Chapter 2

Symmetry, Stability and Scalability

People today see not the moon the ancients saw, and yet this moon shone once upon the ancients.

—Li Bai (701–762), ‘Inquiring of the moon, with wine in hand'

This chapter introduces general models of distributed control systems based on graph theory and linear system theory. The symmetry property of distributed control systems in the frequency domain is defined and investigated. Classic results of the frequency-domain test of the stability of multivariable feedback control systems are reviewed with the emphasis on their extension to the semi-stability and the scalability for distributed control systems.

2.1 System Model

2.1.1 Graph-Based Model of Distributed Control Systems

Consider a distributed control system with n subsystems, each of which is also called an agent in this book. Let the model of the dynamics of the ith agent be given by the following state-space model

(2.1) equation

where img, img and img denote the state, input and output, respectively, of the ith agent, Ti is the input delay. Note that in distributed control systems input delays are usually caused by the processing time for the packets arriving at each agent and/or the communication time between actuators and controllers.

The agents are interconnected according a fixed topology which can be described by a digraph img of order n. Denote by Ni the set of neighbors of agent i. By (1.9), the aggregated measurement of agent i is based on the information from its neighbors and itself:

(2.2) equation

where τij is the communication delay from agent j to agent i which mainly consists of propagation delay and queueing delay, τii is called self-delay which is often introduced to match the communication delays, aij is the weight for the output channel from agent j to agent i and aii is the weight for the self-loop of agent i. To incorporate the aggregated measurement (1.9) and the relative measurement (1.10) in a unified form, throughout this chapter, we assume that aij in (2.2) can be either positive or negative.

Suppose that agent i is manipulated by the following local controller

(2.3) equation

where img is the state of the local controller for agent i, img are parameters of the controller and ri(t) is some reference signal for agent i.

Denote by Gi(s) the transfer function of agent i, i.e.,

(2.4) equation

and by κi(s) the transfer function of the ith controller, i.e.,

(2.5) equation

The feedback control system for agent i is shown in Figure 2.1. Let img and img be the Laplace transformation of the output and input, respectively, of the i-agent. Denote

(2.6) equation

(2.7) equation

(2.8) equation

(2.9) equation

(2.10) equation

(2.11) equation

and

(2.12) equation

A(s) and img can be considered as the weighted adjacency matrix and generalized weighted adjacency matrix, respectively, of topology digraph G. As a matter of fact, in our model the matrix img can be further extended to the following form

(2.13) equation

where the transfer function img describes the possible dynamics of the communication channel from agent j to agent i.

Figure 2.1 Feedback control system of agent i.

img

With notations introduced above, the closed-loop system composed of (2.1), (2.2) and (2.3) can be sketched in Figure 2.2. The transfer function matrix from R(s) to Y(s) is given by

(2.14) equation

Figure 2.2 Diagram of distributed control system.

img

The system shown in Figure 2.2 is an interconnection of two parts. The first part G(s)K(s) is a diagonal matrix, which represents the agent dynamics. The second part A(s) + Λ(s) can be considered as a generalized complex adjacency matrix of the topology graph of the network; it reflects both the topology and the channel dynamics of the network used by the multi-agent system.

Definition 2.1 The dynamics of the agents in the distributed control system shown in Figure 2.1 are said to be homogeneous if κi(s)Gi(s) = κj(s)Gj(s) for all img; the network through which the agents are interconnected is said to be homogeneous if τij = τ0 and αij(s) = α0(s), img, where img and img. The distributed control system is said to be homogeneous if all the agents and the network are all homogeneous. Otherwise, the agents (network) are said to be heterogeneous if they are not homogeneous. The system is said to be heterogeneous if it contains heterogeneous agents or heterogeneous network.

2.1.2 Bipartite Distributed Control Systems

There are many distributed control systems in which information exchanges are conducted through forward and backward channels. The interconnection of agents in such systems can be described by the weighted bipartite digraph shown in Figure 2.3. We call these systems bipartite systems. Denote by (S, L) and A the vertex bipartition and the weighted adjacency matrix of the graph, respectively, where

(2.15) equation

(2.16) equation

(2.17) equation

Let the agents associated with vertex class S be described by transfer functions Gi(s), img; and the agents associated with L be described by transfer functions Pj(s), img. Agent i in S receives output information from its neighbors in L, and agent j in L receives output information from its neighbors in S, i.e.,

(2.18) equation

(2.19) equation

Figure 2.3 Bipartite graph for distributed control systems.

img

where img, img are the communication delay and the weight of the backward channel from agent l in L to agent i in S, respectively; img, img are the communication delay and the weight of the forward channel from agent i in S to agent l in L. Considering (2.17), the following relationship should hold

(2.20) equation

Agent i in S is manipulated by dynamic controller κi(s), i = 1, img, n, and agent l in L is manipulated by dynamic controller γl(s), l img 1, img, m. Denote by

equation

and

equation

the reference input vector and output vector of all the agents in S, respectively; by

equation

and

equation

the reference input vector and output vector of all the agents in L, respectively. Then, the bipartite system with forward and backward channels can be shown by Figure 2.4, where

equation

Figure 2.4 Bipartite distributed control system.

img

The transfer function matrix from R1(s) to Y1(s) is given by

(2.21) equation

and the transfer function matrix from R2(s) to Y2(s) is given by

(2.22) equation

Denote img and img. Then, the open-loop transfer function of W1(s) or W2(s) has the same structure as that of W(s) given by (2.14), i.e., it consists of a diagonal matrix corresponding to the agents' dynamics and a square matrix corresponding to the topology graph. This suggests that it is possible to deal with the stability problem of two kinds of distributed control systems in a unified framework.

2.2 Symmetry in the Frequency Domain

2.2.1 Symmetric Systems

In practice, many multi-agent systems are based on networks with undirected topology graph. In this case, if there is no delay and all the channels have constant dynamics, i.e., τij = 0 αij(s) = 1, img, then the system shown in Figure 2.2 is said to be symmetric because the adjacency matrix of the topology graph is symmetric. Can such a conception of symmetry be extended to the case when there are non-zero delays or non-constant dynamics in network channels? In this book, we define the symmetry of networked-based distributed control systems as follows.

Definition 2.2 Let the open-loop transfer function of a distributed control system be given by img, where img is a diagonal matrix and img is a square matrix. The system is said to be symmetric if img can be decomposed as

(2.23) equation

where Θ(s) is a diagonal matrix and img satisfies the following condition:

(2.24) equation

A starting point for this definition is that a complex generalized adjacency matrix of an undirected graph should be a Hermitian matrix as we defined in Chapter 1. Indeed, letting s = jω, then from (2.24) we have img. So, if we absorb the diagonal matrix Θ(s) in the agents' dynamics which also form a diagonal matrix, then img can be regarded as a generalized adjacency matrix of the undirected topology graph in the frequency domain.

Example 2.3 Symmetry in the frequency domain.

Consider the simplest undirected graph of two nodes, shown by Figure 2.5. Suppose the communication delays from node 1 to node 2 and from node 2 to node 1 are τ21 > 0 and τ12 > 0, respectively. It is natural to set e−τ12s and e−τ21s as the weights in the Laplace domain. But, the matrix

equation

Figure 2.5 Symmetry in the frequency domain.

img

does not satisfy conjugate symmetry even in the case τ12 = τ21 = τ. In the frequency domain, A(jω) is not a Hermitian matrix, either. Now, let us scale the matrix A(s) by multiplying

equation

where

(2.25) equation

which is called the round-trip time between node 1 and node 2. Then, we find that

equation

where img is a diagonal matrix, img satisfies img, and hence, img is a Hermitian matrix. So, by Definition 2.2, the system based on the graph shown in Figure 2.5 is symmetric in the frequency domain even if τ12τ21.

Absorbing Θ(s) into the diagonal transfer function matrix of the agents implies that the agent associated with node 2 obtains an additional time delay T. So, this example shows that sometimes the frequency-domain symmetry of a communication topology can be achieved by introducing a lag in the time domain for agents!

Now, let us consider the general distributed control system composed of (2.1), (2.2) and (2.3) or equivalently the system shown in Figure 2.1. Assume the interconnection topology graph of the system is undirected. Then, to match the symmetry requirement of Definition 2.2, there should exist img, such that

(2.26) equation

It is said that the system has symmetric communication delays if the requirement (2.26) is satisfied. Note that the symmetric communication delay does not imply τij = τji.

2.2.2 Symmetry of Bipartite Systems

Before we study the symmetry property of bipartite systems with forward and backward channels, let us introduce the notion of semi-homogeneousness as follows.

Definition 2.4 Consider a bipartite system which is based on the bipartite graph G(SL). It is said to be semi-homogeneous by vertex L if for each agent i in S, the following conditions are satisfied:

1. all the neighbors of agent i have homogeneous dynamics, i.e.,

(2.27) equation

where img is a constant and αi(s) is a scalar transfer function that is independent of l;
2. the round-trip delay from agent i to its any neighbor in L is a constant, i.e.,

(2.28) equation

Similarly, one can also define the semi-homogeneousness of the system by vertex class S.

Remark. Obviously, if all the agents in vertex class L have homogeneous dynamics, (2.27) holds.

The next theorem shows that the technique used in Example 2.3 can be applied to general semi-homogeneous bipartite systems.

Theorem 2.5 If the system based on the bipartite graph G(SL) is semi-homogeneous by L or by S, then it is symmetric in the frequency domain.

Proof. We just consider the case that the system is semi-homogeneous by L. The proof for the other case is the same.

Denote Γ0 = diag{γ1, img, γm} and Ξ(s) = diag{α1(s), img, αn(s)}. Then, under assumption (2.27) we have

equation

Denote T(s) = diag{eT1s, img, eTns}. Then, by using (2.20) and (2.28), one can see that

equation

So, under the assumption of the semi-homogeneousness of the system by L, we get

equation

So, we get the open-loop transfer function of the system as

equation

where img is a diagonal matrix, img satisfies img. Then, by Definition 2.2, the system is symmetric in the frequency domain. img

2.3 Stability of Multivariable Systems

2.3.1 Poles and Stability

Let the state-space representation of the closed-loop system (2.14) be

(2.29) equation

(2.30) equation

where img is the state-vector of the system, {A0, B, C, D} is the minimal state-space realization of the system with all delays being zero, img, are all possible time-delays in the closed-loop system, img, are the matrices corresponding to time-delayed states. Then, the poles of the system are the roots of the characteristic equation

(2.31) equation

where is ϕ(s) is the characteristic (or pole) quasi-polynomial of the system. We also call ϕ(s) characteristic (or pole) quasi-polynomial corresponding to the transfer function W(s) given by (2.14). By the theory of the time-delayed system (Gu, Kharitonov and Chen 2003), the stability of system (2.29) is fully determined by the characteristic quasi-polynomial ϕ(s). Namely, the system is stable if and only if ϕ(s) has no pole in the closed right half of the complex plane.

In this book we call the open left half of the complex plane simply the LHP and denote it as img, and call the closed right half of the complex plane simply as RHP and denote it as img. For simplicity of statement we also call the closed left half of the complex plane simply the closed LHP and denote it as img, and call the open right half of the complex plane simply the open RHP and denote it as img.

Definition 2.6 The time-delayed system (2.29) is said to be stable if

(2.32) equation

By this definition there is no difference between “stable” and “asymptotically stable”, which are both used in this book. The following notion of semi-stability is also used in this book.

Definition 2.7 The time-delayed system (2.29) is said to be semi-stable if

(2.33) equation

In particular, it is said to be steady semi-stable if

(2.34) equation

By this definition, a linear time-invariant system is semi-stable if all its poles are inside the closed LHP which includes the imaginary axis, and steady semi-stable if all its poles are inside the LHP or at the origin of the complex plane. Obviously, if each root on the imaginary axis is simple, the semi-stability reduces to the so-called marginal stability.

The following theorem from MacFarlane and Karcanias (1976) allows us to determine the characteristic quasi-polynomial of the system directly from the transfer function matrix.

Theorem 2.8 The pole polynomial ϕ(s) corresponding to transfer function W(s) is the least common denominator of all non-identically-zero minors of all orders of W(s).

Let us illustrate the theorem by the following example, which shows how to determine the poles of the system directly from the transfer function matrix.

Example 2.9 Consider the transfer function matrix

(2.35) equation

The minors of order 1 are img, img, img and img. The minor of order 2 is the determinant

equation

The least common denominator of all the minors is

equation

Therefore, the system has four poles: double at s = 0 and double at s = − 2.

For distributed control systems time-delay is inevitable. For a time-delayed system W(s) = G(s)eτs, where G(s) is a rational function matrix and τ > 0 is the delay constant, one can use Theorem 2.8 to handle the rational part G(s) and calculate its “regular” poles. Then, by using a definition of the exponential function we have

(2.36) equation

So, if a transfer function has a factor eτs, equation (2.36) implies that it also has an infinite number of LHP poles at img.

2.3.2 Zeros and Pole-Zero Cancelation

In general, zeros are the values of complex variable s at which the transfer function matrix W(s) loses rank. This is the basis of the following definition of zeros for multivariable systems (or multiple-input multiple-output systems, or briefly MIMO systems) (MacFarlane and Karcanias (1976).

Definition 2.10 zi is a zero of W(s) if the rank of W(zi) is less than the normal rank of W(s). The zero polynomial is defined as img where nz is the number of finite zeros of W(s).

Note that the normal rank of W(s) is the rank of W(s) at all values of s except at a finite number of points (which are the zeros). The zeros defined above are sometimes called transmission zeros.

If the system is described by a state-space model as follows:

(2.37) equation

(2.38) equation

then the transmission zeros of W(s) are the values s = z for which the matrix P(s) loses rank.

Based on this definition of zeros we can introduce the notion of minimum phase systems.

Definition 2.11 (Chen 1984; Davison 1983) The system defined by (2.37) and (2.38) is said to be minimum phase if all its transmission zeros lie inside the LHP.

The following theorem from MacFarlane and Karcanias (1976) is often used for calculating the zeros of a transfer function matrix.

Theorem 2.12 The zero polynomial corresponding to a minimal realization of the system W(s) is the greatest common divisor of all the numerators of all order-r minors of W(s), where r is the nominal rank of W(s), provided that these minors have been adjusted in such a way that the pole polynomial ϕ(s) is their denominator.

For single-input single-output (SISO) systems, if the transfer function has a pole and a zero at the same place on the complex plane, it is said that there is a pole-zero cancelation. For multivariable systems the situation becomes much more complicated. The following example shows a case of pole-zero cancelation.

Example 2.13 Consider the transfer function matrix

(2.39) equation

The minors of order 1 are img, img, img and img. The minor of order 2 is the determinant

equation

The least common denominator of all the minors is

equation

So, by Theorem 2.8, the system has three poles: two at s = 0 and one at s = − 2. And by Theorem 2.12, the zero polynomial of the system is z(s) = s − 5. Note that there is a pole-zero cancelation at s = − 2 when calculating the determinant of W(s).

The next example shows that even if a multivariable system has a pole and a zero at the same place, it does not mean that there must be a pole-zero cancelation.

Example 2.14 Consider the transfer function matrix

(2.40) equation

By Theorem 2.8, W(s) has poles at s = 0 and s = − 2. We also notice that W(s) loses rank at s = 0 and s = − 2. So, s = 0 and s = − 2 are also zeros of W(s), but there is no pole-zero cancelation.

If W(s) loses rank for any value of the complex variable s, by Definition 2.10 or Theorem 2.12, it does not mean that W(s) has (transmission) zeros at any place in the complex plane. However, such a “non-existing” zero may also cause a pole-zero cancelation. Sometimes, we call such a zero an invariant zero.

Example 2.15 Let

equation

By Definition 2.10, both W1(s) and W2(s) have no zero. Now, let us check the poles of W1(s) and W2(s). By Theorem 2.8, W1(s) has only a single pole at s = 0 but W2(s) has double poles at s = 0. Where is another “pole” s = 0 for W1(s)? It may be assumed that the pole is canceled by an invariant zero at s = 0 because det W(s) = 0 for any value of s including zero, although such a zero is not counted by Definition 2.10.

Now, we consider the poles of the cascaded system W1(s)W2(s). At the first look, one may guess that the cascaded system has triple poles at s = 0. But, since

equation

is singular for any complex variable s, by Theorem 2.8, we know that W1(s)W2(s) only has double poles at s = 0. The third pole at s = 0 is canceled again by the invariant zero.

Invariant zeros are actually caused by the singularity of system matrices. Note that in consensus problems, which will be discussed in Chapter 6 and Chapter 7, open-loop transfer function matrices are usually singular for any complex variable since the Laplacian matrix is singular.

2.4 Frequency-Domain Criteria of Stability

The return difference equation of the closed-loop system shown in Figure 2.2 is

(2.41) equation

According to linear system theory (Chen 1984; Desoer and Vidyasagar 1975), it holds that

(2.42) equation

where ϕc(s) is the characteristic polynomial (or quasi-polynomial for time-delayed systems) of the closed-loop system, and ϕo(s) is the characteristic polynomial (or quasi-polynomial for time-delayed systems) of the open-loop system. img is referred to as the open-loop transfer function of the system. For the purpose of the well-posedness of the system we always assume that

equation

Equation (2.42) shows that the solutions of the following equation

(2.43) equation

may be a proper subset of the poles of the closed-loop system since cancelations may occur on the right side of (2.42). Nevertheless, if there is no cancelation between RHP pole(s) of the closed-loop system and RHP pole(s) of the open-loop system, the system is stable if and only if all the solutions of the equation (2.43) have strictly negative real parts.

Exercises2.16 Let the transfer function matrix W(s) be given by (2.14). Show that if

equation

then (1) W(s)→ ∞ as s→ ∞, (2) the closed-loop system does not have a state representation in the standard form (A, B, C, D).

2.4.1 Loop Transformation and Multiplier

For a given closed-loop system, one can make some loop transformation as shown by Figure 2.6 for stability analysis. Obviously, we have

equation

which yields the conclusion that

equation

if and only if

equation

and

equation

where img denotes a region in the complex plane, which can be img, img, or img depending on the stability requirement. Therefore, if the feedback system composed of G1(s) and K(s) is (semi-)stable, the system shown by Figure 2.6(a) has the same (semi-)stability property as the system shown in Figure 2.6(b).

One can also introduce some multiplier into the loop as shown by Figure 2.7 and Figure 2.8.

Figure 2.6 (a) Original feedback system; (b) System with loop transformation.

img

Figure 2.7 (a) Original feedback system; (b) System with multiplier.

img

Figure 2.8 (a) Original feedback system; (b) System with mixed loop transformation.

img

Loop transformation is a very useful tool in the analysis of the stability of feedback systems. When we apply some stability criteria such as the Nyquist criterion, ignorance of pole cancelations between the open-loop transfer function and the closed-loop transfer function may result in incorrect results as we will show later in this chapter. Proper loop transformation can remove such cancelations because it changes open-loop transfer functions. Moreover, since some stability criteria such as spectral radius theorem, small-gain theorem and passivity theorem (positive realness theorem) provide only sufficient conditions of stability and hence are conservative, loop transformations and multipliers change open-loop transfer functions and are helpful in reducing the conservatism of these stability criteria.

2.4.2 Multivariable Nyquist Stability Criterion

The stability test for closed-loop systems can be performed by use of frequency-domain techniques based on the multivariable (or generalized) Nyquist stability criterion. The following theorem is a modified version of the generalized Nyquist stability criterion which has been widely used in references (see, e.g., Vidyasagar (1985).

Theorem 2.17 Consider a system with an open-loop transfer function img which has η RHP poles. Then, the closed-loop system has no RHP poles if and only if the Nyquist diagram of img

1. does not pass through the origin (i.e., img, for all img);
encircles the origin η times anticlockwise.

Remark 1. By the Nyquist diagram or the Nyquist plot of img we mean the image of img as s goes clockwise around the Nyquist D-contour. The Nyquist D-contour includes the entire jω-axis and an infinite semi circle into the right-half plane. If img has jω-axis poles, one can use the D-contour illustrated in Figure 2.9 to avoid these poles by making sufficiently small semi circles into the RHP around them. By using such a D-contour the jω-axis poles of img are obviously excluded from the number of RHP poles of the open-loop system, which are sometimes briefly referred to as open-loop RPH poles (note that the “open-loop RHP pole” is not equal to the “open RHP pole”!). Since this D-contour goes around jω-axis poles of img in the anti-clockwise direction, when drawing the Nyquist plot of img one should add a semi circle with an infinite radius in the clockwise direction to connect the undefined part of the plot corresponding to each jω-axis pole. In this way, the jω-axis poles of the open-loop system are not regarded as “RHP poles”. So, the term “RHP poles” in the theorem actually should be understood as “open RHP poles”. If there is no cancelation of any of the jω-axis poles between the open-loop transfer function and the closed-loop transfer function, conditions (1) and (2) imply stability. Otherwise, only semi-stability is ensured.

Figure 2.9 D-contour excluding jω-axis poles.

img

Remark 2. The theorem itself does not tell if there is any RHP pole cancelation between the closed-loop system and the open-loop system. In Chapter 7 we will discuss how to justify if there exist multiple pole cancelations between the closed-loop system and the open-loop system at s = 0.

Remark 3. Since loop-transformation changes the open-loop transfer function but does not change the closed-loop transfer function, one can also use some loop transformation to avoid pole-zero cancelations of open-loop transfer functions and get the correct conclusion on the stability of the closed-loop system.

We use the following example to illustrate Theorem 2.17 and the associated remarks as well.

Example 2.18 Let img, where img, and img By Theorem 2.8 we know that img has a single pole at s = 0. Actually, as we have shown in Example 2.15, there is another open-loop pole at s = 0 which is canceled because det A = 0. Using the modified Nyquist D-contour given in Figure 2.9, the poles at s = 0 are not regarded as the RHP poles of img. So we have η = 0. Because

equation

we can draw the Nyquist plot of img as shown in Figure 2.10. In the plot we see that it neither passes through nor encircles the origin. Therefore, one may conclude that the closed-loop system has no RHP poles. But this does not lead to the conclusion that the system is stable. Instead, only the steady semi-stability can be derived, i.e., all the poles of the closed-loop system are inside the LHP or at s = 0.

We can investigate the stability of this system by making a loop transformation. The system diagram is sketched in Figure 2.11(a). It can be transformed into the form given by Figure 2.11(c). From Figure 2.11(c) we have the transfer function matrix of the open-loop system as img and img. Therefore, we have η = 0. The Nyquist plot of img is given in Figure 2.12. It passes through the origin, and thus, the system is just marginally stable. The conclusion is the same as that obtained from the original diagram.

Figure 2.10 Nyquist plots without s = 0 as an open-loop RHP pole.

img

Figure 2.11 Transformation of system diagram.

img

Figure 2.12 Nyquist plot of the transformed system.

img

As is well known, a great virtue of the classic Nyquist stability criterion is that it checks the closed-loop system stability by inspection of the Nyquist diagram of the open-loop transfer function. For SISO systems, Theorem 2.17 reduces to the classical Nyquist criterion since img. For the multivariable systems, however, this theorem requires a plot of img. Desoer and Wang (1980) proved a version of the generalized Nyquist stability criterion which checks the stability of the closed-loop system by inspection of the eigenloci of the open-loop system.

The eigenloci are defined as the set of all the eigenvalues of the frequency response of the open-loop transfer function matrix, img, as jω goes clockwise around the Nyquist D-contour. Note that, unlike the SISO case, in general img is not a rational function of s, and hence the conjugate property does not hold, i.e.,

equation

This leads to the fact that a single eigenlocus img may not form a closed path when ω varies from −∞ to ∞. Fortunately, when λi(s) is irrational on s, there always exists another eigenvalue λi+1(s) which is conjugate with λi(s), i.e.,

equation

Hence, the part of the eigenlocus img is conjugate with the part of the eigenlocus img, and the part of the eigenlocus img is conjugate with the part of the eigenlocus img. Therefore, when ω varies from −∞ to ∞, the eigenloci of img and img form two closed paths. In such a way, we can count the times in which the eigenloci img, enclose the point (− 1, j0).

Let N be the sum of the times in which the eigenloci img, enclose the point (− 1, j0) anticlockwise as jω goes clockwise around the Nyquist D-contour. Then, the generalized Nyquist stability criterion can be stated as follows.

Theorem 2.19 Consider a system with an open-loop transfer function img which has η open RHP poles. Then, the closed-loop system has no open RHP poles if and only if

(1) img, for all img;
(2) N = η.

Furthermore, if there is no cancelation of jω-axis poles between the open-loop system and the closed-loop system, the closed-loop system is stable.

Remark. The assumption of no cancelation of jω-axis poles between the open-loop system and the closed-loop system can always be satisfied by introducing a proper loop transformation.

For many network-based distributed control systems, open-loop systems composed of agents have no open RHP poles. The following theorem, which is actually a corollary of Theorem 2.19, is very useful for such systems.

Theorem 2.20 Suppose that the open-loop transfer function img has no open RHP poles. Then, the closed-loop system has no open RHP poles if and only if the eigenloci img, do not enclose the point (− 1, j0) as jω goes clockwise around the Nyquist D-contour. Furthermore, if there is no cancelation of jω-axis poles between the open-loop system and the closed-loop system, the closed-loop system is stable.

Theorem 2.17, Theorem 2.19 and Theorem 2.20 can be easily extended to discrete-time systems after replacing the D-contour by the unit circle in the complex plane. Note that if the z-transfer function matrix img of the open-loop system has a pole just on the unit circle, the unit circle should be modified by making a sufficiently small semi circle into the region outside the unit circle around the pole (see Figure 2.13).

Figure 2.13 Modified unit circle.

img

For convenience of statement, we denote by img the interior of the unit disc (briefly as IUD), i.e., img. The closure of img is denoted by img and briefly referred to as the closed IUD. We also denote by img the closed outer part of the unit disc (briefly as OUD), i.e., img and denote by img the open outer part of the unit disc (briefly as the open OUD), i.e., img.

Definition 2.21 Let ϕ(z) be the pole (quasi-)polynomial of a discrete-time. The system is said to be stable if

(2.44) equation

The system is said to be semi-stable if all its poles are in

(2.45) equation

In particular, it is said to be steady semi-stable if

(2.46) equation

Then, the discrete-time analog of Theorem 2.20 can be stated as follows. The discrete-time analogs of the other versions of the generalized Nyquist stability criterion are similar.

Theorem 2.22 Suppose that the open-loop transfer function img of a discrete-time system has no open OUD poles. Then, the closed-loop system has no open OUD poles if and only if the eigenloci img, do not enclose the point (− 1, j0) as ω varies fromπ to π. Furthermore, if there is no cancelation of unit-circle poles between the open-loop system and the closed-loop system, the closed-loop system is stable.

2.4.3 Spectral Radius Theorem and Small-Gain Theorem

Using the generalized Nyquist stability criterion, Theorem 2.17, it can be proved that the stability of a closed-loop system with a stable open-loop transfer function can be verified by testing the spectral radius of the open-loop transfer function at each frequency, which is defined as the maximum eigenvalue magnitude, i.e.,

(2.47) equation

Theorem 2.23 Consider a system with a stable open-loop transfer function img. Then, the closed-loop system is stable if

(2.48) equation

This theorem which is referred to as the spectral radius theorem was proved in Skogestad and Postlethwaite (1996). With some modification of the proof we can prove the following extended spectral radius theorem for the test of steady semi-stability and marginal stability.

Theorem 2.24 Consider a system with a stable open-loop transfer function img. Then, the closed-loop system is steady semi-stable, if

(2.49) equation

Moreover, if img, and img, then the closed-loop system is marginally stable with s = 0 as a simple pole.

Proof. Firstly, we note that the closed-loop system has no pure imaginary poles because img.

Secondly, we prove that the closed-loop system has no open RHP poles. We prove this by contradiction. Suppose that it has at least one open RHP pole. By the generalized Nyquist theorem (Theorem 2.17), img must enclose the origin, which implies that there exists an img such that img. Then, there exists a gain img img (0, 1) such that

(2.50) equation

because img for img = 0. Then, the following derivation leads a contradiction:

(2.51) equation

(2.52) equation

(2.53) equation

(2.54) equation

(2.55) equation

Based on the above discussion we conclude that the closed-loop system is steady semi-stable.

Thirdly, img implies that the closed-loop system has at least one pole at s = 0. Now, we show that any possible pole of the closed-loop system at s = 0 must be simple. Indeed, let us consider poles of a perturbed system with characteristic equation

(2.56) equation

where img img (0, 1). Obviously, img, for all img including ω = 0 because img = 1. So, by Theorem 2.23 the perturbed system is stable. By the rule of root locus (Nise (2000), root loci for perturbed systems with simple pole, double poles and triple poles, respectively, at s = 0 are shown in Figure 2.14. It is clear that only the case of simple pole can be stable.

Figure 2.14 Root locus near s = 0: (a) simple pole, (b) double poles, (c) triple poles.

img

Note that the open-loop system img in the above theorem may be obtained based on some loop transformation.

Example 2.25 Let us consider the system given in Example 2.18 again. After the loop transformation shown by Figure 2.11 we have

equation

It is easy to check that

equation

and img. So, the closed-loop system has no RHP poles except for a simple pole at s = 0, and hence, is marginally stable.

If we choose a matrix norm satisfying ||AB|| ≤ ||A|| · ||B||, we have img. Then, some small-gain theorems can be directly obtained from Theorem 2.23 and Theorem 2.24 as follows.

Theorem 2.26 Consider a system with a stable open-loop transfer function img. Then, the closed-loop system is stable if

(2.57) equation

Theorem 2.27 Consider a system with a stable open-loop transfer function img. Then, the closed-loop system has all poles in the LHP except for a simple pole at s = 0 and hence is marginally stable, if

(2.58) equation

(2.59) equation

2.4.4 Positive Realness Theorem

According to the generalized Nyquist stability criterion, Theorem 2.17, one can also analyze the stability of the closed-loop system by testing the positive realness of the eigenvalues of the transfer function matrix.

Theorem 2.28 Consider a system with a stable open-loop transfer function img. Then, the closed-loop system is stable if

(2.60) equation

With slight modification of the proof of Theorem 2.24 one can also obtain the following theorem for the test of marginal stability.

Theorem 2.29 Consider a system with a stable open-loop transfer function img. Then, the closed-loop system has all poles in the LHP except for a simple pole at s = 0 and hence is marginally stable, if

(2.61) equation

(2.62) equation

2.5 Scalable Stability Criteria

2.5.1 Estimation of Spectrum of Complex Matrices

To apply Theorem 2.17 or Theorem 2.19 it is necessary to get the eigenloci img. Generally, there is no analytical formula for calculating eigenvalues of matrices. The following lemmas can be used for the estimation of the spectrum of img denoted by σ(A): = {λ1(A), img, λn(A)}.

The first lemma is usually called Gershgorin's disc lemma (Horn and Johnson (1985).

Lemma 2.30 For any img it holds

(2.63) equation

where

(2.64) equation

Remark. The eigenvalues of A are also in the union of the discs:

(2.65) equation

From (2.64) it follows that img, which implies that any complex number in the disk Di has the absolute value less than or equal to the sum of the absolute values of the entries of the same row. So, the following corollary is straightforward.

Corollary 2.31 For any img it holds

(2.66) equation

Define

(2.67) equation

where img. We have ρ(A) = ρ(HAH−1) for img. Based on this fact we get the following corollary of Gershgorin's disc lemma for reducing the conservatism of Corollary 2.31

Corollary 2.32 For any img it holds

(2.68) equation

Let img and ξ*ξ = 1. Then, for any eigenvalue λi of A we have λi = λiξ*ξ = ξ*. So, by defining the field of values of A as follows (Horn and Johnson (1991):

(2.69) equation

the following lemma is obvious.

Lemma 2.33 For any img, it holds

(2.70) equation

If the complex matrix A can be written as a product of a diagonal matrix and a Hermitian matrix, then its spectrum can be further estimated by Vinnicombe's lemma (Vinnicombe (2000), which can be derived from Lemma 2.33.

Lemma 2.34 Let img, Q = Q* ≥ 0 and img. Then,

(2.71) equation

where ρ(·) denotes the matrix spectral radius, and Co(·) the convex hull.

Proof. Denote by N(A) the null space of the complex matrix A. Then,

equation

So, we have

equation

which implies that img. The lemma is thus proved.

Remark. The lemma was proved by Vinnicombe (2000) for the case of Q = Q* > 0. The current form of the lemma and the proof are given by Lestas and Vinnicombe (2005).

The following example illustrates that in most cases Vinnicombe's lemma is much less conservative than Gershgorin's disc lemma if A is a product of a diagonal matrix and a Hermitian matrix.

Example 2.35 Suppose img, and img is a Hermitian matrix with zero as diagonal entries, i.e., qii = 0. Then, by Gershgorin's disc lemma we have

equation

Now, by using Vinnicombe's lemma and Corollary 2.31 of Gershgorin's disc lemma we get

equation

img is a disc centered at the origin with radius as img. Co is a convex polygon contained in the disc as shown by Figure 2.15. Obviously, the higher the diversity degree of θi is, the closer the two regions are.

Figure 2.15 Comparison of Gershgorin's lemma and Vinnicombe's lemma.

img

2.5.2 Scalable Stability Criteria for Asymmetric Systems

Now, we consider the stability of the distributed control system described by (2.1), (2.2) and (2.3), which is shown by Figure 2.2. For convenience, let us re-sketched the diagram of the system as shown by Figure 2.16, where Λ(s) is defined by (2.11) and A(s) is defined by (2.12). The following theorem is a direct consequence of the spectral radius theorem and Gershgorin's disc lemma.

Figure 2.16 Distributed control system.

img

Theorem 2.36 Suppose Gi(s), img, have no open RHP poles. The closed-loop system shown by Figure 2.16 has no open RHP poles, and hence is semi-stable if

(2.72) equation

Furthermore, if there is no cancelation of jω-axis poles between the open-loop system and the closed-loop system, the closed-loop system is stable.

Proof. Denote by H(s) the upper part of the system as shown by Figure 2.16, i.e.,

equation

Then, the open-loop transfer function of the system can be written as diag{Hi(s)}A(s). By Theorem 2.20, the closed-loop system has no open RHP poles if and only if the eigenloci

equation

do not enclose the point (− 1, j0) for m = 1, img, n as ω varies from 0 to infinity. By Lemma 2.30, the spectrum of img is contained in the union of the following discs:

equation

Obviously, the eigenloci img do not enclose (− 1, j0) if

equation

which is equivalent to (2.72). Therefore, the closed-loop system does indeed have no open RHP poles. If there is no cancelation of jω-axis poles between the open-loop system and the closed-loop system, by Theorem 2.20 again, the closed-loop system is stable.

Theorem 2.36 gives a sufficient stability condition which is independent of communication delays τij, ij. But it still depends on the self-delays τii. However, if we replace e−jωτii by a unit disc img, then a sufficient condition for (2.72) is given by the following corollary.

Corollary 2.37 Suppose Gi(s), img, have no open RHP poles. The closed-loop system shown by Figure 2.16 has no open RHP poles if

(2.73) equation

Furthermore, if there is no cancelation of jω-axis poles between the open-loop system and the closed-loop system, the closed-loop system is stable.

This is a more conservative condition but it is independent of self-delays. The geometric meaning of condition (2.73) can be illustrated as follows: all the circle centered at the Nyquist plot of κiaiiGi(jω) with radius img is inside but does not touch the unit disc (see Figure 2.17).

Figure 2.17 Geometrical illustration of Corollary 2.37.

img

Figure 2.17 clearly shows that the condition (2.73) can be satisfied only if img. In the other hand, if img, it can be shown that there always exist local proportional controls stabilizing the system.

Corollary 2.38 Suppose img, img. Then, the system shown by Figure 2.16 is stable if control gains img, are sufficiently small.

Proof. Since Gi(s) has no jω-axis poles, we know that |Gi(jω)| is bounded for all img. Therefore,

equation

This implies that (2.72) is always satisfied if img, are sufficiently small.

Remark. If all self-delays are zero, i.e. τii = 0, the condition (2.72) reduces to

equation

This condition does not require img (see Figure 2.18 for illustration).

Figure 2.18 Geometrical illustration of Theorem 2.36 for the case τii = 0.

img

For the case when the transfer function of the open-loop system is stable, by applying Theorem 2.24 and the argument similar to the proof of Theorem 2.36, we can get the following scalable criterion for the test of the steady semi-stability and the marginal stability.

Theorem 2.39 Suppose img, img. The closed-loop system shown by Figure 2.16 is steady semi-stable if

(2.74) equation

Moreover, if ρ(diag{Hi(0)}A(0)) = 1, and det (I + diag{Hi(0)}A(0)) = 0, then the closed-loop system is marginally stable with s = 0 as a simple pole.

2.5.3 Scalable Stability Criteria for Symmetric Systems

Since img, the system in Figure 2.2 can be re-sketched in Figure 2.19. Recall that the system is said to be symmetric if img.

Figure 2.19 Distributed control system.

img

For symmetric systems one can get some scalable stability criterion that is less conservative than Theorem 2.36.

Theorem 2.40 Suppose img and Gi(s) has no open RHP poles. Let γi be the gain margin of the stable transfer function Gi(s), and img. Then, the closed-loop system shown by Figure 2.19 has no open RHP poles if the following conditions are satisfied for all ω ≥ 0:

(1) img
(2)img

Furthermore, if there is no cancelation of jω-axis poles between the open-loop system and the closed-loop system, the closed-loop system is stable.

Proof. We just prove the semi-stability part. The stability part follows directly from Theorem 2.20.

By Theorem 2.20, the closed-loop system has no RHP poles if and only if the eigenloci

equation

do not enclose the point (− 1, j0) for m = 1, img, n as ω varies from 0 to infinity. Note that

equation

And moreover, img because img. Therefore, by Lemma 2.34, the closed-loop system has no open RHP poles if

equation

The above condition is satisfied if

(2.75) equation

and

(2.76) equation

Since the spectral radius of any matrix is bounded by its maximum absolute row sum, (2.76) is satisfied if condition (2) of the theorem holds. Thus, the theorem is proved.

If we use Corollary 2.32 instead of Corollary 2.31 of Gershgorin's disc lemma in the proof, we can get the following result for reducing the conservatism of Theorem 2.40.

Theorem 2.41 Suppose img and Gi(s) has no open RHP poles. The closed-loop system shown by Figure 2.19 has no open RHP poles if there exist positive real numbers img such that the following conditions are satisfied for all ω ≥ 0:

(1) img
(2)img

Furthermore, if there is no cancelation of jω-axis poles between the open-loop system and the closed-loop system, the closed-loop system is stable.

If the condition

(2.77) equation

in the above theorem is satisfied, the second condition in Theorem 2.40 or Theorem 2.41 gives a local and hence scalable test for the stability. Therefore, condition (2.77) can be considered as a scalability condition. Note that the scalability condition (2.77) does not necessarily hold even if each Nyquist plot img does not contain (− 1, j0). This problem will be studied in the next chapter.

Exercises 2.42 Let

equation

and

equation

in the system shown by Figure 2.19. Derive scalable semi-stability conditions by Theorem 2.36, Theorem 2.40 and Theorem 2.41, respectively. Compare the conservatism of the obtained conditions with the help of a simulation toolbox (e.g., MATLAB®).

2.5.4 Robust Stability in Deformity of Symmetry

Symmetry of is such a fragile property that may be easily destroyed by small perturbations. Many factors in practical systems, such as uncertainty of agent dynamics, diversity of communication delays or heterogeneousness of channel dynamics, may lead to deformity of the symmetry defined by Definition 2.2 Therefore, any analysis or design method based on the assumption of the system symmetry must be robust against perturbations. Fortunately, the robust control theory can help us to extend the scalable stability criteria to most cases of deformity of symmetry.

Let us consider the case when there are some uncertainties in communication channels that cause the asymmetry of the system, as shown by Figure 2.20. Here, we assume that the nominal system, i.e., the system with ΔA(s) = 0, is symmetric. Suppose that the uncertainty matrix ΔA(s) is subject to the following structural constraint:

(2.78) equation

where img and img are some constant matrices that describe the structure of the uncertainty, Δ(s) satisfies the following condition:

(2.79) equation

where img is a known function of ω. Then, by loop transformation we can equivalently transform the system into an interconnection form shown by Figure 2.21 with

equation

According to the robust control theory (see, e.g., Feng, Tian and Xin 1996, Zhou, Doyle and Glover (1996)), under the assumption that

(2.80) equation

the asymmetric uncertain system shown by Figure 2.20 is robustly stable if and only if

(2.81) equation

We summarize the conclusion of the above discussion in the following theorem.

Figure 2.20 Distributed control system with uncertainty in network channels.

img

Figure 2.21 Symmetric system with asymmetric perturbations.

img

Theorem 2.43 Under the conditions (2.78), (2.79) and (2.80), the perturbed system shown by Figure 2.20 is robustly stable if and only if (2.81) holds.

Remark. A sufficient condition for img is that

(2.82) equation

Note that the left-hand side of (2.82) is just the nominal symmetric system. So, its stability can be checked by Theorem 2.40 or Theorem 2.41.

We can also consider the bipartite system with an additive uncertainty ΔP(s) in agent plants, as shown by Figure 2.22. Note that the uncertainty ΔP(s) may destroy the homogeneousness assumption (2.27), and hence, destroy the symmetry of the system. Suppose that ΔP(s) is subject to the following structural constraint:

(2.83) equation

where img and img describe the structure of the uncertainty, and Δ(s) satisfies the following condition:

(2.84) equation

where img is a known function of ω. The system can also be equivalently transformed into the interconnection form shown by Figure 2.21 with

(2.85) equation

Similarly, we assume that

(2.86) equation

Then, we have the following result.

Figure 2.22 Distributed control system with uncertainty in agent plants.

img

Theorem 2.44 Under the conditions (2.83), (2.84) and (2.86), the perturbed system shown by Figure 2.22 is robustly stable if and only if

(2.87) equation

where M(s) is given by (2.85).

2.6 Notes and References

Modeling of systems with forward and backward communication channels are discussed in many references on coordination control and congestion control (see, e.g., Arcak 2007; Low, Paganini and Doyle 2002; Wen and Arcak 2003). Using bipartite graphs to describe interconnection topologies of such systems and the symmetry analysis is new. The definitions of poles and zeros are given following MacFarlane and Karcanias (1976). Note that for multivariable systems there are various definitions of zeros, see MacFarlane and Karcanias (1976) and Rosenbrock (1973). For their computation, see Davison and Wang (1978) and Laub and Moore (1978). The difficulties associated with pole-zero cancelations in multivariable systems can be treated by state-space techniques using the concepts of controllability and observability. A pure algebraic approach to analyzing input-output properties including the equivalence between the systems after/before loop transformations can be found in more detail in the classic book of Desoer and Vidyasagar (1975), where general settings are used. The Nyquist stability was generalized in several ways for multivariable systems in the 1970s (Barman and Katzenelson 1974; Belletrutti and MacFarlane 1971; Callier and Desoer 1972; Davis 1972; Desoer and Vidyasagar 1975; MacFarlane and Postlethwaite 1977; Saeks 1975). Theorem 2.17 is taken from Vidyasagar (1985) while Theorem 2.19 adopts the result of Desoer and Vidyasagar (1975). Gershgorin's disc lemma has been a well-known tool for deriving scalable or decentralized stability results due to the study of frequency-domain analysis and design of multivariable systems by the so-called British school (Belletrutti and MacFarlane 1971; MacFarlane and Karcanias 1976; Rosenbrock 1970, 1974). Lemma 2.34 is a recent important contribution made by Vinnicombe (2000) to the estimation of the spectrum of complex matrices with application to the scalable stability analysis.

References

Arcak M (2007). Passivity as a design tool for group coordination. IEEE Transactions on Automatic Control, 52, 1380–1390.

Barman JF and Katzenelson J (1974). A generalized Nyquist-type stability criterion for multivariable feedback systems. International Journal of Control, 20(4), 593–622.

Belletrutti JJ and MacFarlane AGJ (1971). Characteristic loci techniques in multivariable control systems. Proc. Inst. Elec. Eng., 118, 1291–1297.

Callier FM and Desoer CA (1972). A graphical test for checking the stability of a linear time-invariant feedback systems. IEEE Transactions on Automatic Control, 17, 773–780.

Chen CT (1984). Linear System Theory and Design. CBS College Publishing, New York.

Davis JH (1972). Encirclement conditions for stability and instability of feedback control systems with delays. International Journal of Control, 15, 793–799.

Davison EJ (1983). Some properties of minimum phase systems and “square-down” systems. IEEE Transactions on Automatic Control, 28, 221–222.

Davison EJ and Wang SH (1978). An algorithm for the calculation of transmission zeros of (C,A,B,D) using high-gain output feedback. IEEE Transactions on Automatic Control, 23, 738–741.

Desoer CA and Vidyasagar M (1975). Feedback Systems: Input-Output Properties, Academic Press, New York.

Desoer CA and Wang Y-T (1980). On the generalized Nyquist Stability Criterion. IEEE Transactions on Automatic Control, 25(2), 187–196.

Feng C-B, Tian Y-P and Xin X (1996). Robust Control System Design (in Chinese). Southeast University Press, Nanjing.

Gu K, Kharitonov VL and Chen J (2003). Stability of Time-delayed Systems. Birkhäuser, Boston.

Horn RA and Johnson CR (1985). Matrix analysis, 1st edn. Cambridge University Press, New York.

Horn RA and Johnson CR (1991). Topics in Matrix analysis, 1st edn. Cambridge University Press, New York.

Laub AJ and Moore BC (1978). Calculation of transmission zeros using QZ techniques. Automatica, 14, 557–566.

Lestas I and Vinnicombe G (2005). Scalable robustness for consensus protocols with heterogeneous dynamics. Proceedings of IFAC world Congress, Prague.

Low SH, Paganini F and Doyle JC (2002). Internet congestion control. IEEE control systems Magazine, 22, 28–43.

MacFarlane AGJ and Karcanias N (1976). Poles and zeros of linear multivariable systems: a survey of algebraic, geometric and complex variable theory. International Journal of Control, Vol.24, 33–74.

MacFarlane AGJ and Postlethwaite I (1977). Generalized Nyquist stability criterion and multivariable root loci. International Journal of Control, 25(1), 81–127.

Nise NS (2000). Control Systems Engineering. 3rd ed., John Wiley & Sons, New York.

Rosenbrock HH (1970). State-space and Multivariable Theory. Nelson, London.

Rosenbrock HH (1973). The zeros of a system. International Journal of Control, 18, 297–299.

Rosenbrock HH (1974). Computer-Aided Control System Design. Academic Press, New York.

Saeks R (1975). On the encirclement condition and its generalization. IEEE Transactions on Circuits and Systems, 22, 780–785.

Skogestad S and Postlethwaite I (1996). Multivariable Feedback Control: Analysis And Design. John Wiley & Sons.

Vidyasagar M (1985). Control System Synthesis: A Factorization Approach. MIT Press.

Vinnicombe G (2000). On the stability of end-to-end congestion control for the Internet. Technical report CUED/F-INFENG/TR., No.398, 2000.

Wen JT and Arcak M (2003). A unifying passivity framework for network flow control. Proceedings of IEEE Infocom, San Francisco, California.

Zhou K, Doyle JC and Glover K (1996). Robust and Optimal Control. Prentice-Hall, Upper Saddle River, NJ.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.220.111.87