ArticlePDF Available

Network-Decentralized Control Strategies for Stabilization

Authors:

Abstract and Figures

We consider the problem of stabilizing a class of systems formed by a set of decoupled subsystems (nodes) interconnected through a set of controllers (arcs). Controllers are network–decentralized, i.e. they use information exclusively from the nodes they interconnect. This condition requires a block–structured feedback matrix, having the same structure as the transpose of the overall input matrix of the system. If the subsystems do not have common unstable eigenvalues, we demonstrate that the problem is solvable. In the general case, we provide sufficient conditions for solvability. When subsystems are identical and each input agent controls a pair of subsystems with input matrices having opposite sign (flow networks), we prove that stabilization is possible if and only if the system is connected with the external environment. Our proofs are constructive and lead to structured Linear Matrix Inequalities (LMIs).
Content may be subject to copyright.
Network–Decentralized Control Strategies for Stabilization
Franco Blanchini a, Elisa Franco band Giulia Giordano a
Abstract We consider the problem of stabilizing a class
of systems formed by a set of decoupled subsystems (nodes)
interconnected through a set of controllers (arcs). Controllers
are network–decentralized, i.e. they use information exclusively
from the nodes they interconnect. This condition requires a
block–structured feedback matrix, having the same structure
as the transpose of the overall input matrix of the system. If
the subsystems do not have common unstable eigenvalues, we
demonstrate that the problem is solvable. In the general case, we
provide sufficient conditions for solvability. When subsystems
are identical and each input agent controls a pair of subsystems
with input matrices having opposite sign (flow networks), we
prove that stabilization is possible if and only if the system
is connected with the external environment. Our proofs are
constructive and lead to structured Linear Matrix Inequalities
(LMIs).
Index TermsLinear matrix inequalities (LMIs), network analysis
and control, network decentralized control, state feedback.
I. INTRO DUC TIO N AND MOT IVATI ON
Control and coordination of independent units is rele-
vant in many applications, such as platoons of autonomous
vehicles [13], [16], [17], large data communication net-
works [22], [21], [14], [18], [19], inventory management
and production–distribution systems [5], [6], [7], [9], [10],
[27], [28] and network flows in general [5], [2], [25]. These
systems can be viewed as complex systems composed by
naturally independent subunits that interact through designed
control actions. In these networked control systems it is
often too expensive or physically impossible to implement a
centralized controller deciding an optimal strategy based on
information about all the subsystems. Therefore, controllers
have to be computed based on information about a limited
subset of agents/components. Literature on the topic of
decentralized networked control has flourished in the past
decades, yielding a variety of approaches to stabilize [13],
coordinate [12], or synchronize [24], [26] large sets of
systems using locally computed controllers.
In a wide class of applications, the same controller may
affect simultaneously several subsystems in the network. For
instance, in water distribution networks [20], [4] the flow
controlled in a pipe affects the upstream and the downstream
reservoirs simultaneously; in transportation networks, traffic
control in a communication route affects at once the density
of vehicles at both extremities of the route [3], [23]. If
we associate a graph with this kind of networked systems,
controllers are associated with the arcs connecting the nodes
aDipartimento di Matematica e Informatica, Universit`
a degli
Studi di Udine, 33100 Udine, Italy. blanchini@uniud.it,
giulia.giordano@uniud.it
bDepartment of Mechanical Engineering, University of California at
Riverside, Riverside, CA 92521. efranco@engr.ucr.edu
(dynamically independent subunits). The design and synthe-
sis of this type of controllers has been pioneered in [18],
[17], [19]; more recent work is due to [9], [4], although
essentially limited to the case in which subsystems are first–
order integrators.
In this paper we consider the case in which the nodes
are arbitrary subsystems with their own, possibly unstable,
dynamics. Under stabilizability assumptions, we seek lin-
ear network–decentralized [18], [17], [19], [9], [4] state–
feedback controllers in which each control agent (arc) can
use information only from the subsystems (nodes) it con-
nects. This is equivalent to imposing that the feedback
matrix has the same structure as the transpose of the input
matrix. This type of control is intrinsically different from
decentralized control frameworks where several naturally
interacting subsystems are equipped with their own local
controller [29], since we consider control agents which are
associated not with subsystems, but with flow arcs.
Our main results are:
if the subsystems do not have common unstable eigen-
values, we show that the problem is solvable;
in the case of (possibly) common eigenvalues, we dis-
cuss general structural sufficient conditions, including a
constrained LMI [8];
in the case of a single common eigenvalue (typical in
distribution systems), the problem is solvable if and only
if the LMI is feasible;
in the special case in which all subsystems are equal,
each control agent regulates at most two nodes, and
the input matrices in these nodes have opposite sign
(typical in flow and platoon problems), we prove that
a necessary and sufficient condition for solvability is
that the system is suitably connected with the external
environment.
A. Motivations
In most of the literature on network decentralized dynamic
flow, nodes are buffers modelled by simple integrators [17],
[19], [18]. Some exceptions are first–order node dynam-
ics [18], [9] and systems with a Laplacian state matrix [4].
The general equation for the class of buffer systems is
˙x(t) = Bu(t) + Ed(t)(1)
where Bis the flow matrix, uis the controlled flow and dis
an external signal. Yet, in many cases, the nodes have some
local processing dynamics which are more complex and have
to be taken into account.
Example 1: Consider the model of a water distribution
system, shown in Fig. 1, where each node (circled) represents
© 2014 IEEE. Personal use of this material is permitted. Permission from IEEE must be obtained for all other uses, in any current or future media, including
reprinting/republishing this material for advertising or promotional purposes, creating new collective works, for resale or redistribution to servers or lists, or reuse of
any copyrighted component of this work in other works.
Publisher version DOI: 10.1109/TAC.2014.2331415
Fig. 1: The water distribution network considered in Sections I-A and V
a subsystem with its internal dynamics. Precisely, each node
includes two reservoirs, where water exchange depends on
their relative levels. Different subsystems are connected by
pipes whose flow ucan be controlled. The network has a
constant demand vector d. In Fig. 1, supplementary integra-
tors are added to some reservoirs, so that they asymptotically
achieve exact desired levels. Indeed in previous work [9], [4]
it has been shown that, for systems described by (1), zero
steady state error cannot be assured using static continuous
controllers; yet, discontinuous controllers may not be appli-
cable in flow networks. In this example, zero steady state
error can be guaranteed for all the reservoirs equipped with
a supplementary integrator, for any demand vector d.
II. DEC ENTRALIZED CONTROL O F NETW ORK S:
PROB LEM FO RMU LATIO N
We consider a class of linear, interconnected systems:
˙xi(t) = Aixi(t) + X
j∈Ci
Bij uj(t) + Eid(t)
where xi(t)Rniis the state of the i–th subsystem; Ci
is the set that indexes the control subvectors ujRmj,
j= 1, . . . , M , named agents, affecting the i–th subsystem;
Bij represents the effect of control ujon the i–th subsystem;
dis an external signal affecting the i–th subsystem through
matrix Ei. The overall system can be written as
˙x(t) = Ax(t) + Bu(t) + Ed(t)(2)
where x(t)Rnincludes the state variables associated with
each subsystem, u(t)Rmis the control vector, d(t)Rn
is the vector representing an external, non–controllable signal
affecting the system, Eis a generic matrix, while Aand B
are block–structured: ARn×nis a block–diagonal matrix
A=blockdiag{A1, A2, A3, . . . , AN}(3)
while matrix BRn×mis a suitably structured matrix.
Assumption 1: (A, B)is stabilizable.
System (2) can be naturally represented with a hyper-
graph, where the Nsubsystems are associated with nodes
and control agents are associated with hyperarcs. In the
following, for simplicity, hypergraphs and hyperarcs will
be referred to as graphs and arcs. Each control component
uj,j= 1, . . . , M is a vector in Rmjassociated with a
block column of B. Such a block column has zero blocks
Bij Rni×mjcorresponding to all the nodes not directly
affected by agent uj: formally, Bij = 0 if and only if j6∈ Ci.
Denoting by Njthe set that indexes the nodes affected by
agent j, we also have Bij = 0 if and only if i6∈ Nj. All
the block dimensions must be compatible with the block
structure of A, namely PN
i=1 ni=nand PM
i=1 mi=m.
Example 2: For illustrative purposes, let us consider a
system with 4nodes and 6agents, where
A=blockdiag{A1, A2, A3, A4},
B=
B11 B12 0 0 B15 0
0B22 B23 B24 0 0
000B34 0B36
0 0 B43 B44 B45 B46
,
E=blockdiag{0,0,I, I}.
We have C1={1,2,5},C2={2,3,4},C3={4,6}and
C4={3,4,5,6}. The agents control the following nodes:
N1={1},N2={1,2},N3={2,4},N4={2,3,4},
N5={1,4},N6={3,4}. The graph corresponding to B
(and E) is shown in Fig. 2.
1 2
34
Fig. 2: The graph corresponding to Example 2: pink squares on solid arcs
indicate controllers, dashed arcs represent external, non–controllable signals
We consider controls restricted to the class:
uj=φ(xi, i ∈ Nj).
Thus each agent ujcan have information from nodes in Nj
only. In the case of linear feedback, we can give the following
equivalent definition.
Definition 1: A control of the form u=Kx is decen-
tralized in the sense of networks if Khas the same structural
zero blocks as B.
In Example 2, Khas the following structure:
K=
K
11 K
12 0 0 K
15 0
0K
22 K
23 K
24 0 0
000K
34 0K
36
0 0 K
43 K
44 K
45 K
46
.
III. CAS E OF DI S TI N CT UN STABLE EI GEN VALU E S
In this section we show that under the following assump-
tion, which is a generic property, decentralized stabilizability
is always possible. We refer to all the eigenvalues whose real
part is not strictly negative as “unstable eigenvalues”.
Assumption 2: Two different subsystems do not share
unstable eigenvalues.
Definition 2: The system is node–stabilizable if any sub-
system ican be stabilized using the control inputs in Ci,
namely (Ai,[Bi1Bi2. . . BiM ]) are stabilizable i.
In Example 2 we would have that (A1,[B11 B12 B15]),
(A2,[B22 B23 B24]),(A3,[B34 B36 ]) and
(A4,[B43 B44 B45 B46]) are stabilizable.
We now provide two preliminary results, before we state
our main findings.
Claim 1: Given any state–input matrix pair (F, G), there
exists a Kalman–like transformation (KL-transformation for
short) such that
T1F T =S R
0U, T 1G=V
0,
where (S, V )is a stabilizable pair and Ucontains only
unreachable unstable eigenvalues.
Claim 2: Consider any system of the form
F=blockdiag{F1, . . . , Fr}, G = [G
1. . . G
r],
where Gihave the same number of columns and Fi,Fjdo
not share unstable eigenvalues for i6=j. Then, if (Fi, Gi)
are stabilizable pairs, the system is stabilizable. In particular,
node–stabilizability and stabilizability are equivalent.
The proof of Claim 2 follows from the Popov criterion:
(F, G)is stabilizable iff rank [λI F|G] = nfor all
unstable eigenvalues. Consider an unstable eigenvalue λ, say
of the first block F1, and let ˜
F1=blockdiag{F2, . . . , Fr},
˜
G1= [G
2. . . G
r]. We must have
rank λI F10G1
0λI ˜
F1˜
G1=n
The condition is true, since [λI ˜
F1]has full rank because
λis an eigenvalue of F1only and [λI F1|G1]has full rank
in view of the stabilizability of (F1, G1).
Theorem 1: Under Assumption 2, the following condi-
tions are equivalent:
the system is stabilizable;
the system is node–stabilizable;
the system can be stabilized by means of a network–
decentralized control.
Proof: By Claim 2, if the subsystems do not share
unstable eigenvalues, node–stabilizability is equivalent to
stabilizability. If the system is decentralized–stabilizable,
then it is stabilizable. Therefore, we just need to show that
(node) stabilizability implies decentralized stabilizability.
Assume that the system is node–stabilizable. We consider
the first input u1and we rewrite the system so that the p
subsystems in N1are each KL-transformed (with respect to
input u1) and grouped in the first columns of A and B, as
below:
[A||B] =
S1R1. . . 0 0 0 V1X
0U1. . . 0 0 0 0 X
.
.
..
.
.....
.
..
.
..
.
..
.
..
.
.
0 0 ...SpRp0VpX
0 0 . . . 0Up0 0 X
0 0 . . . 0 0 Λ 0 ˜
B
Matrix Λcontains the subsystems not in N1and Xare
entries we can neglect. By rearranging the blocks we get:
[A||B] =
S1. . . 0R1. . . 0 0 V1X
.
.
.....
.
..
.
.....
.
..
.
..
.
..
.
.
0. . . Sp0. . . Rp0VpX
0. . . 0U1. . . 0 0 0 ˜
B1
.
.
.....
.
..
.
.....
.
..
.
..
.
..
.
.
0. . . 00. . . Up0 0 ˜
Bp
0. . . 0 0 . . . 0 Λ 0 ˜
B
Again by Claim 2, we can stabilize the blocks S1. . . Spby
means of the first input: we feed back the substates associated
with these blocks, xS
1(t), . . . , xS
p(t), so that
S1. . . 0
.
.
.....
.
.
0. . . Sp
+
V1
.
.
.
Vp
˜
K1. . . ˜
Kp= Φ1
is stable. We achieve the following block–triangular form
[ˆ
A||B] = Φ1XΓ11 Γ12
¯
0Φ20 Γ22 .(4)
We will not feed back the substates xS
1(t), . . . , xS
p(t)any-
more, so that a) the term denoted as ¯
0in (4) is preserved,
b) Φ1remains untouched. The procedure is iterated by
considering the remaining part:
2||Γ22] =
U1. . . 0 0 ˜
B1
.
.
.....
.
..
.
..
.
.
0. . . Up0˜
Bp
0. . . 0 Λ ˜
B
Note that this subsystem
includes the inputs we still have to exploit, u2, . . . , uM;
has a block–diagonal structure and meets Assumption 2;
is node–stabilizable, hence stabilizable.
Therefore the problem of its stabilization is exactly as the
one we started with. By exploiting u2and feeding back
components of the state which are not among those of
xS
1(t), . . . , xS
p(t), we reach a new triangular form and so
on. At each step, we deal with the last part k||Γkk ]of a
system of the form
[ˆ
A|| ˆ
B]k=
Φ1X . . . X Γ11 Γ12 . . . Γ1k
0 Φ2. . . X 0 Γ22 . . . Γ2k
.
.
..
.
.....
.
..
.
..
.
.....
.
.
0 0 . . . Φk0 0 . . . Γkk
The assumed node–stabilizability assures that the proce-
dure terminates successfully, because the unstable modes of
the residual system k||Γkk ]are unreachable by the inputs
u1...uk1, therefore they can necessarily be stabilized by
some of the remaining agents.
The procedure provides a control which might take advantage
of only a subset of the control agents. If this is an issue,
we can fully exploit the available arcs: we find a structured
feedback u=¯
Kx, we solve the Lyapunov equation to find
aPfor the closed–loop system, and finally we derive a more
suitable Kwith this Pby solving
min kKk2: (ABK)P+P(ABK)<0, K ∈ S(B)
This convex optimization problem is always feasible and the
obtained control exploits all available links.
IV. CAS E O F SH ARE D UNS TAB LE EI GEN VALU ES
To consider the case of common unstable eigenvalues,
we first provide a general sufficient condition in terms of a
structured LMI [8]. We will show that this condition becomes
sufficient and necessary under additional assumptions.
Proposition 1: Consider system (2), with Ablock–
diagonal and Bblock–structured. If the following LMI
SA+AS 2γBB<0(5)
has a solution S > 0in the form
S=P1=blockdiag{P1
1, P 1
2, . . . , P 1
N},(6)
with Pkof the same dimensions of Ak, then the problem
admits a network–decentralized stabilizing feedback control.
Proof: The LMI is solvable if and only if
(AγBBP)P+P(AγBBP)<0(7)
with P=S1, [11]. The network–decentralized control
K=γBPassures
(ABK)P+P(ABK)<0.(8)
The LMI condition is sufficient, but not necessary, to guar-
antee network decentralized stabilizability [8].
Solvability conditions can be provided under additional
assumptions. The first assumption is motivated by Example
1, in which the subsystems share a single unstable eigenvalue
(λ= 0). In this case, if the system is stabilizable then
the LMI is feasible, hence the problem of decentralized
stabilization can be solved.
Proposition 2: Assume that all the matrices Aihave a
single unstable eigenvalue λ0of ascent 1(i.e. the largest
Jordan block associated with λhas dimension 1). Then the
following conditions are equivalent:
the system is stabilizable;
the system can be stabilized by means of a network–
decentralized control;
the LMI (5) has a structured solution (6).
Proof: We prove that stabilizability implies that the
structured LMI is solvable. The remaining proofs are trivial.
We apply to the blocks Akseparate transformations Tk, so
that T1
kAkTk=blockdiag{λIk,ˆ
AS
k}, where the stable part
ˆ
AS
khas the identity Ias Lyapunov matrix:
(ˆ
AS
k)+ ( ˆ
AS
k) = Qk<0
By rearranging all the blocks and joining all λIk, we can put
the system in the form
ˆ
A=λI 0
0ˆ
ΛS,ˆ
B=ˆ
Bλ
ˆ
BS
where ˆ
ΛS=blockdiag{ˆ
AS
1,ˆ
AS
2, . . . , ˆ
AS
N}is stable.
If the system is stabilizable, then ˆ
Bλhas full row rank, as
it can be immediately seen by means of the Popov criterion.
We consider the candidate block–diagonal matrix
ˆ
S=I0
0µI ,
where µ > 0has to be decided, and the feedback
u=γ[ˆ
B
λ0]x=ˆ
Kx,
which is network–decentralized. Then
ˆ
S(ˆ
Aˆ
Bˆ
K)+ ( ˆ
Aˆ
Bˆ
K)ˆ
S=2(λI γˆ
Bλˆ
B
λ)γˆ
Bλˆ
B
S
γˆ
BSˆ
B
λ
µˆ
Q,
where ˆ
Q=blockdiag{Q1, Q2, . . . , QN}. Since ˆ
Bλˆ
B
λ>0,
for γlarge enough the block 2(λI γˆ
Bλˆ
B
λ)is negative
definite. Let us fix such a γ. Since ˆ
Q > 0, we can
subsequently take µlarge enough to assure that
ˆ
S(ˆ
Aˆ
Bˆ
K)+ ( ˆ
Aˆ
Bˆ
K)ˆ
S < 0.(9)
By using the backward transformations, we restore all the
blocks to the original position and thus we find a structured S
as desired. Then we take P=S1, which is also structured,
to get (7). Thus (5) is satisfied with S > 0structured.
To further investigate the problem, we introduce the follow-
ing definitions.
Definition 3: The network is locally stabilizable if each
agent uican stabilize each of the subsystems in Ni.
Note that this by no means implies that the agent can stabilize
simultaneously more than one subsystem in Ni.
Definition 4: The system is structurally triangularizable if
there exist a) an ordering of the nodes and b) a selection and
ordering of the agents such that the resulting Bhas a block
triangular structure.
For instance, the system in Example 2 is structurally triangu-
larizable by ordering the nodes as 1,2,4,3and disregarding
the last two agents (i.e. selecting the first four). To present
our next result, we need to consider the case in which some
of the subsystems are open–loop stable.
Definition 5: Given the structured system (A, B), we de-
fine the extended system (A, Bext)as follows. For each node
iwhich is asymptotically stable, Bis extended by adding a
fictitious block with nicolumns, which has an identity matrix
corresponding to Aiand zero blocks elsewhere.
For instance, if in the system of Example 2 the second
subsystem is asymptotically stable, we have to extend Bas
Bext := B|[0I0 0 ]
It is understood that this is a fictitious system which cannot
be considered in practice. The following theorem holds.
Theorem 2: If the extended system is triangularizable and
locally stabilizable, then (5)–(6) are feasible.
Proof: If the system can be triangularized, we may
assume Bext in the form
Bext =¯
B˜
B=
¯
B11 X . . . X ˜
B1
0¯
B22 . . . X ˜
B2
.
.
..
.
.....
.
..
.
.
0 0 . . . ¯
BNN ˜
BN
where X are elements we can neglect. By assumption, i,
either (Ai,¯
Bii)is stabilizable or Aiis already asymptotically
stable (then ¯
Bii =I, a fictitious column). Therefore, by
adopting the control
K=¯
K
0,¯
K=blockdiag{K1, K2. . . , KN},
where Kiis such that Ai¯
BiiKiis asymptotically stable, the
resulting closed loop matrix is block triangular. Any asymp-
totically stable block–triangular matrix admits a Lyapunov
matrix which is block diagonal and satisfies (8); this means
that (5)–(6) are feasible.
Theorem 2 has demanding assumptions. However, there
are interesting structural assumptions under which LMI
solvability is guaranteed, such as when each control agent is
associated with an arc of a proper graph (not a hypergraph),
thus affects at most two subsystems.
Definition 6: A network is connected if the corresponding
graph is connected; it is connected with the external envi-
ronment if, in addition, the input matrix Bhas at least one
block–column with a single non–zero block.
Corollary 1: Assume that the system is locally stabiliz-
able and each agent controls at most two subsystems. If
the system is connected with the external environment, then
(5)–(6) are feasible, hence a decentralized stabilizing control
exists.
Proof: It is immediate, because, given a node connected
to the external environment, we can find a spanning tree and
form the triangular structure of Bstarting from this node.
An interesting case is that of flow networks, in which we
have a family of identical subsystems and each network link
connects a pair of nodes in such a way that its action has
opposite effects (inflow and outflow).
Proposition 3: Given a connected network, assume that
all the diagonal blocks of matrix Aare equal, Ai=A1, for
i= 1, . . . , N , and that all non–zero blocks of Bare ±B1.
Assume that there are at most two non–zero blocks in each
block column of matrix Band that, if they are two, they have
opposite sign. Then the following conditions are equivalent:
the system is stabilizable;
the system is locally stabilizable;
the structured LMI is solvable;
either A1is stable or the network is connected with the
external environment.
Proof: Ignoring the trivial case of A1stable, we show
that stabilizability implies both external connection and local
stabilizability; the rest of the proof follows from Corollary 1.
According to the Popov criterion, the system is stabilizable
iff, for all unstable eigenvalues λ,z[λI A|B] = 0 implies
z= 0. Consider now the vector z= [z
1z
1. . . z
1],
where z16= 0 is a left eigenvector associated with A1, hence
z[λI A] = 0. By contradiction, assume the network is not
connected with the external environment. Then, each non–
zero block column would have exactly two blocks of opposite
sign, so we would also have zB= 0 and the system would
not be stabilizable.
Local stabilizability can be proved in a similar way.
Indeed, if the subsystems were not locally stabilizable, there
would be an unstable eigenvalue λsuch that z
1[λI
A1|B1] = 0,z16= 0, but defining again z=
[z
1z
1. . . z
1]6= 0 we would have z[λI A|B] = 0.
V. EX AMP LE
We revisit Example 1 and we assume that the nodes in
Fig. 1 have the following dynamics:
Ai="αiβi0
αiβi0
0 1 0#where βi= 0 for i∈ {1,2,4}.
In each node, the first two states are the reservoirs volumes,
while the third represents the local integrator. Constants αi
[min1] and βi[min1] depend on the size of the reservoirs
and the diameter of the connecting pipes; α1= 15,α2= 20,
α3= 16,α4= 16.7,α5= 14,β3= 12 and β5= 22.
The overall system has the same structure as model (2),
where A=blockdiag{A1, A2, A3, A4, A5},
B=
BuBd0 0 0 0
0BuBd0 0 Bd
0 0 BdBu0 0
0 0 0 BdBu0
0 0 0 0 BuBd
,
Bd=010, Bu=100,
d=[0 1 0 0 1 0 0 1 0 0 1 0 0 1 0]and E=I. Since the
only unstable eigenvalue is λ= 0, which has ascent one and
is common to all the subsystems, we can apply the results
of Proposition 2 to regulate the system with a decentralized
control; Kis obtained through the direct solution of the
LMI (5)–(6), numerically solved using the MATLAB LMI
toolbox [15]. To enforce a certain speed of convergence, in
the LMI Acan be replaced with A+σI,σ > 0, so that
the closed loop eigenvalues have real part less than σ. We
have used σ= 0.15.
In Fig. 3 and 4 the decentralized control is com-
pared with a centralized LQ control, with state weight-
ing matrix Iand input weighting matrix I/2;x(0) =
[8.80 3.63 9.15 8.57 0.43 8.06 6.36 6.35 4.44 7.00 3.19 0.37
9.45 2.97 6.00]. As expected, the equilibrium values of the
states equipped with integral control are smoothly recovered.
VI. CO N CL U DI NG RE MAR KS
We have proved that the problem of designing a network–
decentralized control to stabilize a system formed by a set
of independent subsystems is solvable when the subsystems
have no common unstable eigenvalues. In the general case,
we have provided a structured LMI condition which in
principle is only sufficient, but we have seen that such an
LMI is always feasible in particular cases, e.g. that of flow
0 5 10 15 20
10
5
0
5
10
Time![min]
Reservoir!volumes![m3]
Optimal!LQ!control!(Riccati)
x1x2x4x5x7x8x10 x11 x13 x14
0 5 10 15 20
10
5
0
5
10
Time![min]
Reservoir!volumes![m3]
Network!decentralized!control!(LMI)
x1x2x4x5x7x8x10 x11 x13 x14
Fig. 3: Reservoir volumes evolution: decentralized (top) and optimal (bot-
tom) control.
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
−10
−5
0
5
10
Time [min]
Reservoir volumes [m3]
Comparison between centralized (LQ) and decentralized strategy
x2 LQ x2 LMI x10 LQ x10 LMI x13 LQ x13 LMI
Fig. 4: Detailed simulation for reservoir volumes x2,x10 and x13 (zoomed
from Fig. 3).
networks and that of subsystems with a single common
unstable eigenvalue of ascent 1. Unfortunately, the general
question whether, under possibly common eigenvalues, sta-
bilization implies stabilizability in the sense of networks is
still unsolved and is left as a subject of further investigation.
Another interesting problem is how can we solve the LMI
efficiently: when BBis large and sufficiently sparse, ideas
from chordal decomposition methods [1] could be promising.
REF ER E NC E S
[1] J. Agler, W. Helton, S. McCullough, and L. Rodman, “Positive
semidefinite matrices with a given sparsity pattern,Linear Algebra
and its Applications, vol. 107, pp. 101–149, 1988.
[2] A. Atamturk and M. Zhang, “Two–stage robust network flow and
design under demand uncertainty,Operations Research, vol. 55, no.
4, pp. 662–673, 2007.
[3] B. Ataslar and A. Iftar, “A decentralized control approach for trans-
portation networks,” Proc. of the 8th IFAC Symposium. on Large Scale
Systems, pp. 348–353, 1998.
[4] D. Bauso, F. Blanchini, L. Giarr´
e, and R. Pesenti, ‘The linear saturated
control strategy for constrained flow control is asymptotically optimal,
Automatica, vol. 49, no. 7, pp. 2206–2212, 2013.
[5] D. Bauso, F. Blanchini, and R. Pesenti, “Optimization of Long-
run Average-Flow Cost in Networks with Time-Varying Unknown
Demand,” IEEE Trans. on Autom. Control, vol. 55, no. 1, pp. 20–
31, 2010.
[6] D. Bertsimas and A. Thiele, “A Robust Optimization Approach to
Inventory Theory,” Operations Research, vol. 54, no. 1, pp. 150–168,
2006.
[7] F. Blanchini, F. Rinaldi, and W. Ukovich, “Least Inventory Control of
Multi-Storage Systems with Non-Stochastic Unknown Input,” IEEE
Trans. on Robotics and Automation, vol. 13, no. 5, pp. 633–645, 1997.
[8] F. Blanchini, E. Franco, G. Giordano, “Structured–LMI Conditions for
Stabilizing Network–Decentralized Control”, 52nd IEEE Conference
on Decision and Control, pp. 6880–6885, Florence, 2013.
[9] F. Blanchini, S. Miani, and W. Ukovich, “Control of production-
distribution systems with unknown inputs and system failures,IEEE
Trans. on Autom. Control, vol. 45, no. 6, pp. 1072–1081, 2000.
[10] E. K. Boukas, H. Yang, and Q. Zhang, “Minimax production planning
in failure–prone manufacturing systems,” Journal of Optimization
Theory and Applications, vol. 87, no. 2, pp. 269–286, 1995.
[11] S. Boyd, L. El Ghaoui, E. Feron, and V. Balakrishnan, Linear matrix
inequalities in system and control theory, SIAM Studies in Applied
Mathematics, Philadelphia, 1994.
[12] J. Cortes, S. Martinez, T. Karatas, and F. Bullo. “Coverage control
for mobile sensing networks,” IEEE Transactions on Robotics and
Automation, vol. 20, no. 2, pp. 243–255, 2004.
[13] R. D’Andrea, “A Linear Matrix Inequality Approach to Decentralized
Control of Distributed Parameter System,Proc. of the American
Control Conference, Philadelphia, PA, vol. 3, pp. 1350–1354, 1998.
[14] A. Ephremides and S. Verd´
u, “Control and Optimisation Methods in
Communication Networks,” IEEE Trans. on Autom. Control, vol. 34,
no. 9, pp. 930–942, 1989.
[15] P. Gahinet, A. Nemirovski, A. J. Laub, and M. Chilali, “The LMI
control toolbox,” IEEE Conference on Decision and Control, vol. 3,
pp. 2038–2041, 1994.
[16] A. Iftar, “A decentralized routing control strategy for semi-congested
highways,” Proc. of the 13th IFAC World Congress, vol. P, pp. 319–
324, 1996.
[17] A. Iftar, “A linear programming based decentralized routing controller
for congested highways,” Automatica, vol. 35, no. 2, pp. 279–292,
1999.
[18] A. Iftar and E. J. Davison, “Decentralized Robust Control for Dynamic
Routing of Large Scale Networks,Proc. of the American Control
Conference, San Diego, CA, pp. 441–446, 1990.
[19] A. Iftar and E. J. Davison, “Decentralized control strategies for
dynamic routing,” Optimal Control Applications and Methods, vol.
23, pp. 329–355, 2002.
[20] R. E. Larson and W. G. Keckler, “Applications of Dynamic Program-
ming to the Control of Water Resource Systems,Automatica, vol. 5,
no. 1, pp. 15–26, 1969.
[21] J. C. Moreno and M. Papageorgiou, “A Linear Programming Approach
to Large-scale Linear Optimal Control Problems,” IEEE Trans. on
Autom. Control, vol. 40, no. 5, pp. 971–977, 1995.
[22] F. H. Moss and A. Segall, “An Optimal Control Approach to Dynamic
Routing in Networks,” IEEE Trans. on Autom. Control, vol. 27, no. 2,
pp. 329–339, 1982.
[23] S. Mudchanatongsuk, F. Ord´
o˜
nez, and J. Liu, “Robust Solutions for
Network Design under Transportation Cost and Demand Uncertainty,”
J. Oper. Res. Soc., vol. 59, no. 5, pp. 652–662, 2008.
[24] R. Olfati–Saber, R. M. Murray, “Consensus Problems in Networks of
Agents with Switching Topology and Time–Delays”, IEEE Trans. on
Autom. Control, vol. 49, no. 9, pp. 1520–1533, 2004.
[25] F. Ord´
o˜
nez and J. Zhao, “Robust capacity expansion of network flows,
Network, vol. 50, no. 2, pp. 136–145, 2007.
[26] W. Ren, R. W. Beard, and E. Atkins, “Information Consensus in
Multivehicle Cooperative Control: Collective Group Behavior through
Local Interaction,” IEEE Control Systems Magazine, vol. 27, no. 2,
April, pp. 71–82, 2007.
[27] H. Sarimveisa, P. Patrinos, C. D. Tarantilis, and C. T. Kiranoudis,
“Dynamic modeling and control of supply chain systems: A review,”
Computers & Operations Research, vol. 35, pp. 3530–3561, 2008.
[28] E. A. Silver and R. Peterson, Decision System for Inventory Manage-
ment and Production Planning, Wiley, New York, NY, 1985.
[29] S.H. Wang and E.J. Davison, “On the stabilization of decentralized
control systems,” IEEE Trans. on Autom. Control, vol. 18, no. 5,
pp. 473–478, 1973.
... Although centralized strategies are able to achieve the optimal system performance for any prespecified evaluation index [11], they are heavily dependent on the central authority, leading to high communication resource consumption and weak system robustness. To avoiding using the authority, various decentralized strategies have been developed in the literature [14]- [17]. Wang et al. [14] and Corfmat et al. [15] introduced a class of local control frameworks for stabilizing multi-channel systems. ...
... It is worth noting that decentralized strategies always confront structural constraints such that controllers may not exist, especially in the output-feedback control scheme. Although some structural linear matrix inequalities have been adopted to overcome this drawback [17], they are still restricted to simple application scenarios. To completely bypass the structural issues, distributed strategies are proposed by introducing communication between channels [18]- [22]. ...
... and ϵ and π i are defined below (17). Then, K i is selected as ...
Preprint
Full-text available
Cooperative behavior design for multi-agent systems with collective tasks is a critical issue to promote swarm intelligence. This paper investigates cooperative control for a multi-channel system, where each channel is managed by an agent that can communicate with neighbors in a network. Each agent is expected to self-organize a controller based only on local information and local interaction to stabilize the multi-channel system collaboratively. A novel cooperative control strategy is designed for each agent by leveraging a decomposing technique and a fusion approach. Then, a privacy-preserving mechanism is incorporated into this strategy to shield all private information from eavesdropping. Moreover, a fully distributed designing method for the strategy parameters is developed. As a result, agents can self-design and self-perform their controllers with private information preserved. It is proved that the multi-channel system stability can be ensured by the proposed strategy with finite fusion steps during each control interval. In addition, the cost of introducing the privacy-preserving mechanism and the effect of adding more channels on the system performance are quantitatively analyzed, which benefits mechanism design and channel placement. Finally, several comparative simulation examples are provided to demonstrate the effectiveness of the theoretical results.
... However, they pay a great price for communication and computation when the number of inputs increases. As a remedy, decentralized methods have been designed in [11]- [14]. Earlier, a class of local control laws, depending on partial system information, were proposed for stabilizing multi-variable systems in [11], [12]. ...
... Later, these results were extended to cases with rate constraints on collecting local information [13]. Recently, in [14], structured linear matrix inequalities were adopted to obtain solutions to the control problem of multi-input systems. It should be noted that, although the above decentralized methods can be performed based only on local information, the controller gains have to be designed by a central unit with global information. ...
... Theorem 2: Considering system (1) with inputs designed by (14), if Assumptions 1, 2 and 3 hold, for any finite positive integer M , the optimized cost function J M is upper bounded by ...
Preprint
Full-text available
In this paper, a cooperative Linear Quadratic Regulator (LQR) problem is investigated for multi-input systems, where each input is generated by an agent in a network. The input matrices are different and locally possessed by the corresponding agents respectively, which can be regarded as different ways for agents to control the multi-input system. By embedding a fully distributed information fusion strategy, a novel cooperative LQR-based controller is proposed. Each agent only needs to communicate with its neighbors, rather than sharing information globally in a network. Moreover, only the joint controllability is required, which allows the multi-input system to be uncontrollable for every single agent or even all its neighbors. In particular, only one-time information exchange is necessary at every control step, which significantly reduces the communication consumption. It is proved that the boundedness (convergence) of the controller gains is guaranteed for time-varying (time-invariant) systems. Furthermore, the control performance of the entire system is ensured. Generally, the proposed controller achieves a better trade-off between the control performance and the communication overhead, compared with the existing centralized/decentralized/consensus-based LQR controllers. Finally, the effectiveness of the theoretical results is illustrated by several comparative numerical examples.
... Dans ce cas, l'interaction entre les noeuds du réseau est linéaire et en configuration cascade, c'est-à-dire les ordres de production passent successivement du noeud en amont à celui en aval, pour approvisionner le niveau de stockage qui sert directement la demande externe. Ces deux modèles sont étudiés profondément dans les chapitres 3 et 4. Les divers travaux de Blanchini et de ses coauteurs concernant la commande par invariance de systèmes logistiques, que nous introduisons dans le chapitre 2, sont basés sur le même modèle que nous utilisons [Blanchini (1990), Blanchini et al. (2014)]. De même, Ignaciuk et Bartoszewicz ont utilisé ce modèle dans leurs travaux [Ignaciuk et Bartoszewicz (2011), Ignaciuk (2012], pour l'étude des systèmes logistiques complexes constitués de plusieurs fournisseurs. ...
Thesis
Ce travail de thèse porte sur l’étudeanalytique de la dynamique des réseaux logistiqueset des systèmes logistiques élémentaires,soumis aux retards d’acquisition desproduits, aux pertes de stockage vu la naturepérissable des produits, et aux contraintes depositivité et de saturation des ressources dusystème. La finalité est de maîtriser la productionafin de satisfaire la demande des clientsincertaine mais bornée, sans risque ni de ruptureni de débordement de stockage, et en tenantcompte des diverses spécifications notammentles capacités de production et destockage. En se basant sur les formalismesdes systèmes Entrée-Sortie et les propriétésd’invariance, un premier objectif consiste àproposer des méthodes de calcul de bornes,afin d’identifier les ensembles de sortie atteignables,du réseau logistique. Un objectif secondaire,repose sur l’étude de nouvelles méthodesde synthèse et de conception de commandesdu type feedback-prédicteur, linéairesou saturées, pour déterminer des conditionsd’existence de lois de commande satisfaisantespour le système élémentaire. Un troisièmeaspect guidant ce sujet est de prendreen compte les incertitudes sur certains paramètresintrinsèques du système ou des perturbationssupplémentaires, d’où l’introduction denouvelles techniques de commandes robusteset réactives
... Another LMI method to design stable column sparse feedback controllers is proposed in (Polyak et al., 2013). Also, Blanchini et al. (2013Blanchini et al. ( , 2015 study a set of decoupled subsystems interconnected through a block-structured feedback matrix, whose stability is guaranteed via LMIs. ...
Article
Full-text available
This work presents a method based on linear matrix inequalities (LMIs) to design a feedback controller that guarantees stability regardless of the network topology considered. The proposed controller has a modular structure, that is, it is composed of blocks associated with communication links in a control network, obtaining the feedback of each topology by inserting zeros in the blocks related to disabled links. Conditions to obtain the desired modular structure are provided, as well as a numerical example to assess the feasibility of the design method.
Article
In this paper, a cooperative Linear Quadratic Regulator (LQR) problem is investigated for multi-input systems, where each input is generated by an agent in a network. The input matrices are different and locally possessed by the corresponding agents respectively, which can be regarded as different ways for agents to control the multi-input system. By embedding a fully distributed information fusion strategy, a novel cooperative LQR-based controller is proposed. Each agent only needs to communicate with its neighbors, rather than sharing information globally in a network. Moreover, only a joint controllability is required, which allows the multi-input system to be uncontrollable for every single agent. In particular, only one-time information exchange is necessary at every control step, which significantly reduces the communication consumption. It is proved that the boundedness (convergence) of the controller gains is guaranteed for time-varying (time-invariant) systems. Furthermore, the performance of the entire system is ensured. Generally, the proposed controller achieves a better trade-off between the control performance and the communication overhead compared with existing distributed LQR-based controllers. Finally, the effectiveness of the theoretical results is illustrated by several comparative numerical examples.
Article
We proposed network-decentralized control strategies, in which each actuator can exclusively rely on local information, without knowing the network topology and the external input, ensuring that the flow asymptotically converges to the optimal one with respect to the $p$ -norm. For $1 < p < \infty $ , the flow converges to a unique constant optimal $u_{p}^{*}$ . We show that the state converges to the optimal Lagrange multiplier of the optimization problem. Then, we consider networks where the flows are affected by unknown spontaneous dynamics and the buffers need to be driven exactly to a desired set-point. We propose a network-decentralized proportional-integral controller that achieves this goal along with asymptotic flow optimality; now it is the integral variable that converges to the optimal Lagrange multiplier. The extreme cases $p=1$ and $p=\infty $ are of some interest since the former encourages sparsity of the solution while the latter promotes fairness. Unfortunately, for $p=1$ or $p=\infty $ these strategies become discontinuous and lead to chattering of the flow, hence no optimality is achieved. We then show how to approximately achieve the goal as the limit for $p \rightarrow 1$ or $p \rightarrow \infty $ .
Article
Coalitional control groups dynamically local controllers into clusters that jointly determine their control actions to maximize control performance while minimizing the cooperation burden. This work presents linear matrix inequalities decision methods to set state-space regions where the switchings between network topologies satisfy properties of interest. In particular, convexity guarantees ellipsoidal switching sets with back-switchings avoided via a dwell time. Also, the convexity property is exploited to analyze the need for coordination at different points of the state space, leading to coordination effort zones. Moreover, by considering invariance , new ellipsoids that confine the state between topologies transitions are provided. Finally, we introduce additional conditions to attain submodularity and, hence, reduce the effort to find optimal solutions. A numerical example is given to illustrate the feasibility of the proposed approach.
Conference Paper
Full-text available
In this paper we consider a set of dynamically decoupled systems interconnected by control agents. We can visualize this architecture as a graph where subsystems are associated to nodes and control agents are associated to arcs. The interconnection between subsystems is determined by their input matrix. The decisions of each control agent can directly affect only the nodes connected by the corresponding arc. We seek a stabilizing control framework in which each control agent has information only about the state components associated with the nodes it influences; we say this control architecture is decentralized in the sense of networks. This problem setup involves block-structured feedback matrices, with structural zero blocks. We provide a constructive, sufficient condition based on an LMI with block-diagonal constraints, which guarantees stabilizability through a network-decentralized state-feedback control law. We show that under some structural conditions, concerning local stabilizability and connection with the external environment, the LMI condition we provide is always feasible. Thus, the desired controller can be found in an efficient way.
Article
A decentralized routing controller for multi-destination transportation networks is proposed. The basic aim of the introduced approach is to minimize the queue lengths at the nodes. The proposed controller is decentralized in the sense that all the on-line computations are done locally at the nodes with minimum information transfer from the adjacent downstream nodes. The proposed controller clears the queues and leads each vehicle to its destination in finite time, does not over saturate the links, and avoids looping. It is believed that the proposed strategy is especially suited for "semi-congested" networks where certain nodes may become congested from time to time.
Article
A decentralized control approach for transportation networks to prevent traffic congestion is presented. The approach was developed by considering flow control and routing control together. The control algorithm is based on a previously proposed congestion measure. In this paper, in order to illustrate the performance of the proposed controller, some simulation results are presented.
Article
We consider production-distribution systems with buffer and capacity constraints in the presence of uncertain inputs. We face the problem of finding a control which drives the buffer levels to a prescribed value. We consider a continuous-time model and show that some important properties which fail in general in the discrete-time case. In particular we show that the case in which the controller knows the external input is “equivalent” to the case in which such information is not available. Furthermore, a decentralized strategy can be determined. We also show that in the case of network failures the controller does not need to know the current network configuration in order to guarantee the tracking goal as long as a certain necessary and sufficient condition is satisfied
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.
Conference Paper
This paper describes a new MATLAB-based toolbox for control design via linear matrix inequality (LMI) techniques. After a brief review of LMIs and of some of their applications to control, the toolbox contents and capabilities are presented.
Article
We present an algorithm for constrained network flow control in the presence of an unknown demand. Our algorithm is decentralized in the sense that it is implemented by a team of agents, each controlling just the flow on a single arc of the network based only on the buffer levels at the nodes at the extremes of the arc, while ignoring the actions of other agents and the network topology. We prove that our algorithm is also stabilizing and steady-state optimal. Specifically, we show that it asymptotically produces the minimum-norm flow. We finally generalize our algorithm to networks with a linear dynamics and we prove that certain least-square optimality properties still hold.
Article
A decentralized controller for dynamic routing in multidestination trasportation networks is presented. The proposed algorithm is especially suited for congested networks, where the traffic demand is close to or more than (at least for some duration of time, e.g., during a rush hour) the total capacity of the network. The basic aim of the introduced approach is to avoid queue build-up for as large external arrival rates as possible. It is shown that this problem can be formulated as a linear programming problem, and can be solved off-line. The proposed controller is decentralized in the sense that all the on-line computations are done locally at the nodes without any information transfer from the other nodes. Furthermore, no synchronization among the nodes is needed. The controller guarantees stability and clears the queues of the network in finite time assuming that the external arrival rates do not become larger than a maximum value. The controller also avoids looping.