Vous êtes sur la page 1sur 28

JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL Pages 123 – 150

Active vibration control of flexible plate


structures with distributed disturbances
S. Julai1 and M. O. Tokhi2
1Department of Mechanical Engineering, University of Malaya, Kuala Lumpur,

Malaysia
2Department of Automatic Control and Systems Engineering, The University of

Sheffield, United Kingdom


Email: sabsz@um.edu.my

Received 8th March 2011

ABSTRACT
This paper presents the development of an active vibration control (AVC)
system with distributed disturbances using genetic algorithms, particle swarm
optimization, and ant colony optimization. The approaches are realized with
multiple-input multiple-output and multiple-input single-output control
configurations in a flexible plate structure. A simulation environment
characterizing a thin, square plate, with all edges clamped, is developed using
the finite difference method as a platform for test and verification of the
developed control approaches. Simulations are carried out with random
disturbance signal. The control design comprises a direct minimization of the
error (observed) signal by allowing a collective determination of detection and
secondary source locations together with controller parameters. The algorithms
are formulated with a fitness function based on the mean square of the observed
vibration level. In this manner, knowledge of the input/output characterization
of the system is not required for design of the controller. The performance of the
system is assessed and analyzed both in the time and frequency domains and it
is demonstrated that significant vibration reduction is achieved with the
proposed schemes.

Keywords: Active vibration control, genetic algorithms, particle swarm optimization,


ant colony optimization, flexible plate structure.

1. INTRODUCTION
Interest has increased in the active control of vibration in mechanically flexible
systems. These in aerospace and aircraft structures include space-based radar
antennal and solar panels, space robotics, propeller and aircraft fuselage and wings,
in electromechanical systems that include turbo generator shafts, gas turbine rotors
and electric transformer cores, and in civil engineering applications include
skyscrapers and bridges [1-6]. Such structures may be damaged or become
ineffective under the influence of undesired vibrational loads they constantly
experience. Due to these problems, an effective control mechanism is required to
attenuate the vibration in order to preserve structural integrity of such systems. The
conventional form of external passive damping to change the dynamic characteristics
of the structure is not preferred as the addition of a damper adds to the overall system
weight, which is undesirable and makes the system less transportable especially for
space applications. This has led to extensive research into active control techniques
where the disturbance to be cancelled or the properties of the controlled system vary
with time. Active control solutions are known to achieve good vibration suppression
performance in comparison to passive control [7].

Vol. 31 No. 2 2012 123


Active vibration control of flexible plate structures with distributed disturbances

Active vibration control (AVC) consists of artificially generating cancelling


sources to interfere destructively with the unwanted source and thus result in a
reduction in the level of vibration (disturbance) at desired location(s). This
unwanted source can be either from one excitation source or a number of excitation
sources that can be modelled as a set of point sources distributed around the surface
of the structure. To suppress the vibration from one excitation source, a single
detector is generally sufficient to obtain the required signal information needed to
generate the cancelling signal. In the case where the vibration is from distributed
sources, a multiple set of detectors is required to obtain sufficient signal information
in either single-output or multiple-output control configuration. This will lead to the
realization of a multiple-input single-output (MISO) or multiple-input multiple-
output (MIMO) control system [8-9].
AVC is realized with actuators, sensors and electronic control to reduce vibration
without necessarily adding damping to the system. Due to the broadband nature of
the disturbances, it is required that the control strategy realizes suitable frequency-
dependent characteristics so that cancellation over a broad range of frequencies is
achieved [8]. Remarkable advances in smart materials, such as piezoelectric
transducers which are used extensively as distributed sensors and actuators, and
computing technology have lead AVC to provide cost-effective solutions to most
sound- and vibration-control problems [10]. However, careful consideration must be
taken in positioning the sensors and actuators to ensure good control [11]. This
means that the locations of these transducers have significant influence on the
performance of the control system as well as the controlled response. Many
techniques have been reported to find optimal locations of sensors and actuators.
Methods for optimal placement of sensors and actuators have been investigated by
many researchers [12-15]. All the research works have shown the importance of
optimal placement of sensor and actuator for achieving a significant level of
vibration suppression.
The aforementioned researches have utilized GA and PSO in finding optimal
locations of sensors and actuators for vibration control. However, the use of ant
colony optimization (ACO) has not been reported in this context. GA was
introduced by John Holland [16] based on the principle of the Charles Darwinian
Theory of evolution to natural biology. It is a search procedure based on the
mechanics of natural selection and natural genetics. A possible solution to an
optimization problem can be encoded as an individual (or a chromosome), and later
ranked by comparison to a particular fitness function. New individuals are created
using genetic operators such as mutation and crossover that produce the genetic
composition of a population (finite number of individuals). PSO was developed by
Eberhart and Kennedy [17] in 1995, inspired by the social behaviour of bird
flocking or fish schooling. It is a population based stochastic optimization technique
with a set of randomly generated potential solutions (initial swarm) referred to as
particles that propagate in the search space towards the optimal solution. Each
particle adjusts its position in the search space according to its own flying
experience and the flying experience of other particles. ACO was developed by
Dorigo and his colleagues in 1991, inspired by the cooperative search behaviour of
real ants where the problem of interest is to mimic almost blind ants to establish
shortest paths from their nest to food source and back, through local message
exchange via the deposition of pheromone trails [18].
Traditional optimization methods emphasize accurate and exact computation, but
may fail on achieving the global optimum. GA, PSO, and ACO on the other hand,
have appeared as promising algorithms for handling optimization in complex real-
world problems, and are widely used as design tools and problem solvers in
optimization problems due to their versatility and ability to optimize in complex
multimodal search spaces with nondifferentiable cost functions. These algorithms
are nature inspired with population based stochastic search and have been very
popular in the field of computational intelligence. Solving an optimization problem
using stochastic search can often out-perform classical methods of optimization

124 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

when applied to difficult real-world problems. They offer better chances to achieve
the global optimum since they do not use gradient information and are useful in
solving problems where such information is unavailable or very costly to obtain
[19]. Classical methods based on gradient information, on the other hand,
emphasize accurate and exact computation, but may fail on achieving the global
optimum.
In this work, a multi-input-AVC system is developed using a non-model based
approach where the controller is designed directly based on minimization of the
observed signal with GA, PSO, and ACO, by allowing a combined determination of
detection and secondary source locations together with controller parameters. The
goal of this paper is to propose an approach for vibration reduction of flexible plate
structures by using GA-AVC, PSO-AVC, and ACO-AVC strategies based on
minimization of the observed signal. This approach is used to determine the
controller parameters in such a way that the value of the objective function assigned
is minimized. The approach does not require knowledge of the input/output
characterization of the system for controller design. The performances of
optimization techniques are presented in terms of convergence rate, computational
time, and spectral attenuation achieved at the dominant modes of vibration of the
plate.
The rest of the paper is structured as follows: Section 2 presents classical
governing dynamic equations of a thin rectangular plate and the corresponding
numerical simulation algorithm based on finite difference (FD) method. The GA,
PSO, and ACO algorithms are described in Section 3. The AVC strategies are
presented in Section 4. Section 5 presents performance results of the AVC
approaches within the flexible plate simulation environment. The results are
analyzed and discussed from the time- and frequency-domain perspectives. The
paper is concluded in Section 6.

2. THE FLEXIBLE PLATE STRUCTURE


Flexible structures offer several advantages including fast response, low energy
consumption, reduced mass and low cost, in comparison to their rigid and bulky
counterparts. Besides the advantages these structures offer, a major constraint to
take into consideration is the system vibration arising from structural flexibility.
Accordingly, control of flexible systems has been a challenge for researchers and
engineers [20]. Vibration suppression is focused more on flexible structures due to
their potential widespread use as they are capable of being operated at high speeds
and handling of larger payloads with the same actuator capabilities [22].
Dynamic modeling and simulation of a flexible plate structure using the FD
method has been reported in [23], where a flat, square plate with all edges clamped
has been considered. The classical dynamic equation of a thin rectangular plate is
described with a partial differential equation (PDE) formulation as [24, 25]:

∂w4
∂w ∂w ρ ∂w q
4 4 2

+2 + + = (1)
∂x ∂x ∂ y ∂ y
4 2 2
D ∂t D 4 2

where w is the lateral deflection, r is the mass density per unit area, q=q(x, y) is the
transverse external force at point (x, y) on the plate and has dimensions of force per
unit area, ∂w/∂t2 is the lateral acceleration, D=[EH3/12(1-v)] is the flexural rigidity
with υ representing the Poisson ratio, h the thickness of the plate, and E the Young’s
modulus. A simulation algorithm characterizing the dynamic behaviour of the plate
is developed through discretisation of the PDE. The plate can be divided into n × m
sections, along the x and y axes as x = i∆x and y = j∆y. Using central difference
approximations for the first-, second-, third- and fourth-order derivatives of the
response, a linear relation for the deflection of each section (mesh) can be developed
using FD approximations, as

Vol. 31 No. 2 2012 125


Active vibration control of flexible plate structures with distributed disturbances

D∆t
(
2

w =− 20 w + 8  w +w +w +w  +
i , j ,k +1
ρ∆ 4
xy
i , j ,k i+1, j ,k i , j +1,k i−1, j ,k i , j −1,k

+ 2  w i+1, j +1,k
+w i−1, j +1,k
+w i−1, j −1,k
+w i+1, j −1,k
 (2)
∆t q
2

+w i + 2 , j ,k
+w i , j + 2 ,k
+w i − 2 , j ,k
+w i , j − 2 ,k ) + 2w i , j ,k
−w i , j ,k −1
+
ρ
i, j

where the x-axis is represented with the reference index i, y-axis with the reference
index j, and time, t represented with a reference index k. Here, Wi,j,k+1 is the
deflection of nodal point (xi, yj) of the plate at time step k+1. For the case of all
edges clamped, the deflection is always zero along the edges, and the tangent of the
deflection at the edge is equal to zero, e.g. at y = a a, wy=a = ∂w/∂yy=a = 0. This
condition needs to be satisfied at every nodal point along the clamped edge within
the FD formulation.
The plate is divided into 20 × 20 sections with a sampling time of 0.0016 sec. An
aluminium type plate with boundary condition of all edges clamped and
specifications given in Table I is considered. The simulation algorithm developed is
used as a platform for test and performance evaluation of AVC approaches
developed in this work.

Table 1
Parameters of the flexible plate structure.

Parameter Value
Length, L 1.0 m
Width, W 1.0 m
Thickness, T m
Moment of inertia, I m
Mass density per area,
Young s Modulus, E
Simulation time, t 4 seconds
Poisson s ratio, υ 0.3

3. OPTIMIZATION TECHNIQUES
Traditional optimization methods based on gradient information emphasize accurate
and exact computation, but may fail in achieving the global optimum. Optimization
techniques such GA, PSO, and ACO offer better chances to achieve the global
optimum since they do not use gradient information and are very useful in solving
problems where such information is unavailable or very costly to obtain [19].
Although the algorithms are population-based, but due to different mechanisms,
they have different features and operators.

3.1. Real-coded genetic algorithm


Genetic algorithms are search and optimization techniques that are motivated by the
principle of natural evolution. In traditional GA, all the variables of interest must
first be encoded as binary digits (genes) forming a string (chromosome).This
representation is known as binary-coded genetic algorithm (BCGA). Each
chromosome (individual) has an associated fitness which is the value of the
objective function for that set of parameters and contains sub-strings or genes as
units that contribute in different ways to the overall fitness of the individual. BCGA

126 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

is found to be a robust search technique avoiding local optima, but the major
drawback is the difficulty faced when it is applied to problems with large search
space and requiring high precision. Large string lengths result in more precise
solution but lead to increase in computational cost [26]. To overcome the difficulties
related to binary representation, a floating-point representation of parameters as
chromosomes known as real-coded genetic algorithm (RCGA) is used. All genes in
a chromosome used in RCGA are real numbers. The use of this floating point
representation outperforms binary representations in real-valued optimization
problems because they are more consistent, precise, and lead to faster convergence
[27]. In RCGA, the length of chromosomes becomes shorter than those with the
equivalent binary representation. This implies that computer programming for such
algorithms can be easily performed. The tuning mechanism for mutation and
crossover operations is also performed using floating point numbers instead of long
strings of zeros and ones. In RCGA, three basic operations are used: selection,
crossover, and mutation.

(i) Selection
Selection is an important aspect of evolutionary computation. It dictates what
member of the current population affects the next population. More fit individuals
are generally given a higher chance to participate in the recombination process.
Stochastic universal sampling (SUS) provides zero bias and minimum spread. The
individuals are mapped to contiguous segments of a line, such that each individual’s
segment is equal in size to its fitness exactly as in roulette-wheel selection. Here
equally spaced pointers are placed over the line as many as there are individuals to
be selected [28]. After the selection, the parent chromosomes are combined and
mutated to form the offspring chromosomes.

(ii) Crossover
Let x = (x1, ..., xn) and y = (y1, ..., yn) be the parent strings, where the generic xi and
yi are real variables. In extended intermediate recombination

z = x +α (y − x )
k k k k k
(3)

where ak is chosen uniform randomly in the interval . Intermediate recombination is


capable of producing any point within a hypercube slightly larger than that defined
by the parents. Fig. 1 (a) shows the possible area of offspring after intermediate
recombination.

(ii) Mutation
This function takes a vector containing the real representation of the individuals in
the current population, mutates the individuals with probability pm and returns the
resulting population. The mutation operator is able to generate most points in the
hypercube defined by the variables of the individual and the range of the mutation,
as in Fig. 1 (b).

area of possible
offspring before
variable mutation
variable possible offspring
2 2
parents after
mutation

variable 1 variable 1
(a) (b)
Fig. 1. (a) Possible areas of the offspring after intermediate recombination, and (b) effect of mutation.

Vol. 31 No. 2 2012 127


Active vibration control of flexible plate structures with distributed disturbances

3.2 . Particle swarm optimization


Particle swarm optimization shares many similarities with evolutionary
computation techniques such as GA. It is inspired by the ability of flocks of birds,
schools of fish, and herds of animals to adapt to their environment, find rich sources
of food by implementing an information sharing approach, and it possesses the
properties of easy implementation and fast convergence [17, 29]. The algorithm has
been widely applied to continuous and discrete optimization problems and has
received great attention in systems and control engineering, automatic recognition,
radio systems, etc. Starting with a randomly initialized population, each particle in
the PSO flies through the d-dimensional problem space and remembers the best
position it has seen. The particles evaluate their positions relative to a global fitness
during each iteration and use their memorized best positions to adjust their own
velocities and subsequent positions. In this way, the particles tend to fly towards
better and better searching areas through the search process.
In a PSO algorithm, the position vector and the velocity vector of the i-th particle
in a d-dimensional search space can be represented as Xi = (xi1, xi2, ..., xid) and Vi =
(vi1, vi2, ..., vid), respectively. The best position of each particle (which corresponds
to the best fitness value obtained by that particle at time t) is denoted as Pi = (pi1,
pi2, ..., pid) , and the fittest particle found so far at time t as Pg = (pg1, pg2, ..., pgd) .
Then the new velocities and positions of the particles for the next fitness evaluation
are calculated using the following equations:

v (t + 1) = ω × v (t ) + c × rand(⋅) × ( p − x (t ))
id id 1 id id
(4)
) + c × Rand(⋅) × ( p − x (t ))
2 gd id

x (t + 1) = x (t ) + v (t + 1)
id id id
(5)

where ω is the inertia weight, c1 is the cognition factor, c2 is the social factor, and
rand (·) and Rand (·) are two separately generated uniformly distributed random
numbers in the range [0, 1] [30].
Abd. Latiff and Tokhi, [31] have proposed a strategy to guarantee fast
convergence as an improvement to the concept of time-varying acceleration
coefficients (TVAC), developed by Ratnaweera et al., [32], to effectively control the
global search and convergence to the global best solution. They have established that
by continuously modifying the value of inertia weight, superior results can be
achieved as compared to the case when the inertia weight is fixed. For fast
convergence purposes, the particles have to know their location and relative distance
from each other in exploring the search space. To do so, the spread factor (SF) has
been introduced, which measures the distribution of particles in the search space as
well as the precision and accuracy of the particles with respect to the global optimum.
Precision (or spread of particles in the search space) refers to the maximum distance
between particles in the best and worst positions while accuracy (or deviation of the
particles) refers to the distance of average particle position from the global best
particle. The value of spread factor varies from the maximum range of the search
space down to the desired convergence precision and can be calculated as

 spread + deviation 
SF = 0.5 ×  id id
(6)
 range − range 
id
max min

where spreadid = xmax id – xmin id and deviationid = ∑(xid – gbestid)/N, with xmax id and
xmin id representing maximum and minimum values of the ith particle’s position, N is
the number of particles, and rangemax and rangemin are maximum and minimum

128 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

range of the specified variables. This factor is used to modify the inertia weight,

ω = exp( − iter / ( SF × max _iter )) (7)

with
c = 2 × (1 − iter / max _iter )
1
(8)
c =2
2

where c1 is linearly reduced to zero from its initial value of 2, and c2 is maintained
at 2 to ensure all particles are pulled towards global optimum.

3.3. Ant colony optimization


The basic ACO algorithm actually fits a discrete problem only and is not suitable
for solving continuous optimization problems such as linear or non-linear
programming. Applying this algorithm to continuous domains was not
straightforward where the main issue is how to model a continuous nest
neighbourhood with a discrete structure. As the design problem can always be
formulated as optimization problem in continuous design space, the ACO algorithm
applied in any field should be modified accordingly [33 – 34]. Continuous ACO was
proposed by Wang and Wu, [35] and Quan and Chao, [36]. This algorithm is
referred to as ACO1 throughout this work. Here, the optimization problem is solved
by a cooperation of artificial ant colony by exchanging information via pheromone
deposited on graph edges. The essence of ACO1 is composed of two parts:
1. A state transition rule used by an ant to determine the next destination of a
complete tour, and
2. A pheromone updating rule in allocating a greater amount of pheromone to
shorten the tour.
Let the vector X = [x1, x2, ..., xn] be the parameters to be optimized, where n
represents the total number of parameters in the AVC system, along with upper and
lower bounds to be xi∈D(xi) = [xi_low, xi_up] with i = 1,2, ..., n. . The definition field
D(xi) is divided into M subspaces, and the middle of each subspace defines a node.
A single artificial ant k (k = 1, 2, ..., Nant) where Nant is the maximum number of
ants, would choose to move from one node to the other, in the total of P nodes in
each D(xi). The length of each sub-space hi can be expressed by

x −x
h=
i
i _ up i _ low
(9)
M

For each level which has P nodes on it, there are M × n nodes in total. A
modification has been made to this algorithm by the author to the state vector of the
ant k that completed its tour, as shown in Fig. 2, with travel index [i8, i7, i6, ...,i4].
This index depends on the entries of cumulative probability (CP) from the
probability Pij of the ant k to move to the ith node on the jth level. For example, if M
= 10, CP = [0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, 0.9, 1.0] and the generated random
number lies between 0.7 and 0.8, the first travel index, i8, is chosen as 8 (eighth
column of the CP). This process is done until all travel indexes are found. The
values of the parameters X, held by ant k, are as

[ x , x , x …, x ] = [ x
1 2 3 n 1_ low
+i ×h, x8 1 2 _ low
+ i × h ,x
7 2 3_ low
(10)
+ i × h ,…, x
6 3 n_low
+i ×h] 4 n

Vol. 31 No. 2 2012 129


Active vibration control of flexible plate structures with distributed disturbances

The state transition rule of the ant k is expressed as

P =τij ij ∑τ i=1
ij
(11)

where Pij is the probability of the ant k to move to the ith node on the jth level, and
τij is the amount of the pheromone at the node. When all the ants have finished their
tours, the pheromone is updated by using

τ = (1 − ρ )τ + Q / f
ij ij best
(12)

where 0 < ρ < 1 is a pheromone decay parameter, Q is the quantity of pheromone


laid by an ant per iteration cycle, τ0 is a constant for the initial value of τij (for
initialization τij on the RHS is set to be τ0), and is the objective function value given
by the best ant of each searching period.

Fig. 2. State space graph for ACO1.

4. ACTIVE VIBRATION CONTROL DESIGN WITH DISTRIBUTED


SOURCES OF DISTURBANCE
4.1. MIMO-AVC
A schematic diagram of the geometric arrangement of a MIMO feedforward AVC
structure considered in this study is shown in Fig. 3 (a). The disturbance signals,
UD emitted by n primary point sources are detected by a set of n detectors. These
detection signals, UM are processed by the controller and fed to a set of k secondary
sources, UC. The aim of the controller design is to minimize the observed signals Y,
where Y =[y1, y2, ..., yk], via UC by generating anti-phase control signals to
counteract the vibration produced by UD. This can be achieved through optimization
of the assigned objective function using GA, PSO or ACO to obtain the best value
of the controller parameters.
Fig. 3 (b) shows a block diagram of the MIMO-AVC system [8], where E is an
n × n matrix representing transfer characteristics of the propagation paths
between the primary sources and the detectors, F is a k × n matrix representing
transfer characteristics of the paths between the secondary sources and the
detectors, G is an n × k matrix representing transfer characteristics of the paths
between the primary sources and the observers, H is a k × k matrix representing
transfer characteristics of the acoustic paths between the secondary sources and
the observers, M is an n × n diagonal matrix representing transfer characteristics
of the detectors, L is a k × k matrix representing transfer characteristics of the
secondary sources, UD is a 1 × n matrix representing the primary signals at the
source points, P0 is a 1 × k matrix representing the primary signals at the
observation points, UC is a 1 × k matrix representing the secondary signals at the
source points, S0 is a 1 × k matrix representing the secondary signals at the

130 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

observation points, UM is a 1 × n matrix representing the detected signals, and O is


a 1 × k matrix representing the combined primary and secondary signals at the
observation points.

Optimization

Y
C

UM UC
Detectors Secondary
sources

UD Observed
Primary sources signals

(a)

F
P0
+ + UM + +
E • M C L H •
UD UC S0 O

Optimization

(b)
Fig. 3. AVC structure for flexible plate: (a) schematic diagram, and (b) block diagram.

To suppress the vibration at the observation points, it requires the observed primary
and secondary signals at each observation point to be equal in magnitudes and have
a phase difference of 180˚:

S = −P
0 0
(13)

Using the block diagram in Fig. 3(b), P0 and S0 can be expressed as

P = PG , S = PEMCL[I − FMCL] H
0 0
−1
(14)

where I is the identity matrix. The required controller transfer function, C, can be
obtained by substituting equ. (14) into equ. (13), and simplifying as:

C = M ∆ GH L −1 −1 −1 −1
(15)

where ∆ is an n × n matrix given by

∆ = GH F − E −1
(16)

From equ. (15), the matrix dimension for C can be obtained as n × k, given by

Vol. 31 No. 2 2012 131


Active vibration control of flexible plate structures with distributed disturbances

 c  c  c
 11
 12 1k

 c c  c 
C= 21

22 2k

(17)
     
 c c  c 
 n1
 n2 nk

where cim(i = 1, 2, ..., n; m = 1, 2, ..., k) represents the controller transfer function


along the secondary path from detector i to secondary source m. This controller can
be realized in MIMO form as shown in Fig. 4. In this work, the number of primary
sources and secondary sources are set to n = 2 and k = 2, as presented in Fig. 5 (a).
The controller transfer function in equ. (17) then becomes

 c c 
= 
11 12
C (18)
 c 
MIMO
c
 21 22

where each transfer function is realized in a linear parametric form, as

p q

U (t ) = ∑ b ⋅ U (t − j ) − ∑ a ⋅ U (t − i )
C 11 j 11 M1 j 11 C 11
(19)
j =0 i=1

p q

U (t ) = ∑ b ⋅ U (t − j ) − ∑ a ⋅ U (t − i )
C 12 j 12 M1 j 12 C 12
(20)
j =0 i=1

p q

U (t ) = ∑ b ⋅ U (t − j ) − ∑ a ⋅ U (t − i )
C 21 j 21 M2 j 21 C 21
(21)
j =0 i=1

p q

U (t ) = ∑ b ⋅ U (t − j ) − ∑ a ⋅ U (t − i )
C 22 j 22 M2 j 22 C 22
(22)
j =0 i=1

where ai, bj are the controller parameters and q ≥ p represents the order of the
controller. The number of parameters to be estimated is (q + p + 1) × q × p.
Randomly selected controller parameters, i.e. a1,...,aq and b0, ..., bp are identified
for different, arbitrarily chosen orders to fit to the system. The stability of the
obtained controller must be ensured. In the discrete-time case, pole-zero diagram of
the corresponding controller transfer function provides a simple and effective means
of assessing its stability. Using, for example equ. (19), the transfer function of the
controller can be formed as

b + b z + b z + ... + b z
−1 −2 −p

C (z) = 011 111 211 p 11


(23)
11
1 + a z + a z + a z + ... + a z
111
−1
211
−2
311
−3
q 11
−q

Optimization techniques are carried out to achieve vibration reduction by feeding


the observed signal Y to the controller C via optimization routines. The
optimization process is based on minimizing the mean-squared error (MSE) of the
observed signal, Y. This is formulated as:

1
( )
S

∑ Y (i) + Y (i) + ... + Y (i) ; subject to –1 ≤ ai ≤ 1 and –1 ≤ bj ≤ 1 (24)


2

f error
( MIMO )
= 1 2 k
S i=1

where S represents the number samples.

132 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

4.2. MISO-AVC
Unlike MIMO-AVC system as mentioned above, the MISO-AVC system has more
than one input but only one output. The MISO-AVC structure for the flexible plate
with two primary sources is illustrated in Fig. 5 (b). The controller is designed to
minimize the deflection at observation point 1 via UC by generating anti-phase
control signal to counteract the vibration produced by UD. For this case the number
of primary sources and secondary source are set to n = 2 and k = 1. The controller
transfer function in equ. (17) then becomes

 c 
=  (25)
11
C
 c 
MISO

 21

where each of the transfer functions is realized in a linear parametric form as

p q

U (t ) = ∑ b ⋅ U (t − j ) − ∑ a ⋅ U (t − i )
C 11 j 11 M1 j 11 Ci 1
(26)
j =0 i=1

p q

U (t ) = ∑ b ⋅ U (t − j ) − ∑ a ⋅ U (t − i )
C 21 j 21 M2 j 21 C 21
(27)
j =0 i=1

The number of parameters to be estimated here is (q + p + 1) × n. The transfer


function of controller can be formed from equ. (26) and (27). The performance of
the MISO-AVC system can be assessed with the mean-squared error (MSE) of the
observed signal, Y. The objective function is thus used is

1
( )
S

∑ Y (i) ; subject to –1 ≤ ai ≤ 1 and –1 ≤ bj ≤ 1


2

f error
( MISO )
= (28)
S i=1

Observation
UM1 + + point-1
C11 C12 C1k
+

Observation
UM2 + + point-2
C21 C22 C2k
+

Observation
UMn + point-k
+
Cn1 Cn2 Cnk
+

Fig. 4. MIMO controller.

Vol. 31 No. 2 2012 133


Active vibration control of flexible plate structures with distributed disturbances

UM1 UC11 + Secondary


C11 C12 signal 1 UM1 UC11
C11
UC21 + +
Secondary
UC12 signal 1
+
UM2
UM2 UC22 + Secondary C21
C21 C22 signal 2 UC21
+

(a) (b)
Fig. 5. AVC with multiple-input: (a) multiple-output, and (b) single-output.

4.3. System geometry


The effectiveness of AVC for controlling the vibration of a flexible structure is
directly affected by the geometrical arrangement of the system components:
detection, observation and secondary source points. A typical implementation
comprises a sensor, to detect the vibration mode to be suppressed, and control
electronics driving actuators to generate cancelling sources to interfere destructively
with the unwanted source from primary point. To control the vibration on the
structure requires that the actuators are optimally situated. In this work, two
mechanisms of geometrical arrangement for detection and secondary source points
are investigated, variable geometry and fixed geometry. For both mechanisms,
points of primary and observation for both cases are kept at specific mesh-points.

4.3.1. Variable geometry


In this mechanism, the locations for the detection and secondary source points as
well as the parameters of the controller are determined by the global search
technique of GA, PSO, and ACO. Since the algorithms use stochastic random
searches, there are a few possibilities that need to be considered. In MIMO- and
MISO-AVC cases, the algorithms may converge to primary and observation points
for the detection and secondary source points, or to the same point for detection
point 1 and secondary source 2, for example. To ensure such situations are avoided,
penalty has been added to the algorithms. On the other hand, when the location of
detection point 1 is the same as secondary point 1 or detection point 2 is the same
as secondary point 2 (MIMO case), or secondary source 1 is the same as detection
point 1 or detection point 2 (MISO case), such a situation is allowable. This is
known as collocation, where the advantage is reduction of the required space to
install actuator and sensor in mechanical design of the system [37].
Given that the plate is divided into 20 × 20 sections and the deflection along the
boundaries is zero, the range of coordinates for the detection and secondary source
points must be within 1 and 19. Here, about possible detection and secondary
source locations are to be explored to achieve the highest degree of effectiveness to
produce the lowest value of the objective function. However, after several
investigations, it was found that points located in the smaller range [4, 16] have
produced more vibration attenuation than the range [1,19]. For this reason, the
execution of the RCGA-AVC, PSO-SF-AVC, and ACO1-AVC of the flexible plate
are used to find these points in the range of [4, 16], as shown in the white area of
the plate section in Fig. 6. The optimization procedures are described as below:

(i) RCGA-AVC
step 1 Initialize parameters of the flexible plate and location of disturbance
points: P1 = (7, 7) is primary point 1, P2 = (10, 6) is primary point 2,and
observation points; O1 = (9, 14) is observation point 1 (MISO and MIMO
cases), and O2 = (13,8) is observation point 2 (MIMO case only).
step 2 Initialize range of variables for optimal location of detection and
secondary source points, in the range , and range of variables for
controller parameters, in the range [-1, 1].
step 3 Generate random population of X chromosomes where the initial
chromosome for ith individual is represented as below:

134 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

MISO case:

X = ( d , d , s , s , d , d , a ,…, a
i i (1) i(2) i ( 3) i(4) i ( 5) i(6) i(7) i ( q × k +6 )
,b i ( q × k +7 )
,…, b i (( p + q )× k +8 )
) (29)

where the first six variables are coordinates for detection point 1, D1 = (di1, di2) and
secondary source point 1, S1 = (si3, si4), and detection point 2, D2 = (di5, di6). The
remaining variables are the controller parameters to form transfer functions as in
equ. (26) and (27).

MIMO case:

X = ( d , d , s , s , d , d , s , s , a ,…, a
i i (1) i(2) i ( 3) i(4) i ( 5) i(6) i(7) i(8) i(9) i (( p + q )× k +10 )
,
(30)
b i (( p + q )× k +11)
,…, b i (( p + q )× k × n+12 )
)

where the first eight variables are coordinates for detection point 1, D1 = (di1, di2),
secondary source point 1, S1 = (si3, si4), and detection point 2, D2 = (di5, di6), and
secondary source point 2, S1 = (si3, si4), and detection point 2, S2 = (si7, si8). The
remaining variables are the controller parameters to form transfer functions as in
equ. (19) – (22).

step 4 Add penalty to Xiif if any of the following conditions occur:

( d , d ) + penalty, if ( d , d ) = P or ( d , d ) = P or

i1 i2 i1 i2 1 i1 i2 2

 ( d , d ) = O or ( d , d ) = O or
D = i1 i2 1 i1 i2 2
(31)
( d , d ) = D or ( d , d ) = S
1

 i1 i2 2 i1 i2 2

( d , d ),
 i1 i2
otherwise

( s , s ) + penalty, if ( s , s ) = P or ( s , s ) = P or

i3 i4 i3 i4 1 i3 i4 2

 ( s , s ) = O or ( s , s ) = O or i3 i4 1 i3 i4 2

S = [( s , s ) = S or ( s , s ) = D ]

1

 i3 i4 2 i3 i4 2


MIMO case only
(32)
( s , s ),
i3 i4
otherwise

step 5 Evaluate the fitness value, i.e. MSE of the observed signal, for each
chromosome in the population according to equ. (24) or (28).
step 6 Create a new population by repeating the following steps until the new
population is complete:
6.1. For selection operator, select two parent chromosomes from a
population using SUS algorithm, according to their fitness.
Parents that produced the smallest MSE value are considered as
the best fit individuals.
6.2. For crossover operator, cross over the parents to form a new
offspring (children) according to equ. (3). If no crossover was
performed, offspring is an exact copy of parents.
6.3. For mutation operator, mutate the new offspring to alter the genes
of some of the children, according to the probability of mutation,
pm.

Vol. 31 No. 2 2012 135


Active vibration control of flexible plate structures with distributed disturbances

step 7 Add penalty to the offspring as in step 4.


step 8 Evaluate the fitness values of all offspring.
step 9 Reinsert the offspring into the original population using fitness-based-
reinsertion, with the new offspring replacing the least fit member of the
original population.
step 10 If the specified maximum number of generation is not reached repeat
step 6 until convergence, else, end the algorithm.
step 11 Output the values of the decision variables for final evaluation of the
AVC system. End of procedure.

(ii) PSO-SF-AVC
step 1-2 Same as in RCGA-AVC procedure.
step 3 Initialize the population of particles with random positions and
velocities, and maximum iteration, where the initial position vector for
ith particle is represented as in step 3 in RCGA-AVC procedure.
step 4 Same as RCGA-AVC procedure.
step 5 Evaluate the fitness value, i.e. MSE of the observed signal, for each
particle of the population according to equ. (24) or (28). The best
particle that produces the smallest fitness value with its position will be
stored as global best position.
step 6 Calculate spread, deviation, SF, ω, c1, and, c2 as in equ. (6), (7), and (8),
for all the variables and particles.
step 7 Update the position and velocity of particle according to equ. (4) and
(5).
step 8 Add penalty to the updated position as in step 4.
step 9 Evaluate the fitness values of all particles and determine the best
particle for the current iteration. If the MSE observed signal is smaller
than the MSE value of the global best position, then update the global
best position and its MSE value with the position and MSE value of the
current best particle.
step 10-11 Same as RCGA-AVC procedure

(iii) ACO1-AVC
step 1-2 Same as RCGA-AVC procedures.
step 3 Initialize randomly the tours of ants within the predefined range of
variables, initial pheromone, initial probability, and maximum iteration,
where the initial tour vector for kth ant is represented as in step 3 in
RCGA-AVC procedure.
step 4 Same as RCGA-AVC procedure.
step 5 Evaluate the initial fitness value, i.e. MSE of the observed signal, for
each ant of the colony according to equ. (24) or (28). The best ant that
produces the smallest fitness value with its vector will be stored as the
best vector.
step 6 Evaluate the MSE value of every ant using the solution vector
constructed from equ. (10). Add penalty to the updated position as in
step 4. If the MSE value of the observed signal is smaller than the MSE
value of the best vector, then update the best vector and its fitness value.
Repeat until all ants finish the tour.
step 7 Update the pheromone and calculate the probability of the ant k to
move to the ith node on the jth level as in equ. (11) and (12), for all the
variables and ants. The probability constructed here will be used in the
next iteration.
step 8-9 Same as RCGA-AVC procedures.

4.3.2. Fixed geometry


For the case of fixed geometry, the locations of detection and secondary source
points are fixed. This is done after variable-geometry-mechanism has been carried

136 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

out for all the algorithms for both MISO and MIMO cases. Since all the algorithms
are stochastic search, the solutions obtained are different from one another, giving
more than one possible location for the search points. Therefore, in order to choose
the best locations, the levels of attenuation using different algorithms are compared
and the searched points giving the maximum mean of the attenuation for the first
five dominant vibration modes are used. Here, the optimization process is carried
out only on the parameters of the controller so as to minimize the objective function
defined in equ. (24) for the MIMO case, or equ. (28) for the MISO case.

Fig. 6. Predefined region of arrangement for systems components.

5. IMPLEMENTATION AND RESULTS


This section presents an assessment of the performances of the MISO- and MIMO-
AVC systems with RCGA, PSO-SF, and ACO1 design approaches. The random
disturbance signal was applied to the plate at t = 0.2 to t = 0.5 sec. For the detection
and secondary source points within the predefined range, it was observed that the
first five dominant modes of the plate were at 10.35 rad/sec, 33.76 rad/sec, 64.94
rad/sec, 81.88 rad/sec, and 99.37 rad/sec. The results presented in this section are
assessed on a comparative basis in terms of spectral density attenuation for both
cases of variable-geometry-AVC and fixed-geometry-AVC with RCGA-, PSO-SF-,
and ACO1-optimization techniques.

5.1. Simulation results for MISO-AVC


5.1.1. Variable-geometry
A simulation environment was developed using m-file in MATLAB software. The
simulation was carried out to find the optimal locations of detection (4 variables)
and secondary source (2 variables) points together with controller parameters (8
variables). The best (minimum) MSE levels obtained using RCGA-AVC, PSO-SF-
AVC, and ACO1-AVC are as shown in Fig. 7 and Table II.
Since the algorithms are stochastic in nature, the solutions found may vary at
each run. Stability of the algorithm was carried by running the program 10 times
with the corresponding best parameters for each algorithm. The results tabulated in
Table II are based on statistical performance measures, i.e. minimum, maximum,
mean, and standard deviation of the objection function. The smallest MSE value
was obtained using ACO1-AVC. It is noted that RCGA achieved consistent MSE
values over multiple runs, followed by ACO1 and PSO-SF.
It is observed from Fig. 7 that the MSE convergence for all algorithms was large
for iterations less than 5, and rapid convergence was achieved in the first 10
iterations, where the algorithms tried to locate the optimum region for detection and
secondary source points. However, no or minor improvement was made after 30
iterations. As shown in Fig. 8, the computational time for RCGA was significantly
lower than PSO-SF and ACO1 mechanisms.

Vol. 31 No. 2 2012 137


Active vibration control of flexible plate structures with distributed disturbances

Fig. 9 shows the performance results of RCGA-AVC, PSO-SF-AVC, and ACO1-


AVC. The best controller order that produced the smallest MSE was 2 for all
algorithms. All algorithms were run with 20 individuals/particles/ants and
maximum generation/iteration of 40. For RCGA-AVC, the specific parameters used
were generation gap of 0.80, crossover rate of 0.67, and mutation rate of 0.0714. For
ACO1-AVC, the initial parameters used were τ0 0.01, ρ = 0.99, M = 10 and Q = 10.
It is noted that all algorithms significantly reduced vibration at the first mode
with minor reduction or reinforcements at the remaining modes. The amounts of
cancellation/reinforcement with the corresponding best location obtained with the
optimization techniques are summarized in Table III. It is noted that ACO1-AVC
has achieved the highest mean attenuation followed by RCGA-AVC, and PSO-SF-
AVC. Therefore, the best locations for detection and secondary source points were
found with ACO1-AVC at D1(11, 7), D2(6, 10), and S1(7, 6).
These three locations obtained were nearer to the primary source 1 and primary
source 2, within the range [6, 11]. The distance between the primary source and the
detection point was thus shorter, enabling the sensor to detect most of the dynamic
characteristics of the disturbance signal for the controller. The shorter distance
between primary and secondary source results in shorter travelling path between
cancelling and unwanted signals, thus enhancing the vibration suppression. Thus, in
this case, the physical extent of vibration cancellation around the observation point
is higher when the detection and secondary source points are located close to the
primary source.

Table II
MSE value for 10 trials for variable-geometry MISO-AVC.
MSE value
Algorithm
Minimum Maximum Mean Standard deviation
RCGA 14 14 14 15
2.93 10 6.68 10 4.98 10 9.26 10
PSO-SF 14 13 13 13
4.18 10 7.02 10 1.50 10 2.02 10
ACO1 14 13 14 14
1.95 10 1.44 10 5.41 10 3.57 10

-8
x 10
8
MISO-RCGA-AVC
MISO-PSO-SF-AVC
6 MISO-ACO1-AVC
-14
x 10
MSE

4 6
4
2 2
20 30 40
0
0 10 20 30 40
No. of iterations

Fig. 7. Convergence graph for variable-geometry MISO-AVC

150
MISO-RCGA-AVC
MISO-PSO-SF-AVC
MISO-ACO1-AVC
100
Time (min)

50

0
0 10 20 30 40
No. of iterations

Fig. 8. Computational time for MISO-AVC with RCGA, PSO-SF, and ACO1.

138 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

Table III
Spectral attenuation achieved at resonance modes with variable-geometry MISO-AVC.
Spectral attenuation (dB) Mean
Algorithm Position
Mode 1 Mode 2 Mode 3 Mode 4 Mode 5 (dB)
RCGA D1 (13, 6), D2 (8, 9) , S1 (7, 6) 22.63 2.93 -2.80 4.93 0.55 5.65
PSO-SF D1 (9, 8), D2 (10, 10) S1(7, 9) 24.23 0.87 -4.84 0.28 -1.54 3.80
ACO1 D1 (11, 7), D2 (6, 10) , S1(7, 6) 34.57 4.54 -3.56 2.51 1.05 7.82

Note: Negative value indicates spectral reinforcement.

-6
x 10 uncontrolled uncontrolled
RCGA-controlled RCGA-controlled
PSO-SF-controlled -60 PSO-SF-controlled
1.5 ACO1-controlled ACO1-controlled
-70
1

Magnitude (dB)
-80
Deflection (m)

0.5 -90
0 -100

-0.5 -110

-1 -120

-130
0 1 2 3 4 0 50 100 150
Time (sec) Frequency (rad/sec)
(a) Time-domain at observation point (b) Spectral density at observation point

Fig. 9. Performance of variable-geometry MISO-AVC.

5.1.2. Fixed-geometry
The best location obtained for detection and secondary source points as mentioned
in section 5.1.1, defined as fixed-geometry, are used here where the algorithms will
only find the controller parameters (8 variables) to minimize the objective function.
The best (minimum) MSE levels obtained using RCGA-AVC, PSO-SF-AVC, and
ACO1-AVC are as shown in Fig. 10 and Table IV.
The simulation was run 10 times for each algorithm to check its stability in
searching the controller parameters. From the results presented in Table IV, it is
noted that the minimum MSE values was obtained by PSO-SF-AVC. It is noted
from the values of maximum, mean, and standard deviation that the algorithms
yielded consistent MSE value over multiple runs. The standard deviation values
were significantly lower than these in variable-geometry MISO-AVC, indicating
that they are clustered closely around the mean value, i.e. the algorithms are stable
in finding the controller parameters.
The level of vibration reduction achieved at the observation points is
demonstrated with the time-domain responses and spectral densities of the system
responses before and after cancellation of RCGA-, PSO-SF-, and ACO1-AVC
performances, as shown in Fig. 11. The deflection throughout the plate is further
shown with the 3-dimensional plots of system response before and after control at
t = 0.5 sec. It is noted that significant reduction in plate vibration was achieved. The
spectral attenuations achieved are summarized in Table V. All algorithms achieved
spectral attenuation at the first, second, third, and fifth modes. At the fourth mode,
in contrast, minor reinforcement occurred. The mean spectral attenuations obtained
with RCGA-, PSO-SF-, and ACO1-AVC were 8.65 dB, 9.93 dB, and 8.85 dB,
respectively.
All controllers were stable as revealed by the pole-zero plots in Fig. 12 where all
the poles were inside the unit circle. Furthermore, the controlled signal in time
domain at both observation points was at consistent level from 0.3 sec to 4 sec. The
discrete transfer functions of the controllers at a sampling time of 0.0016 sec thus
obtained are

Vol. 31 No. 2 2012 139


Active vibration control of flexible plate structures with distributed disturbances

 −0.199 z + 0.179 0.011z − 0.171


C (z) =  C = C = (33)
RCGA
 z + 0.301z + 0.102
11 2
z + 0.129 z − 0.251 21 2

 0.019 z − 0.244 −0.016 z + 0.091 (34)


C (z) =  C = C =
PSO-SF
 z − 0.107 z + 0.049
11 2
z + 0.124 z + 0.233 21 2

 0.046 z − 0.103 −0.021z − 0.141 (35)


C (z) =  C = C =
ACO1
 z + 0.211z + 0.183
11 2
z + 0.188 z − 0.214 21 2

Table IV
MSE value for 10 trials for fixed-geometry MISO-AVC.
MSE value
Algorithm
Minimum Maximum Mean Standard deviation
RCGA 14 14 14 16
1.90 10 2.11 10 1.99 10 6.09 10
PSO-SF 14 14 14 15
1.79 10 2.21 10 1.95 10 1.23 10
ACO1 14 14 14 16
1.88 10 2.15 10 1.97 10 7.73 10

Table V
Spectral attenuation achieved at resonance modes with fixed-geometry MISO-AVC.
Spectral attenuation (dB) Mean
Algorithm
Mode 1 Mode 2 Mode 3 Mode 4 Mode 5 (dB)
RCGA 29.13 4.63 8.64 -0.70 1.57 8.65
PSO-SF 32.36 4.35 14.81 -2.74 0.89 9.93
ACO1 31.63 4.70 6.38 -0.01 1.57 8.85

Note: Negative value indicates spectral reinforcement.

-13 MISO-RCGA-AVC
x 10
2 MISO-PSO-SF-AVC
MISO-ACO1-AVC

1.5 -14
x 10
2.2
MSE

1 2

1.8
0.5 10 20 30 40

0 10 20 30 40
No. of iterations
Fig. 10. Convergence graph for fixed-geometry MISO-AVC.

140 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

-6 uncontrolled uncontrolled
x 10
RCGA-controlled RCGA controlled -7
PSO-SF-controlled -60 PSO-SF-controlled x 10
1.5 ACO1-controlled ACO1-controlled
-70
10

Magnitude (dB)
1

Deflection (m)
-80

Deflection (m)
0.5 -90 5

0 -100
0
-0.5 -110

-1 -120 20
20
-130 10 15
0 1 2 3 4 0 50 100 150 10
5
Time (sec) Frequency (rad/sec) y (j) x (i)

(a) Time-domain at observation point (b) Spectral density at (c) System response before
observation point cancellation
-7 -7 -7
x 10 x 10 x 10

10 10 10
Deflection (m)

Deflection (m)

Deflection (m)
5 5 5

0 0 0

20 20 20
20 20 20
10 15 10 15 10 15
10 10 10
5 5 5
y (j) x (i) y (j) x (i) y (j) x (i)

(d) System response after (e) System response after (f) System response after
cancellation for RCGA-AVC cancellation for PSO-SF-AVC cancellation for ACO1-AVC

Fig. 11. Performance of fixed-geometry MISO-AVC.

(a) Pole zero plot C(z)11 (b) Pole zero plot C(z)21 (c) Pole zero plot C(z)11
for RCGA for RCGA for PSO-SF

(d) Pole zero plot C(z)21 (e) Pole zero plot C(z)11 (f) Pole zero plot C(z)21
for PSO-SF for ACO1 for ACO1
Fig. 12. Pole-zero plots of the fixed-geometry MISO-AVC controllers.

5.2. Simulation results for MIMO-AVC


The control structure in this case comprises two excitation points (P1 and P2), two
detection points (D1 and D2), two secondary source points (S1 and S2) and the
vibration is observed at two observation points (O1 and O2). For the detection and
secondary source points within the predefined range, it was observed that the first
five dominant modes of the plate were at 10.35 rad/sec, 33.76 rad/sec, 64.94 rad/sec,
81.88 rad/sec, and 99.37 rad/sec. The same parameters for the algorithms as in
section 5.1.1 were used

5.2.1. Variable-geometry
The simulation was carried out to find the optimal locations of detection (4 variables)
and secondary source (4 variables) points together with controller parameters (16
variables). The best (minimum) MSE levels obtained using RCGA-AVC, PSO-SF-
AVC, and ACO1-AVC are shown in Fig. 13 and Table VI. The minimum MSE value
was obtained using ACO1-AVC. From the maximum, mean, and standard deviation
values, it is noted that the algorithms have performed consistently at similar levels.
A similar behaviour as with variable-geometry MISO-AVC can be observed where
RCGA achieved consistent MSE values over multiple runs, followed by ACO1 and

Vol. 31 No. 2 2012 141


Active vibration control of flexible plate structures with distributed disturbances

PSO-SF. The MSE values for all algorithms were large number for iterations less
than 5, and rapid convergence was made in the first 10 iterations. For most
algorithms, no or minor improvement was made after 30 iterations. In terms of the
computational times, RCGA has performed faster than ACO1 and PSO-SF.
Fig. 14 shows the performance results of RCGA-AVC, PSO-SF-AVC, and
ACO1-AVC. The best controller order that produced the smallest MSE was 2 for all
algorithms. The results of attenuation/reinforcement achieved with the
corresponding best locations of the detection and secondary source points are listed
in Table VII. It is observed that ACO1-AVC has achieved the highest mean
attenuation followed by RCGA-AVC, and PSO-SF-AVC.
As summarized in Table VII, the maximum mean attenuation achieved was
achieved using ACO1-AVC with best locations at D1(7, 10), D2(11, 7), S1(6, 7), and
S2(8, 8). All the four points were located within the range , nearer to the primary
source 1 and primary source 2, which was similar to variable-geometry-MISO-AVC
case. It can be concluded that the short distance between the primary source and the
detection point enables the sensor to detect most of the dynamic characteristics of
the disturbance signal, while short distance between primary and secondary source
results in shorter travelling path between cancelling and unwanted signals, thus
enhancing the vibration suppression. Thus, in this case, the physical extent of
vibration cancellation around the observation point is higher when the detection and
secondary source points are located closer to the primary source. Again, the physical
extent of vibration cancellation around the observation point is higher when the
detection and secondary source points are located closer to the primary source.

Table VI
MSE value for 10 trials for variable-geometry MIMO-AVC.
MSE value
Algorithm
Minimum Maximum Mean Standard deviation
14
RCGA 9.49 10 5.41 10 13 1.87 10 13 1.51 10 13
13 12 13
PSO-SF 1.02 10 1.80 10 6.20 10 6.62 10 13
14 13 13
ACO1 6.29 10 6.32 10 2.57 10 1.91 10 13

Table VII
Spectral attenuation achieved at resonance modes with variable-geometry MIMO-AVC.

Algorithm Spectral attenuation (dB) Mean


Position
(observation point) Mode 1 Mode 2 Mode 3 Mode 4 Mode 5 (dB)
RCGA (1) 26.19 8.64 -0.57 1.45 -2.41 6.66
D1 (10, 9), D2 (8, 11)
S1 (7, 8), S2 (9, 7) RCGA (2) 26.14 4.94 3.88 1.62 -3.05 6.71
PSO-SF (1) 23.65 5.26 -9.32 -1.27 -4.45 2.77
D1 (11, 8), D2 (10,9)
S1 (8, 6), S2 (7, 11) PSO-SF (2) 23.52 1.91 -5.12 -3.17 -5.00 2.43
ACO1 (1) 32.13 18.00 -4.47 5.83 -0.47 10.20
D1 (7, 10), D2 (11, 7)
S1 (6, 7), S2 (8, 8) ACO1 (2) 32.96 15.69 -1.38 3.20 -0.97 9.90

Note: Negative value indicates spectral reinforcement.

-7
x 10
MIMO-RCGA-AVC
1 MIMO-PSO-SF-AVC
MIMO-ACO1-AVC
0.8 -13
x 10
MSE

0.6 1.4
1.2
0.4 1
0.8
0.2 0.6
20 30 40
0
0 10 20 30 40
No. of iterations
Fig. 13. Convergence graph for variable-geometry MIMO-AVC.

142 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

-6 uncontrolled uncontrolled
x 10
RCGA-controlled RCGA-controlled
PSO-SF-controlled -60 PSO-SF-controlled
1.5 ACO1-controlled ACO1-controlled
-70
1

Magnitude (dB)
-80

Deflection (m)
0.5 -90
0 -100

-0.5 -110

-1 -120

-130
0 1 2 3 4 0 50 100 150
Time (sec) Frequency (rad/sec)
(a) Time-domain at observation point-1 (b) Spectral density at observation point-1
-6 uncontrolled uncontrolled
x 10
2 RCGA-controlled RCGA-controlled
PSO-SF-controlled -60 PSO-SF-controlled
1.5 ACO1-controlled ACO1-controlled

1 -80

Magnitude (dB)
Deflection (m)

0.5
-100
0

-0.5
-120
-1

-1.5 -140
0 1 2 3 4 0 50 100 150
Time (sec) Frequency (rad/sec)
(c) Time-domain at observation point-2 (d) Spectral density at observation point-2
Fig. 14. Performance of variable-geometry MIMO-AVC.

5.2.2. Fixed-geometry
The best locations obtained for detection and secondary source points in 5.2.1,
defined as fixed-geometry, are used here where the algorithms will only find the
controller parameters that minimize the objective function. The convergence graphs
thus obtained are shown in Fig. 15 and listed in Table VIII. It is noted that the
minimum MSE values was obtained by ACO1-AVC. From the values of maximum,
mean, and standard deviation for the algorithms, it is noted that when the locations
of all points are fixed, the algorithms yield consistent MSE value over multiple runs.
The standard deviation values are significantly lower those that in variable-
geometry MIMO-AVC, indicating that they are clustered closely around the mean
value, i.e. the algorithms are stable in finding the controller parameters.
The system performance results with RCGA-, PSO-SF-, and ACO1-AVC are
shown in Fig. 16 and summarized in Table IX. The level of vibration reduction
achieved at the observation points is demonstrated with the time-domain responses
and spectral densities of the system responses before and after cancellation. The
deflection throughout the plate is further shown with the 3-dimensional plots of
system response before and after control. Vibration at most of the dominant modes
was reduced at both observation points, except a minor reinforcement at the third
mode at the observation point 1 with RCGA-AVC. The mean attenuations achieved
were 9.19 dB and 9.17 dB, at observation point 1 and point 2, respectively. For PSO-
SF-AVC, it is noted that there were a few minor reinforcements at both points. The
mean spectral attenuations achieved at observation point 1 and observation point 2
were 10.91 dB and 9.63 dB, respectively. For ACO1-AVC, the mean spectral
attenuations achieved at observation point 1 and observation point 2 were 10.35 dB
and 10.83 dB, respectively. There were few minor reinforcements at both points.
The controlled signal in time domain at both observation points has shown stable
vibration amplitude from 0.3 sec to 4 sec. The discrete transfer functions of the
controllers at a sampling time of 0.0016 sec thus obtained are given below, which
are stable as all the poles inside the unit circle, as shown in Fig. 17.

Vol. 31 No. 2 2012 143


Active vibration control of flexible plate structures with distributed disturbances

 −0.792 z − 0.793 0.044 z − 0.434 


 C = C = 
 z − 0.272 z − 0.379
11 2
z − 0.595z − 0.343 12 2

C (z) =  
RCGA
  (36)
 C = −0.839 z + 0.169 C =
0.478 z − 0.354 
 z − 0.701z − 0.131
21 2
z + 0.358 z + 0.996 22 2


 −0.301z + 0.099 −0.301z + 0.256 


 C = C =
 z + 0.268 z + 0.292
11 2 12
z + 0.295z − 0.301
2

C (z) = 
PSO-SF
 (37)
 C = −0.301z + 0.299 C =
−0.299 z − 0.298
 z + 0.045z − 0.267
21 2
z + 0.207 z + 0.167 
22 2

 −0.271z + 0.121 0.059 z − 0.159 


 C = C = 
 z + 0.065z − 0.024
11 2
z + 0.115z − 0.101 12 2

C (z) =  
ACO1
  (38)
 C = −0.119 z − 0.014 C =
−0.165z − 0.070 
 z + 0.216 z − 0.276
21 2
z + 0.059 z + 0.167 22 2


Table VIII
MSE value for 10 trials for fixed-geometry MIMO-AVC.
MSE value
Algorithm
Minimum Maximum Mean Standard deviation
RCGA 6.17 10 14 7.05 10 14 6.68 10 14 3.30 10 15
14 14 14
PSO-SF 6.67 10 7.43 10 6.97 10 2.10 10 14
14 14 14
ACO1 6.15 10 7.12 10 6.63 10 4.12 10 14

Table IX
Spectral attenuation achieved at resonance modes with fixed-geometry MIMO-AVC.

Algorithm Spectral attenuation (dB) Mean


(observation point) Mode 1 Mode 2 Mode 3 Mode 4 Mode 5 (dB)
RCGA (1) 32.08 12.57 -1.59 2.11 0.79 9.19
RCGA (2) 33.17 9.08 2.04 1.28 0.29 9.17
PSO-SF (1) 37.70 15.41 -3.05 5.23 -0.74 10.91
PSO-SF (2) 38.01 12.36 -0.19 -0.82 -1.21 9.63
ACO1 (1) 35.73 14.61 -5.09 7.19 -0.71 10.35
ACO1 (2) 36.85 12.24 -1.76 8.08 -1.24 10.83

Note: Negative value indicates spectral reinforcement.

144 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

x 10
-12 MIMO-RCGA-AVC
6 MIMO-PSO-SF-AVC
MIMO-ACO1-AVC
5

4 x 10
-14

MSE
3 7
6.8
6.6
2 6.4
6.2
1 20 30 40

0
0 10 20 30 40
No. of iterations

Fig. 15. Convergence graph for fixed-geometry MIMO-AVC.

-6 uncontrolled uncontrolled
x 10 RCGA-controlled RCGA-controlled
PSO-SF-controlled -60
PSO-SF-controlled
1.5 ACO1-controlled ACO1-controlled
-80

Magnitude (dB)
1
Deflection (m)

0.5
-100
0

-0.5 -120
-1
-140
0 1 2 3 4 0 50 100 150
Time (sec) Frequency (rad/sec)

(a) Time-domain at observation point-1 (b) Spectral density at observation point-1


-6 uncontrolled uncontrolled
x 10 RCGA-controlled RCGA-controlled
2 PSO-SF-controlled -60 -7
PSO-SF-controlled x 10
ACO1-controlled ACO1-controlled
1.5
-80 10
Magnitude (dB)

1
Deflection (m)
Deflection (m)

0.5 5
-100
0
0
-0.5
-120
-1 20
20
10 15
-140 10
0 1 2 3 4 0 50 100 150 5
Time (s) Frequency (rad/sec) y (j) x (i)

(c) Time-domain at observation (d) Spectral density at observation (a) System response before
point-2 point-2 cancellation
-7 -7 -7
x 10 x 10 x 10

10 10 10
Deflection (m)

Deflection (m)

Deflection (m)

5 5 5

0 0 0

20 20 20
20 20 20
10 15 10 15 10 15
10 10 10
5 5 5
y (j) x (i) y (j) x (i) y (j) x (i)

(b) System response after (c) System response after (d) System response after
cancellation for RCGA-AVC cancellation for PSO-SF-AVC cancellation for ACO1-AVC

Fig. 16. Performance of fixed-geometry-MIMO-AVC.

Vol. 31 No. 2 2012 145


Active vibration control of flexible plate structures with distributed disturbances

(a) Pole-zero plot (b) Pole-zero plot (c) Pole-zero plot


C(z)11_RCGA C(z)12_RCGA C(z)21_RCGA

(e) Pole-zero plot (f) Pole-zero plot (g) Pole-zero plot


C(z)11_PSOSF C(z)12_PSOSF C(z)21_PSOSF

(i) Pole-zero plot (j) Pole-zero plot (k) Pole-zero plot


C(z)11_ACO1 C(z)12_ACO1 C(z)21_ACO1
Fig. 17. Pole-zero plots of the fixed-geometry MIMO-AVC controllers.

5.3. Comparative assessment of system performances


From the results presented in the previous sections, ACO1-AVC has successfully
achieved the highest mean spectral attenuation both in variable-geometry-MISO-
and -MIMO-AVC. For fixed-geometry case, PSO-SF-AVC has achieved the highest
mean spectral attenuation with MISO-AVC, while ACO1-AVC has performed well
with MIMO-AVC. Thus, ACO1 has performed well at finding optimal controller
parameters, outperformed RCGA and PSO-SF for suppressing the vibration of the
flexible plate. The best location for detection and secondary source points were
found to be within the x- and y-section of on the flexible plate. This indicates that
the vibration can be controlled better if detection and secondary source(s) points are
located nearer to the primary source.
As presented in Tables II and VI, the standard deviations of the MSE value for
variable-geometry cases were relatively higher than those in fixed-geometry cases
shown in Table IV and VIII. In variable-geometry, MISO-AVC case utilized 14
design variables (three sets of (x, y) coordinates for detection and secondary source
points and 8 controller parameters) while in MIMO-AVC case, 24 design variables
(four sets of (x, y) coordinates for detection and secondary source points and 16
controller parameters) were utilized. In contrast, MISO- and MIMO-AVC fixed
geometry cases dealt with fewer design variables, i.e. 8 for MISO- and 16 for
MIMO-AVC, resulting in consistent value of MSE and relatively low standard
deviation. The large number of design variables has contributed to the high standard
deviation in variable-geometry. Also, the effect of the location of detection and
secondary source points has significant influence on the performance of the
controlled system.
In terms of level spectral attenuation achieved as listed in Tables III and V,
variable-geometry-AVC systems have shown significant vibration reduction with all
the systems at the first dominant mode for all the disturbances. As presented in
Tables V and IX, the vibration reduction obtained with fixed-geometry-AVC

146 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

systems at this mode was considerably higher than those in variable-geometry cases.
This is followed by minor attenuation or reinforcement at the other modes which has
insignificant effect to the overall reduction in vibration level. Considering that most
of the vibration energy is in the first dominant mode, attenuation obtained within
22.63 – 34.57 dB with variable-geometry MISO-AVC, 29.13 – 32.36 dB with fixed-
geometry MISO-AVC, 23.52 – 32.96 dB with variable-geometry MIMO-AVC, and
32.08 – 38.01 dB with fixed-geometry MIMO-AVC system, at this mode reduced
the system vibration considerably.
It is noted that the performance of the system with the MIMO controller was
significantly superior to MISO controller. The improvement in mean spectral
attenuation by utilizing a multiple set of cancelling sources, i.e. MIMO case, for
RCGA-AVC, PSO-SF-AVC, and ACO1-AVC, at observation point 1 was 6.24%,
9.87%, and 16.95%. These improvements have shown that the utilization of
multiple set of cancelling sources enhances the performance of the system in terms
of vibration reduction.

6. CONCLUDING REMARKS
In this paper, the development of an AVC strategy with GA, PSO, and ACO
techniques with multiple inputs has been presented. The controllers have been
designed based on direct optimization of the location of the detector and secondary
source, and the controller parameters based on minimizing the MSE level of the
error (observed) signal. The approach does not require knowledge of the dynamic
characterization of the plant for controller adaptation. The performances of GA,
PSO, and ACO based AVC systems have been assessed within the simulation
environment of a flexible plate system. It has been demonstrated that all the
developed control systems perform successfully in suppressing the vibration of the
system. The utilization of multiple set of cancelling sources in MIMO case has been
shown to enhance the performance of the system in terms of vibration reduction
compared to MISO case. Overall, the results indicate that GA, PSO, and ACO
algorithms can be used effectively for optimal placement of system components and
optimization of controllers for vibration suppression in flexible structures.

ACKNOWLEDGMENTS
S. Julai acknowledges the financial support of the Ministry of Higher Education
Malaysia and University of Malaya, Kuala Lumpur, Malaysia.

REFERENCES
[1] Hu Q. (2009) A composite control scheme for attitude maneuvering and
elastic mode stabilization of flexible spacecraft with measurable output
feedback. Journal of Aerospace Science and Technology, 13 (2-3), 81-91.

[2] Suleman A. and Costa A. P. (2004) Adaptive control of an aeroelastic flight


vehicle using piezoelectric actuators, Journal of Computers & Structures, 82
(17-19), 1303-1314.

[3] Liu L. K. and Zheng G. T. (2007) Parameter analysis of PAF for whole-
spacecraft vibration isolation, Journal of Aerospace Science and Technology,
11 (6), 464-472.

[4] Sinha J. K. (2008) Vibration-based diagnosis techniques used in nuclear power


plants: An overview of experiences, Journal of Nuclear Engineering and
Design, 238 (9), 2439-2452.

[5] García B., Burgos J. C. and Alonso Á. (2005) Winding deformations detection
in power transformers by tank vibrations monitoring, Journal of Electric
Power Systems Research, 74 (1), 129-138.

Vol. 31 No. 2 2012 147


Active vibration control of flexible plate structures with distributed disturbances

[6] Siringoringo M. and Fujino Y. (2008) System identification of suspension


bridge from ambient vibration response, Journal of Engineering Structures, 30
(2), 462-477.

[7] Kang Y. K., Park H. C., Kim J., and Choi S. B. (2002) Interaction of active and
passive vibration control of laminated composite beams with piezoceramic
sensors/actuators, Materials & Design, 23 (3), 277-28.

[8] Tokhi M. O. (1999) The design of active noise control systems for compact
and distributed sources, Journal of Sound and Vibration, 225 (3), 401-424.

[9] Framptom K. D. (2005) Distributed group-based vibration control with a


networked embedded system, Smart Materials and Structures, 14 (2), 307-314.

[10] Rashid A. and Mihai Nicolescu C. (2006) Active vibration control in


palletised workholding system for milling, International Journal of Machine
Tools and Manufacture, 46 (12-13), 1626-1636.

[11] Kumar K. R. and Narayanan S. (2008) Active vibration control of beams with
optimal placement of piezoelectric sensor/actuator pairs, Smart Materials and
Structures, 17 (5), 1-15.

[12] Wei L., Zhiku H. and Demetriou M. A. (2006) A computational scheme for
the optimal sensor/actuator placement of flexible structures using spatial H2
measures, Mechanical Systems and Signal Processing, 20 (4), 881-895.

[13] Montazeri A., Poshtan J. and Yousefi-Koma A. (2008) The use of ‘particle
swarm’ to optimize the control system in a PZT laminated plate, Smart
Materials and Structure, 17 (4), 1-15.

[14] Halim D. and Reza Moheimani S. O. (2003) An optimization approach to


optimal placement of collocated piezoelectric actuators and sensors on a thin
plate, Mechatronics, 13 (1), 27-47.

[15] Hongwei Z., Lennox B., Goulding P. R. and Leung A. Y. T., A float-encoded
genetic algorithm technique for integrated optimization of piezoelectric
actuator and sensor placement and feedback gains, Smart Materials and
Structures, 9, 552–557.

[16] Holland J. (1975) Adaptation in natural and artificial systems, University of


Michigan Press, USA.

[17] Kennedy J. and Eberhart R. C. (1995) Particle swarm optimization,


Proceedings of IEEE Conference on Neural Networks IV, Piscataway, NJ,
1942 – 1948.

[18] Dorigo M., Maniezzo V., and Colorni A. (1991) Positive feedback as a search
strategy, Technical Report 91-016, Dipartimento di Elettronica, Politecnico di
Milano, IT, 1-20.

[19] Braik M., Sheta A., and Arieqat A. (2008) A Comparison between GAs and
PSO in Training ANN to Model the TE Chemical Process Reactor,
Proceedings of Symposium on Affective Language in Human and Machine,
AISB Convention, 1 – 7.

[20] Mat Darus I. Z. and Tokhi M. O. (2005) Soft computing-based active vibration
control of a flexible structure, Engineering Applications of Artificial
Intelligence, 18 (1), 93-114.

148 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL


S. Julai and M. O. Tokhi

[21] Mohd Hashim S. Z. and Tokhi M. O. (2004) Genetic Modelling and


Simulation of Flexible Structures. Studies in Informatics and Control, 13 (3),
253-264.

[22] Shaheed M. H., and Tokhi M. O. (2002) Dynamic modelling of single-link


flexible manipulator: parametric and non-parametric approaches, Robotica,
20, 93–109.

[23] Mat Darus I. Z. and Tokhi M. O. (2004) Finite Difference Simulation of a


Flexible Plate Structure, Journal of Low Frequency Noise, Vibration and
Active Control, 23 (1), pp. 27-46.

[24] Timoshenko S. and Woinowsky-Krieger S. (1959) Theory of plates and shells,


McGraw-Hill Book Company, New York.

[25] Leissa A. W. (1969) Vibration of plates, Scientific and Technical Information


Division Office of NASA, NASA SP-160, Washington, D.C.

[26] Goldberg D. E. (1991) Real-coded genetic algorithms, virtual alphabets, and


blocking, Complex Systems 5, 139-168.

[27] Herrera F., Lozano M., Verdegay J. L. (1998) Tackling real-coded genetic
algorithms: operators and tools for behavioural analysis, Artificial Intelligence
Review, 12 (4), 265-319.

[28] Baker J. E. (1987) Reducing bias and inefficiency in the selection algorithm,
Proceedings of the Second International Conference on Genetic Algorithms
and their Application, Lawrence Erlbaum, Associates, Inc. Mahwah, NJ,
USA, 14-21.

[29] Eberhart R. C. and Yuhui S. (2001) Tracking and optimizing dynamic systems
with particle swarms, Proceedings of the IEEE of Congress on Evolutionary
Computation, 94-100.

[30] Shi Y. and Eberhart R. (1998) A modified particle swarm optimizer,


Proceedings of the IEEE Congress on Evolutionary Computation, Anchorage,
AK, USA, 69-73.

[31] Abd Latif I. and Tokhi M. O. (2009) Fast convergence strategy for particle
swarm optimization using spreading factor, Proceedings of the IEEE Congress
on Evolutionary Computation, 2693-2700.

[32] Ratnaweera A., Halgamuge S. K., and Watson H. C. (2004) Self-organizing


hierarchical particle swarm optimizer with time-varying acceleration
coefficients, IEEE Transactions on Evolutionary Computation, 8 (3), 240-255.

[33] Chen L., Shen J., Qin L., and Chen H. J. (2003) An improved ant colony
algorithm in continuous optimization, Journal of Systems Science and Systems
Engineering, 12 (2), 224-235.

[34] Socha K. and Dorigo M. (2008) Ant colony optimization for continuous
domain, European Journals of Operational Research, 185 (3), 1155-1173.

[35] Wang L. and Wu Q. (2001) Ant system algorithm for optimization in


continuous space, Proceedings of the IEEE International Conference on
Control Applications, 395-400.

Vol. 31 No. 2 2012 149


Active vibration control of flexible plate structures with distributed disturbances

[36] Quan H. and Chao H. (2007) Satellite Constellation Design with Adaptively
Continuous Ant System Algorithm, Chinese Journal of Aeronautics, 20 (4),
297-303.

[37] Lottin J., Formosa F., Virtosu M., and Brunetti L. (2006) About optimal
location of sensors and actuators for the control of flexible structures.
Proceedings of Research and Education in Mechatronics, 1-4.

150 JOURNAL OF LOW FREQUENCY NOISE, VIBRATION AND ACTIVE CONTROL

Vous aimerez peut-être aussi