Chapter 7

Consensus in Heterogeneous Multi-Agent Systems

Gentlemen can live harmoniously together even though they have heterogeneous characters, but non-gentlemen of the same character live in discord.

Confucius (551–479 BC), Analects of Confucius

Multi-agent systems (MASs) may contain heterogeneous network channels and/or heterogeneous agent dynamics. Even for systems with identical plants in agents, diverse input delays generate heterogeneous agent dynamics, and diverse communication delays result in heterogeneous network channels. Consensus problems in heterogeneous MASs are studied in this chapter. High-order consensus is defined for a class of high-order heterogeneous MASs. A necessary and sufficient condition is given for the existence of high-order consensus solutions to the considered class of systems. The condition shows that for systems with diverse communication delays, high-order consensus does not require that the self-delay of each agent to be equal to the corresponding communication delay. The frequency-domain scalability analysis method developed in Chapter 3 is applied to first-order and second-order MASs with diverse input delays and communication delays.

7.1 Integrator Agent System with Diverse Input and Communication Delays

In this section, we first consider the consensus problem for the integrator agent system with diverse input delays based on undirected graphs. Due to the heterogeneousness caused by the diverse input delays, the consensus problem of such a system can not be converted into an ordinary stability problem by using the technique developed in Chapter 6. However, we will show that it can still be considered as a semi-stability problem if the interconnection topology graph is connected. Therefore, using the frequency-domain analysis theory developed in Chapter 3, we develop various scalable consensus conditions, which uses only local information of each agent. Finally, by considering the consensus problem for digraph-based systems with both diverse communication delays and diverse input delays, we will show that consensus condition is dependent on input delays but independent of communication delays when the digraph contains a globally reachable node.

7.1.1 Consensus in Discrete-Time Systems

Consider a discrete-time multi-agent system with topology graph G = (V, E, A) and integrator agents given by

(7.1) equation

where img and img denote the state and the control input of agent i at time instant k, respectively. The topology graph G = (V, E, A) of the system can be directed or undirected, depending on context. The consensus protocol for the system is given by

(7.2) equation

where Ni denotes the neighbors of agent i, and aij > 0 is the adjacency element of A in the graph G = (V, E, A).

When each agent is subject to an input delay Di, system (7.1) becomes

(7.3) equation

Under diverse communication delays, the consensus protocol becomes

(7.4) equation

where τij represents the communication delay from agent j to agent i.

Under protocol (7.4), multi-agent system (7.3) is said to achieve a consensus asymptotically if

equation

where img is a constant.

The closed-loop system of (7.3) and (7.4) is

(7.5) equation

Let x(k) = [x1(k), img, xn(k)]T, and

equation

Then, equation (7.5) can be rewritten as a time-delayed system in a vector form

(7.6) equation

where img, and nd = n(n + 1). Obviously, img, which is the Laplacian matrix of the topology graph.

The characteristic equation of system (7.6) is given by

(7.7) equation

The equilibrium set of system (7.6) is defined by

equation

When L is singular, Xe is a continuum of equilibrium points. Assume that the interconnection topology of the system is described by a connected undirected graph or a digraph containing a globally reachable node. Then, by Theorem 1.9, the Laplacian matrix L has a simple eigenvalue 0, i.e., det (L) = 0 and rank(L) = n − 1. By the definition of L we also have L1n = 0. So, all the elements in Xe can be represented as c1n where c is any constant. Therefore, system (7.6) achieves a consensus asymptotically, if the solution of the system starting from any given initial states img, asymptotically converges to an element in Xe. According to this analysis the following lemma can be easily proved.

Lemma 7.1 If the solutions of equation (7.7) have modulus less than unity except for a root at z = 1, then system (7.6) with a connected undirected graph or a digraph containing a globally reachable node achieves a consensus asymptotically.

This lemma implies that under the assumption that the graph is connected or the digraph contains a globally reachable node the first-order agent system with diverse input delays and communication delays achieves a consensus asymptotically if the closed-loop system is steady semi-stable with z = 1 as a simple pole.

7.1.2 Consensus under Diverse Input Delays

In this subsection we consider the consensus problem for multi-agent systems with input delays only. In this case, the closed-loop form (7.5) of the system reduces to

(7.8) equation

The following theorem gives a scalable consensus condition for multi-agent systems with diverse input delays

Theorem 7.2 Assume that system (7.8) of n agents is based on an undirected and connected graph G = (V, E, A) with symmetric weights. The system achieves a consensus asymptotically if

(7.9) equation

Proof. Taking the z-transformation of system (7.8) and writing it in vector form, we get

(7.10) equation

Note that L in (7.10) is a positively semi-definite matrix since an undirected graph is considered. The characteristic equation is

(7.11) equation

Define img. Then, we will prove that all the zeros of p(z) have modulus less than unity except for a zero at z = 1.

Let z = 1, then p(1) = det (L). Since G = (V, E, A) is connected, by Theorem 1.9, zero is a simple eigenvalue of L, i.e., det (L) = 0 and rank(L) = n − 1. Thus, p(z) indeed has a zero at z = 1.

To prove that the system is steady semi-stable with z = 1 as a simple pole it suffices to prove that the zeros of img have modulus less than unity. By the general Nyquist stability criterion for discrete-time systems (Theorem 2.22), this is the case if the eigenloci of

equation

do not enclose the point (− 1, j0) for all ω img [− π, π]. To show this we rewrite F(jω) as

equation

where img. Since

equation

by Lemma 2.34 we have

equation

Since the spectral radius of any matrix is bounded by its maximum absolute row sum according to Corollary 2.31 of Gershgorin's disc lemma, it follows from the condition (7.9) that

equation

Now, from Theorem 3.16 we conclude that eigenloci of F(jω) do not enclose the point (− 1, j0) for all ω img [− π, π], which implies that the zeros of p(z) have modulus less than unity except for a zero at z = 1. Theorem 7.2 is thus proved by Lemma 7.1. img

Remark. By using Corollary 2.32 instead of Corollary 2.31 of Gershgorin's disc lemma in the proof, we know that system (7.8) achieves a consensus asymptotically if there exists img such that

(7.12) equation

Of course, this condition can be less conservative than (7.9) and reduces to it if H = I.

Theorem 7.2 gives a scalable delay-dependent consensus condition. This condition suggests that under large input delays, small interconnection weights and small numbers of neighbors increase the possibility of achieving consensus if the interconnection topology graph is connected.

Now, we apply the result of Theorem 7.2 to study the effect of diverse input delays on some important systems that have been extensively investigated in literature.

First, let us consider Vicsek's model, which describes a group of agents moving in the plane with the same line velocity (Vicsek et al. 1995). When the headings of the agents are close to each other, the local updating rule for the headings may be approximated by the following linearized equation (Jadbabaie, Lin and Morse 2003)

(7.13) equation

with img and where ni denotes the number of the neighbors of agent i. By applying Theorem 7.2 to this model, it is easy to get its consensus condition as follows:

(7.14) equation

Obviously, when n > 1, condition (7.14) holds only if Di = 0. This implies that Vicsek's model is very sensitive to input delays. To overcome this problem we proposed a modified linearized Vicsek's model with input delays as

(7.15) equation

where εi > 0 is an adjustable interconnection gain. From Theorem 7.2, we get the following corollary.

Corollary 7.3 Suppose that the interconnection topology graph of system (7.15) is connected. Then, the system achieves a consensus asymptotically if

(7.16) equation

Remark 1. When Di = 0, from (7.16) it follows that

equation

which always holds if εi ≤ 1. This implies that the linearized Vicsek's model in its original form (εi = 1) can achieve a consensus asymptotically if and only if the interconnection topology graph of the system is connected. When Di ≥ 1, inequality (7.16) holds only if εi < 1. So, the introduction of small εi is necessary for enhancing the robustness of the linearized Vicsek's model against input delays.

Remark 2. Corollary 7.3 clearly shows the relationship of input delays, interconnection weights and number of neighbors: for large input delays one should use small interconnection weights or have small numbers of neighbors when the graph is kept connected.

Similarly, we can also apply Theorem 7.2 to Moreau's model (Moreau 2005) with input delays

(7.17) equation

where img denotes the positive weight corresponding to the edge eij in the weighted graph G. The following result is a direct corollary of Theorem 7.2.

Corollary 7.4 Suppose that the interconnection topology graph of system (7.17) is connected and has symmetric weights. Then, the system achieves a consensus asymptotically if

(7.18) equation

Remark. Obviously, (7.18) always holds if Di = 0. This implies that a symmetric Moreau's model can achieve a consensus asymptotically if and only if the interconnection topology graph of the system is connected. For Di ≥ 1, from (7.18) we get a sufficient consensus condition of Moreau's model as

(7.19) equation

7.1.3 Consensus under Diverse Communication Delays and Input Delays

In this subsection, we consider multi-agent systems with both communication delays and input delays. Actually, if the communication delays τij are symmetric, i.e., they satisfy the requirement (2.26), then all the results given in Section 7.1.2 can be extended to systems with both communication delays and input delays without any difficulty.

In general, however, the diversity of communication delays destroys the symmetry of the system even if the graph is undirected with symmetric weights. This implies that the tools used for Theorem 7.2, which are mainly referred to Lemma 2.34 and Theorem 3.16, are no longer applicable. So, in the following analysis we will use Greshgorin's disk theorem (Lemma 2.30) to estimate matrix eigenvalues, which does not require the symmetry. Note that the interconnection topology studied in this subsection can be a digraph with asymmetric weights.

Let us first introduce the following lemma as a preliminary result.

Lemma 7.5 The following inequality

(7.20) equation

holds for all non-negative integers D and all ω img [− π, π].

Proof. First of all, we claim that

(7.21) equation

holds for any non-negative integer D. Indeed, by denoting img, we have x img (0, 1] for any non-negative integer D. Thus, inequality (7.21) is equivalent to the well-known inequality img, where x img (0, 1].

Now, we note that img We just need to prove (7.20) for all ω img (0, π] because the left-hand side of (7.20) is an even function for ω img [− π, π].

When img, let img. Calculating the derivative of h(ω) with respect to ω yields img Obviously, we have img, i.e., h(ω) is not increasing for all img. Since h(0) = 0, we have h(ω) ≤ 0, i.e., img for all img. Since img, we get img.

When img, we have img for all non-negative integers D. So, from (7.21), we get

equation

for all img and all non-negative integers D. The lemma is proved. img

Theorem 7.6 Consider multi-agent system (7.3) with protocol (7.4). Assume that the interconnection topology digraph G = (V, E, A) of the system has a globally reachable node. Then the system achieves a consensus asymptotically if

(7.22) equation

Proof. The closed-loop system of (7.3) with (7.4) is given by (7.5). Taking the z-transformation of the system (7.5), we get

(7.23) equation

where Xi(z) is the z-transformation of xi(k). Define an n × n matrix img as follows:

equation

Obviously, img, which is the Laplacian matrix. Then, (7.23) can be written as img where X(z) = [X1(z), img, Xn(z)]T. Define

equation

Then, we will prove that all the zeros of p(z) have modulus less than unity except for a zero at z = 1 in the following.

Let z = 1, img. Since G = (V, E, A) has a globally reachable node, by Theorem 1.9, zero is a simple eigenvalue of L, i.e., det (L) = 0 and rank(L) = n − 1. Thus, p(z) indeed has a simple zero at z = 1.

Now, we prove that the zeros of img have modulus less than unity. Based on the general Nyquist stability criterion (Corollary 2.20), the zeros of f(z) have modulus less than unity, if the eigenloci of img, i.e., img, do not enclose the point (− 1, j0) for ω img [− π, π]. By Greshgorin's disk lemma (Lemma 2.30), we have img for all ω img [− π, π], where

equation

Further, we can show that

equation

where img. Now, define

(7.24) equation

The Nyquist plot of Gi(ω) for ω img [− π, π] is illustrated by Figure 7.1. Note that Gi(ω) is just the center of the disc img. So, img does not enclose the point (− 1, j0) for all ω img [− π, π] as long as the point (− a, j0) with a ≥ 1 is not in the disc img for all ω img [− π, π], i.e., img holds for all ω img [− π, π] when a ≥ 1.

Figure 7.1 Nyquist plot of Gi(ω). (Reproduced with permission from Tian Y.-P. and Liu C.-L., “Consensus of multi-agent systems with diverse input and communication delays,” IEEE Transactions on Automatic Control, 53, 9, 2122–2128, 2008. © 2008 IEEE.)

img

From (7.24), we have

equation

Because img holds for ω img [− π, π] by Lemma 7.5, it follows from (7.22) that

equation

Thus,

equation

i.e.,

equation

holds for all ω img [− π, π] when a ≥ 1.

Now, we have proved that the zeros of p(z) have modulus less than unity except for a zero at z = 1. Therefore, Theorem 7.6 is proved by Lemma 7.1. img

Remark 1. Noticing the inequality (7.21), we know that the condition (7.22) is more conservative than the condition (7.9) given by Theorem 7.2, which is even necessary and sufficient for the case of a single-link, two-node network with equal delay. However, it is still scalable, and moreover, it is applicable to the systems based on digraphs with asymmetric weights.

Remark 2. When there are no input delays, i.e., Di = 0, the consensus condition (7.22) reduces to

(7.25) equation

which implies that the system can achieve a consensus asymptotically if and only the interconnection topology digraph of the system has a globally reachable node and (7.25) holds regardless of the existence of diverse communication delays.

Now, let us apply Theorem 7.6 to study the linearized Vicsek model and Moreau's model.

With communication delays and input delays, the linearized Vicsek model given in Jadbabaie, Lin and Morse (2003) becomes

(7.26) equation

where ni denotes the number of the neighbors of agent i.

From Theorem 7.6 we get the following corollary for Vicsek's model (7.26).

Corollary 7.7 Assume that the interconnection topology of system (7.26) has a globally reachable node. System (7.26) achieves a consensus asymptotically if

(7.27) equation

Remark. On the one hand, when Di = 0, inequality (7.27) holds automatically. This implies that the convergence of the consensus protocol given by Vicsek's model is independent of communication delays provided the graph has a globally reachable node. This coincides with the result given in Wang and Slotine (2006). Cao et al. (2006) extended this result to the case when the graph is jointly rooted. On the other hand, inequality (7.27) holds only for Di = 0. This may suggest that Vicsek's model is very sensitive to input delays. To enhance its robustness against input delays, one should introduce small weights as shown in model (7.15).

Similarly, Moreau's model given in (Moreau 2005) with communication delays and input delays can be written as

(7.28) equation

The following result is a direct corollary of Theorem 7.6.

Corollary 7.8 If the interconnection topology digraph of system (7.28) has a globally reachable node, then the system achieves a consensus asymptotically if

(7.29) equation

Remark. Obviously, (7.29) can be rewritten as img. So, it always holds if Di = 0. But for Di > 0 it holds only for some appropriately designed weights img.

7.1.4 Continuous-Time System

Consider a continuous-time multi-agent system with diverse input delays and communication delays

(7.30) equation

It is easy to show that under the assumption that the topology graph is connected or the topology digraph contains a globally reachable node, the system achieves a consensus asymptotically if it is steady semi-stable with s = 0 as a simple pole.

Using the theory developed in Chapter 3 (Theorem 3.10), in a similar way to that shown in the proof of Theorem 7.2 one can get the consensus condition for continuous-time systems with input delays.

Theorem 7.9 Suppose that the topology graph of system (7.30) is connected with symmetric weights. Then, the system achieves a consensus asymptotically if

(7.31) equation

Remark. It is easy to see that in the case when all the delays Ti are the same for all img, the condition (7.31) reduces to the result given by Olfati-Saber and Murray (2004).

The continuous-time system with diverse input and communication delays can be written as

(7.32) equation

Through a similar procedure to that used in the proof of Theorem 7.6, one can get a consensus condition for continuous-time systems with diverse communication and input delays.

Theorem 7.10 If the interconnection topology digraph of system (7.32) has a globally reachable node, then the system achieves a consensus asymptotically if

(7.33) equation

Remark. When there is no input delays, i.e., Ti = 0, the consensus condition (7.33) always holds, which implies that the system can achieve a consensus asymptotically if and only the interconnection topology digraph of the system has a globally reachable node regardless of the existence of diverse communication delays. This conclusion coincides with existing results in references such as Blondel et al. (2005) and Wang and Slotine (2006).

7.1.5 Simulation Study

Example 7.1 Symmetric system.

Consider a system of eighty agents described by the modified linearized Vicsek model (7.15). The interconnection topology for the agents is a closed ring (Figure 7.2). Note that under non-zero input delays, the linearized Vicsek's model with unity weights (εi = 1) has no consensus. In simulations, we choose the coupling weights as

equation

By Corollary 7.3, the admissible values of the input delays are

equation

Simulations validate that under the admissible input delays the system achieves consensus asymptotically. We depict the boundary values of the input delay versus the coupling weights in Figure 7.3.

Figure 7.2 Undirected graph: a closed ring. (Reproduced with permission from Tian Y.-P. and Liu C.-L., “Consensus of multi-agent systems with diverse input and communication delays,” IEEE Transactions on Automatic Control, 53, 9, 2122–2128, 2008. © 2008 IEEE.)

img

Figure 7.3 Bounds of input delays. (Reproduced with permission from Tian Y.-P. and Liu C.-L., “Consensus of multi-agent systems with diverse input and communication delays,” IEEE Transactions on Automatic Control, 53, 9, 2122–2128, 2008. © 2008 IEEE.)

img

Note that the bound of input delay determined by Corollary 7.3 may not apply to asymmetric systems. For example, with each agent's input delay at boundary, we introduce an identical communication delay, T ≥ 1, which destroys the conjugate symmetry of the system matrix in the frequency domain, between each pair of neighboring agents. Then, the simulation shows that the system has no asymptotic consensus.

Now, we can use Theorem 7.6 to get a consensus condition for the system with both input delays and communication delays. By Theorem 7.6, the admissible values of the input delays are

equation

The boundary values of input delays estimated by Theorem 7.6 are also shown in Figure 7.3. The boundary determined by Theorem 7.6 is lower than the boundary given by Corollary 7.3, which implies that Theorem 7.6 is more conservative than Corollary 7.3 for symmetric systems. However, with input delays in the area determined by Theorem 7.6, the asymptotic consensus of the system is robust to arbitrary communication delays.

Example 7.2 Asymmetric system.

Consider the multi-agent system (7.5) with an interconnection digraph shown by Figure 7.4. The weights of the directed paths are: a12 = 0.1, a16 = 0.05, a23 = 0.15, a36 = 0.1, a43 = 0.05, a45 = 0.1, a56 = 0.15, a62 = 0.15, and the corresponding communication delays are: τ12 = 5, τ16 = 3, τ23 = 4, τ36 = 4, τ43 = 4, τ45 = 6, τ56 = 6, τ62 = 5. Simulation shows that the system is quite sensitive to input delays, and it cannot converge to any consensus when Ti > 3. This is an asymmetric system to which Theorem 7.2 does not apply. Using Theorem 7.6, we get that Ti ≤ 2 is a sufficient consensus condition which is independent of communication delays. We choose Ti = 2, i = 1, 2, 3, 4, 5, 6, in the simulation. The multi-agent system converges to a consensus as shown by Figure 7.5.

Figure 7.4 Digraph of a group of six agents.

img

Figure 7.5 Consensus with communication and input delays. (Reproduced with permission from Tian Y.-P. and Liu C.-L., “Consensus of multi-agent systems with diverse input and communication delays,” IEEE Transactions on Automatic Control, 53, 9, 2122–2128, 2008. © 2008 IEEE.)

img

7.2 Double Integrator System with Diverse Input Delays and Interconnection Uncertainties

7.2.1 Leader-Following Consensus Algorithm

Consider a multi-agent system with interconnection topology graph G = (V, E, A). There are n agents with diverse input delays

(7.34) equation

where img, img, img, and Ti > 0 are the position, velocity, acceleration and input delay, respectively, of agent i.

For system (7.34), the leader-following coordination control strategy is adopted in this section. Let the dynamics of the leader be determined by

(7.35) equation

where img is the position of the leader, and img is a constant which represents the desired velocity for all the agents.

Then, the consensus protocol for the first-order multi-agent system (Olfati-Saber and Murray (2004) can be easily extended to the leader-following system as follows

(7.36) equation

where ki > 0 and γ > 0, Ni denotes the neighbors of agent i, aij > 0 is the adjacency element of A in the digraph G = (V, E, A), and bi is the linking weight from agent i to the leader (7.35). Note that bi > 0 if there is a directed edge from agent i to the leader; otherwise, bi = 0. Let img for notation convenience.

Remark. Protocol (7.36) can be used only for following a leader with a constant velocity, or a leader with a velocity which is time-varying but asymptotically approaching to a constant. If the leader's velocity is not converging to any constant, then each agent should estimate its neighbors' accelerations, and the stability analysis of that kind of consensus protocol will be much more complicated.

With consensus protocol (7.36), the closed-loop form of system (7.34) is given by

(7.37) equation

The following lemma gives some structural property of the leader-following system.

Lemma 7.13 Assume that the interconnection topology graph of n agents together with the leader in system (7.37) has the leader as a globally reachable node. Then, the matrix L + B has no zero eigenvalues, where L is the Laplacian matrix of the interconnection topology of n agents without the leader.

Proof. Consider the interconnection topology graph with n + 1 nodes corresponding to the n agents of system (7.37) and the leader. Obviously the Laplacian matrix of this topology is given by

equation

where img. Since the leader is a globally reachable node in the graph, we have img, by Theorem 1.9. Taking elementary column transforms for img by adding all the other columns to the first column as follows

equation

we get rank(L + B) = n. The lemma is proved. img

7.2.2 Consensus Condition under Symmetric Coupling Weights

Let

equation

Then, from (7.37) it follows that

(7.38) equation

Taking the Laplace transform of (7.38), one gets

(7.39) equation

Denote

(7.40) equation

Then, using the framework of Lee and Spong (2006) one can get a sufficient condition of consensus as

(7.41) equation

However, it can be shown that this condition is so conservative that it gives an empty set of available control parameters for the second-order multi-agent system with input delays. Let us rewrite (7.40) as

equation

where img, and

(7.42) equation

Then, the condition (7.41) is equivalent to

(7.43) equation

Actually, such a condition never holds for any κi > 0 when γ > 0, Ti > 0.

Let us derive a less conservative consensus condition for system (7.37) with symmetric coupling weights.

For all img, we denote

equation

equation

As shown in Chapter 3 (Proposition 3.26), ω0(i) is the critical point of the frequency response of Wi(s) from clockwise part to anti-clockwise part.

Let img be the agent which has the maximal input delay constant img, i.e.,

(7.44) equation

Now, we are in a position to present some sufficient consensus conditions for the second-order multi-agent system with input delays.

Theorem 7.14 Assume that system (7.37) is composed of n agents and a leader with a static interconnection topology that has the leader as a globally reachable node, and the topology graph has symmetric weights, i.e., aij = aji. For each agent the following preconditions are assumed:

(7.45) equation

(7.46) equation

Then, all the agents in the system asymptotically converge to the leader's state, if

(7.47) equation

where img is the gain margin of the transfer function Wi(s) defined in (7.42).

Proof. Writing (7.39) in the vector form, we can get the characteristic equation of system (7.38) as

(7.48) equation

where L is the Laplacian matrix corresponding to the interconnection topology for all the agents without the leader.

Define img. To prove Theorem 7.14 it suffices to prove that all the zeros of F(s) are in the open left half of the complex plane.

Let s = 0. Then F(0) = det (γdiag{ki}(L + B)). Because the interconnection topology composed of the n agents together with the leader has the leader as a globally reachable node, F(0) ≠ 0 by Lemma 7.13.

Now, define img. We will prove that all the zeros of p(s) are inside the LHP. Based on the general Nyquist stability criterion (Corollary 2.20), all the zeros of p(s) lie inside the LHP, if the eigenloci

equation

do not enclose the point (− 1, j0) for all img.

For the symmetric weights (aij = aji), L + B = (L + B)T. Hence, based on Lemma 2.34, we have

equation

Since the spectral radius of any matrix is bounded by its largest absolute row sum, it follows from the condition (7.47) that

equation

Therefore, from Theorem 3.32 it follows that

equation

i.e., the eigenloci of img do not enclose the point (− 1, j0) for all img, which implies that the zeros of F(s) are all inside the LHP. Theorem 7.14 is thus proved. img

Remark. The result of Theorem 7.14 can be extended to systems with both communication delays and input delays without any difficulty if the diverse communication delays τij are symmetric, i.e., they satisfy the requirement (2.26).

7.2.3 Robust Consensus under Asymmetric Perturbations

The consensus condition given by Theorem 7.14 depends on the strict symmetry of the Laplacian matrix L. In practice, however, perturbations of coupling weights may occur and destroy the symmetry. In the following, we study the robustness of the consensus protocol against asymmetric perturbations.

Suppose that the symmetric coupling weights of system (7.37) are subject to some asymmetric perturbations, denoted by img for each one. Then the system becomes

(7.49) equation

where aij = aji, and aij + δij > 0 hold for j img Ni.

A robust consensus condition of the perturbed system is given by the following theorem.

Theorem 7.15 Assume that the nominal part of system (7.49), i.e., the system without asymmetric weight perturbations δij, converges to the leader's states asymptotically. Let

(7.50) equation

where img and

equation

Then, the agents in the perturbed system (7.49) converge to the leader's states asymptotically, if

(7.51) equation

where img denotes the largest singular value of matrix, and Δ = {Δij} is the asymmetric perturbation matrix, which is defined as follows

equation

Proof. Under the same variable transformation as used in the previous subsection

equation

it is easy to get the characteristic equation of system (7.49) as

(7.52) equation

Since the system (7.49) without asymmetric weight perturbations δij converges to the leader's states asymptotically, the roots of the characteristic equation (7.48) all lie inside the LHP, i.e., the zeros of det (s2I + s2KD(s)(L + B)) lie inside the LHP, and det (L + B) ≠ 0.

In the following, we will prove that the roots of equation (7.52) are all inside the LHP.

First we show that equation (7.52) has no roots at s = 0. Indeed, by setting ω = 0 we get from (7.51) that img. This implies that

equation

So, it follows that

equation

or equivalently,

equation

This proves that equation (7.52) has no roots at s = 0. Therefore, the characteristic equation (7.52) can be equivalently rewritten as

(7.53) equation

The feedback diagram corresponding to the characteristic equation (7.53) is demonstrated in Figure 7.6.Using the linear fractional transformation, the diagram in Figure 7.6 can be equivalently transformed into the form shown by Figure 7.7, where M(s) is given by equation (7.50).

Figure 7.6 System with asymmetric perturbation.

img

Figure 7.7 Transformed system.

img

The characteristic equation of the closed-loop system in Figure 7.7 is

(7.54) equation

Obviously, D(s) has no poles in the open RHP. Thus, ΔM(s) has no poles in the open RHP. According to the general Nyquist stability criterion (Corollary 2.20), the roots of the characteristic equation (7.54) all lie inside the LHP, as long as the eigenloci of ΔM(s), i.e., λM(jω)), do not enclose the point (− 1, j0) for img.

From condition (7.51) it follows that

(7.55) equation

Hence, λM(jω)) does not enclose the point (− 1, j0) for all img, i.e., the roots of the characteristic equation (7.54) all lie inside the LHP. Therefore, the closed-loop system in Figure 7.7 is asymptotically stable, and the agents in (7.49) converge to the leader's states asymptotically. Theorem 7.15 is proved. img

7.2.4 Simulation Study

Example 7.16 Design procedure based on Theorem 7.14

Consider a system (7.37) of five agents and one leader described by (7.35). The interconnection topology is described in Figure 7.8. Obviously, the leader is globally reachable. Assume that the input delays for the agents are: T1 = 0.5(s), T2 = 1.0(s), T3 = 0.7(s), T4 = 0.6(s) and T5 = 0.8(s). The weights of the edges are: a12 = a21 = 0.30, a25 = a52 = 0.70, a13 = a31 = 0.10, a34 = a43 = 1.10, a42 = a24 = 0.50, b5 = 1.50.

In the following, we design parameters γ and ki in the consensus protocol (7.36) so that the agents converge to the leader's state asymptotically.

Figure 7.8 Network of five agents and a leader. (Reprinted from Automatica, 45, 5, Tian Y.-P. and Liu C.-L., “Robust consensus of multi-agent systems with diverse input delays and asymmetric interconnection perturbations,” 1347–1353, 2009, with permission from Elsevier.)

img

Step 1: Choosing γ.

The condition (7.45) requires γ img (0, 0.4495/Ti), img, which implies that

(7.56) equation

Since T2 is the maximal input delay, we have img. Denote img, and the condition (7.46) can be represented as Ei > 0, i = 1, 2, 3, 4, 5. With the given input delays, Ei is a function of γ. The curves of Ei on γ are shown in Figure 7.9.

Figure 7.9 Choosing parameter γ. (Reprinted from Automatica, 45, 5, Tian Y.-P. and Liu C.-L., “Robust consensus of multi-agent systems with diverse input delays and asymmetric interconnection perturbations,” 1347–1353, 2009, with permission from Elsevier.)

img

From Figure 7.9 it is clear that the condition Ei > 0, i = 1, 2, 3, 4, 5, holds if

(7.57) equation

According to (7.56) and (7.57), we can choose γ = 0.10 to guarantee the conditions (7.45) and (7.46).

Step 2: Choosing ki.

For the transfer functions

equation

using the MATLAB® simulator we obtain the inverses of their gain margins as img, img, img, img, img. From the condition (7.47), the constraints on ki can be calculated as k1 img (0, 3.788), k2 img (0, 0.498), k3 img (0, 0.906), k4 img (0, 0.801), k5 img (0, 0.651), We choose k1 = 3.4, k2 = k3 = k4 = k5 = 0.4 for the simulation.

With the parameters chosen above and the initial states generated randomly, the agents in the system (7.37) asymptotically converge to the leader's state as shown in Figure 7.10.

Figure 7.10 Positions and velocities of the agents under symmetric weights. (Reprinted from Automatica, 45, 5, Tian Y.-P. and Liu C.-L., “Robust consensus of multi-agent systems with diverse input delays and asymmetric interconnection perturbations,” 1347–1353, 2009, with permission from Elsevier.)

img

Since there are no other theoretic results to compare with, we test the conservatism of our results by simulation. The procedure is as follows. Setting k2, k3, k4 and k5 at our theoretic boundary values as 0.498, 0.906, 0.801 and 0.651 respectively, we increase k1 from our boundary value 3.788 until the system has no consensus. Then we find the computational margin for k1 is k1m = 7.519. Using similar procedures we can obtain the other marginal gains as k2m = 0.797, k3m = 1.088, k4m = 0.954, k5m = 0.773. Note that unlike our theoretical results, these computational margins cannot be used in the consensus protocol simultaneously.

Example 7.17 System with interconnection uncertainties.

Consider the multi-agent systems (7.49) of five agents and one leader described by (7.35) with the same interconnection topology as Example 7.16 (see Figure 7.8). For simplicity, we choose the same aij, bi, input delays Ti and control parameters γ and img as given in Example 7.16. From Theorem 7.14, the system (7.49) without asymmetric weight perturbations converges to the leader's states asymptotically, and the zeros of det (s2I + s2KD(s)(L + B)) lie inside the LHP. Using the MATLAB® simulator, we obtain that the largest value of img on ω img (− ∞, ∞) is img. From Theorem 7.15, if the largest singular value of the asymmetric disturbance matrix Δ, i.e., img, satisfies img, the closed system in Figure 7.6 with Δ is asymptotically stable. For example, when

equation

one can check that img, and aij + δij > 0 for j img Ni. Therefore, with the Laplacian matrix L + Δ and the initial states generated randomly, the agents in (7.49) converge to the leader's states asymptotically as shown in Figure 7.11.

Figure 7.11 Positions and velocities of the agents under asymmetric weights. (Reprinted from Automatica, 45, 5, Tian Y.-P. and Liu C.-L., “Robust consensus of multi-agent systems with diverse input delays and asymmetric interconnection perturbations,” 1347–1353, 2009, with permission from Elsevier.)

img

7.3 High-Order Consensus in High-Order Systems

7.3.1 System Model

Suppose the interconnection topology of the system is described by a digraph G = (V, E, A) with |V| = n. We assume that the interconnection topology of the system is a connected undirected graph or a digraph containing a globally reachable node. Then, by Theorem 1.9, the Laplacian matrix L has a simple eigenvalue 0, i.e., det (L) = 0 and rank(L) = n − 1. Moreover, the definition of L implies that L · 1n = 0.

Let the model of the ith agent (img) be given by the following transfer function

(7.58) equation

where img and img denote the Laplace transformation of the output and input, respectively, of the ith agent; Ti is the input delay; ν and mi are positive integers satisfying miν; img are system parameters. For any non-negative integer k, denotes by img the kth-order derivative of the output yi(t) of the ith agent.

Definition 7.18 Multi-agent system (7.58) is said to reach the rth-order consensus asymptotically if

(7.59) equation

(7.60) equation

for system solutions from any admissible initial conditions, and

(7.61) equation

for system solutions from some admissible initial conditions.

Note that when r = 0, the above definition implies that

equation

where img is a constant. In this case we say the system achieves a constant consensus. So, constant consensus can be regarded as the zeroth-order consensus by Definition 7.18. We also note by this definition the constant c can not identically be zero for all initial conditions, i.e., it excludes the case of trivial consensus which actually implies that each agent is asymptotically stabilized.

Let τij be the communication delay from agent j to agent i. Then, at time instance t the information obtained by agent i from agent j is img instead of img. Let the consensus protocol be

(7.62) equation

where κi > 0, img are some control parameters, and τ'ij denotes the estimation of communication delay τij used by agent i. Sometimes τ'ij, j img Ni are also referred to as self-delays of agent i. We will show that the high-order consensus can be achieved even though the self-delays are not equal to the communication delays.

7.3.2 Consensus Condition

It is easy to get the closed-loop form of system (7.58) with protocol (7.62) as

(7.63) equation

Taking the Laplace transform under zero initial condition for the above equation yields

(7.64) equation

Let

(7.65) equation

(7.66) equation

(7.67) equation

(7.68)–(7.69) equation

and

(7.70) equation

(7.71) equation

(7.72) equation

Then, the closed-loop system can be shown by Figure 7.12.The frequency-domain model of the closed-loop system is given by

(7.73) equation

and the return difference equation of the system is given by

equation

According to (2.42), direct use of the equation

(7.74) equation

may ignore cancelation between the poles of the closed-loop system and the poles of the open-loop system at s = 0. To avoid these cancelations we consider the following equation

(7.75) equation

as the characteristic equation of the system. Obviously,

(7.76) equation

Therefore, all the non-zero solutions of (7.75) are contained in the solutions of (7.76), and vice versa.

Figure 7.12 Interconnected system.

img

The following proposition gives an obvious condition for justifying if the open-loop poles at s = 0 enter into the set of the closed-loop poles.

Proposition 7.19 Suppose H(s)KL(s) is analytic at s = 0. If rank[H(s)KL(s)] = n − 1, for all img at which H(s)KL(s) is analytic, then

equation

where g(s) has neither poles nor zeros at s = 0.

Proof. The proposition can be easily proved by converting H(s)KL(s) into a diagonal matrix through a similar transformation. img

Proposition 7.19 shows that s = 0 is a repeated pole with multiplicity r + 1 of the closed-loop system. This also implies that the solution of the closed-loop system can be expressed as

(7.77) equation

where img are constant vectors, img are vectors whose elements are polynomials of t, λk are zeros of g(s), and M is a finite or infinite positive integer. Furthermore, if Reλi < 0, we have img as t→ ∞. We call img the steady state of the system.

Lemma 7.20 Assume that det (sr+1I + H(s)KL(s)) = sr+1g(s), where g(s) has neither poles nor zeros at s = 0. Then, for any s ≠ 0, g(s) = 0 if and only if

(7.78) equation

Proof. The lemma is obvious due to the equation (7.76).img

By Definition 2.7, a linear time-invariant system is said to be steady semi-stable if all its characteristic roots are inside the LHP or at the origin of the complex plane. By Lemma 7.20, to verify the steady semi-stability of the system, we just need to check if all the zeros of (7.78) have negative real parts. Note that the generalized Nyquist stability criterion (Theorem 2.19 and Theorem 2.20) can be used for this purpose.

Denote img for convenience, and denote by img the kth-order derivative of img.

Theorem 7.21 Assume that the multi-agent system (7.58) with consensus protocol (7.62) is steady semi-stable, and the transfer function img is analytic in a neighborhood of s = 0. Then, the rth-order consensus is reached at the steady state if img, for all img, and img for all img at which img is analytic.

Proof. By Proposition 7.19 we know that the system has zeros at s = 0 with multiplicity r + 1, and the system solution can be expressed by (7.77). With the assumption of the steady semi-stability, from (7.77) it is easy to see that

(7.79) equation

and

(7.80) equation

when t→ ∞. So, conditions (7.60) and (7.61) in Definition 7.18 are already satisfied. We just need to check (7.59).

Let s be in the neighborhood D of s = 0 in which img is analytic. By the Taylor formula for functions of a complex variable we have

(7.81) equation

where

(7.82) equation

(7.83) equation

Substituting (7.81) into (7.73) yields

equation

Note that the above equation holds in the neighborhood D of s = 0. When the system is steady semi-stable, Y(s) is analytic for Res > 0. Then, by taking the inverse Laplace transformation of the equation, we have

(7.84) equation

for t→ ∞. Applying (7.79) to the above equation we get

(7.85) equation

when t→ ∞.

Differentiating (7.80)r times yields

(7.86) equation

Substituting (7.80) and (7.86) into (7.85) yields

(7.87) equation

where

(7.88) equation

Equation (7.87) holds if and only if img.

From Er = 0 we get Cr img span(1n) because img and img. Using the result Cr img span(1n) and the assumption img, from Er−1 = 0 it follows that img which implies Cr−1 img span(1n). Conducting this procedure to the end, i.e., E0 = 0, we get img. Thus, we have y(t) img span(1n) when t→ ∞, which, by (7.80), also implies that img. img

7.3.3 Existence of High-Order Consensus Solutions

Theorem 7.21 gives some sufficient consensus conditions for the system. Here we try to find a necessary and sufficient condition of the existence of high-order consensus solutions. First, we show that the condition given by Proposition 7.19 can be further weakened as follows.

Proposition 7.22 Suppose img is analytic in a neighborhood of s = 0. Then,

equation

with g(s) having neither poles nor zeros at s = 0, if and only if img, and there exists a non-zero constant vector α img span(1n) such that img, img, and img.

Proof. Let s be in the neighborhood of s = 0, denoted by D, in which img is analytic. By the Taylor formula for functions of a complex variable, we have

(7.89) equation

where

(7.90) equation

(7.91) equation

It is clear that

equation

if and only if there exists a normalized non-singular matrix img (i.e., det (T) = 1) such that one column (say without loss of generality, the first one) of T−1L1(s)T is zero, i.e., [T−1L1(s)T](:, 1) = 0. Denote by α the first column of T. It is easy to get

equation

Obviously, [T−1L1(s)T](:, 1) = 0, ∀ s img D if and only if img, img.

Denoting by pij the (i, j)th elements of T−1L1(s)T for img, we have

(7.92) equation

And denoting by gij the (i, j)th elements of T−1L2(s)T, we have

(7.93) equation

Since H(s)KL(s) is analytic at s = 0, all the elements pij (img) are analytic at s = 0. Hence, g(s) is also analytic at s = 0.

Now, we prove g(s) has no zeros at s = 0, i.e., g(0) ≠ 0 if and only if img and img. Denote img. From (7.93) we have

equation

because det (T−1) = 1. Obviously, g(0) ≠ 0 if and only if

equation

and

equation

From (7.92) we know that

equation

and hence img if and only if img. Simple calculation shows that img. So g(0) ≠ 0 if and only if img, and img. Finally, we note that the non-zero vector α img span(1n). This is indeed the case. Since we have proved img, the null space of img is one-dimensional. From the property of Laplacian matrix, L(0) · 1n = 0, we get img. So, α img span(1n). The proposition is proved. img

Remark. When img, a sufficient condition for img is img, where c is a non-zero constant. And a more sufficient condition for is img.

Exercises 7.23 Suppose img and img. Show α : = c1n with c ≠ 0 is not in img.

When s = 0 is a repeated pole with multiplicity r + 1 of the closed-loop system, the solution of the closed-loop system can be expressed by equation (7.77)

Now, we are ready to present the following theorem.

Theorem 7.24 Assume that the multi-agent system (7.58) with consensus protocol (7.62) is steady semi-stable and the transfer function img is analytic in a neighborhood of s = 0. Then, the rth-order consensus is reached at the steady state if and only if img, img, img, and img.

Proof. (Sufficiency) Suppose img, img, img, and img. Then, by Proposition 7.22 we know that the system has zeros at s = 0 with multiplicity r + 1, and the system solution can be expressed by (7.77). The rest of the sufficiency part of the this proof is as the same as given in the proof of Theorem 7.21.

(Necessity) Suppose that the system reaches the consensus defined by (7.59), (7.60) and (7.61). Note that (7.60), (7.61) and the steady semi-stability assumption imply that the system has zeros at s = 0 of multiplicity r + 1. So, by Proposition 7.22, the necessity is obvious. img

7.3.4 Constant Consensus

For the case of constant consensus, i.e., the case when r = 0, without loss of generality, we let bi0 = 1 for all img. Then, the protocol (7.62) reduces to the following form

(7.94) equation

the closed-loop system equation becomes

equation

and H(s) defined in (7.70) takes the form

equation

where

(7.95) equation

Let us apply Theorem 7.24 to the case of constant consensus. For this case, Theo-rem 7.24 requires: (1) the marginal stability of the closed-loop system, (2) img and img. Since κi ≠ 0, requirement (2) implies that rank[L(0)] = n − 1 and L(0) · 1 = 0. It is well known that this is equivalent to the connectivity condition for the interconnection graph, i.e., the digraph has a globally reachable node.

To check the marginal stability of the closed-loop system, we may make some loop transformation on the system as shown by Figure 2.8. Denote by img the open-loop transfer function matrix after the loop transformation. Then, based on the extended spectral radius theorem for steady semi-stability (Theorem 2.24), we get the following result immediately.

Theorem 7.26 Consider the multi-agent system (7.58) with the protocol (7.94). Assume the interconnection digraph has a globally reachable node, then, the system achieves a constant consensus, if

(7.96) equation

(7.97) equation

Since L(s) defined in (7.72) can be rewritten as

equation

where

equation

and

equation

the system diagram shown in Figure 7.12 can be equivalently transformed as shown in Figure 7.13. In this case we have

(7.98) equation

where

(7.99) equation

Straightforward calculation shows that

equation

So, the requirement (7.97) of Theorem 7.25 is satisfied. Therefore, under the assumption that the interconnection digraph has a globally reachable node, a sufficient condition for achieving a constant consensus is

(7.100) equation

By using Corollary 2.31 of Gershgorin's disc lemma, a more conservative but scalable condition is

(7.101) equation

This condition was first given by Lee and Spong (2006).

Figure 7.13 Equivalent diagram for constant consensus.

img

Similarly, from Theorem 2.27 and Theorem 2.29 we can also get the following two theorems for the constant consensus problem.

Theorem 7.26 Consider the multi-agent system (7.58) with the protocol (7.94). Assume the interconnection digraph has a globally reachable node, then, the system achieves a constant consensus, if

(7.102) equation

(7.103) equation

Theorem 7.27 Consider the multi-agent system (7.58) with the protocol (7.94). Assume the interconnection digraph has a globally reachable node, then, the system achieves a constant consensus, if

(7.104) equation

(7.105) equation

7.3.5 Consensus in Ideal Networks

Before applying Theorem 7.24 to ideal networks, i.e., networks with zero communication delays and constant channel dynamics, let us review the following fact from graph theory.

Let G be a digraph having at least one globally reachable node. If we choose only one globally reachable node, say img, of G, and cut off all the edges from img, then obviously img is still a globally reachable node of G. But, if we do this operation for a node which is not globally reachable in G, or simultaneously do this operation for two or more globally reachable nodes, then G has no globally reachable node anymore. This is equivalent to saying, if img is a Laplacian of a digraph having at least one globally reachable node, then rank[diag{b1, . . . , bn}L] = n − 1 if and only if img, or img and img is a globally reachable node of G, where img denotes the set of integers from 1 to n excluding j.

Now, let us consider a multi-agent system based on an ideal network. Suppose the topology digraph G contains at least one globally reachable node. Then, we have rank[L] = n − 1 and L · 1n = 0. In this case we have τij = τ'ij = 0 and L(s) = L. From L · 1n = 0 it is easy to get img, img and img. Finally, by denoting img, we know that img if and only if gi(0) ≠ 0, or img and img is a globally reachable node of digraph G. Thus, we get the following result as a corollary of Theorem 7.24.

Theorem 7.28 In an ideal network with zero communication delays and constant channel dynamics, the rth-order consensus will be achieved for a steady semi-stable system at its steady state if and only if the topology graph contains at least one globally reachable node, and the agents's dynamics and the protocol satisfy gi(0) ≠ 0, or img and img is a globally reachable node of digraph G.

7.4 Integrator-Chain Systems with Diverse Communication Delays

7.4.1 Matching Condition for Self-Delay

Consider the multi-agent system, whose agents are chains of integrators, i.e., mi = ν, and hi(s) = 1. In this case gi(s) = bi(s), and thus we have

equation

Since κi > 0 and bi0 ≠ 0, we have img when the topology graph contains a globally reachable node. Now, we check the condition img, img. For the second-order consensus (ν = 2), this condition implies that

(7.106) equation

And generally, the consensus condition for the νth-order consensus can be obtained as

(7.107) equation

Equation (7.106) or (7.107) uncovers a very interesting fact that the second-order or high-order consensus does not necessarily require τ'ij = τij. We call condition (7.107) the matching condition for self-delays.

7.4.2 Adaptive Adjustment of Self-Delay

Since (7.106) does not require each self-delay to be equal to the communication delay, for the second-order consensus each agent can set an identical self-delay, i.e., τ'ij = τ'i, ∀ j img Ni, and adjust τ'i so that (7.106) is satisfied. In the following, we propose a simple algorithm for the adaptive adjustment of the self-delay.

The performance index can be proposed as

(7.108) equation

where t0 is an initial time instant from which the adaptive adjustment begins, T > 0 is a large enough constant of time interval.

The algorithm of adaptive adjustment of τ'i is as follows:

1. Set the tolerance error img, sampling period length Δt, and an initial value τ ' (0). Discretize the performance index (7.108) and write it in the recursive form given as

(7.109) equation

where tm = mΔt.
2. Compute the adjustment to τ'i by

(7.110) equation

where β > 0 is a proper adaptation parameter, and img can be obtained from (7.109) as follows:

(7.111) equation

where

(7.112) equation

(7.113) equation

Set m = m + 1.
3. If Ei > img, go to step 2; otherwise, the algorithm is stopped.

The proposed algorithm is essentially a gradient algorithm which is usually just locally convergent. To get the equilibria of the algorithm, one can set img. Then, it follows

(7.114) equation

Since yi(t) → C0 + C1t, from (7.114) it follows that

(7.115) equation

or

(7.116) equation

Obviously, (7.115) is exactly (7.106) corresponding to the desired equilibrium; (7.116) implies that the velocities of all the agents go to zero, which is the zeroth-order consensus (constant consensus) state for the second-order system.

7.4.3 Simulation Study

In this subsection several simulation experiments are conducted through two numerical examples to verify the obtained theoretical results.

Example 7.29 Consensus of a system of heterogeneous agents.

Consider a system of five agents described by

equation

Obviously, agent 1 is absolutely different from the other four agents which are all double-integrators. The agents are interconnected by a digraph shown in Figure 7.14. The weighted adjacency matrix is given as follows

equation

Figure 7.14 Interconnection graph.

img
Case 1: no communication delays
Firstly, we consider the ideal network without communication delays. Thus, the consensus protocol is simply given by

equation

Let img. Then, for a random set of initial states the simulation results are presented in Figures 7.15 to 7.18. From these figures one can see that both positions and velocities of all the agents of the system reach consensus solutions.

Figure 7.15 Positions of the agents (without delays).

img

Figure 7.16 Velocities of the agents (without delays).

img

Figure 7.17 Position error between agent 1 and agent 2.

img

Figure 7.18 Velocity error between agent 1 and agent 2.

img
Case 2: with diverse communication delays
The agents' dynamics, interconnection topology graph and adjacency matrix are all the same as given in Case 1. Now, we assume that there are diverse communication delays in the system, which are given as follows

equation

Let the self-delays of each agent be equal to the corresponding communication delays, i.e., img. Then, the consensus protocol becomes

equation

Let img. For a random set of initial states simulation results are presented in Figures 7.19 and 7.20, which show the positions (velocities) that all the agents of the system reach at a consensus solution in spite of diverse communication delays.

Figure 7.19 Positions of the agents (with diverse delays).

img

Figure 7.20 Velocities of the agents (with diverse delays).

img

Example 7.30 Consensus with unknown communication delays.

Now, we study the consensus protocols for unknown communication delays. To show the correctness of the theoretical results, we consider the case of identical double-integrator agents, which are extensively studied in the literature. Let

equation

The interconnection topology graph is assumed to be the same as Figure 7.14. The adjacency matrix is given by

equation

The diverse communication delays are set as

equation

Case 1: unequal but matched self-delays
In this case the self-delays of the agents in the protocol are not equal to the corresponding communication delays. But we assume that they satisfy the match condition (7.106). Let

equation

It is easy to see that the self-delays of agent 1 are unequal to the corresponding communication delays. However, the constraint (7.106) is still satisfied because 2 × 0.2 + 1 × 0.5 = 2 × 0.3 + 1 × 0.3. Let the control gains in the protocol be img. The simulation result (Figure 7.21) shows that the positions and velocities of the five agents tend to consensus asymptotically.

Figure 7.21 Positions and velocities of the agents (with unmatched but constrained delays).

img
Case 2: Unmatched self-delays with an adaptive adjustment
Let us change the value of τ'12 and τ'14 from 0.3 to 0.8. Of course, the match condition (7.106) is no longer satisfied. Firstly, we conduct a simulation with this set of self-delays. The simulation result presented in Figure 7.22 shows that both the position error and velocity between each pair of agents diverge when time goes to infinity, i.e., no consensus is achieved. From the figure one can also observe the phenomenon that, due to the consensus protocol, the system states tend to some common value before they eventually diverge. This gives the chance to adjust the delay estimation (self-delay). Now, let us introduce the adaptive adjustment mechanism (7.110) for τ'12 = τ'14 img τ'1. Choosing β = 0.05 and τ'1(0) = 0.8 we conduct the simulation again. As shown by Figures 7.23 and 7.24, the positions and the velocities of the five agents tend to the consensus solution asymptotically in this case.

Figure 7.22 Position errors and velocity errors (with unconstrained delays).

img

Figure 7.23 Positions of the agents (with an adaptive adjustment of self-delay).

img

Figure 7.24 Position errors and velocity errors (with an adaptive adjustment of self-delay).

img

It should be noted that the convergence of both the adaptive algorithm and the stability of the overall system heavily depend on the values of the adaptive gain β and the initial value of the delay estimation τ'1(0). Generally speaking, the larger β is, the smaller the upper bound of admissible initial value of τ'1 is; when the initial value of the delay estimation is sufficiently close to the value satisfying the matching condition, the adaptive gain can be arbitrarily large (see Table 7.1).

Table 7.1 Parameter values under which the adaptive algorithm converges.

img

7.5 Notes and References

The study on consensus problems with communication delays can be traced in the research of distributed computation algorithms back to as early as the 1980s (Tsitsiklis, Bertsekas and Athans 1986). Similar problems are also studied in the context of synchronization of coupled oscillators (Yeung and Strogatz 1999). Recently, Olfati-Saber and Murray (2004) reviewed the consensus problem of the first-order multi-agent system and proposed a consensus protocol in which each agent delays its own measurement of state by the same value as the communication delay so that it could be matched by the delayed states of its neighbors. Olfati-Saber and Murray (2004) obtained a delay-dependent sufficient condition of constant consensus for the system with a uniform communication delay. Lin et al. (2007) extended the consensus protocol of Olfati-Saber and Murray (2004) to the second-order multi-agent system, and also obtained a delay-dependent sufficient condition of the second-order consensus. However, since each self-delay is equal to the corresponding communication delay in their protocol and only the unified delay bound is considered, the multi-agent systems studied in Olfati-Saber and Murray (2004) and Lin et al. (2007) are essentially homogeneous.

Lee and Spong (2006) considered a consensus protocol without self-delay and obtained a delay-independent sufficient condition of constant consensus for high-order heterogeneous systems with diverse communication delays by using Gershgorin's disc lemma and the small-gain theorem shown in Chapter 2. By using a similar technique, Wang and Elia (2008) also considered the consensus protocol without self-delay and gave a delay-independent sufficient condition of constant consensus for the first-order multi-agent systems with heterogeneous dynamic communication channels. Arcak (2007), Chopra and Spong (2006) studied heterogeneous MASs based on the passivity theory, and developed a general framework for the design of group coordination control of systems with nonlinear dynamical agents, which is applicable to the constant-consensus problem. Lestas and Vinnicombe (2006, 2007, 2010) also considered the constant-consensus problem for heterogeneous systems, and introduced the notion of S-hull, a relaxation of the convex hull of a set in the complex plane, to overcome the conservativeness of small-gain-like or passivity-like stability results. Using the frequency-domain analysis theory developed in Chapter 3, Section 7.1 gives scalable delay-dependent conditions of constant consensus for the first-order multi-agent system with diverse input delays and communication delays while Section 7.2 gives scalable delay-dependent conditions of the second-order consensus for the second-order multi-agent system with diverse input delays and asymmetric perturbations in communication channels. Sections 7.1 and 7.2 are mainly based on the results of Tian and Liu (2008, 2009).

Actually, the existence of constant consensus depends only on the connectivity of the interconnection topology of MASs (Ren and Beard 2005; Wang, Cheng and Hu 2008; Xiao and Wang 2007). The values of self-delays introduced by agents in consensus protocols may lead to instability of the consensus solution (Papachristodoulou, Jadbabaie and Mijünz 2010) but they do not influence the existence of the constant-consensus solution. Therefore, the main focus of the above-mentioned references on high-order heterogeneous MASs is on the stability instead of the existence of the set of consensus solutions.

It can be shown that an inappropriate value of self-delay may lead to the non-existence of high-order consensus solution. To guarantee the existence of high-order consensus solutions, currently existing consensus protocols introduce self-delays which are exactly equal to the corresponding communication delays (see, e.g., Hu and Hong 2007). In practice, however, communication delays can only be estimated approximately. Section 7.3 proposes a high-order consensus protocol based on the estimation of communication delays and investigates the existence of high-order consensus solution of high-order heterogeneous MASs under such a protocol. Section 7.4 presents a simple algorithm for on-line adjusting self-delays to guarantee the existence of high-order consensus solutions. These two sections are mainly taken from Tian and Zhang (2012) except for Section 7.3.4.

References

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

Blondel VD, Hendrickx JM, Olshevsky A and Tsitsiklis JN (2005). Convergence in multiagent coordination, consensus, and flocking. Proceeding of the Joint 44th IEEE Conference on Decision and Control and European Control Conference, Seville, Spain, 2996–3000.

Cao M, Morse AS and Anderson BDO (2006). Reaching an agreement using delayed information. IEEE Conference on Decision and Control, San Diego, CA, USA, 3375–3380.

Chopra N and Spong MW (2006). Passivity-based control of multi-agent systems. Advances in Robot Control: From Everyday Physics to Human-like Movements, S. Kawamura and M. Svinin, Editors, 107–134, Spinger-Verlag, Berlin.

Hu J and Hong Y (2007). Leader-following coordination of multi-agent systems with coupling time delays. Physica A, 374, 853–863.

Jadbabaie A, Lin J and Morse AS (2003). Coordination of groups of mobile autonomous agents using nearest neighbor rules. IEEE Transactions on Automatic Control, 48, 988–1001.

Lee D and Spong MW (2006). Agreement with non-uniform information delays. Proceedings of the American Control Conference, Minneapolis, Minnesota, USA, 756–761.

Lestas I and Vinnicombe G (2006). Scalable decentralized robust stability certificates networks of interconnected heterogeneous dynamical systems. IEEE Transactions on Automatic Control, 51, 1613–1625.

Lestas I and Vinnicombe G (2007). The S-hull approach to consensus. Proceedings of the 46th IEEE Conference on Decision and Control, New Orleans, USA, 182–187.

Lestas I and Vinnicombe G (2010). Heterogeneity and scalability in group agreement protocols: Beyond small gain and passivity approaches. Automatica, 46, 1141–1151.

Lin P, Jia Y, Du J and Yuan S (2007). Distributed consensus control for second-order agents with fixed topology and time-delay. Proceeding of the 26th Chinese Control Conference, Zhangjiajie, Hunan, China, 577–581.

Moreau L (2005). Stability of multiagent systems with time-dependent communication links. IEEE Transactions on Automatic Control, 50, 169–182.

Olfati-Saber R and Murray RM (2004). Consensus problems in networks of agents with switching topology and time-delays. IEEE Transactions on Automatic Control, 49, 1520–1533.

Papachristodoulou A, Jadbabaie A and Münz U (2010). Effect of delay in multi-agent consensus and oscillator synchronization. IEEE Transactions on Automatic Control, 55, 1471–1477.

Ren W and Beard RW (2005). Consensus seeking in multiagent systems under dynamically changing interaction topologies. IEEE Transactions on Automatic Control, 50, 655–661.

Tian Y-P and Liu C-L (2008). Consensus of multi-agent systems with diverse input and communication delays. IEEE Transactions on Automatic Control, 53, 2122–2128.

Tian Y-P and Liu C-L (2009). Robust consensus of multi-agent systems with diverse input delays and asymmetric interconnection perturbations. Automatica, 45, 1347–1353.

Tian Y-P and Zhang Y (2012). High-order consensus of heterogeneous multi-agent systems with unknown communication delays. Automatica, 48, 1205–1212.

Tsitsiklis JN, Bertsekas DP and Athans M (1986). Distributed asynchronous deterministic and stochastic gradient optimisation algorithms. IEEE Transactions on Automatic Control, 31, 803–812.

Vicsek T, Czirok A, Ben Jacob E, et al.(1995). Novel type of phase transitions in a system of self-driven particles. Physical Review Letters, 75, 1226–1229.

Wang J, Cheng D and Hu X (2008). Consensus of multi-agent linear dynamic systems. Asian Journal of Control, 10, 144–155.

Wang J and Elia N (2008). Consensus over network with dynamic channels. Proceeding of the American Control Conference, Seattle, WA, 2637–2642.

Wang W and Slotine JJE (2006). Contraction analysis of time-delayed communications and group cooperation. IEEE Transactions on Automatic Control, 51, 712–717.

Xiao F and Wang L (2007). Consensus problems for high-dimensional multi-agent systems. IET Control Theory and Applications, 1, 830–837.

Yeung MKS and Strogatz SH (1999). Time delay in the Kuramoto model of coupled oscillators. Physical Review Letter, 82, 648–651.

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.15.186.79