Some Control and Observation Issues in
Cellular Automata
Samira El Yacoubi
Théo Plénet
IMAGES-ESPACE-DEV, Univ. Perpignan Via Domitia, Perpignan, France
ESPACE-DEV, Univ. Montpellier, IRD, Montpellier, France
Sara Dridi
University of Setif, Setif, Algeria
Franco Bagnoli
Dept. Physics and Astronomy and CSDC, University of Florence
via G. Sansone 1, 50019 Sesto Fiorentino (FI), Italy also INFN, sez. Firenze
Laurent Lefèvre
Clément Raïevsky
Univ. Grenoble Alpes, Grenoble INP, Institute of Engineering Univ.
Grenoble Alpes, LCIS, 26000 Valence, France
This review article focuses on studying problems of observability and
controllability of cellular automata (CAs) considered in the context of
control theory, an important feature of which is the adoption of a statespace model. Our work first consists in generalizing the obtained
results to systems described by CAs considered as the discrete counterpart of partial differential equations, and in exploring possible
approaches to prove controllability and observability. After having
introduced the notion of control and observation in cellular automata
models, in a similar way to the case of discrete-time distributed parameter systems, we investigate these key concepts of control theory in the
case of complex systems. For the controllability issue, the Boolean class
is particularly studied and applied to the regional case, while the observability is approached in the general case and related to the reconstructibility problem for linear or nonlinear CAs.
Keywords: cellular automata; control systems; controllability;
observability
1. Introduction
Controllability and observability are among the most prominent and
main considered issues in control theory. They were introduced by
Kalman and well studied during the second half of the last century. A
wide variety of works related to controllability and observability of
distributed parameters systems (DPS) has been achieved [1]. The
https://doi.org/10.25088/ComplexSystems.30.3.391
392
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
study of these notions on DPS via the structure of actuators (inputs)
and sensors (outputs) was the subject of intense research activity
[1, 2]. The traditional and most-used models in these studies are
based on a set of partial differential equations (PDEs) for the description of system input–output dynamics. Whereas controllability concerns the ability to steer the processes so as to bring them toward
desired profiles through specific actions, observability deals with the
ability to reconstruct the initial system state, taking into account a sufficient knowledge of the system dynamics based on specified output
measurements. These two major concepts have already been studied
for continuous systems described by PDEs as reported in the literature
[1, 3, 4]. In the case of deterministic linear systems analysis, the socalled Kalman condition [5, 6] is essential and has been widely used
to obtain the main characterization results regarding the choice of
actuator/sensor structures, locations, number and types (mobile or
fixed). See, for example, [1, 7] and the references therein.
The aim of this paper is to summarize some recent advances in
controllability and observability of systems described by cellular
automata (CAs), considered as the discrete counterpart of PDEs, and
explore other suitable approaches to prove controllability and observability for such systems. CAs are widely used mathematical models
for studying dynamical properties of discrete systems and constitute
very promising tools for describing complex natural systems in terms
of local interactions [8–10].
CAs are the simplest models of spatially extended systems that may
provide a good description for complex phenomena. They are discrete
dynamical systems where space and time, as well as states related to
physical quantities, are all discrete. Evolution is governed by a set of
simple local and microscopic transition rules that may exhibit a complex behavior. In the macroscopic limit (i.e., after space and time
coarse graining), CAs can reproduce the same phenomena usually
modeled with PDEs. A wide range of applications in biology,
chemistry, physics and ecology was successfully developed using the
cellular automaton (CA) paradigm as reported in the large dedicated
literature [11–13].
Since CA models were so far considered as autonomous systems,
the idea was first to introduce the notion of control and observation
in these models, in a similar way as was previously done for discretetime distributed parameter systems, in order to be able to study some
concepts of control theory related to inputs and outputs [14]. We
then started to study the two key and most popular concepts of control theory, namely controllability and observability, in the case of
complex systems modeled by CAs.
For the controllability problem, we focused on a particular case of
the so-called regional controllability. In the context of distributed
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
393
parameter systems, the term regional has been used to refer to control
problems in which the desired state is only defined, and may be reachable, on some portion of the domain. In many physical problems, the
regional controllability is naturally chosen in order to shape a natural
phenomenon just in a subregion of the whole domain. The case of
Boolean CAs has been particularly examined to investigate the boundary regional controllability [15]. It consists in considering objective
functions defined on a subregion of the domain and exerting control
actions on the boundary of the target region. This problem has been
dealt with using several tools: namely the Kalman theorem, Markov
chains and graph theory [16, 17]. The extension to nonlinear CAs has
also been studied in these works.
For the problem of observability, we assume that the studied system is autonomous and we apply the tools mentioned to prove observability as a dual notion of controllability. Where in controllability
analysis the state of the system has to be steered to a desired value
using an unknown input signal with minimal energy, in the case of
observability we seek to observe through a given output signal the
maximum energy of the unknown state of the system. The first results
were obtained for affine CAs and a rank condition for observability
was proved in [18]. This condition is reminiscent of deterministic linear systems theory. Several criteria to assess the observability and the
reconstructibility of CAs were formulated according to the choice of
sensor structures, locations and types (mobile or fixed). Some examples were given to illustrate the theoretic results. The nonlinear case
as well as the probabilistic case is currently under investigation.
2. Controllability of Cellular Automata
2.1 Introduction to Distributed Parameters Systems
Several phenomena exhibit spatially distributed behavior and are distributed parameter systems (DPS) where the state variables depend on
time and space and describe the system’s behavior in terms of inputs
and outputs that also depend on space and time.
Although DPS are more common in industry, their applications
have been expanded to include biological, ecological or economic systems and are becoming extremely related to the study of complex systems. Depending on the application, various representations of DPS
can be considered with different input and output structures. The
usual form they take is based on a state-space representation
described as a set of PDEs that provide detailed descriptions of the
internal behavior of the system.
https://doi.org/10.25088/ComplexSystems.30.3.391
394
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
The mathematical description of DPS is usually given by three operators: A describing the dynamics, B and C determining how the controls act on the system and how it is observed.
z′ (t) Az(t) + Bu(t); 0 < t < T
z0 z0 ∈ DA
(1)
y(t) Cz(t).
and augmented by the output equation:
(2)
Two key concepts for analyzing such systems are controllability
and observability, which are studied through operators B and C. The
controllability concerns the ability to steer the system from any initial
state to any desired state by acting on inputs that are involved in the
operator B. Whereas, the observability deals with the capability to
reconstruct the initial state of the system, taking into account sufficient knowledge of the system’s dynamics through certain output measurements according to the operator C.
The problem of controllability of DPS has been widely studied in
recent years [1]. Various types of controllability have been considered
for DPS: exact, weak or regional [19, 20]. The regional case was introduced by Zerrik et al. [4] as a special case of output controllability
[19, 21]. It consists in achieving an objective only on a region ω of the
spatial domain on which the governing partial differential system is
considered. This regional idea appears naturally in many real-world
dynamical systems, when studying a natural phenomenon only in a
specific area. This concept has been widely developed and interesting
results have been proved, in particular, the possibility to reach a state
on an internal subregion or on a part of the boundary of the domain
when some specific actions are exerted on the system, in its domain
interior or on its boundaries. CA models are particularly suitable for
simulating natural phenomena that are usually highly nonlinear and
better described in terms of discrete units rather than by means of
PDEs [11, 22–24]. The CA approach has been recently promoted for
the study of control problems on spatially extended systems for which
the classical approaches cannot be used. The addressed question: can
we consider CAs as a possible alternative to DPS for modeling and
analysis of spatially extended systems?
2.2 Discrete-Time Distributed Parameters Systems Statement by
Means of Cellular Automaton Formalism
CAs are spatially extended systems that are widely used for modeling
various problems ranging from physics to biology, engineering,
medicine, ecology and economics. An ultimate understanding of such
systems gives us the ability to control them in order to achieve desired
behavior.
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
A CA is
ℒ, , , f where
Definition 1.
classically
defined
395
by
a
quadruple
◼ ℒ is a d-dimensional lattice of cells c that are arranged depending on
space dimension and cell shape. In the infinite case, ℒ ℤd .
◼ denotes a discrete state set. It is a finite commutative ring given
by 0, 1, … , k - 1 in which the usual operations use modular
arithmetic.
◼ is a mapping that defines the cell’s neighborhood.
The neighborhood is usually given by:
: ℒ ⟶ℒn
c ⟶(c) {c′ ∈ ℒ c′ - ci ≤ r}
where ci , i ∈ 1, ∞ indicates the sum and the maximum, respectively, of the absolute value of the components of cell c (for d 2,
c′ - c1 c′i - ci + c′j - cj and c′ - c∞ maxc′i - ci , c′j - cj .
◼ f is a transition function that can be defined by:
f : n
⟶
st ((c)) ⟶ st+1 (c) f(st ((c))
where st (c) designates the c cell state at time
st ((c)) {st (c′ ), c′ ∈ (c)} is the state of the neighborhood.
t
and
n a x , where the sum is taken
In the linear case, f(x1 , …, xn ) ∑i1
i i
modulo k. In the affine case, f(x1 , …, xn ) ∑N
i1 ai xi + c.
We can also have cases in which the function f is linear only for a
certain set of variables. As we shall see, for control purposes, the case
of peripherally linear CAs is particularly interesting, for which
f(x1 , …, xn ) a1 x1 + h(x2 , …, xn-1 ) + an xn ,
where h is an arbitrary function (linear or nonlinear).
In order to consider CAs in the context of DPS, a description in the
form of a state equation is necessary.
Consider the case ℒ ℤd , (d ≥ 1) and introduce a metric over
d
X ℤ as:
2.2.1 New State Equation
dδ (x, y)
c∈ℤd
δ(x(c), y(c))
2c∞
https://doi.org/10.25088/ComplexSystems.30.3.391
396
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
where δ : ⨯ → 0, 1 is defined by:
δ(i, j)
0
1
if i j
if i ≠ j
The set X ℒ equipped with the distance dδ is a compact metric
space and the global dynamics F
F: X ⟶ X
s ⟶ F(s)
is continuous according to the topology induced by dδ .
Proposition 1.
[14]
◼ The compact configurations set X defines the state space of the
autonomous CA.
◼ The sequence of continuous global maps Fi defined as the ith iteration
under F plays the same role as the semi-group, usually denoted by (Φt ),
generated by the operator A.
In a way similar to discrete-time DPS, the evolution of an
autonomous CA starting from a given initial configuration s0 can be
defined in terms of the global dynamics by the state equation:
st+1 Fst
s0 ∈ X
In the linear case, the operator F is simply a circulant band matrix
J (of width n),
st+1 Jst ,
where the matrix sum-product is taken modulo k. In the affine case,
st+1 Jst + C.
2.2.2 Control and Observation in Cellular Automata
The CA model will be completed by control and measurement functions. For the control aspects, it is done via inputs (actuators) that
have a spatial structure (number, spatial location and distribution).
Let us consider the following general hypothesis.
◼ ℒ ℤd is a cellular domain whose elements c i1 , … , id .
◼ IT 0, 1, … , T is a discrete time horizon.
◼ ℒp is a subdomain that defines the region of the lattice ℒ where the CA
is excited. It contains p cells that may be connected or not.
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
397
◼ The control operator G that defines the way the control excites the CA
through the cells of ℒp is given by:
G : ⟶ ℤ
d
u ⟶ Gu
where U ℓ2 ℒp , u : ℒp ⟶
∑c∈ℒp u2 (c) < ∞ is the control
space.
G is an operator that transforms the physical actions u to a realization Gu in the state space. Its action on cell c is denoted as gc (u).
◼ The CA is then considered as a controlled system, defined by the local
transition function:
st+1 (c) f (st ((c))) + gc (ut )
where again the sum is taken modulo k. Considering for simplicity the
Boolean case, this means that where gc (u) is one, the actions of the fc
are reversed, and where gc (u) is zero, it is not modified.
◼ The corresponding state equation is:
st+1 Fst + Gut , t ∈ IT
s0
∈
X.
◼ The observation problem can be considered by duality where an obser-
vation space and a global observation operator have to be defined.
For Im 0, 1, … , Tm and ℓ2 ℒq , , the observation space
consists of all bounded measurements made in ℒq ⊂ ℒ and is given by a
measurement variable (output) denoted by θ : Im ⟶ that defines the
measurement at time t. The global observation operator H defined by:
H : ℤ ⟶
d
s
⟶ Hs
associates a measurement to each configuration s.
◼ This leads to a complete description of CAs in terms of inputs and out-
puts where the state equation is augmented with
θt Hst , t ∈ Im
and then defines the so-called distributed CA.
◼ The obtained CA statement is very close to the usual discrete-time dis-
tributed parameter systems formulation augmented by the output function.
The problem that we want to address here is that of forcing the
appearance of a given pattern inside a region by imposing a suitable
set of values onto some specific sites that could be in the lattice or in
its boundary.
https://doi.org/10.25088/ComplexSystems.30.3.391
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
398
The idea is to explore different formalisms and approaches, some
of which are specific to PDEs, that can prove the regional controllability of CAs, focusing on Boolean CAs.
2.3 Regional Controllability of Cellular Automata: Kalman Condition
Let us recall the classical Kalman rank condition [5] as stated for the
finite dimensional systems:
Definition 2. The controllability matrix related to equation (1) is a
matrix of dimension n ⨯nm defined by:
Mc B, AB, …, An-1 B.
(3)
The determination of the matrix gives information about whether
the system is controllable or not. We have the following theorem
(Kalman condition):
Theorem 1.
Equation (1) is controllable if and only if the controllability matrix is of full rank; in other words:
rank(Mc ) n.
(4)
This was generalized to CAs and allowed to prove the regional controllability.
Definition 3.
Let us consider:
◼ A CA defined on a discrete lattice ℒ with state set and described by a
transition function f
◼ ω⊂ℒ
◼ sω the restriction to ω of the CA configuration s
◼ ω {s : ω → }
The CA is said to be regionally controllable if for a given sd ∈ ω
there exists a control sequence u (u0 , …, uT-1 ) with ui ∈ such
that:
sT sd on ω
where sT is the final configuration at time T and is the control
space.
Let us consider a Boolean CA defined on a lattice ℒ that is assumed
to be finite and composed of N interior cells and two boundary cells
where the actions will be exerted, denoted by cl and cr for the left and
right boundary cells, respectively. We are interested in finding the suitable sequences of controls acting on the boundary of the lattice,
and u0r , url , …, uT-1
, so as to steer the system from
u0l , u1l , …, uT-1
l
r
2.3.1 Special Case of Boolean Cellular Automata
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
399
a given initial state s0 to a desired configuration sd on the subregion ω
at a given time T, such that sT (ci ) sd (ci ), ∀ ci ∈ ω. The desired configuration sd is assumed to be reachable in the evolution of the
CA�rule.
In the following, to make notation more clear, we will indicate
with the symbol ⊕ the sum modulo two.
The state equation for linear Boolean CAs is
st+1 Jst ⊕ But ; 0 < t < T
where
s0 ∈ X
◼ J is the circulant band matrix, each line of which contains the coefficients a1 , … , an in the positions corresponding to the neighborhood .
◼ B is an n ⨯ 2 matrix that represents the control operator.
◼ ut utl , utr is the control matrix at time t consisting of a two-compo-
nent vector in this particular case.
Theorem 2.
Kalman condition for controllability.
A one-dimensional linear CA is regionally controllable via boundary actions if and only if:
Rank(Mc ) RankB JB J2 B…JT-1 B T N - 1.
Where T is the time horizon, N is the size of the CA lattice and J is
the Jacobian matrix.
Proof. Let s0 be the initial configuration of . Assume that the CA is
regionally controllable in ω1 by acting on cl such as ω1 ⊂ ω and
ω1 ω - {cl }, then a sequence of control utl , utr exists such as
t 0, …T - 1. By using
st+1 Jst ⊕ But
(5)
it follows that
s1 Js0 ⊕ Bu0l
s2 Js1 ⊕ Bu1l
s2 JJs0 ⊕ Bu0l ⊕ Bu1l
s2 J2 s0 ⊕ JBu0l ⊕ Bu1l
s3 Js2 ⊕ Bu2l
s3 JJ2 s0 ⊕ JBu0l ⊕ Bu1l ⊕ Bu2l
s3 J3 s0 ⊕ J2 Bu0l ⊕ JBu1l ⊕ Bu2l .
https://doi.org/10.25088/ComplexSystems.30.3.391
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
400
Therefore:
sT JT s0 + B JB J2 B…JT-1 BuT-1
uT-2
…u0l .
l
l
tr
We define the controllability matrix Mc :
B JB J2 B…JT-1 B.
We get the regional controllability when
Rank(Mc ) T.
For peripheral linear CAs, we also know that in order to be able to
change any state of region ω1 {c1 , c2 , …, cN } from one boundary cl
(if the CA is left-linear), the time T should equal N - 1 [25], where N
is the size of the CA. Hence, we can get that:
Rank(Mc ) T N - 1.
Now suppose Rank(Mc ) T N - 1 where Rank(Mc )
dim(Image(Mc )); then for each initial configuration s0 , we can associate a desired configuration sd . Hence the proof. □
Remark 1.
It is difficult to develop algorithms to decide whether a
generic rule is controllable. But for linear rules or peripherally linear
ones (that depend linearly on a peripheral site), this can be done iteratively: the least external site can control a site in the target region, flipping its value if wrong. And then this can be repeated for all sites to
be controlled.
Figure 1. Control of the region ω by applying the control on one boundary
c1 . The change in the initial condition on the controlled cell c1 will be propagated to the cell c4 at time T 3.
Example 1.
Consider the elementary cellular automaton (ECA)
rule�150. We suppose that we act only on the left boundary cell of a
region ω {c10 , …, c30 } and show that there exists a sequence of
controls u u0c , …, uT-1
c that steers the system in ω from the initial
configuration to a desired one constituted of ones, sd 1, …, 1 at
time T 39.
l
Complex Systems, 30 © 2021
l
Some Control and Observation Issues in Cellular Automata
(a)
401
(b)
Figure 2. The evolution of CA rule 150 (a) with and (b) without the applica-
tion of the left boundary controls.
2.4 Alternative Methods to Kalman Condition for Regional
Controllability of Cellular Automata
The notion of controllability was identified by Kalman as one of the
central properties determining system behavior. His simple rank condition is prevalent in the analysis of linear systems. In this section, we
aim at exploring new approaches to prove the regional controllability
of CAs in more general cases.
2.4.1 Markov Chains Approach for Controllability
The first method comes from the idea that CAs and Markov chain
modeling are of great interest when merged and applied in practical
situations. The evolution of all possible configurations of a probabilistic CA can be written as a Markov chain. Since deterministic CAs are
limit cases of probabilistic ones, they also can be seen as particular
Markov chains. A Markov chain such that for some t, Mj > 0 for all
i, j is said to be regular, and this implies that any configuration can be
reached by any configuration in time t. The evolution of a controlled
CA can be seen as a Markov chain where the states are the possible
configurations of region ω. Two CA configurations restricted to the
target region ω, s0 ω and s1 ω are related to each other if there exists
a boundary control l, r such that ps , s 1, where ps , s denotes the
0
1
0
1
probability for jumping from s0 to s1 in one step. In that way, the socalled transition matrix can be constructed in order to describe and
analyze a Markov chain.
Example 2.
Consider the CA rule 150 and its transition matrix and
associated graph, as shown in Figure 3.
https://doi.org/10.25088/ComplexSystems.30.3.391
402
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
(a)
(b)
Figure 3. (a) The transition matrix 150 for ω 3. (b) Graph of 150 .
A CA (linear or nonlinear) is regionally controllable if and
only if there exists a power of matrix T such that all the components
are strictly positive.
Theorem 3.
2.4.2 Graph Theory Approach for Regional Controllability of Cellular
Automata
Markov chains can be described using directed graphs where the
nodes represent the different possible states and the edges represent
the probability of the system moving from one state to the other in
the next time instance. The graph theory seems to be a good and
appropriate tool to study the problem of regional controllability of
CAs. We give the important characterization result in terms of graph
theory.
A CA is regionally controllable if and only if there exists a
t such that the graph associated to the transformation matrix t contains a Hamiltonian circuit; that is, a circuit of a graph G V, AR
is a simple directed path of G that includes every vertex exactly
once [17].
Theorem 4.
3. Observability and Reconstructibility of Cellular Automata
3.1 Introduction to Observability and Reconstructibility
Observability, as defined by Kalman [5, 6], allows us to determine if a
system is observable by one or more sensors, that is, if it is possible to
reconstruct the state of the system based on the measurements of
these sensors. This criterion makes it possible to choose and place efficiently the different sensors needed for the observation. Once the sensors are placed so that they can observe the system, it is necessary to
build the observer, an entity that estimates the state of the system
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
403
from the measurements of the sensors. As the observability criterion is
verified, we know that such an observer exists.
The estimated state of the system, denoted z , is constructed from
the y measurements taken by the sensors and from the dynamics of
the physical system. Its mathematical description is similar to the augmented definition (equations (1) and (2)) of the physical system
except for the addition of an output error feedback via the observer
operator L. The estimated state dynamics satisfy therefore
z (t) Az (t) + Ly(t) - y (t) + Bu(t); 0 < t < T
(6)
y (t) Cz (t)
z0 z ∈ DA.
0
The L operator represents the process by which the estimated state
converges toward the state of the system; the way it is constructed
influences the performance or the robustness of the observer.
In most studies, observability, and by extension all the aspects
related to observation, are studied through controllability by the principle of duality that binds one to the other [5]. In this section, we will
pay particular attention to the use of observation for monitoring and
estimation of initial state for complex systems modeled by CAs, specifically autonomous CAs, that is, those that are not controlled
(∀ t ∈ 0, T, u(t) 0). In the following, we present separately our
study on CA monitoring and CA initial state estimation. The first consists of determining the current state of the system, while the second
focuses on finding out the initial state. These two notions are called,
respectively, reconstructibility and observability [26]. The latter is
stronger than the former because if the initial state can be estimated,
then it is possible to compute the current state from this initial state.
Reconstructibility is rarely addressed in the study of linear continuoustime invariant (LTI) systems or continuous-time DPS because of its
equivalence to observability (which is not valid any longer for discretetime systems; see [27]). It is, however, starting to be investigated in
the case of the Boolean control network [26, 28].
3.2 Observability and Reconstructibility of Cellular Automata
CAs are mathematical models where time, space and state are discrete. This total discretization prevents the direct use of the observability and controllability results of continuous-time control theory, or
even discrete-time-only control theory. Indeed, the existence and
uniqueness of an inverse requires the state space to be a field (in the
mathematical sense). To build a finite (or Galois) field from only a
finite number of possible state values requires this number to be
prime. On the other hand, the finiteness of the state values makes it
https://doi.org/10.25088/ComplexSystems.30.3.391
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
404
possible to use discrete mathematics results such as, for instance,
those from graph theory. The behavior of the system can be represented in the form of a graph, called a configuration graph, that links
the different possible states of the system [29]. This discrete approach
to observability and controllability is recent in control theory, but not
in computer science nor in applied mathematics. There are numerous
methods of state reconstruction, but they are not based on observability and reconstructibility criteria.
From Section 2.2.2 given the global dynamics (F) and observation
(H) operators, a sequence of output Θ around a time horizon T can
be reconstructed from the initial state, through some extended observability (or Gramian) operator:
ΘT : s0 ↦ {θ0 , θ1 , …, θT-1 } Hs0 , HFs0 , …, HFT-1 s0
(7)
where ΘT represents the T first outputs of the initial configuration
s0 ∈ X. Plenet et al. [29] proposed a definition for the observability
and reconstructibility that can both be mathematically linked to the
property of injectivity of the output sequence ΘT .
Definition 4.
Observability and reconstructibility.
Let A be a CA with a global transition function F and an output
sequence ΘT ; then the two following propositions hold:
◼ A is observable
′
′′
′
′′
⟺ ∀ s′0 , s′′
0 ∈ X, ΘT (s0 ) ΘT (s0 ) ⟹ s0 s0 .
◼ A is reconstructible
′
′′
T ′
T ′′
⟺ ∀ s′0 , s′′
0 ∈ X, ΘT (s0 ) ΘT (s0 ) ⟹ F (s0 ) F (s0 ).
From Definition 4, it is clear that both can be used to reconstruct a
configuration from the measurements, as every output sequence is the
result of only one configuration. Moreover, each proposition deals
with a specific configuration: the observability deals with the initial
configuration (s0 ), whereas the reconstructibility deals with the cur-
rent configuration (st FT (s0 )). In addition, observability is stronger
than reconstructibility, as sT can be calculated from s0 ; thus the
observability must be assessed before the reconstructibility.
The observability of CAs is scarcely studied, even for systems with
stationary sensors. The following sections present two methods to
determine CA observability: a function-based method and a
configuration-based method. In the first, the algebraic properties of
the transition function (local or global) are studied to derive observability properties. In the second, the discrete aspect of CAs is
combined with the notion of relationship between the automata configurations to get information about the observability of the CA. The
function-based method has the advantage of a linear complexity with
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
405
the number of cells, since each cell has a finite number of states. Conversely, the configuration-based method has exponential complexity,
since the number of configurations depends on the number of cells
and the number of states (e.g., a Boolean CA with N cells has 2N possible configurations).
3.2.1 Observability of Cellular Automata: Kalman Condition
This section presents a function-based method to determine observability for affine CAs through the affine property of the transition
function. It relies on the Kalman condition presented in Theorem 1,
and on the controllability/observability duality. A CA is said to be
affine if its F transition function is affine; that is, it uses only addition
and multiplication by constants. The transition function can then be
expressed with a linear map and a constant vector.
Definition 5.
Affine CAs.
If a CA and an output operator are affine, then F (respectively H)
can be written in the form of a linear map plus a constant, which can
in turn be written as a matrix A and a constant η F0 (respectively
C and γ H0). The evolution of the CA can therefore be written as:
st+1 F Ast + η
θt Cst + γ
s0 ∈ X.
(8)
The output sequence ΘT is also an affine map composed
with OT the observability matrix (the dual of the controllability
matrix Mc presented in Definition 2) and ΓT a constant vector
Definition 6.
ΘT
θ0
θ1
θT-1
…
C0
CA
…
CAT-1
OT
s0 +
γ
CJ0 η + γ
…
CJT-2 η + γ
OT s0 + ΓT
(9)
ΓT
with Jt ∑tk0 Ak .
The Kalman rank condition for the observability of an affine CA is
the following:
Theorem 5.
Kalman rank condition.
An affine CA F, observed by an affine output operator H, is observable under a time horizon T, if and only if OT is full rank.
This Kalman rank condition is usually used for continuous and
discrete LTI systems and was extended to affine CAs [18, 29]. A
https://doi.org/10.25088/ComplexSystems.30.3.391
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
406
corollary is also given, allowing the reconstruction of the initial state
from the measurements by inverting (or pseudo inverting) the full
rank matrix OT .
Corollary 1. If the Kalman criterion is verified, then it is possible to
recover the initial state by inverting the observability matrix:
x0 O†T (YT - ΓT ).
(10)
Theorem 5 and Corollary 1 are applicable for an observable system
but can be extended in the case of a reconstructible system with the
following theorem and corollary.
Theorem 6.
Reconstructibility condition.
An affine CA F, observed by an affine output operator H, is reconstructible if and only if there exists a finite time horizon T such that:
ker OT ⊂ ker AT .
(11)
Corollary 2. If the reconstructibility criterion is verified, then it is possible to find a matrix R such that:
xT R(YT - Γ) + JT-1 η.
(12)
These two theorems and corollaries allow us to assess the observability and reconstructibility only for affine CAs. This class of
automata is in fact quite small. For instance, among Wolfram’s 256
elementary CA, only 16 are affine: the eight additive rules (0, 60, 90,
102, 150, 170, 204, 240) and their complementary rules (255, 195,
165, 153, 105, 85, 51, 15). Therefore, a second method has been proposed to investigate observability and reconstructibility of CA control
systems. It is presented hereafter.
3.2.2 Observability of Cellular Automata: Configuration Relation
This configuration-based method makes use of a binary relation
representation of the F, H and Θ operators, respectively denoted ℛF ,
ℛH and ℛΘ .
The transition binary relation ℛF associated to the global
transition function F is defined as:
Definition 7.
ℛF (s0 , s1 ) s0 , s1 ∈ X and s1 F(s0 ).
(13)
The binary relation can be represented in the form of a matrix
called a logical matrix.
The logical matrix M associated to a binary relation
ℛ ∈ X⨯Y is defined as:
Definition 8.
Mi,j
1 xi , yj ∈ ℛ
0 else.
Complex Systems, 30 © 2021
(14)
Some Control and Observation Issues in Cellular Automata
407
This binary relation representation allows us to assess the observability and reconstructibility of all kinds of CAs, independently of
their algebraic nature. From Definition 4, it is clear that the observability and the reconstructibility are related to the injectivity of ΘT .
This property can be studied through the relation ℛΘ for the observability and ℛΘ F-T (constructed from ℛΘ and the converse of ℛFT ) for
T
the reconstructibility. This results in the following theorem using the
logical matrices of such relations:
A CA is observable (reconstructible) if the logical matrix
related to the binary relation ℛΘ (ℛΘ F-T ) has at most one nonzero ele-
Theorem 7.
T
ment per column.
These logical matrices have very few nonzero elements, making it
possible to use sparse matrices to represent them numerically. This
reduces not only the memory allocation but also the algorithmic complexity for the assessment of observability and reconstructibility.
The main problem with this method is the exponential increase in
the number of configurations in relation with the number of cells.
Even when the algorithms used to evaluate observability are optimized, the exponential complexity coming from the number of configurations makes this method less efficient when the particular case of
affine CAs is investigated.
3.3 Observation of Distributed Parameters Systems through
Mobile Sensors
In the case of “classic” observation problems, only the sensors’ types
are chosen according to the natures of the output variables to be measured and the state variables to be observed (e.g., a displacement,
velocity or force for a mechanical system). But for spatially distributed systems, the position of the sensors also becomes a crucial
problem in the design of the observer. The placement of sensors
remains, even today, a major concern in many areas such as water
networks, forest fire detection and telemonitoring of human physiological data. Sensor placement is often related to the wireless sensor
network (WSN) research area [30, 31], which studies strategies to
place sensors to get optimal information about an observed system.
The main difference with our approach, based on control theory
and observation, is that WSNs do not seek to reconstruct the state of
the system using its dynamics. Sensor placement is carried out using
different algorithmic methods such as genetic algorithms [32] and
Bayesian optimization. In these studies, the sensors are generally considered fixed. This reduces their performance for the observation of
spatially distributed systems. Indeed, the use of mobile sensors makes
it possible to obtain observability, whereas stationary sensors would
https://doi.org/10.25088/ComplexSystems.30.3.391
408
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
have failed [29]. However, this poses new problems such as the calculation of sensor trajectories, which is still under investigation. Some
researchers are directly studying the problem of observability with
mobile sensors. Demetriou et al. [34, 35] and Hussein et al. [36] are
considering networks of mobile actuators and/or sensors. Related
problems occurred when mobile sensors (or actuators) are considered,
such as obstacle avoidance or formation coordination to get maximum coverage. Such problems are intensively investigated, independently from the observation problem. For instance, among the many
different existing methods to handle collision avoidance for mobile
sensors, we can cite the predictive control approach [37] or the potential field approach [38].
As the sensors are moving in space, the cells they observe change
through time. The output operator C therefore becomes time dependent (it may then be denoted C(t)). The observability criteria become
time dependent and it is therefore necessary to evaluate the observability for all the different sensor trajectories. The two observability
conditions presented in Section 3 can both be assessed with mobile
sensors but for only one trajectory at a time. They can be used as a
constraint in constrained trajectory planning to find a mobile sensor
trajectory that respects observability and other constraints external to
the observability problem, such as obstacle avoidance. This trajectory
planning can be carried out either prior to system deployment or in
real time.
4. Discussion
Controllability and observability are two fundamental properties that
characterize the behavior of a given system and determine the relationships between state and input and output variables. As the real
physical systems become more and more complex, the need to design
controllers and observers for complex nonlinear systems is real. The
study of controllability of linear systems was first performed in detail
by Kalman and his collaborators in [6]. The first part of this paper
discussed an analog of the Kalman controllability rank condition for
systems described by cellular automata (CAs) considered as the discrete counterpart of partial differential equations (PDEs). It also
explored some alternative approaches to this condition that could
work for linear and nonlinear CAs. The paper focused on the special
case of regional controllability of Boolean one-dimensional CAs. A
necessary and sufficient condition for regional controllability of CAs
was proved through original approaches such as Markov chains and
graph theory.
This work can be extended in many tracks that can explore in particular the case of non-Boolean CAs and the highly nonlinear CAs.
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
409
The case of large lattice size is also to be considered. The examples
considered concerned only a very limited number of cells because of
the enormous computational cost generated in terms of spatial and
temporal complexity. The control obtained is generally not unique at
this stage, and the problem of optimality will be addressed later. A
first problem of regional controllability in minimum time is currently
under study.
Another perspective concerns the investigation of more adequate
algorithms for the calculation of preimages that could be of low complexity, depending on the dimensions of the lattice.
For the observability issue, two methods for assessing observability
and reconstructibility were presented. The first, using the Kalman condition, only applies to affine CAs, while the second allows us to study
nonlinear CAs, but its exponential complexity makes it difficult to
use. Several approaches can be considered to reduce the algorithmic
complexity linked to the calculation of observability and reconstructibility. The first would be a new function-based method but for
nonlinear CAs. Previous strategies already used in the case of linear
continuous-time invariant (LTI) systems, such as the use of linearization or the use of nonlinear Jacobians, could be considered. The second method to make observability and reconstructibility computation
tractable on CAs with a large number of cells is to use multiple
observers. Each of these smaller and simpler observers is then in
charge of the observation of a part of the whole cellular automaton
(CA). The limited number of possible configurations for each
observed region of the CA implies a dimensional reduction, making it
tractable to determine observability or reconstructibility on these
parts of the CA. For example, with a Boolean controller of 1000 cells,
there are 21000 configurations. If this automata is observed by 10
observers of 100 cells, then we must evaluate 10 times the observability for 2100 configurations. However, a new notion of observability
will still have to be defined to account for these distributed observers
in order to ensure that the local observability of the local observers
can be combined to obtain the observability of the whole system.
The use of a mobile sensor network not only ensures observability
where static sensors would have failed, but also makes the topology
of the observation network dynamic and allows the sensors to focus
on a specific part of the system. However, the use of mobile sensors
coupled with the distributed observers requires strong coordination of
the sensors. These two criteria make it extremely complex to calculate
trajectories under constraints using the classical methods of control
theory. One approach for improvement would be to use a notion
already well studied in computer science: the multi-agent systems
(MAS) paradigm. The agents (here observers and sensors) are
https://doi.org/10.25088/ComplexSystems.30.3.391
410
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
autonomous and communicate with each other. Each agent would
calculate its own trajectory according to the constraints (observability
and other) and the information exchanged between the agents. In the
same way that the distributed observer manages to reduce the complexity of the observability calculation by distributing the task
between several entities, the use of MAS produces the same effect but
with respect to the trajectory calculation. The organization of the network and the communications between agents will have to be carefully designed to ensure global observability of the system based on
local computations by the agents.
It should be noted that the controllability and observability problems were considered separately in this paper. For controllability
issues, we assumed that the state of the system can be measured
at�each time, and for observability problems, the system is supposed
to be autonomous. The global objective of considering a complete
state equation with control and observation will be investigated in
future work.
References
[1] A. El Jai, “Distributed Systems Analysis via Sensors and Actuators,”
Sensors and Actuators A: Physical, 29(1), 1991 pp. 1–11.
doi:10.1016/0924-4247(91)80026-L.
[2] A. El Jai and M. Amouroux, “Sensors and Observers in Distributed
Parameter Systems,” International Journal of Control, 47(1), 1988
pp.�333–347. doi:10.1080/00207178808906013.
[3] R. F. Curtain and H. Zwart, An Introduction to Infinite-Dimensional
Linear Systems Theory, New York: Springer-Verlag, 1995.
[4] E. Zerrik, A. Boutoulout and A. El Jai, “Actuators and Regional Boundary Controllability of Parabolic Systems,” International Journal of Systems Science, 31(1), 2000 pp. 73–82. doi:10.1080/002077200291479.
[5] R. E. Kalman, “On the General Theory of Control Systems,” in IFAC
Proceedings Volumes, 1(1), 1960 pp. 491–502.
doi:10.1016/S1474-6670(17)70094-8.
[6] R. E. Kalman, “Mathematical Description of Linear Dynamical Systems,” Journal of the Society for Industrial and Applied Mathematics,
Series A: Control, 1(2), 1963 pp. 152–192. doi:10.1137/0301010.
[7] A. El Jai and S. El Yacoubi, “On the Relations between Actuator
Structures and Final-Constraint Minimum-Energy Problem,” Sensors
and Actuators A: Physical, 33(3), 1992 pp. 175–182.
doi:10.1016/0924-4247(92)80164-X.
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
411
[8] B. Chopard and M. Droz, Cellular Automata Modeling of Physical Systems, New York: Cambridge University Press, 1998.
[9] A. Wuensche and M. Lesser, “The Global Dynamics of Cellular
Automata,” An Atlas of Basin of Attraction Fields of One-Dimensional
Cellular Automata, Avalon Publishing, 1992.
[10] S. Wolfram, Cellular Automata and Complexity: Collected Papers, Reading, MA: Addison-Wesley Pub. Co., 1994.
[11] G. B. Ermentrout and L. Edelstein-Keshet, “Cellular Automata
Approaches to Biological Modeling,” Journal of Theoretical Biology,
160(1), 1993 pp. 97–133. doi:10.1006/jtbi.1993.1007.
[12] X. Xiao, S. Shao, Y. Ding, Z. Huang, X. Chen and K.-C. Chou, “Using
Cellular Automata to Generate Image Representation for Biological
Sequences,” Amino Acids, 28(1), 2005 pp. 29–35.
doi:10.1007/s00726-004-0154-9.
[13] L. B. Kier, P. G. Seybold and C.-K. Cheng, Cellular Automata Modeling
of Chemical Systems: A Textbook and Laboratory Manual, Dorderecht:
Springer, 2005.
[14] S. El Yacoubi, “A Mathematical Method for Control Problems on Cellular Automata Models,” International Journal of Systems Science, 39(5),
2008, pp. 529–538. doi:10.1080/00207720701847232.
[15] S. Dridi, S. El Yacoubi and F. Bagnoli, “Boundary Regional Controllability of Linear Boolean Cellular Automata Using Markov Chain,”
Recent Advances in Modeling, Analysis and Systems Control: Theoretical Aspects and Applications (E. Zerrik, S. Melliani and O. Castillo,
eds.), Cham, Switzerland: Springer, 2020 pp. 37–48.
doi:10.1007/978-3-030-26149-8_4.
[16] S. Dridi, F. Bagnoli and S. El Yacoubi, “Markov Chains Approach for
Regional Controllability of Deterministic Cellular Automata, via
Boundary Actions,” Journal of Cellular Automata, 14(5/6), 2019
pp.�479–498.
[17] S. Dridi, S. El Yacoubi, F. Bagnoli and A. Fontaine, “A Graph Theory
Approach for Regional Controllability of Boolean Cellular Automata,”
International Journal of Parallel, Emergent and Distributed Systems,
35(5), 2020 pp. 499–513. doi:10.1080/17445760.2019.1608442.
[18] T. Plénet, S. El Yacoubi, C. Raïevsky and L. Lefèvre, “Observability of
Affine Cellular Automata through Mobile Sensors,” Cellular Automata
(ACRI 2020), Lodz, Poland (T. M. Gwizdałła, L Manzoni, G. C. Sirakoulis, S. Bandini and K. Podlaski, eds.), Cham, Switzerland: Springer,
2021 pp. 36–45.
link.springer.com/chapter/10.1007/978-3-030-69480-7_5.
[19] J.-L. Lions, “Controlabilité exacte des systèmes distribués: remarques
sur la théorie générale et les applications,” Analysis and Optimization
of Systems (A. Bensoussan and J. L. Lions, eds.), Berlin, Heidelberg:
Springer, 1986 pp. 3–14. doi:10.1007/BFb0007542.
https://doi.org/10.25088/ComplexSystems.30.3.391
412
S. El Yacoubi, T. Plénet, S. Dridi, F. Bagnoli, L. Lefèvre and C. Raïevsky
[20] I. Kaarer, A. Ayadi and I. Rezzoug, “Weak Controllability and the New
Choice of Actuators,” Global Journal of Pure and Applied Mathematics, 14(2), 2018 pp. 325–330.
www.ripublication.com/gjpam18/gjpamv14n2_13.pdf.
[21] D. L. Russell, “Controllability and Stabilizability Theory for Linear
Partial Differential Equations: Recent Progress and Open Questions,”
SIAM Review, 20(4), 1978 pp. 639–739. doi:10.1137/1020095.
[22] S. A. Kauffman, “Homeostasis and Differentiation in Random Genetic
Control Networks,” Nature, 224, 1969 pp. 177–178.
doi:10.1038/224177a0.
[23] C. Damiani, R. Serra, M. Villani, S. A. Kauffman and A. Colacci, “CellCell Interaction and Diversity of Emergent Behaviours,” IET Systems
Biology, 5(2), 2011 pp. 137–144. doi:10.1049/iet-syb .2010.0039.
[24] A. Deutsch and S. Dormann, Cellular Automaton Modeling of Biological Pattern Formation: Characterization, Examples, and Analysis,
Boston: Birkhäuser, 2005.
[25] F. Bagnoli, S. El Yacoubi and R. Rechtman, “Toward a Boundary
Regional Control Problem for Boolean Cellular Automata,” Natural
Computing, 17(3), 2018 pp. 479–486. doi:10.1007/s11047-017-9626-1.
[26] E. Fornasini and M. E. Valcher, “Observability, Reconstructibility and
State Observers of Boolean Control Networks,” IEEE Transactions on
Automatic Control, 58(6), 2013 pp. 1390–1401.
doi:10.1109/TAC.2012.2231592.
[27] A. C. Antoulas, Approximation of Large-Scale Dynamical Systems,
Philadelphia: Society for Industrial and Applied Mathematics, 2005.
[28] K. Zhang, L. Zhang and R. Su, “A Weighted Pair Graph Representation
for Reconstructibility of Boolean Control Networks,” SIAM Journal
on �Control and Optimization, 54(6), 2016 pp. 3040–3060.
doi:10.1137/140991285.
[29] T. Plénet, S. El Yacoubi, C. Raïevsky and L. Lefèvre, “Observability and
Reconstructibility of Cellular Automaton,” International Journal of Systems Science, forthcoming.
[30] M. Younis and K. Akkaya, “Strategies and Techniques for Node Placement in Wireless Sensor Networks: A Survey,” Ad Hoc Networks, 6(4),
2008 pp. 621–655. doi:10.1016/j.adhoc.2007.05.003.
[31] J. Yick, B. Mukherjee and D. Ghosal, “Wireless Sensor Network
Survey,” Computer Networks, 52(12), 2008 pp. 2292–2330.
doi:10.1016/j.comnet.2008.04.002.
[32] L. Yao, W. A. Sethares and D. C. Kammer, “Sensor Placement for OnOrbit Modal Identification via a Genetic Algorithm,” AIAA Journal,
31(10), 1993 pp. 1922–1928. doi:10.2514/3.11868.
Complex Systems, 30 © 2021
Some Control and Observation Issues in Cellular Automata
413
[33] R. Garnett, M. A. Osborne and S. J. Roberts, “Bayesian Optimization
for Sensor Set Selection,” in Proceedings of the 9th ACM/IEEE International Conference on Information Processing in Sensor Networks (IPSN
’10), Stockholm, Sweden, New York: Association for Computing
Machinery, 2010 pp. 209–219. doi:10.1145/1791212.1791238.
[34] M. A. Demetriou, “Guidance of Mobile Actuator-Plus-Sensor Networks
for Improved Control and Estimation of Distributed Parameter
Systems,” IEEE Transactions on Automatic Control, 55(7), 2010
pp.�1570–1584. doi:10.1109/TAC.2010.2042229.
[35] M. A. Demetriou and I. I. Hussein, “Estimation of Spatially Distributed
Processes Using Mobile Spatially Distributed Sensor Network,” SIAM
Journal on Control and Optimization, 48(1), 2009 pp. 266–291.
doi:10.1137/060677884.
[36] I. I. Hussein and D. M. Stipanovic, “Effective Coverage Control for
Mobile Sensor Networks with Guaranteed Collision Avoidance,” IEEE
Transactions on Control Systems Technology, 15(4), 2007 pp. 642–657.
doi:10.1109/TCST.2007.899155.
[37] J.-M. Park, D.-W. Kim, Y.-S. Yoon, H. J. Kim and K.-S. Yi, “Obstacle
Avoidance of Autonomous Vehicles Based on Model Predictive Control,” in Proceedings of the Institution of Mechanical Engineers, Part D:
Journal of Automobile Engineering, 223(12), 2009 pp. 1499–1516.
doi:10.1243%2F09544070JAUTO1149.
[38] O. Khatib, “Real-Time Obstacle Avoidance for Manipulators and
Mobile Robots,” Autonomous Robot Vehicles (I. J. Cox and G. T. Wilfong, eds.), New York: Springer, 1986 pp. 396–404.
doi:10.1007/978-1-4613-8997-2_29.
[39] A. Elnagar and A. Hussein, “On Optimal Constrained Trajectory Planning in 3D Environments,” Robotics and Autonomous Systems, 33(4),
2000 pp. 195–206. doi:10.1016/S0921-8890(00)00095-6.
[40] G. Y. Vichniac, “Boolean Derivatives on Cellular Automata,” Physica
D: Nonlinear Phenomena, 45(1-3), 1990 pp. 63–74.
doi:10.1016/0167-2789(90)90174-N.
[41] S. El Yacoubi, A. El Jai and N. Ammor, “Regional Controllability with
Cellular Automata Models,” in Cellular Automata (ACRI 2002),
Geneva, Switzerland (S. Bandini, B. Chopard and M. Tomassini, eds.),
Springer, 2002 pp. 227–253. doi:10.1007/3-540-45830-1_34.
https://doi.org/10.25088/ComplexSystems.30.3.391