Conference PaperPDF Available

Abstract and Figures

Synchronization among autonomous agents via local interactions is one of the benchmark problems in multi-agent control. Whereas synchronization algorithms for identical agents have been thoroughly studied, synchronization of heterogeneous networks still remains a challenging problem. The existing algorithms primarily use the internal model principle, assigning to each agent a local copy of some dynamical system (internal model). Synchronization of heterogeneous agents thus reduces to global synchronization of identical generators and local synchronization between the agents and their internal models. The internal model approach imposes a number of restrictions and leads to sophisticated dynamical (and, in general, nonlinear) controllers. At the same time, passive heterogeneous agents can be synchronized by a very simple linear protocol, which is used for consensus of first-order integrators. A natural question arises whether analogous algorithms are applicable to synchronization of agents that do not satisfy the passivity condition. In this paper, we study the synchronization problem for heterogeneous agents that are not passive but satisfy a weaker input feedforward passivity (IFP) condition. We show that such agents can also be synchronized by a simple linear protocol, provided that the interaction graph is strongly connected and the couplings are sufficiently weak. We demonstrate how stability of cooperative adaptive cruise control algorithms and some microscopic traffic flow models reduce to synchronization of heterogeneous IFP agents.
Content may be subject to copyright.
Simple synchronization protocols for
heterogeneous networks: beyond passivity
(extended version)
Anton V. Proskurnikov ,∗∗ Manuel Mazo Jr
Delft Center for Systems and Control (DCSC), Delft University of
Technology, The Netherlands
∗∗ ITMO University & Institute for Problems of Mechanical
Engineering (IPME RAS), St. Petersburg, Russia
Abstract: Synchronization among autonomous agents via local interactions is one of the
benchmark problems in multi-agent control. Whereas synchronization algorithms for identical
agents have been thoroughly studied, synchronization of heterogeneous networks still remains
a challenging problem. The existing algorithms primarily use the internal model principle,
assigning to each agent a local copy of some dynamical system (internal model). Synchronization
of heterogeneous agents thus reduces to global synchronization of identical generators and local
synchronization between the agents and their internal models. The internal model approach
imposes a number of restrictions and leads to sophisticated dynamical (and, in general,
nonlinear) controllers. At the same time, passive heterogeneous agents can be synchronized
by a very simple linear protocol, which is used for consensus of first-order integrators. A natural
question arises whether analogous algorithms are applicable to synchronization of agents that
do not satisfy the passivity condition. In this paper, we study the synchronization problem
for heterogeneous agents that are not passive but satisfy a weaker input feedforward passivity
(IFP) condition. We show that such agents can also be synchronized by a simple linear protocol,
provided that the interaction graph is strongly connected and the couplings are sufficiently
weak. We demonstrate how stability of cooperative adaptive cruise control algorithms and some
microscopic traffic flow models reduce to synchronization of heterogeneous IFP agents.
1. INTRODUCTION
As the influential monograph (Strogatz, 2003) states, “the
tendency to synchronize is one of the most pervasive
drives in the universe, extending from atoms to animals”.
Synchrony among subsystems (agents, cells) of a complex
system is a basic principle, which explains many natu-
ral phenomena (Strogatz, 2003) and has found numer-
ous applications in engineering (Mesbahi and Egerstedt,
2010; Olfati-Saber et al., 2007; Wu, 2007; Ren and Beard,
2008). Establishing synchronization (consensus) is consid-
ered now as a benchmark problem in multi-agent control
and has been thoroughly examined in the recent decades.
Most of the attention has been paid to synchronization
among identical agents. The protocols establishing syn-
chronization among single integrators are usually based
on the idea of contraction: the convex hull, spanned by the
agents’ states, is shrinking until it collapses into a single-
ton (M¨unz et al., 2011). An alternative approach is based
on convergence criteria for infinite matrix products (Ren
and Beard, 2008). The protocols for synchronization of
agents, obeying higher order equations, are similar in
spirit to first-order algorithms. Synchronization of linear
and linearly coupled agents is often analyzed via the
spectral decomposition of the Laplacian matrix (Olfati-
Saber et al., 2007; Li et al., 2010; Ren and Beard, 2008;
?Partial funding was provided by STW (project 13712). E-mails:
anton.p.1982@ieee.org, m.mazo@tudelft.nl
Ren and Cao, 2011). Nonlinear protocols are usually
examined by Lyapunov methods (Ren and Cao, 2011),
employing, among others, the Kalman-Yakubovich-Popov
lemma (Zhang et al., 2014; Proskurnikov and Matveev,
2015), contraction theory (DeLellis et al., 2011) and the
idea of incremental passivity (Stan and Sepulchre, 2007;
Proskurnikov et al., 2015; Liu et al., 2015b).
However, in practice autonomous agents are usually het-
erogeneous. Algorithms for output synchronization of
non-identical agents have been proposed quite recently
and most of them employ the internal model princi-
ple (Wieland et al., 2011; De Persis and Jayawardhana,
2014; Isidori et al., 2014; Bidram et al., 2014; Liu et al.,
2015a), assigning to each agent a virtual copy of some
dynamical system, referred to as the internal model or the
local reference generator. The control algorithm then con-
sists of two layers: a protocol, synchronizing the (identical)
reference generators and local model-matching controllers,
synchronizing the agents to their generators.
The general internal model approach has, however, several
disadvantages. Being formally decentralized, its implemen-
tation assumes that the agents share the same internal
model and are able to match it (e.g. in the case of
linear agents the Francis regulator equations should be
solvable (Wieland et al., 2011; Liu et al., 2015a)). Hence
design of an algorithm requires to know the global infor-
mation about the network. Unlike many synchronization
arXiv:1703.02937v1 [cs.SY] 8 Mar 2017
algorithms for identical agents (Olfati-Saber et al., 2007;
Li et al., 2010; Ren and Beard, 2008; Ren and Cao, 2011)
that use only relative measurements, that is, the deviations
between an agent’s output and the outputs of its neigh-
bors, the model-matching controllers need access to the
absolute outputs of the agents. Dealing with mobile robots,
this implies that agents have to measure their positions
and/or velocities in the global frame of reference.
At the same time, synchronization among heterogeneous
passive agents (e.g. mechanical systems in the Euler-
Lagrange form) can be established by the same simplest
protocols (Pogromsky and Nijmeijer, 2001; Arcak, 2007;
Hatanaka et al., 2015) as used to synchronize single inte-
grator agents (Olfati-Saber et al., 2007). Such a protocol
does not require any knowledge of the agents’ dynamics
(except for their passivity) and uses only deviations be-
tween the agents’ outputs, but not the outputs themselves.
Thus a visible gap exists between the problems of syn-
chronization in networks of passive heterogeneous agents,
provided by a very simple algorithm, and synchroniza-
tion among general heterogeneous agents, which requires
sophisticated model-based controllers. In this paper, we
make a step towards filling this gap and show that
the conventional synchronization algorithm for passive
agents (Hatanaka et al., 2015) is applicable also to input-
feedforward passive (IFP) (Khalil, 1996; Torres et al.,
2015) agents, provided that the couplings among them
are sufficiently weak. The class of IFP systems is much
broader than the class of passive systems (and contains,
in particular, all asymptotically stable linear systems). We
demonstrate applications of our results to the design of
cooperative adaptive cruise control (CACC) for platoons
of automated vehicles and stability of a microscopic traffic
flow model with delayed drivers’ responses, both of which
can be reduced to synchronization of IFP agents.
2. PRELIMINARIES
In this section, we introduce basic concepts from graph
theory and define input-feedforward passivity (IFP).
2.1 Graphs and their connectivity properties
A (weighted directed) graph is a triple G= (V,E, A),
where V={v1, . . . , vN}stands for the set of nodes,E ⊂
V × V is a set of arcs and A= (ajk )N
j,k=1 is a non-negative
adjacency matrix, such that ajk >0 if (vk, vj)Eand
otherwise ajk = 0. We always assume that the number of
nodes Nand their indices are fixed, so V={1, . . . , N },
there is a one-to-one correspondence between such graphs
and their adjacency matrices A7→ G[A]
= (V, E [A], A),
where E[A]
={(j, k) : akj 6= 0}. Henceforth all graphs
have no self-loops ajj = 0 j. A graph is called undirected
if A=A>. For any node jwe introduce the weighted in-
and out-degrees d+
j[A]
=PN
k=1 ajk and d
j[A]
=PN
k=1 akj .
Awalk connecting nodes vand v0is a sequence of
nodes vi0
=v, vi1, . . . , vis1, vis
=v0(n1) such that
(vik1, vik)Efor k= 1, . . . , s. A graph is strongly
connected if a walk between any two distinct nodes exists.
A graph is quasi-strongly connected, or has a directed
spanning tree, if one of its nodes is connected by walks to
all other nodes. For an undirected graph these conditions
are equivalent (such a graph is simply called connected).
2.2 Passivity and input-feedforward passivity
Consider the dynamical system
˙x(t) = f(x(t), u(t)), y(t) = h(x(t), u(t)), t 0,(1)
where x(t)Rn,u(t)Rmand y(t)Rmstand,
respectively, for the state, control and output.
The system (1) is passive (Khalil, 1996; Willems, 1972) if
there exists a storage function V(x)0 such that
V(x(T)) V(x(0)) ZT
0
y(t)>u(t)ds T0 (2)
(here Tvaries in the interval where the solution exists).
Assuming Vto be C1-smooth, (2) can be rewritten as
˙
V(x, u) = V
∂x f(x, u)h(x, u)>uxRn, u Rm.(3)
In this paper, we primarily deal with systems, satisfying a
“relaxed” passivity condition, defined as follows.
Definition 1. The system (1) is IFP(α) (input-feedforward
passive with the passivity index α) if it is passive with
respect to the output ˜y=y+αu, i.e.
V(x(T)) V(x(0)) ZT
0y(t)>u(t) + α|u(t)|2dt. (4)
In the case α= 0 an IFP(α) system is passive; if α < 0
the condition (4) is referred to as the strict input passivity.
In this paper, we are primarily interested in systems that
are not passive but IFP(α) with α > 0. Examples of such
systems are discussed in Section 3.3.
Although this is not required by the formal definition, the
conditions of passivity and IFP usually hold for systems
with zero equilibrium: f(0,0) = 0 and h(0,0) = 0. For
systems without equilibria points a modification of passiv-
ity condition exists, referred to as the incremental passiv-
ity (De Persis and Jayawardhana, 2014; Liu et al., 2015b).
Similarly, we introduce the incremental IFP condition.
Definition 2. A dynamical system is said to be iIFP(α)
(incrementally IFP(α)) if for any two solutions (x1, u1, y1)
and (x2, u2, y2) the respective deviations δx =x2x1,
δu =u2u1,δy =y2y1satisfy the inequality
V(δx(T)) V(δx(0)) ZT
0δy>δu +α|δu|2dt, (5)
where Tbelongs to the interval where both solutions exist.
The function Vis called the incremental storage function.
Obviously, for linear systems IFP(α)iIFP(α).
3. MAIN RESULTS: SYNCHRONIZATION
PROTOCOLS FOR IFP AGENTS
Consider a group of Nagents obeying the equations:
˙xj(t) = fj(xj(t), uj(t)), yj(t) = hj(xj(t)), t 0,(6)
for j∈ {1, . . . , N }. Here xj(t)Rnj,uj(t)Rm,
yj(t)Rmstand respectively for the jth agent’s state,
control and output.
In this paper, we study distributed protocols, synchroniz-
ing the outputs yjasymptotically or in L2-norm.
Definition 3. Solutions {(xj(t), uj(t), yj(t))}N
j=1 of the sys-
tems (6), defined on t[0; ), are output synchronized if
|yi(t)yj(t)| −
t→∞ 0i, j = 1, . . . , N. (7)
More specifically, the solutions are output synchronized
with a predefined reference signal ¯y: [0; )Rmif
|yi(t)¯y(t)| −
t→∞ 0i= 1, . . . , N. (8)
Definition 4. Solutions {(xj(t), uj(t), yj(t))}N
j=1 of the sys-
tems (6), defined for t0, are output L2-synchronized if
Z
0
|yi(t)yj(t)|2dt < ∞ ∀i, j = 1, . . . , N. (9)
The solutions are output L2-synchronized with a prede-
fined reference signal ¯y: [0; )Rmif
Z
0
|yi(t)¯y(t)|2dt < ∞ ∀i= 1, . . . , N. (10)
In practice, the difference between the asymptotical
and L2-synchronization is minor. Mathematically, none
of these conditions implies the other one. However, in
some special situations it is possible to prove that L2-
synchronization implies asymptotical synchronization.
Proposition 5. Let yj(t) be absolutely continuous and
( ˙yi˙yj)Lp[0; ] for some p > 1 and for any i, j.
Then (9) implies (7). If, additionally, ¯y(t) is absolutely
continuous and (˙yi˙
¯y)Lp[0; ]ithen (10) entails (8).
Proposition (5), as well as all other statements of this
paper, is proved in Appendix. In the following subsections
we examine synchronization algorithms.
3.1 Synchronization without reference signal
We start examinating the linear controller:
uj(t) =
N
X
k=1
ajk (yk(t)yj(t)),(11)
where ajk 0 are the coupling gains. The matrix A=
(ajk ) determines the interaction graph (or the network’s
topology)G[A], where node kis connected to node jby
an arc if and only if ajk 6= 0, that is, the control input of
agent jis directly influenced by the output of agent k.
It is widely known (Olfati-Saber et al., 2007; Ren and
Beard, 2008; M¨unz et al., 2011) that single integrators ˙yj=
uj, coupled via the protocol (11) reach consensus (that
is, a common limit y= limt→∞ yj(t) exists) whenever
G[A] has a directed spanning tree. Output synchroniza-
tion (7) is retained replacing single integrators by general
passive systems (6) and assuming strong connectivity of
G[A] (Hatanaka et al., 2015, Theorem 8.3). Our first result
extends this to IFP agents.
Theorem 6. Assume that agent j(for j= 1, . . . , N ) is
IFP(αj) with a storage function Vj(xj)0. Let G[A] be
strongly connected and the couplings be “weak”, i.e.
αjd+
j[A] = αj
N
X
k=1
ajk <1/2j= 1, . . . , N . (12)
Then the following statements hold.
(1) Any solution of the system (6),(11), which is pro-
longable to , is output L2-synchronized (9);
(2) Suppose that for any jthe function Vjis radially
unbounded lim|xj|→∞ Vj(xj) = , the map fjis
continuous and hjis C1-smooth. Then, any solution
of the closed-loop system (6),(11) is prolongable to
, bounded, and output synchronized (7).
The proofs of Theorem 6 and other results of this section
are given in Section A. Note that in the case of αj= 0 the
inequalities (12) hold for any matrix A, and Theorem 6
coincides with Theorem 8.3 in (Hatanaka et al., 2015). We
proceed with two remarks, regarding the assumptions.
Remark 7. Unlike passive agents, for general IFP agents
the requirement of weak coupling (12) cannot be disre-
garded, as demonstrated by the following example. For
any p, q > 0 the system:
...
yj(t) + p¨yj(t) + q˙yj(t) = uj(t)R, t 0,(13)
is IFP(α) with some α=α(p, q)>0 (c.f. Subsect. 3.3).
Applying the protocol (11) with all-to-all coupling aij =
κ>0i, j to a group of identical agents (13), output
synchronization is guaranteed (Olfati-Saber et al., 2007;
Li et al., 2010) only when the polynomial s3+ps2+qs +
κ(N1) = 0 is Hurwitz. Accordingly to the Routh-
Hurwitz criterion, this is possible only if κ(N1) < pq,
i.e. the gain κis small.
Remark 8. Dealing with general heterogeneous agents, the
condition of strong connectivity cannot be replaced by the
existence of a directed spanning tree in G[A]. Consider, for
instance, a pair (N= 2) of harmonic oscillators
¨
ξ1+ω2
1ξ1=u1,¨
ξ2+ω2
2ξ2=u2, ω16=ω2.
that are passive with respect to the outputs y1=˙
ξ1and
y2=˙
ξ2. Consider the protocol u1=k(˙
ξ2˙
ξ1), u2=
0, which corresponds to the graph with N= 2 nodes
and the only arc 2 7→ 1. It can be shown that the
system has a family of solutions ξ1(t) = Re[W(ıω2)ceıω2t],
ξ2=Re[ceıω2t], where cCis constant and W(s) =
ks/(s2+ks +ω2
1). The corresponding outputs are y1(t) =
Re[ıω2W(ıω2)ceıω2t] and y2(t) = Re[ıω2ceıω2t]. Since
|W(ıω2)|=
kıω2
(ω2
1ω2
2) + kıω2
<1,
the outputs are harmonic signals with the same frequency
ω2but different amplitudes and cannot be synchronous.
Under some additional assumptions, synchronization over
a quasi-strongly connected graph may be established by
using the internal model control (Isidori et al., 2014).
3.2 Reference-tracking synchronization
We now consider the more complex problem of output
synchronization with a reference signal (8). In this paper
we confine ourselves to a special situation: when the
desired trajectory is generated as the output of an agent
for some appropriate control input and initial condition.
Assumption 9. For any jsystem (6) has a solution
xj(t),¯uj(t),¯yj(t)) such that ¯yj(t)¯y(t)t0 (in partic-
ular, the solution is prolongable to ). At any time agent
jis aware of the value 1¯uj(t), however the reference ¯y(t)
may be available only to a few “dedicated” agents.
1If ¯uj(·) is not unique, agent jknows one of such solutions.
Assumption 9 is often adopted implicitly or explicitly
in reference-tracking synchronization problems. For linear
agents (Li et al., 2010; Liu et al., 2015a) the reference
signal ¯y(t) is usually supposed to be an output of a
reference system, whose model is known and included
by the models of other agents. Dealing with first-order
integrator agents ˙yj=uj, Assumption 9 implies that the
agents know the derivative ˙
¯y(t); this holds e.g. if ¯y(t) =
t¯v+ ¯y(0), where ¯vis known, but the initial condition ¯y(0)
is uncertain. A practical example of this type is discussed
in Section 4. Note that the solution (¯xj(t),¯uj(t),¯yj(t)) is
not assumed to be asymptotically stable, so the control
uj(t) = ¯uj(t)does not guarantee the reference signal
tracking (8). In general, only some of the agents are able
to measure the tracking error ¯y(t)yj(t), whereas the
remaining agents measure only deviation between theirs
and their neighbors’ outputs.
Consider the following modification of the algorithm (11)
ui(t) = ¯ui(t)+bi(¯y(t)¯yi(t))+
N
X
j=1
aij (yj(t)yi(t)).(14)
Here bi>0 if agent ihas access to the reference signal,
and otherwise bi= 0. The following result is a counterpart
of Theorem 6 for reference-tracking synchronization.
Theorem 10. Let Assumption 9 hold and further assume
that: for all j∈ {1, . . . , N }agent jis iIFP(αj), G[A] is
strongly connected, at least one agent has access to the
reference signal, i.e. Pibi>0, and the couplings are
sufficiently weak, i.e.
αj(d+
j[A]+2bj)<1/2j= 1, . . . , N . (15)
Then, the following two statements hold:
(1) Any solution of the system (6),(14), prolongable to
, is output L2-synchronized with the reference
signal (10); in particular, R
0|¯u(t)uj(t)|2dt < .
(2) If for all jthe functions Vjare radially unbounded,
the maps fjare C1-smooth, the Jacobians ∂fj
∂xj,fj
∂uj
are uniformly bounded, and the maps hjare linear:
hj(ξ1ξ2) = hj(ξ1)hj(ξ2); then, any solution of the
closed-loop system (6),(14) is prolongable to and
output synchronized (8) with the reference signal.
3.3 Examples of IFP agents
In this subsection, examples of IFP agents are provided.
SISO agents with a pole at zero
Consider a SISO system
(s)ζ(t) = u(t)R, s
=d
dt, ρ(λ) =
r
X
k=0
ρkλk;
y(t) = η(s)ζ(t), η(λ) =
r
X
k=0
ηkλk
(16)
Lemma 11. Assume that ρ(s) is a Hurwitz polynomial and
η0ρ00. Then the system (16) is IFP(α) for sufficiently
large α0. Denoting the transfer function from uto yby
W(λ) = η(λ)/(λρ(λ)), the passivity index can be found as
α=inf
ωRRe W (ıω).(17)
For instance, Lemma 11 implies that the system (13) is
IFP (in this case, ρ(λ) = λ2++qis Hurwitz since
p, q > 0 and y(t) = ξ(t)).
First-order delayed integrators
Consider now a delayed system:
˙y(t) = u(tα)Rm.(18)
Here α0 is a constant delay and we assume, by
definition, that u(t)u0(t) for t[α; 0], where
u0L2([α; 0] Rm) is a given function. The vector
y(0) and the function u0are the initial conditions for
the system (18). Formally, our definition of IFP deals
with ordinary differential equations (1) only and is not
applicable to delay systems. However, the following weaker
condition holds for (18), see (Proskurnikov, 2016, p.141,
proof of Lemma 7.3).
Lemma 12. For any solution of (18) one has
ZT
0
(y(t)>u(t) + α|u(t)|2)dt ≥ −V T0,(19)
where V=V(y(0), u0(·)) 0 is independent of T.
Lemma 12 allows to extend the synchronization criteria to
ensembles of agents (20).
Theorem 13. For a group of linear delayed agents
˙yi(t) = ui(tαi), i = 1, . . . , N , (20)
the protocol (11) provides output synchronization (7)
and L2-synchronization (9), whenever the graph G[A] is
strongly connected and (12) holds.
Remark 14. In the monograph Tian (2012) a more general
result is formulated without a complete proof (Theorem
7.10), stating that under assumptions of Theorem 13
synchronization is retained if the graph is not strongly
connected but has a directed spanning tree.
Remark 15. Theorem 10 also holds for agents (20). How-
ever, Assumption 9 becomes impractical since each agent
has to be aware of ¯ui(t) = ˙
¯y(t+αj) at time t, which makes
the controller (14) non-causal. The protocol (14) may still
be used in the case where the reference signal is linear
¯y(t) = v0t+ ¯y(0) and v0is known, but ¯y(0) is uncertain.
4. SYNCHRONIZATION IN VEHICLE PLATOONING
AND TRAFFIC FLOW MODELING
In this section, we consider two practical applications of
the synchronization criteria from Section 3.
4.1 Stability of a microscopic traffic flow model
A basic problem in vehicular traffic is the prevention of
congestions and accidents. Microscopic traffic flow models
are often employed to represent the traffic flow as a result
of cooperation between individual drivers. Since the pio-
neering work of Chandler et al. (1958), the delay in drivers
reaction has been recognized as a crucial factor participat-
ing into the overall flow dynamics. The simplest model of
this kind (Chandler et al., 1958; Sipahi et al., 2007) deals
with Nvehicles, indexed 1 through N, traveling along a
common straight or circular single lane road (their order
remains unchanged since overtaking is not possible). Each
Fig. 1. Platoon of vehicles with bidirectional coupling.
driver is aiming to equalize his velocity of his own vehicle
with that of its predecessor:
˙vi(t) = ui(tα), ui(t) = K(vi1(t)vi(t)).(21)
Here vi(t) is the speed of the i-th vehicle, αis the
delay in its driver’s action, and Kstands for the driver’s
“sensitivity” to alterations of the relative velocity of the
predecessor vehicle. In the case of straight road, v0(t)v0
is the desired velocity with respect to the leading vehicle
1; for a circular road, v0(t)vN(t), i.e. vehicle 1 follows
vehicle N. A key issue addressed via this model Chandler
et al. (1958) is that of the stability of the “synchronous”
manifold: v1=. . . =vN.
For the straight road case a necessary and sufficient
condition for such a synchronization: 2αK < 1 was found
in (21). We extend this classical result to the traffic flow
model with a general directed interaction topology and
heterogeneous delays and sensitivities of the drivers.
˙vi(t) = ui(tαi), ui(t) =
N
X
j=1
aij (vj(t)vi(t)),i. (22)
The model (22) allows, in particular some drivers to
respond to the change not only in the predecessor’s, but
also in the follower’s velocity, or use the information about
several predecessors and followers. The following theorem
gives a criterion of velocity synchronization in (22) under
the assumption of a strongly connected topology, which
holds e.g. for uni- and bidirectional ring coupling (circular
road). The gain aij 0 in (22) stands for the sensitivity of
driver ito changes in the speed of vehicle j. Theorem 13,
applied to yi=vi, yields in the following corollary:
Corollary 16. Suppose that the graph G[A] is strongly
connected and (12) holds. Then the vehicles’ velocities are
asymptotically synchronized vi(t)vj(t)
t→∞ 0.
4.2 An application to cooperative adaptive cruise control
In this subsection we demonstrate an application of The-
orem 10 to the stability of a platoon of vehicles (Fig. 2),
constituted by the leading vehicle 0 and Nfollower vehi-
cles, indexed 1 through N(Fig. 1). Cooperative adaptive
cruise control (CACC) system implements a control algo-
rithm, making each vehicle keep the safe distance to the
predecessor and, provided that this safety constraint is sat-
isfied, follow the leader’s velocity. The interaction topology
between the vehicles may be different (Zheng et al., 2016);
the most studied is a unidirectional topology, where each
vehicle has information only about the predecessor.
In this subsection, we examine a CACC algorithm with
bidirectional interactions. The advantages of bidirectional
platooning algorithms over unidirectional ones are dis-
cussed e.g. in (Zhang et al., 1999; Barooah et al., 2009;
Zheng et al., 2016) (see also references therein); in many
senses such algorithms are more robust against distur-
bances propagating through the platoon (“string-stable”).
We examine the CACC algorithm, proposed in (Barooah
et al., 2009). The leader’s speed v0(t)v0is broadcasted
to every follower (Fig. 1). Besides this, the vehicles 1
through N1 measure the distances to both their pre-
decessors and followers, and the rear vehicle Nmeasures
the distance to its predecessor. Denoting the position of
vehicle i’s rear bumper by qiR(see Fig. 2), the goal of
the CACC algorithm is to keep the desired distance to the
predecessor and the desired velocity, i.e.
qi1(t)qi(t)
t→∞ si, vi(t) = ˙qi(t)
t→∞ v0.(23)
Fig. 2. Platoon of vehicles. Notation used in the text
As usual in CACC problems (Zhang et al., 1999; Zheng
et al., 2016), the follower vehicles obey linear models
τi...
qi+ ¨qi=ai,des (t),(24)
where ai,des is the desired acceleration and τiis a time con-
stant, depending on the vehicle’s powertrain. The vehicles
1 through N1 apply the following controller:
ai,des(t) = µi(v0vi(t)) + ηi(qi1(t)qi(t)si)
+νi(qi+1(t)qi(t) + si+1 ),1iN1,(25)
Vehicle Nis controlled similarly, but has no follower
aN,des(t) = µN(v0vN(t)) + ηN(qN1(t)qN(t)sN).
(26)
Theorem 17. Let µiτi<1
2and ηi, νi>0 satisfy
µ2
i
2>
ηi+νi,1< i < N;
2η1+ν1, i = 1;
ηN, i =N.
i(27)
Then the algorithm (25), (26) provides (23).
The result of Theorem 17 can be extended to some
cases of nonlinear vehicles’ dynamics, where the inner-loop
engine and torque controllers (Zhang et al., 1999) fail to
attenuate the nonlinearities. Notice that Theorem 17 does
not address the string stability problem, i.e. the robustness
of CACC against small disturbances in measurements as
Nbecomes large; the analysis of string stability is based
on other techniques and is beyond the scope of this paper.
5. CONCLUSIONS AND FUTURE WORK
In this paper, we offer simple distributed protocols
for synchronization of heterogeneous non-passive agents
that satisfy an IFP property. We apply the obtained
results to analysis of microscopic traffic flow models
and CACC algorithms for heterogeneous platoons. The
results can be extended to nonlinearly coupled net-
works, where the couplings satisfy the conditions of anti-
symmetry and sector inequalities (Hatanaka et al., 2015;
Proskurnikov, 2016; Proskurnikov and Matveev, 2015),
time-varying graphs and antagonistic interactions among
the agents (Proskurnikov and Cao, 2016). The results may
be extended to discrete-time IFP agents. The robustness
of synchronization against measurements noises and com-
munication delays are subjects of ongoing research.
REFERENCES
Arcak, M. (2007). Passivity as a design tool for group
coordination. IEEE Trans. Autom. Control, 52(8),
1380–1390.
Barooah, P., Mehta, P., and Hespanha, J. (2009).
Mistuning-based control design to improve closed-loop
stability margin of vehicular platoons. IEEE Trans.
Autom. Control, 54(9), 2100–2113.
Bidram, A., Lewis, F., and Davoudi, A. (2014). Synchro-
nization of nonlinear heterogeneous cooperative systems
using input-output feedback linearization. Automatica,
50(10), 2578–2585.
Chandler, R., Herman, R., and Montroll, E. (1958). Traffic
dynamics: Analysis of stability in car following. Oper.
Res., 7, 165184.
De Persis, C. and Jayawardhana, B. (2014). On the
internal model principle in the coordination of nonlinear
systems. IEEE Trans. Control Netw. Syst., 1(3), 272–
282.
DeLellis, P., di Bernardo, M., and Russo, G. (2011).
On QUAD, Lipschitz, and contracting vector fields
for consensus and synchronization of networks. IEEE
Trans. Circuit Syst. - I, 58(3), 576–583.
Hatanaka, T., Chopra, N., Fujita, M., and Spong, M.
(2015). Passivity-Based Control and Estimation in
Networked Robotics. Springer.
Isidori, A., Marconi, L., and Casadei, G. (2014). Robust
output synchronization of a network of heterogeneous
nonlinear agents via nonlinear regulation theory. IEEE
Trans. Autom. Control, 59(10), 2680–2691.
Khalil, H. (1996). Nonlinear systems. Prentice-Hall,
Englewood Cliffs, NJ.
Li, Z., Duan, Z., Chen, G., and Huang, L. (2010). Con-
sensus of multiagent systems and synchronization of
complex networks: A unified viewpoint. IEEE Trans.
on Circuits and Systems - I, 57(1), 213–224.
Liu, H., De Persis, C., and Cao, M. (2015a). Robust
decentralized output regulation with single or multiple
reference signals for uncertain heterogeneous systems.
Int. J. Robust Nonlinear Control, 25, 1399–1422.
Liu, T., Hill, D., and Zhao, J. (2015b). Output syn-
chronization of dynamical networks with incrementally-
dissipative nodes and switching topology. IEEE Trans.
Circuits Syst.-I, 62(9), 2312–2323.
Mesbahi, M. and Egerstedt, M. (2010). Graph Theoretic
Methods in Multiagent Networks. Princeton University
Press, Princeton and Oxford.
unz, U., Papachristodoulou, A., and Allg¨ower, F. (2011).
Consensus in multi-agent systems with coupling delays
and switching topology. IEEE Trans. Autom. Control,
56(12), 2976–2982.
Olfati-Saber, R., Fax, J., and Murray, R. (2007). Consen-
sus and cooperation in networked multi-agent systems.
Proceedings of the IEEE, 95(1), 215–233.
Olfati-Saber, R. and Murray, R. (2004). Consensus prob-
lems in networks of agents with switching topology and
time-delays. IEEE Trans. Autom. Control, 49(9), 1520–
1533.
Pogromsky, A. and Nijmeijer, H. (2001). Cooperative oscil-
latory behavior of mutually coupled dynamical systems.
IEEE Trans. Circuits Syst. - I, 48(2), 152–162.
Proskurnikov, A., Zhang, F., Cao, M., and Scherpen,
J. (2015). A general criterion for synchronization of
incrementally dissipative nonlinearly coupled agents.
In Proceedings of European Control Conference (ECC
2015), 581–586. Linz, Austria.
Proskurnikov, A. (2016). Delay robustness of nonlinear
consensus protocols: Analytic criteria. In O. Sename,
E. Witrant, and E. Fridman (eds.), Recent results on
time-delay systems: analysis and control, chapter 7, 125–
146. Springer.
Proskurnikov, A. and Cao, M. (2016). Polarization in
coopetitive networks of heterogeneous nonlinear agents.
In Proc. IEEE Conf. Decision and Control (CDC). Las-
Vegas, NV, USA.
Proskurnikov, A. and Matveev, A. (2015). Popov-type
criterion for consensus in nonlinearly coupled networks.
IEEE Trans. Cybernetics, 45(8), 1537–1548.
Ren, W. and Beard, R. (2008). Distributed Consensus in
Multi-vehicle Cooperative Control: Theory and Applica-
tions. Springer-Verlag, London.
Ren, W. and Cao, Y. (2011). Distributed Coordination of
Multi-agent Networks. Springer.
Sipahi, R., Atay, F., and Niculescu, S.I. (2007). Stability
of traffic flow behavior with distributed delays modeling
the memory effects of the drivers. SIAM J. Appl. Math.,
68(3), 738–759.
Stan, G.B. and Sepulchre, R. (2007). Analysis of intercon-
nected oscillators by dissipativity theory. IEEE Trans.
Autom. Control, 52(2), 256–270.
Strogatz, S. (2003). Sync: The Emerging Science of
Spontaneous Order. Hyperion Press, New York.
Tian, Y.P. (2012). Frequency-Domain Analysis and Design
of Distributed Control Systems. John Wiley & Sons.
Torres, L., Hespanha, J., and Moehlis, J. (2015). Syn-
chronization of identical oscillators coupled through a
symmetric network with dynamics: A constructive ap-
proach with applications to parallel operation of invert-
ers. IEEE Trans. Autom. Control, 60(12), 3226–3241.
Wieland, P., Sepulchre, R., and Allg¨ower, F. (2011). An
internal model principle is necessary and sufficient for
linear output synchronization. Automatica, 47(5), 1068–
1074.
Willems, J. (1972). Dissipative dynamical systems (in 2
parts). Arch. Rational Mech. and Anal., 45(5), 321–393.
Wu, C. (2007). Synchronization in complex networks of
nonlinear dynamical systems. World Scientific, Singa-
pore.
Zhang, F., Trentelman, H., and Scherpen, J. (2014). Fully
distributed robust synchronization of networked lure
systems with incremental nonlinearities. Automatica,
50(10), 2515–2526.
Zhang, Y., Kosmatopoulos, E., Ioannou, P., and Chien,
C. (1999). Autonomous intelligent cruise control using
front and back information for tight vehicle following
maneuvers. IEEE Trans. Vehicul. Tech., 48(1), 319–328.
Zheng, Y., Li, S.E., Wang, J., Cao, D., and Li, K. (2016).
Stability and scalability of homogeneous vehicular pla-
toon: Study on the influence of information flow topolo-
gies. IEEE Trans. Intell. Transp. Syst., 17(1), 14–26.
Appendix A. TECHNICAL PROOFS
We start with several technical lemmas, used to prove
Theorems 6 and 10, and then proceed with proofs of
Proposition 5, Theorems 6, 10, 13 and 17 and Lemma 11.
A.1 Auxiliary lemmas
The proofs of Theorems 6,10 are based on the following
lemma. Consider two groups of vectors y1, . . . , yNRm
and u1, . . . , uNRm, such that
ui=
N
X
j=1
aij (yjyi)i= 1, . . . , N. (A.1)
Lemma 18. Let the graph G[A] be strongly connected.
Then the system of equalities (A.1) entails that
N
X
i=1
piy>
iui=1
2
N
X
i,j=1
piaij |yjyi|2,(A.2)
where (pi)N
i=1 are constants, determined 2by A= (aij ).
The proof of Lemma 18 is a part of the proof of Theo-
rem 8.5 in (Hatanaka et al., 2015) and omitted here.
Corollary 19. Let the graph G[A] be strongly connected
and the conditions (12) hold. Then (A.1) entails that
N
X
i=1
piy>
iui+αi|ui|2≤ −ε
N
X
i,j=1
|yjyi|2,(A.3)
where ε > 0 is a constant, determined by the matrix A.
Proof. Using the Cauchy-Schwartz inequality, one has
d+
i[A]αi
N
X
j=1
aij |yjyi|2=αi
N
X
j=1
aij
N
X
j=1
aij |yjyi|2
αi
N
X
j=1
a1/2
ij a1/2
ij (yjyi)
2
(A.1)
=αi|ui|2i.
By multiplying these inequalities by pi, summing them up
over iand using (A.2), it can be easily shown that
N
X
i=1
piy>
iui+αi|ui|2≤ −
N
X
i,j=1
κiaij |yjyi|2,(A.4)
where κi
=pi(1/2d+
i[A]αi)(12)
>0. The inequality (A.3)
with some ε=ε(A)>0 now follows from (A.4) since the
matrix (κiaij )N
i,j=1 corresponds to a strongly connected
graph (Olfati-Saber and Murray, 2004, Theorem 8).
The proof of Theorem 10 also will use the following lemma.
Lemma 20. Suppose that the system (1) is IFP(α) with
some α0 and storage function V. Consider the constants
b(0; 1/(2α)) and ˆα
=α/(1 2αb). Then the system (1)
is IFP(ˆα) for the same output yand the new input ˆu=u+
by with the storage function ˆ
V(x) = V(x)/(1 2αb).
Moreover, for γ
=b(1 αb)/(1 2αb)0 one has
2In fact, p>= (p1,...,pN) stands for the non-negative left eigen-
vector of the weighted Laplacian matrix L=L[A], corresponding
to zero eigenvalue p>L= 0. The strong connectivity of the graph
implies that piare strictly positive (Hatanaka et al., 2015).
ˆ
V(x(T)) ˆ
V(x(0)) ZT
0y>ˆu+ ˆα|ˆu|2γ|y|2dt (A.5)
for any solution and T0. The statement retains its
validity for iIFP system; in the latter case the vectors
x(t), y(t), u(t) in (A.5) should be replaced by the respective
deviations δx(t), δy (t), δu(t) between two solutions.
Proof. The proof is straightforward from Definition 1,
substituting u= ˆuby into (4) and noticing that
y>u+α|u|2= (1 2αb)y>ˆu+ ˆα|ˆu|2γ|y|2.
The statement for iIFP system is proved in the same way,
substituting δu =δˆubδy into (5).
A.2 Proof of Proposition 5
Note first that an absolutely continuous function ξ, such
that ˙
ξLp[0; ] with p > 1, is uniformly continuous
on [0; ]. This follows e.g. from the H¨older inequality,
entailing that for t0 and t0t
|ξ(t)ξ(t0)|=
Zt0
t
˙
ξ(s)ds
(t0t)qk˙
ξkLp[t;t0],
where q=p/(p1) (by definition, q= 1 if p=). If,
additionally, ξL2, the Barbalat lemma (Khalil, 1996)
implies that ξ(t)
t→∞ 0. Proposition 5 now follows,
applying this to, respectively, ξ=yjyiand ξ= ¯yyi.
A.3 Proof of Theorem 6.
Let the conditions of Theorem 6 hold. Introducing the
stack vector X(t) = [x1(t)>, . . . , xN(t)>]>and denoting
V(X)
=PN
i=1 piVi(xi), the IFP(αi) property of the
agents (6) and Corollary 19 imply that
V(X(T)) V(X(0))
N
X
i=1 ZT
0
piy>
iui+αj|ui|2dt
ε
N
X
i,j=1 ZT
0
|yj(t)yi(t)|2dt 0.
(A.6)
This implies statement (1): if the solution is defined for any
t0, the solution is output L2-synchronized. To prove
statement (2), notice that radial unboundedness of all
storage functions Vi(xi) implies that V(X) is also radially
unbounded since pi>0i. Since V(x(T)) V(x(0)) for
any T, the state vectors xi(t) are uniformly bounded, in
particular, the solution does not escape to infinity in finite
time and thus is infinitely prolongable. Recalling that the
maps hiare continuous, the outputs yi(t) are bounded;
the same holds for ui(t) due to (11). Since the maps fiare
continuous, ˙xi(t) are bounded. Recalling that hjis C1-
smooth, ˙yj(t) = h0(xj(t)) ˙xj(t) is also bounded. Output
synchronization (7) now follows from Proposition 5.
A.4 Proof of Theorem 10
Denoting ˜xi(t)
=xi(t)¯xi(t), ˜ui(t)
=ui(t)¯ui(t),
˜yi(t)
=yi(t)¯yi(t), where ( ¯xi(t),¯ui(t),¯yi(t)) is the solution
from Assumption 9. Recalling that ¯yi(t)¯y(t), one has
˜ui(t) + bi˜yi(t)(14)
=X
j=1
aij yj(t)˜yi(t)).(A.7)
Denoting ˆui
= ˜ui+bi˜yiand applying Lemma 20 to the
system (6), α=αjand b=bj, one arrives at
ˆ
Vixi(T)) ˆ
Vixi(0)) ZT
0˜y>
iˆui+ ˆαi|ˆui|2γi|˜yi|2dt,
(A.8)
where ˆαi=αi/(1 2biαi) and γi>0 if and only if bi>0.
The inequalities (15) imply that d+
j[A]ˆαj<1/2. Thus
retracing the proof of Corollary 19, (A.7) implies that
N
X
i=1
pi˜y>
iˆui+ ˆαi|ˆui|2≤ −ε
N
X
i,j=1
|yjyi|2,(A.9)
where pi>0 and ε > 0 depend on A,αiand bi.
Introducing the stack vector ˜
X(t) = [˜x1(t)>,...,˜xN(t)>]>
and the storage function ˆ
V(˜
X) = Pipiˆ
Vixi), we obtain
ˆ
V(˜
X(T)) ˆ
V(˜
X(0)) ≤ −ε
N
X
i,j=1 ZT
0
|˜yj(t)˜yi(t)|2dt
X
i
piγiZT
0
|˜yi(t)|2dt =ZT
0
Fy1(t),...,˜yN(t))dt.
(A.10)
Here F(y1, . . . , yN)
=εPN
i,j=1 |yjyi|2+Pipiγi|yi|2is a
quadratic form; since γi>0 for at least one i, this form is
positive definite and thus F(y1, . . . , yN)ε0(|y1|2+. . . +
|yN|2) for sufficiently small constant ε0>0. The end of
the proof retraces the proof of Theorem 6. If a solution
exists, (A.10) implies that ˜yi(t) = ¯y(t)yi(t) is L2-
summable and hence ˜ui(t) = ¯u(t)ui(t) is L2-summable,
which proves statement (1). To prove statement (2), notice
that since Viare radially unbounded, (A.10) implies that
the deviation of the solutions ˜
X(t) remains bounded; since
¯xi(t) is globally defined, X(t) cannot grow unbounded
in finite time. Thus the solution is prolongable to .
Recalling that hiis a linear map, the function ˜yi(t) =
yi(t)¯y(t) = hi(xi(t)¯xi(t)) = hixi(t)) is uniformly
bounded for any i. Thus ˜ui(t) is bounded due to (A.7)
and
|˙
˜xi(t)|=|fi(xi(t), ui(t)) fi( ¯xi(t),¯ui(t))| ≤
| ˜xi(t)|sup
xi,ui
∂fi(xi, ui)
∂ui
+|˜ui(t)|sup
xi,ui
∂fi(xi, ui)
∂ui
(here suprema are taken over the space of all possible
vectors xiRni,uiRm). Thus the functions ˙
˜xi(t) are
also uniformly bounded; using the linearity of hi, the same
holds for ˙
˜yi(t) = hi(˙
˜xi(t)). Applying Proposition 5, the
solutions are output synchronized (8).
A.5 Proof of Lemma 11
Lemma 11 is immediate from a more general result, which
in turn is implied by the standard positive real lemma.
Consider a linear SISO system
˙x=P x +Qu, y =Rx +Su (A.11)
and let W(λ)
=S+R(λI P)1Qstand for its scalar
transfer function.
Lemma 21. Let the system (A.11) be controllable and
observable. Then it is IFP(α) for some αif and only if
the following conditions hold
(1) the matrix Phas no strict unstable eigenvalues:
det(λI P)6= 0 when Re λ > 0;
(2) all imaginary eigenvalues (if they exist) are simple,
at any such eigenvalue λ=ıω0the residual is non-
negative limλıω0(λıω0)W(λ)0;
(3) Re W (ıω) + α0 for any ωRsuch that det(ıωI
P)6= 0.
Proof. As was noticed in Section 2, the IFP(α) condition
is equivalent to passivity of system (A.11) with respect to
the new input ˜y=y+αu. The statement of Lemma 21 is
immediate, applying the result of (Willems, 1972, Theo-
rem 1 in Part 2) (the positive real lemma) to the respective
transfer function ˜
W(ıω) = αIm+W(ıω).
Notice that if the condition (1) and (2) in Lemma 21 hold
then (3) is valid for sufficiently large α > 0. Indeed, the
function Re W (ıω) = [W(ıω) + W(ıω)]/2 (where ωR)
is bounded as ω→ ∞ and in the vicinity of any imaginary
pole due to statement (2), thus this function is bounded
and, in particular, semi-bounded from below.
Proof of Lemma 11 is now obvious from Lemma 21 since
the system (16) can be rewritten as a controllable and
observable system (A.11) with a single imaginary pole
λ= 0. Such a system satisfies condition (1) in Lemma 21,
and (2) also holds since limλ0λW (λ) = η000.
A.6 Proof of Theorem 13
Any solution of the linear time-invariant closed-loop sys-
tem (20), (11) is infinitely prolongable. A closer look at
the proof of statement (1) in Theorem 6 reveals that the
IFP(αi) condition can be replaced by a weaker condition
ZT
0
(yi(t)>ui(t) + αi|ui(t)|2)dt ≥ −ViT0,
where Vi0 is some constant (for an IFP(αi) agent (6)
with storage function Vi, the latter inequality holds for
Vi=Vi(xi(0))). Thanks to Lemma 12, statement (1) is
valid for the agents (20) and thus any solution is output L2
synchronized. Since uiL2[0; ] for any i, ˙yiL2[0; ]
due to (20) and thus outputs are also asymptotically
synchronized due to Proposition 5.
A.7 Proof of Theorem 17
Introducing the control inputs ui
=ai,des +µiviand
outputs yi=qi+ (si+si1+. . . +s1), the closed-loop
system (24), (25), (26) boils down to a group of agents
τi...
yi+ ¨yi+µi˙yi=ui, i = 1, . . . , N, (A.12)
coupled through (14). Here ¯ui(t) = µiv0, ¯y(t) = q0(t) and
bi=η1, i = 1
0, i > 1, aij =
ηi, i > 1, j =i1
νi, i < N, j =i+ 1
0,otherwise.
Using Lemma 11, the agent (A.12) is IFP(12
i) since
Re 1
τi(ıω)3+ (ıω)2+µi(ıω)=1
µ2
i+ (1 2τiµi)ω2+ω4.
A straightforward computation shows that (27) im-
plies (15). Obviously, the graph G[A] is a bidirectional
chain and thus is strongly connected. Hence, the outputs
are L2-synchronized (10) and ui¯uiL2. Denoting
εi
=viv0, one has ˙yi=vi=εi+v0, ¨yi= ˙εi,...
yi= ¨εiand
hence
τi¨εi+ ˙εi+µiεi=uiµv0=ui¯uiL2[0; ],
so that εi= ˙yi˙
¯yL2and ˙εiL2. Applying
Proposition (5), one proves that (7) holds and εi(t)0
as t→ ∞, which implies (23).
... In [5], agents are strictly G-passifiable via output feedback while [6] deals with linear agents which are either passive or passifiable via state/output feedback agent. In [10], [19], [20], input feedforward passivity is studied in connection with output synchronization. Nonlinear inputaffine passive agents are considered in [3], [27], [35], [39], [41] while general nonlinear passive agents are studied in [9], [16], [36]. ...
... The robust H ∞ almost state disturbance decoupling problem with bounded input via static output feedback is equivalent to the above with the only modification being that instead of (20) and (21), the closed-loop transfer functions ...
... ωy , T λ uy , T λ ωu and T λ uu , are given by (20), (21), (22) and (23) respectively. ...
... Further, [15] develops a squared-down passivity to establish state synchronization result for MAS with non-square passive or passifiable agent. Nonlinear input-affine passive agents are considered in [16,17,18,19,20] while general nonlinear passive agents are studied in [21,22,23,24]. Passivity based on complex dynamic network (or neural network) is developed in [25,26]. ...
... Remark 2. When 1 = 2 = and = , squared-down passifiable via static input feedforward is reduced to static input feedforward passivity as used in [24]. ...
Article
Full-text available
This paper studies state synchronization of homogeneous multiagent systems (MAS) via a static protocol with partial‐state coupling in the presence of a time‐varying communication topology, which includes general time‐varying graphs as well as switching graphs. If the agents are squared‐down passive or squared‐down passifiable (via static output feedback or static input feedforward), then a static protocol can be designed for balanced, time‐varying graphs. Moreover, this static protocol works for arbitrary switching directed graphs if the agents are squared‐down minimum phase with relative degree one. The static protocol is designed for each agent such that state synchronization is achieved without requiring exact knowledge about the time‐varying network.
... For example, [36] and [6] considered linear agents which are either passive or passifiable via state/output feedback while [5] dealt with agents which are strictly G-passifiable via output feedback. In [11,22,21], input feed forward passivity was considered in connection with output synchronization. Nonlinear input-affine passive agents were considered in [3,29,37,41,43] while general nonlinear passive agents were studied in [10,18,38]. ...
Article
Full-text available
This paper studies H∞ and H2 almost state or output synchronization of homogeneous multi-agent systems (MAS) with partial-state coupling via static protocols in the presence of external disturbances. We provide solvability conditions for designing static protocols. We characterize three classes of agents for which we can design linear static protocols for state or output synchronization of a MAS such that the impact of disturbances on the network disagreement dynamics, expressed in terms of the H∞ and H2 norms of the corresponding closed-loop transfer function, is reduced to arbitrarily small value. Meanwhile, the static protocol only needs rough information on the network graph, that is a lower bound for the real part and an upper bound for the modulus of the non-zero eigenvalues of the Laplacian matrix associated with the network graph. Our study focuses on three classes of agents which are squared-down passive, squared-down passifiable via output feedback and squared-down minimum-phase with relative degree 1.
... The proofs of Theorem 7 and other results of this section are omitted due to the page limit and can be found in (Proskurnikov and Mazo Jr, 2017). In the case of α j = 0 the inequalities (12) hold for any matrix A, and Theorem 7 coincides with Theorem 8.3 in (Hatanaka et al., 2015). ...
Article
Full-text available
Synchronization among autonomous agents via local interactions is one of the benchmark problems in multi-agent control. Whereas synchronization algorithms for identical agents have been thoroughly studied, synchronization of heterogeneous networks still remains a challenging problem. The existing algorithms primarily use the internal model principle, assigning to each agent a local copy of some dynamical system (internal model). Synchronization of heterogeneous agents thus reduces to global synchronization of identical generators and local synchronization between the agents and their internal models. The internal model approach imposes a number of restrictions and leads to sophisticated dynamical (and, in general, nonlinear) controllers. At the same time, passive heterogeneous agents can be synchronized by a very simple linear protocol, which is used for consensus of first-order integrators. A natural question arises whether analogous algorithms are applicable to synchronization of agents that do not satisfy the passivity condition. In this paper, we study the synchronization problem for heterogeneous agents that are not passive but satisfy a weaker input feedforward passivity (IFP) condition. We show that such agents can also be synchronized by a simple linear protocol, provided that the interaction graph is strongly connected and the couplings are sufficiently weak. We demonstrate how stability of cooperative adaptive cruise control algorithms and some microscopic traffic flow models reduce to synchronization of heterogeneous IFP agents.
Article
This paper studies semiglobal and global state synchronization of homogeneous multiagent systems with partial-state coupling (ie, agents are coupled through part of their states) via a static protocol. We consider 2 classes of agents, ie, G-passive and G-passifiable via input feedforward, which are subjected to input saturation. The proposed static protocol is purely decentralized, ie, without an additional channel for the exchange of controller states. For semiglobal synchronization, a static protocol is designed for an a priori given set of network graphs with a directed spanning tree. In other words, the static protocol only needs rough information on the network graph, ie, a lower bound for the real part and an upper bound for the modulus, of the nonzero eigenvalues of the corresponding Laplacian matrix. Whereas for global synchronization, only strongly connected and detailed balanced network graphs are considered. In this case, for G-passive agents, the static protocol does not need any network information, whereas for G-passifiable agents via input feedforward, the static protocol only needs an upper bound for the modulus of the eigenvalues of the corresponding Laplacian matrix.
Conference Paper
Full-text available
The mechanisms of regular cooperative behavior in multi-agent networks, such as consensus and synchronization, have been thoroughly studied. However, many natural and engineered networks do not synchronize, exhibiting persistent disagreement or clustering. One of the reasons for this “irregular” behavior is competition among some pairs of agents. Whereas cooperative interactions are usually represented by attractive couplings, bringing the trajectories closer, competition between two agents is naturally modeled by a repulsive coupling, leading to disagreement among the agents and preventing their trajectories from convergence. Such couplings may e.g. describe interactions of antagonistic individuals in social groups, competing economic agents and repelling particles. Networks where agents can both cooperate and compete are said to be coopetitive. To study the dynamics of general coopetitive networks, in particular, mechanisms of agents' splitting into several clusters, remains a challenging problem. A simple yet insightful model of polarization under coopetitive interactions was proposed in [1], [2]. These papers address consensus-type dynamics over signed graphs, where arcs of positive and negative weights correspond, respectively, to cooperative and competitive couplings between the agents. If the graph is structurally balanced, these protocols lead to either consensus or “bipartite consensus” (polarization): the agents split into two competing “camps”, and the values (opinions) of agents from different camps agree in modulus but differ in signs. The results from [1], [2] are limited to single integrator agents, interacting over static signed graphs. In this paper, we extend these results to time-varying graphs and nonlinear heterogeneous agents that satisfy a relaxed passivity condition.
Chapter
Full-text available
This chapter addresses consensus under bounded delays for nonlinearly coupled multi-agent networks, where the agents have the single-integrator dynamics. The network topology is time-varying, and the couplings are uncertain and satisfy a conventional sector condition with known sector slopes. The delays are uncertain, time-varying and obey known upper bounds. Our goal is to estimate the margin for the delay, under which the consensus is established. Most existing results in this direction apply only to linear networks and lead to high-dimensional systems of LMI . We confine ourselves to undirected networks which satisfy symmetry condition, resembling the Newton’s Third Law. Explicit analytical conditions for the delay robust consensus over such networks are offered that employ only the known upper bounds for the delays and the sector slopes. Applications to microscopic traffic flow models with delayed drivers’ reactions are also discussed.
Article
Full-text available
This paper studies asymptotic output synchronization for a class of dynamical networks with identical nonlinear nodes and switching topology. The node dynamics are characterized by a quadratic form of incremental-dissipativity. The output synchronization problem of the switched network is first converted into the set stability problem for the interconnected nonlinear system with a particular selection of input-output pair, which preserves dissipativity. Then, synchronization under arbitrary switching among self-synchronizing subnetworks and synchronization by design of switching among subnetworks, where none of them is necessarily self-synchronizing, are investigated by using the common Lyapunov function method and the single Lyapunov function method, respectively. Algebraic synchronization criteria for both cases are established, and the obtained results are applied to the investigation of coupled biochemical oscillators and coupled Chua's circuits.
Book
Highlighting the control of networked robotic systems, this book synthesizes a unified passivity-based approach to an emerging cross-disciplinary subject. Thanks to this unified approach, readers can access various state-of-the-art research fields by studying only the background foundations associated with passivity. In addition to the theoretical results and techniques,the authors provide experimental case studies on testbeds of robotic systems including networked haptic devices, visual robotic systems, robotic network systems and visual sensor network systems.The text begins with an introduction to passivity and passivity-based control together with the other foundations needed in this book. The main body of the book consists of three parts. The first examines how passivity can be utilized for bilateral teleoperation and demonstrates the inherent robustness of the passivity-based controller against communication delays. The second part emphasizes passivitys usefulness for visual feedback control and estimation. Convergence is rigorously proved even when other passive components are interconnected. The passivity approach is also differentiated from other methodologies. The third part presents the unified passivity-based control-design methodology for multi-agent systems. This scheme is shown to be either immediately applicable or easily extendable to the solution of various motion coordination problems including 3-D attitude/pose synchronization, flocking control and cooperative motion estimation.Academic researchers and practitioners working in systems and control and/or robotics will appreciate the potential of the elegant and novel approach to the control of networked robots presented here. The limited background required and the case-study work described also make the text appropriate for and, it is hoped, inspiring to students.
Article
In this paper, we discuss consensus problems for networks of dynamic agents with fixed and switching topologies. We analyze three cases: 1) directed networks with fixed topology; 2) directed networks with switching topology; and 3) undirected networks with communication time-delays and fixed topology. We introduce two consensus protocols for networks with and without time-delays and provide a convergence analysis in all three cases. We establish a direct connection between the algebraic connectivity (or Fiedler eigenvalue) of the network and the performance (or negotiation speed) of a linear consensus protocol. This required the generalization of the notion of algebraic connectivity of undirected graphs to digraphs. It turns out that balanced digraphs play a key role in addressing average-consensus problems. We introduce disagreement functions for convergence analysis of consensus protocols. A disagreement function is a Lyapunov function for the disagreement network dynamics. We proposed a simple disagreement function that is a common Lyapunov function for the disagreement dynamics of a directed network with switching topology. A distinctive feature of this work is to address consensus problems for networks with directed information flow. We provide analytical tools that rely on algebraic graph theory, matrix theory, and control theory. Simulations are provided that demonstrate the effectiveness of our theoretical results.
Article
This book presents a unified frequency-domain method for the analysis of distributed control systems. The following important topics are discussed by using the proposed frequency-domain method: (1) Scalable stability criteria of networks of distributed control systems; (2) Effect of heterogeneous delays on the stability of a network of distributed control system; (3) Stability of Internet congestion control algorithms; and (4) Consensus in multi-agent systems. This book is ideal for graduate students in control, networking and robotics, as well as researchers in the fields of control theory and networking who are interested in learning and applying distributed control algorithms or frequency-domain analysis methods.
Article
We consider the problem of synchronizing a group of N identical oscillators, coupled by a symmetric network that is modelled by a multiple-input/multiple-output dynamical system. We provide results that can be used to establish asymptotic synchronization of a given system and also to construct identical feedback oscillators for which synchronization is guaranteed. These results are based on a new notion of passivity with respect to manifolds defined in the input and output spaces of a dynamical system. The problem under consideration is motivated by the design of control algorithms for the parallel operation of power inverters. Simulation results have shown that synchronization could be achieved in less than two AC cycles, depending on the electrical interconnection network.