a} such that eAtcaeλat||e^{At}||\leq c_{a}e^{-\lambda_{a}t}. By treating the term involving xx in (63) in the same manner as the term involving e¯\bar{e} in (51) was treated, one can easily conclude that for a suitably defined constant hh

|e(s)|eλsT(|e(0)|+d|e¯(0)|+ϵhx(0)),s1|e(s)|\leq e^{-\lambda sT}(|e(0)|+d|\bar{e}(0)|+\epsilon h||x(0)||),\;\;s\geq 1

if λa>λ\lambda_{a}>\lambda, or

|e(s)|eλsT(|e(0)|+d|e¯(0)|)+ϵheλasTx(0),s1|e(s)|\leq e^{-\lambda sT}(|e(0)|+d|\bar{e}(0)|)+\epsilon he^{-\lambda_{a}sT}||x(0)||,\;\;s\geq 1

if λaλ\lambda_{a}\leq\lambda. If the former is true, then the same arguments as were used in the last section can be used to show that the state estimations errors ei(t)e_{i}(t) converge to zero as fast as eλte^{-\lambda t} does. On the other hand, if the latter is true, by similar reasoning ei(t)e_{i}(t) can easily be shown to converge to zero as fast as eλate^{-\lambda_{a}t} does. Note that in this case, if λa\lambda_{a} is small, the effect of the resulting slow convergence of xx will to some extent be mitigated by the smallness of ϵ\epsilon, so even with small λa\lambda_{a}, the performance of the hybrid observer may be acceptable for sufficiently small perturbations of the start times of the event time sequences from 0.

In the other situation, which is when AA is not a stability matrix, the hybrid observer cannot perform acceptably except possibly if finite time state estimation is all that is desired and ϵ\epsilon is sufficiently small.

Key Point: This limitation applies not only to the hybrid observer discussed in this paper, but to all state estimators, centralized or not, including Kalman filters which are not being used in feedback loops.444Some of the adaptive observers developed in the past may be an exception to this, but such observers invariably require persistent excitation to achieve exponential convergence.

Experience has shown that this limitation is not widely recognized, despite its simple justification. Here is the justification.
Suppose one is trying to obtain an estimate x^\widehat{x} of the state xx of a single-channel, observable linear system y=Cxy=Cx, x˙=Ax\dot{x}=Ax using an observer but approximately correct values of AA and CC - say A^\widehat{A} and C^\widehat{C} - upon which to base the observer design are known. The observer would then be a linear system of the form

x^˙=A^x^+K(C^x^y)\dot{\widehat{x}}=\widehat{A}\widehat{x}+K(\widehat{C}\widehat{x}-y) (64)

with KK chosen to exponentially stabilize A^+KC^\widehat{A}+K\widehat{C}. Then it is easy to see that the state estimation error e=x^xe=\widehat{x}-x must satisfy

e˙=(A^+KC^)e+(A^A+K(C^C))x\dot{e}=(\widehat{A}+K\widehat{C})e+(\widehat{A}-A+K(\widehat{C}-C))x

Therefore if AA is not a stability matrix and either A^\widehat{A} is not exactly equal to AA or C^\widehat{C} is not exactly equal to CC, then instead of converging to zero, the state estimation error ee will grow without bound for almost any initialization. In other words, with robustness in mind, the problem of trying to obtain an estimate of the state of a linear system with an “open-loop” state estimator, does not make sense unless AA is a stability matrix.  Of course, if one is trying to use a state estimator generate an estimate x^\widehat{x} of the state xx of the forced linear system

x˙=Ax+BFx^\dot{x}=Ax+BF\widehat{x}

where A+BFA+BF is a stability matrix, this problem does not arise, but to accomplish this one has to change the estimator dynamics defined in (64) to

x^˙=A^x^+K(C^x^y)+BFx^\dot{\widehat{x}}=\widehat{A}\widehat{x}+K(\widehat{C}\widehat{x}-y)+BF\widehat{x}

While this modification works in the centralized case, it cannot be used in the decentralized case as explained in [36]. In fact, until recently there appeared to be only one of distributed observer which could be used in a feedback configuration thereby avoiding the robustness issue just mentioned [36]. However, recent research suggests other approach may emerge [37].

5 Resilience

By a (passively) resilient algorithm for a distributed process is meant an algorithm which, by exploiting built-in network and data redundancies, is able to continue to function correctly in the face of abrupt changes in the number of vertices and arcs in the inter-agent communication graph upon which the algorithm depends. In this section,it will be shown that the proposed estimator can cope with the situation when there is an arbitrary abrupt change in the topology of the neighbor graph such as the loss or addition of an arc or a vertex provided connectivity is not lost in an appropriately defined sense.

Consider first the situation when there is a potential loss or addition of aa arcs in the neighbor graph. Assume the neighbor graph is a¯\bar{a}-arc redundantly strongly connected in that the graph is strongly connected and remains strongly connected after any aa¯a\leq\bar{a} arcs are removed. With this assumption, strong connectivity of the neighbor graph and jointly observability of the system are ensured when any aa¯a\leq\bar{a} arcs are lost. Alternatively, if any number of new arcs are added, strong connectivity and joint observability are clearly still ensured. Thus, in the light of Theorem 22, whenever aa¯a\leq\bar{a} arcs are lost from or added to the neighbor graph, the hybrid estimator under consideration will still function correctly without the need for any “active” intervention such as redesign of any of the KiK_{i} or readjustment of qq. In fact, Theorem 22 guarantees that correct performance will prevail, even if arcs change over and over, no matter how fast, just so long as strong connectivity is maintained for all time.

Consider next the far more challenging situation when at some time tt^{*} there is a loss of v<mv<m vertices from the neighbor graph (t).\mathbb{N}(t). For this situation, only preliminary results currently exist. One possible way to deal with this situation is as follows.

As a first step, pick the KiK_{i} as before, so that all mm local observer state estimator errors converge to zero as fast as eλte^{-\lambda t} does. Next, assume that the neighbor graph is v¯<m\bar{v}<m-vertex redundantly strongly connected in that it is strongly connected and remains strongly connected after any vv¯v\leq\bar{v} vertices are removed. Assume in addition that the system described by (1), (2) is v¯\bar{v} redundantly jointly observable in that the system which results after any vv¯v\leq\bar{v} output measurements yiy_{i} have been deleted, is still jointly observable. Let 𝒟\mathcal{D} denote the family of all nonempty subsets 𝐝𝐦\mathbf{d}\subset\mathbf{m} such that each subset 𝐝𝒟\mathbf{d}\in\mathcal{D} contains at least mv¯m-\bar{v} vertices. Thus each loss of at most v¯\bar{v} vertices results in a strongly connected subgraph of (t)\mathbb{N}(t) for some subset 𝐝𝒟\mathbf{d}\in\mathcal{D}; call this subgraph 𝐝(t)\mathbb{N}_{\mathbf{d}}(t). Correspondingly, let Σ𝐝\Sigma_{\mathbf{d}} denote the multi-channel linear system which results when those outputs yiy_{i} i𝐝i\not\in\mathbf{d} are deleted from (1), (2). Thus Σ𝐝\Sigma_{\mathbf{d}} is a jointly observable multi-channel linear system whose channel outputs are the yi,i𝐝y_{i},\;i\in\mathbf{d}. Fix λ>0\lambda>0.

Fix 𝐝𝒟\mathbf{d}\in\mathcal{D} and let m𝐝m_{\mathbf{d}} denote the number of vertices in 𝐝\mathbb{N}_{\mathbf{d}}. Since Σ𝐝\Sigma_{\mathbf{d}} is jointly observable it is possible to compute a number ρ𝐝\rho_{\mathbf{d}} which satisfies (17). Using the pair (ρ𝐝,m𝐝)(\rho_{\mathbf{d}},m_{\mathbf{d}}) in place of the pair (ρ,m)(\rho,m) in (18) and (22), it is possible to calculate a value of qq, for which (22) holds. In other words, for this value of qq, henceforth labelled q𝐝q_{\mathbf{d}} Theorem 22 holds for the multichannel system Σ𝐝\Sigma_{\mathbf{d}} and neighbor graph 𝐝(t)\mathbb{N}_{\mathbf{d}}(t). By then picking

q=max𝐝𝒟q𝐝q^{*}=\max_{\mathbf{d}\in\mathcal{D}}q_{\mathbf{d}}

one obtains a value of qq for which Theorem 22 holds for all pairs (Σ𝐝,𝐝(t))(\Sigma_{\mathbf{d}},\mathbb{N}_{\mathbf{d}}(t)) as 𝐝\mathbf{d} ranges over 𝒟\mathcal{D}. Suppose a hybrid observer using q=qq=q^{*} is implemented. Suppose in addition that at some time tt^{*}, for some specific 𝐝𝒟\mathbf{d}\in\mathcal{D}, agents with labels in 𝐦𝐝\mathbf{m}-\mathbf{d} stop functioning. Clearly the remaining agents with labels in 𝐝\mathbf{d} will be able to deliver the desired state estimates with the prescribed convergence rate bounds. In this sense, the observer under consideration is resilient to vertex losses. However, unlike the loss or addition of edges mentioned above, no claim is being made at this point about what might happen if some or all of the lost vertices rejoin the network, especially if this loss-gain process is rapidly reoccurring over and over as time evolves.

A similar approach can be used to deal with the situation when at some time tt^{*}, the network gains some additional agents. In this case one would have to specify all possibilities and make sure that for each one, one has a strongly connected graph and a jointly observable system.

A little thought reveals that what makes it possible to deal with a change in the number of vertices in this way, is the fact that there is a single scalar quantity, namely qq, with the property that for each possible graphical configuration resulting from an anticipated gain or loss of vertices, there is a value of qq large enough for the distributed observer to perform correctly and moreover if qq is assigned the maximum of these values then the distribute observer will perform correctly no matter which of the anticipated vertex changes is actually encountered. Since the distributed observers described in [23, 24, 26] also require the adjustment of only a single scalar-valued quantity for a given neighbor graphs, the same basic idea just described can be used to make the observers in [23, 24, 26] resilient to a one-time gain or loss of the number of vertices on their associated neighbor graph. On the other hand, some distributed observers such as the ones described in [15, 2, 38] are not really amenable to this kind of generalization because for such observers changes in network topology require completely new designs involving the change of many of the observer’s parameters. There are also papers [39, 40] deal with sensor attacks, where a malicious attacker can manipulate their observations arbitrarily when each sensor only has one dimensional measurement.

6 Simulation

The following simulations are intended to illustrate (i) the performance of the hybrid observer in the face of system noise, (ii) the robustness of the hybrid observer with respect to variations of event time sequences, and (iii) resilience of the hybrid observer to the loss or gain of an agent. Consider the four channel, four-dimensional, continuous-time system described by the equations x˙=Ax,yi=Cix,i{1, 2, 3, 4}\dot{x}=Ax,\;y_{i}=C_{i}x,\;i\in\{1,\;2,\;3,\;4\}, where

A=[0.10.4000.10.100000.20.20020.1]A=\begin{bmatrix}-0.1&0.4&0&0\\ -0.1&-0.1&0&0\\ 0&0&-0.2&0.2\\ 0&0&-2&0.1\end{bmatrix}

and CiC_{i} is the iith unit row vector in IR1×4{\rm I\!R}^{1\times 4}. Note that AA is a stable matrix with two eigenvalues at 0.1±j0.2-0.1\pm j0.2 and a pair of complex eigenvalues at 0.05±j0.6144-0.05\pm j0.6144. While the system is jointly observable, no single pair (Ci,A)(C_{i},A) is observable. However the system is “redundantly jointly observable” in that what remains after the removal of any one output yiy_{i}, is still jointly observable. For the first two simulations (t)\mathbb{N}(t) is switching back and forth between Figure 2(a) and Figure 2(b), and for the third simulation the neighbor graph is as shown in Figure 2(a). Both graphs are strongly connected, and the graph in Figure 2(a) is redundantly strongly connected in that it is strongly connected and remains strongly connected after any one vertex is removed.

Refer to caption
(a)  
Refer to caption
(b)  
Figure 2: (t)\mathbb{N}(t)

Suppose T=1T=1 for this system. To achieve a convergence rate of λ=2\lambda=2, λ¯\bar{\lambda} and qq are chosen to be q=50q=50 and λ¯=3\bar{\lambda}=3 respectively.
For agent 1: C¯1=[0 1]\bar{C}_{1}=[0\;1],

A¯1=[0.10.10.40.1],L1=[01001000],K1=[13.74.8]\bar{A}_{1}\!\!=\!\!\begin{bmatrix}-0.1&-0.1\\ 0.4&-0.1\end{bmatrix},\;L_{1}\!\!=\!\!\begin{bmatrix}0&1&0&0\\ 1&0&0&0\end{bmatrix},\;K_{1}\!\!=\!\!-\begin{bmatrix}13.7\cr 4.8\end{bmatrix}

For agent 2: C¯2=[0 1]\bar{C}_{2}=[0\;1],

A¯2=[0.10.40.10.1],L2=[10000100],K2=[54.74.8]\bar{A}_{2}=\begin{bmatrix}-0.1&-0.4\\ 0.1&-0.1\end{bmatrix},L_{2}=\begin{bmatrix}-1&0&0&0\\ 0&1&0&0\end{bmatrix},K_{2}=-\begin{bmatrix}54.7\\ 4.8\end{bmatrix}

For agent 3: C¯3=[0 1]\bar{C}_{3}=[0\;1],

A¯3=[0.120.20.2],L3=[00010010],K3=[30.64.9]\bar{A}_{3}=\begin{bmatrix}0.1&-2\\ 0.2&-0.2\end{bmatrix},L_{3}=\begin{bmatrix}0&0&0&1\\ 0&0&1&0\end{bmatrix},K_{3}=-\begin{bmatrix}30.6\\ 4.9\end{bmatrix}

For agent 4: C¯4=[0 1]\bar{C}_{4}=[0\;1],

A¯4=[0.20.220.1],L4=[00100001],K4=[2.324.9]\bar{A}_{4}=\begin{bmatrix}-0.2&-0.2\\ 2&0.1\end{bmatrix},L_{4}=\begin{bmatrix}0&0&-1&0\\ 0&0&0&1\end{bmatrix},K_{4}=-\begin{bmatrix}2.32\cr 4.9\end{bmatrix}

In all four cases the local observer convergence rates are all 22.

This system was simulated with x(0)=[3 2 4 1]x(0)=[3\;2\;4\;1]^{\prime} as the initial state of the process, w1(0)=[5 5]w_{1}(0)=[5\;5]^{\prime}, w2(0)=[5 5]w_{2}(0)=[5\;5]^{\prime}, w3(0)=[5 5]w_{3}(0)=[5\;5]^{\prime}, and w4(0)=[5 5]w_{4}(0)=[5\;5]^{\prime} as the initial states of the four local observers, x1(0)=x2(0)=[5 5 5 5]x_{1}(0)=x_{2}(0)=[5\;5\;5\;5]^{\prime}, and x3(0)=x4(0)=[4 4 4 4]x_{3}(0)=x_{4}(0)=[4\;4\;4\;4]^{\prime} as the initial estimates of the four local estimators.

Three simulations were performed. The first is intended to demonstrate performance in the face of system noise. For this a modified process dynamic of the form x˙=Ax+bν\dot{x}=Ax+b\nu is assumed where b=[1 1 1 1]b=[1\;1\;1\;1]^{\prime} and ν=cos10t\nu=\cos 10t is system noise. Traces of this simulation are shown in Figure 3 where x1(3)x_{1}^{(3)} and x(3)x^{(3)} denote the third components of x1x_{1} and xx respectively. Only the trajectory of x1(3)x_{1}^{(3)} is plotted because for agent 11 only the third component is unobservable, and all the other components are observable.

Refer to caption

Figure 3: Performance in the face of system noise

The second simulation, which is without system noise, is intended to demonstrate the hybrid observer’s robustness against a small change in the event time sequence of one of the agents. The change considered presumes that the event times of agent 4 occur .2T.2T time units before the the event times of the other three agents. Traces of this simulation are shown in Figure 4.

Refer to caption

Figure 4: Performance in the face of a perturbed event time sequence

The third simulation, also without system noise, is intended to demonstrate the hybrid observer’s resilience against the disappearance of agent 4 at time t=5t=5 and also against agent 44’s re-emergence at time t=7t=7. Traces of this simulation are shown in Figure 5.

Refer to caption

Figure 5: Performance in the face of abrupt node changes

Disruption appearing at the beginning of the traces for all three simulations are due to initial conditions and are not important. In the third simulation, the loss of agent 4 at t=5t=5 does not appear to have any impact whereas the trace shows that the re-emergence of agent 4 at t=7t=7 briefly effects performance. While the claims in this paper do not consider the possibility of of agent re-emergence, it is not surprising that this event does not cause misbehavior because the time between the loss and gain of the agent, namely 22 time units, is large compared to the time constants of the observer. Clearly much more work needs to be done here to better understand rapidly occurring and re-occurring losses and gains of agents.

7 Concluding Remarks

One of the nice properties of the hybrid observer discussed in this paper is that it is resilient. By this we mean that under appropriate conditions it is able to continue to provide asymptotically correct estimates of xx, even if communications between some agents break down or if one or several of the agents joins or leaves the network. The third simulation provides an example of this capability. As pointed out earlier, further research is needed to more fully understand observer resilience, especially the situation when agents join or leave the network.

Generally one would like to choose TT “small” to avoid unnecessarily large error overshooting between event times. Meanwhile it is obvious from (20) that the larger the number pp and consequently the number of iterations qq on each event-time interval, the faster the convergence. Two considerations limit the value of qq - how fast the parameter estimators can compute and how quickly information can be transmitted across the network. We doubt the former consideration will prove very important in most applications, since digital processors can be quite fast and the computations required are not so taxing. On the other hand, transmission delays will almost certainly limit the choice of qq. A model which explicitly takes such delays into account will be presented in another paper.

A practical issue is that the development in this paper does not take into account measurement noise. On the other hand, the observer provides exponential convergence and this suggests that if noisy measurements are considered, the observer’s performance will degrade gracefully with increasing noise levels. Of course one would like an “optimal” estimator for such situations in the spirit of a Kalman filter. Just how to formulate and solve such a problem is a significant issue for further research.

{ack}

This work was supported by NSF grant 1607101.00, AFOSR grant FA9550-16-1-0290, and ARO grant W911NF-17-1-0499.

References

  • [1] L. Wang, A. S. Morse, D. Fullmer, and J. Liu. A hybrid observer for a distributed linear system with a changing neighbor graph. In Proceedings of the 2017 IEEE Conference on Decision and Control, pages 1024–1029, 2017.
  • [2] L. Wang and A. S. Morse. A distributed observer for an time-invariant linear system. IEEE Transactions on Automatic Control, 63(7):2123–2130, 2018.
  • [3] Reza Olfati-Saber. Distributed Kalman filter with embedded consensus filters. Proceedings of the 44th IEEE Conference on Decision and Control, and the European Control Conference, CDC-ECC ’05, pages 8179–8184, 2005.
  • [4] Reza Olfati-Saber. Distributed Kalman filtering for sensor networks. In Proceedings of the 46th IEEE Conference on Decision and Control, pages 5492–5498, 2007.
  • [5] Reza Olfati-Saber. Kalman-Consensus Filter : Optimality, stability, and performance. Proceedings of the 48h IEEE Conference on Decision and Control, and the 28th Chinese Control Conference, pages 7036–7042, dec 2009.
  • [6] Usman A Khan, Soummya Kar, Ali Jadbabaie, and José M.F. Moura. On connectivity, observability, and stability in distributed estimation. In Proceedings of the IEEE Conference on Decision and Control, pages 6639–6644, 2010.
  • [7] U. A. Khan and Ali Jadbabaie. On the stability and optimality of distributed Kalman filters with finite-time data fusion. In Proceedings of the 2011 American Control Conference, pages 3405–3410, 2011.
  • [8] Jaeyong Kim, Hyungbo Shim, and Jingbo Wu. On distributed optimal Kalman-Bucy filtering by averaging dynamics of heterogeneous agents. In Proceedings of the 55th IEEE Conference on Decision and Control, pages 6309–6314, 2016.
  • [9] Jingbo Wu, Anja Elser, Shen Zeng, and Frank Allgöwer. Consensus-based Distributed Kalman-Bucy Filter for Continuous-time Systems. IFAC-PapersOnLine, 49(22):321–326, 2016.
  • [10] Reza Olfati-Saber and Parisa Jalalkamali. Coupled distributed estimation and control for mobile sensor networks. IEEE Transactions on Automatic Control, 57(10):2609–2614, 2012.
  • [11] Mohammadreza Doostmohammadian and Usman A. Khan. On the genericity properties in distributed estimation: Topology design and sensor placement. IEEE Journal of Selected Topics in Signal Processing, 7(2):195–204, 2013.
  • [12] Usman A. Khan and Ali Jadbabaie. Collaborative scalar-gain estimators for potentially unstable social dynamics with limited communication. Autom., 50:1909–1914, 2014.
  • [13] Shinkyu Park and NC Martins. An augmented observer for the distributed estimation problem for LTI systems. In Proceedings the 2012 American Control Conference, pages 6775–6780, 2012.
  • [14] Shinkyu Park and Nuno C. Martins. Necessary and sufficient conditions for the stabilizability of a class of LTI distributed observers. In Proceedings of the 51st IEEE Conference on Decision and Control, pages 7431–7436, 2012.
  • [15] Shinkyu Park and Nuno C Martins. Design of Distributed LTI Observers for State Omniscience. IEEE Transactions on Automatic Control, 62(2):561–576, 2017.
  • [16] Aritra Mitra and Shreyas Sundaram. Distributed Observers for LTI Systems. IEEE Transactions on Automatic Control, 63(11):3689–3704, 2018.
  • [17] Behçet Açikmeşe, Milan Mandić, and Jason L Speyer. Decentralized observers with consensus filters for distributed discrete-time linear systems. Automatica, 50(4):1037–1052, 2014.
  • [18] Valery A. Ugrinovskii. Distributed robust estimation over randomly switching networks using hh_{\infty} consensus. Automatica, 49:160–168, 2013.
  • [19] Álvaro Rodríguez del Nozal, Pablo Millán, Luis Orihuela, Alexandre Seuret, and Luca Zaccarian. Distributed estimation based on multi-hop subspace decomposition. Automatica, 99:213–220, 2019.
  • [20] Francisco Castro Rego, Ye Pu, Andrea Alessandretti, A. Pedro Aguiar, António M. Pascoal, and Colin N. Jones. A distributed luenberger observer for linear state feedback systems with quantized and rate-limited communications. IEEE Transactions on Automatic Control, 66(9):3922–3937, 2021.
  • [21] Shih-Ho Wang Shih-Ho Wang and E Davison. On the stabilization of decentralized control systems. IEEE Transactions on Automatic Control, 18(5):473–478, 1973.
  • [22] J. P. Corfmat and A. S. Morse. Decentralized control of linear multivariable systems. Automatica, 12(5):479–497, September 1976.
  • [23] Taekyoo Kim, Hyungbo Shim, and Dongil Dan Cho. Distributed Luenberger Observer Design. In Proceedings of the 55th IEEE Conference on Decision and Control, pages 6928–6933, Las Vegas, USA, 2016.
  • [24] Weixin Han, Harry L. Trentelman, Zhenhua Wang, and Yi Shen. A simple approach to distributed observer design for linear systems. IEEE Transactions on Automatic Control, 64(1):329–336, 2019.
  • [25] Weixin Han, Harry L. Trentelman, Zhenhua Wang, and Yi Shen. Towards a minimal order distributed observer for linear systems. Systems and Control Letters, 114:59–65, 2018.
  • [26] L. Wang, J. Liu, and A. S. Morse. A distributed observer for a continuous-time linear system. In Proceedings of 2019 American Control Conference, pages 86–89, Philadelphia, PA, USA, July 2019.
  • [27] Jin Gyu Lee and Hyungbo Shim. A distributed algorithm that finds almost best possible estimate under non-vanishing and time-varying measurement noise. IEEE Control Systems Letters, 4(1):229–234, 2020.
  • [28] L. Wang, J. Liu, A. S. Morse, and B. D. O. Anderson. A distributed observer for a discrete-time linear system. In Proceedings of the 2019 IEEE Conference on Decision and Control, pages 367–372, 2019.
  • [29] Y. Li, S. Phillips, and R.G. Sanfelice. Robust Distributed Estimation for Linear Systems under Intermittent Information. IEEE Transactions on Automatic Control, 63(4):1–16, 2018.
  • [30] Aritra Mitra, John A Richards, Saurabh Bagchi, and Shreyas Sundaram. Finite-Time Distributed State Estimation over Time-Varying Graphs : Exploiting the Age-of-Information. 2018. arXiv:1810.06151 [cs.SY].
  • [31] L. Wang and A. S. Morse. A distributed observer for a continuous-time, linear system with a time-varying graph. arXiv:2003.02134v1, mar 2020.
  • [32] L. Wang, D. Fullmer, and A. S. Morse. A distributed algorithm with an arbitrary initialization for solving a linear algebraic equation. In Proceedings of the 2016 American Control Conference (ACC), pages 1078–1081, July 2016.
  • [33] S. Mou, J. Liu, and A. S. Morse. An distributed algorithm for solving a linear algebraic equation. IEEE Transactiona on Automatic Control, pages 2863–2878, 2015.
  • [34] Mohammadreza Doostmohammadian, Usman A. Khan, Mohammad Pirani, and Themistoklis Charalambous. Consensus-based distributed estimation in the presence of heterogeneous, time-invariant delays. IEEE Control Systems Letters, 6:1598–1603, 2022.
  • [35] L. Wang, D. Fullmer, and A. S. Morse. A distributed algorithm with an arbitrary initialization for solving a linear algebraic equation. In Proceedings of the 2016 American Control Conference, pages 1078–1081, 2016.
  • [36] F. Liu, L. Wang, D. Fullmer, and A. S. Morse. Distributed feedback control of multi-channel linear systems. arXiv:1912.03890, 2020. under revision.
  • [37] T. Kim, D. Lee, and H. Shim. Decentralized design and plug-and-play distributed control for linear multi-channel systems. arXiv:2011.09735, 2020.
  • [38] Mohammadreza Doostmohammadian, Hamid R. Rabiee, Houman Zarrabi, and Usman A. Khan. Distributed estimation recovery under sensor failure. IEEE Signal Processing Letters, 24(10):1532–1536, 2017.
  • [39] Chanhwa Lee, Hyungbo Shim, and Yongsoon Eun. On redundant observability: From security index to attack detection and resilient state estimation. IEEE Transactions on Automatic Control, 64(2):775–782, 2019.
  • [40] Xingkang He, Xiaoqiang Ren, Henrik Sandberg, and Karl Henrik Johansson. How to secure distributed filters under sensor attacks. IEEE Transactions on Automatic Control, 67(6):2843–2856, 2022.