Anda di halaman 1dari 27

Some experiences on vibrations and noise

suppression with piezoelectric active control systems


Gian Luca Ghiringhelli 1
Politecnico di Milano
Dipartimento di Ingegneria Aerospaziale
via La Masa, 34 20158 MILANO - Italy
Abstract
In this paper research activities, aimed at verifying the possibility of designing active
vibration control systems based on piezoelectric devices, are described. A full analogic
controller has been developed to suppress the vibrations in a clamped rectangular at
panel using piezoceramic patches. Vibration reduction tests and their correlations to nu-
merical predictions, based on an integrated piezo-structural model, are presented. Only
simple co-located direct feed-back laws have been used to point out problems related to
measurement conditioning circuits. Self sensing actuators have been also tested, the the-
oretical possibilities o ered by this kind of devices being signi cant, con rming that their
actual implementation entails some practical problems. The correlation with numerical
prediction shows that the adopted approach is capable of supplying a suitable design tool.
Encouraging results have been achieved, con rming the well known possibilities of piezo-
electric active control of thin panels. The reduction of the noise radiated by the vibrating
panel has been also measured, con rming the possibility of active vibration control in this
task.
1. INTRODUCTION
Often engineers are asked to improve mechanical designs either to match ever more strin-
gent requirements or to face the problem of no more acceptable performances; the active
control is an important tool suitable in these cases. Such an approach have been recog-
nised to be more exible and less invasive than interventions based on passively modi ed
structures. In fact better performances can usually be obtained for a wide range of oper-
ating conditions along with a signi cant save in weight.
Piezoelectrics are more and more often adopted to realise active control systems, direct
1Associate Professor in Design of Aerospace Structures,
E-mail: gianluca.ghiringhelli@polimi.it
Tel.: +(39)-02-2399-8356 Fax:+(39)-02-2399-8334

1
and converse piezoelectric e ects being exploitable to sense the structural response and
to apply the control action. A wide amount of literature is devoted to the basic problems
of this technology and only a very partial list is hereafter summarized. Theoretical and
mathematical models of piezoelectric materials have been available in the literature from
the beginning of their use, e.g. [1],[2]. The associated analytical solutions, however, are
restricted to simple geometries and boundary conditions; several papers describe more
general approaches, most of them base on nite element approximation, e.g. Tzou and
Tseng [3], Hwang and Park [4] and Wang and Rogers [5], while a simpler but popular
approach is represented by the pin force method, e.g. [6]. Furthermore a number of
applications, referring to numerical and experimental evaluation of piezo active systems,
are available, ranging from the vibration active control of beams and panels, e.g. Crawley
and Anderson [7], Gaudenzi [8], Yang and Lee [9]or Yang and Chiu [10], to the active
control of noise due to vibrations, e.g. Clark and Fuller [11] or Dimitradis, Fuller and
Rogers [12].
In this paper some of the experiences, gathered in testing piezoelectric control systems are
summarised. These activities are aimed at controlling the vibration of a clamped rectan-
gular at panel, mechanically excited and actuated by piezoceramic patches. Correlations
with numerical predictions are also presented. These are based on a integrated nite el-
ement program capable of describing the dynamic behaviour of the system composed by
a generic structure embedding piezoelectric devices (Ghiringhelli et Al.[22]). It allows a
natural integration of structural and piezoelectric models with actual signal conditioning
circuits and any other electronic devices used to realise the control system.
A completely analogic active control system and simple direct feedback laws have been
used to point out problems related to measurement conditioning circuits. Self sensing
actuators, the usefulness of which is well known (Dosch et Al.[13], Saunders et Al.[14],
Brusa et Al.[15], Akishita et Al.[16], Cole and Clark[17], Anderson and Hagood[18], Tsou
and Hollkamp[19], Ko and Tongue[20], Chang-qing et Al.[21]), have also been tested using
appropriate bridge and conditioning circuits. The theoretical possibilities o ered by this
kind of devices are evident but the actual realisation entails non trivial problems and
some of these papers pointed-out diculties in keeping the balance of sensor bridges.
The choice of a fully analogic system is due to two main considerations. First of all the
need of testing and tuning electronic devices suitable to be used in actual implementation
of the control system (the testing of a digital multi-input multi-ouput active control and of
adaptive control laws is in progress). The second reason is related to the diculties that

2
can be envisaged when implementing a digital control the band of which reaches relatively
high frequencies: in this case the availability of well tested analogic control systems could
ease the solution of some problems.
2. INTEGRATED MODELISATION OF PIEZOELECTRIC
In this section the main features of an integrated modeling techniques, based on a nite
element discretisation, are presented (Ghiringhelli et Al.[22]).
2.1 Piezoelectric e ect
The direct and inverse piezoelectric e ects, i.e. the generation of an electric eld due to
mechanical strains and the ability to induce a strain when subjected to an electric eld,
are described by the linear constitutive laws of piezoelectric materials, in IEEE standard
notation [23]:
h i
fT g = cE fS g , [e]T fE g (1)
h i
fDg = [e] fS g + S fE g (2)
where fT g is the stress vector, fS g is the strain vector, fE g is the electric eld, fDg is the
h i
electric displacement, cE is the elastic matrix at constant electric eld, [e] is the piezo-
n o
electric characteristic matrix and S is the dielectric matrix at constant strain. Indeed
Eqs. 1 and 2 represent both a simpli cation and a linearisation of generally nonlinear
constitutive equations that characterise the more general electro-magneto-thermo-elastic
problem. These laws show a simultaneous dependence on stress fT g, heat ux fqg, cur-
rent density fj g, electric displacement fDg and magnetising eld fH g from strain fS g,
temperature T , electric and magnetic elds fE g,fB g (Eringen[2]), i.e.:
fj g = J (fS g ; fE g ; fB g ; T ) ; fH g = H (fS g ; fE g ; fB g ; T )
fqg = Q (fS g ; fE g ; fB g ; T )
fT g = T (fS g ; fE g ; fB g ; T ) ; fDg = D (fS g ; fE g ; fB g ; T )
A completely coupled formulation can be obtained, but usually no dependence of stresses
and electric displacement on magnetic eld and temperature is assumed, i.e. a simple
electrostatic assumption. Small changes, with respect to a reference temperature, are
then considered, as well as frequencies low enough to avoid heating of the piezoelectrics,

3
due to the hysteresis associated to the direct and inverse e ects. Finally a linearisation
of the constitutive laws in a reference con guration is made, i.e.:
" # " # " # h i " @ fDg #
= @@ ffTS gg [e] = @@ ffD g = @@ ffET gg
h i
cE Sg [e]T S = @ fE g
leading to Eqs. 1 and 2.
2.2 Response of a controlled system
Starting from Eqs. 1 and 2 the electro-structural model can be obtained using the inte-
grated nite element approach described in Ghiringhelli et Al.[22], the results of which
are brie y recalled here. Once the structural displacement d and the electric ux  , that
is the voltage time integral Z
 = V dt
are assumed as primary eld unknowns, they are approximated by means of a nite
element approach in terms of the corresponding nodal values fdg and fg.
The use of an appropriate nite element discretization allows to write a generalised Virtual
Work Principle equation and the exploitation of this principle leads to a linear form of
the dynamics equations of the coupled system, i.e.:
" #( ) " #( ) " #( ) ( )
M 0 d + D ,A d_ + K 0 d = f (3)
0 ,CP  ,ST 0 _ 0 0  iP
where [M ] ; [D] ; [K ] are the mass, damping and sti ness matrices, fdg is the displacements
vector, ff g is the loads vector, fiP g is the nodal currents vector, [A] is the electro-elastic
coupling matrix related to actuators, [S ] is the electro-elastic coupling matrix related to
sensors and [CP ] is the dielectric matrix. In this equation the second row is changed in
sign to enforce the symmetry of the problem. The analytical de nition of these terms and
other details on the nite element development can be found in [22] and [24] respectively.
When piezoelectrics are used for control purposes, the electric unknowns  are naturally
limited to a relatively small number of voltages, related to controlled electrodes. On the
contrary, Eq. 3 can contain too many structural unknowns to be directly used in the
design of active control systems. Thus it is generally necessary to take a condensation of
the structural part of Eq. 3. Here a modal transformation of the structural part alone
has been adopted, based on the eigenvectors matrix [X ] of the structural problem, ( i.e.
fdg = [X ] fqg ), fqg being the generalised structural dofs., so that we can write the

4
generalised form of Eq. 3:
" T #( ) " #( )
X MX 0 q + X T DX ,X T A q_ +
0 ,CP  ,ST X 0 _
" #( ) ( )
+ X 0KX 00 q = XT f
T
 iP (4)
The previous modal condensation does not change the structure of the dynamics equation,
in the following the previously de ned symbols will take the generalised meaning.
In the second row of Eq. 3 we can immediately recognise the matrix form of the sensing
equation, that can be alternatively derived from the second equation of the piezoelectric
constitutive law. The array arranging the currents due to the piezoelectric e ect is given
by: fiP g = [S ]T fq_g .
The assumption of the time integral of the voltage as electric primary unknown, instead
of the most commonly used voltage, allows a natural formulation of the coupled elec-
tric/elastic problem in presence of external circuits. In fact a complete analogy between
structural displacements and electric uxes is established: networks of electric components
can be easily modelled, as properly connected generalised lumped capacitors (masses), re-
sistors (dampers) and inductors (springs), and an integrated linear problem is obtained
by simply appending the electric model to Eq. 3; by expanding the array of uxes with
the values at the nodes of the external electric network we can write:
" #( ) " #( )
M 0 q + D ,A q_ +
0 CL , CP  ,S GL
T _
" #( ) ( )
+ K0 R0 q =
 i
f
+ (5)
L P iL
where [CL], [GL], and [RL] are respectively the lumped symmetric capacitance, conduc-
tivity and reluctance matrices. They are assembled from local matrices, related to each
side of the electric circuit, while iL are current lumped inputs.
2.3 Direct feedback with co-located control
A co-located control can be easily implemented provided the two elements of a pair of
piezoelectrics are used to sense and actuate respectively. In fact, if the host structure is
thin enough, we can assume the measurements on the top surface to be fully equivalent
to those on the bottom, one once its sign is properly assumed: in this way a co-located
control is obtained, in the limits of two identical and perfectly placed devices. The bene ts
of a co-located control are well known, in particular a direct velocity feedback scarcely
5
su er spillover problems even if the actuator dynamics must be carefully evaluated to grant
unconditional stability, e.g. [26], [27]; a feedback on the position, i.e. the positive position
feedback [32], also showed interesting properties of insensitivity to spillover furthermore
being not destabilized by nite actuator dynamics.
The measurement of the piezoelectric charge has been frequently used as the input of a
direct feed-back; the conditioning circuit needed for this job can be simply built using a
voltage follower: the circuit is a high pass lter that behaves like an ideal displacement
transducer in the frequency range of interest by measuring the piezoelectric charge. The
corresponding sensor equation is:
fVS g = [CP ],1 [S ]T fqg (6)
While the control voltage is de ned as:
fVC g = , [G] fVS g (7)
In Eq. the control voltage has been de ned by means of the gain matrix, [G]; it can be
fully populated but, in the applications here presented, only a diagonal form has been
tested.
 
[M ] fqg + [C ] fq_g + [K ] + [A] [G] [CP ],1 [S ]T fqg = ff g (8)
This kind of control is usually referred to as a position feed-back, e.g. [8] and [29],
because the sensing voltage, in Eq. 6, is related to the structural displacement through
the array of modal amplitudes. Indeed this measurement senses the surface strain, i.e. the
local curvature of the panel, and a priori it cannot be linearly related to the transversal
displacement at the piezoelectric location. So in the following it will be referred to as
strain feedback.
By the use of a measurement of the piezoelectric current, a strain rate feedback can be
implemented: the second row of Eq. 3 is straightforwardly exploited for this:
fVS g = [R] [S ]T fq_g (9)
leading to the following controlled system dynamics equation:
 
[M ] fqg + [C ] + [A ] [G] [R] [S ]T fq_g + [K ] fqg = ff g (10)
In these equations the matrix [R] is diagonal and is representative of the resistive branches
along with the piezoelectric currenties fiP g ow. In analogy with the previous discussion
6
this control is known as a velocity feedback, e.g. [8],[28], [25]. It has been also tested,
using a current to voltage conversion circuit, but this conditioning circuit failed: high
frequency instabilities occurred as the control gain was increased (analoguous problems
have been already outlined in the literature, e.g. Akishita et Al.[16] and in Hong el Al.
[25]).
It is well known that a colocated control (both in position and in velocity) is stable. But
practical applications lead to imperfect systems and some of the possible causes can be
cited: the actual frequency properties of each element of the control system, e.g. sen-
sor conditioning circuits and the driving devices, and the imperfect co-location, due to a
wrong position of sensors and actuators. The use of piezoelectric patches worsen these
problems. In fact piezos represent high capacitive loads for driving devices, in this way
limiting their frequency properties. Furthermore they are more sensitive to high frequency
modes, when used to sense, as well as they have more authority on them, when used to
actuate, in this way entailing problems in an ecient ltering. At last the piezos of a pair
can show some di erences, e.g. in sizes, piezoelectric and dielectric properties, possibly
worsening the aforementioned sensitivity problems.
Then instabilities are expected, especially with high gains, but they arise after a good
e ectivenes has been achieved, during tests on strain feedback control, while, using the
strain-rate feedback, the instability onset is very early, then preventing any positive re-
sults. This fact is worth studying in depth by exploiting the analysis of the properties of
the control matrix, in relation to stability theorems. In fact in the ideal case it is sym-
metric and positive semi-de nite, but this is no more true in practical applications: when
the actual properties of devices are introduced, independent sensing and control coupling
matrices should be used. It is apparent that it can now be no more symmetric and its
symmetric part, fundamental for the stability of the control, no more positive de nite.

2.4 The self-sensing actuator


A self-sensing actuator allows to build a real and more e ective co-located control: the
availability of the self-sensing actuator allows the use of a couple of symmetric elements
as actuators, so that a more e ective bending moment is applied.
For such a device it is mandatory to remove the e ects of the control voltage from the
measurements. This task can be accomplished by using the measurement bridge circuit
proposed by Dosch et Al. [13]. The bridge, depicted in Fig. 1, is realised by means of
7
two branches, each made with two series impedances, the rst one is capacitive, while
the second one, i.e. Z1 and Z2, is generic. The piezoelectric devices plays the role of
a capacitive impedance Cp, while usual capacitance, the value of which C is nominally
equal to the piezoelectric capacitance, Cp, is placed on the second branch. The di erential
voltage across points 1 and 2 leads to a measure of the piezoelectric current, canceling
the control voltage e ects. In the Laplace domain the sensing voltage becomes:
 
Z sC P Z1
VS = V1 , V2 = 1 + sC Z IP + 1 + sC Z , 1 + sCZ VC
1 sCZ 2
(11)
P 1 P 1 2

By introducing the following assumptions:


a) CP Z1 = CZ2 b) CP Z1 << 1 (12)
the sensor equation becomes:
Vs = V1 , V2 = Z1IP (13)
Depending on the kind of the complex impedance Z , a displacement, velocity or accel-
eration measurement could be obtained, respectively, with a capacitor, a resistance and
an inductance. The Eq. 12a states the balance condition of the bridge, that leads to
the canceling of the control voltage e ect, from which both stability and measurement
precision depend. Due to diculties in satisfactorily balancing a capacitive bridge and in
satisfying the 12b condition in the case of inductances, only resistive bridges have been
experimented. The measurement equation in this case is given by:
VS = Rm iP = Rms []T fqg (14)
Once a resistive impedance is chosen, the Eq. 13 leads to a measurement proportional to
the piezoelectric current, that is the strain time rate of the host structure at the position
of the piezoelectric.
It must be outlined that, due to the presence of a resistance between the piezoelectric
device and the ground, the e ective voltage acting on the piezoelectric itself, is:
Veff = VC , V1 (15)
If a direct feedback is used, i.e. Eq. 7, the e ective voltage is:
Veff = ,[(G + 1)V1 , G V2]  ,G Vs (16)
that approximates the nominal control voltage only if the gain G is high enough. It is
possible to observe that, in absence of active control, i.e. G = 0, a control action is still
8
present simply due to the presence of a closed loop.
The dynamics equations of a structure equipped with a self-sensing actuator measuring
the strain time rate are given by:
38
2
M 0 0 >
< q 9
>
=
2 38
D  (I + G)  G > < q_ 9>
=
6
4 0 R1CP (I + G) ,R1 CP G 75 > 1 > + 64 ,R1 T I 0 75 > _ 1 > +
0 R2 CP G R2 C (I , G) : 2 ; 0 0 I : _ 2 ;
2 38 9 8 T 9
6
K 0 0 >
7 < q > = > < [X ] ff g >=
4 0 0 0 5 > 1 > = > 0 >
0 0 0 : 2 ; : 0 ;
It can be seen that the sensitivity is not changed while the authority of the actuator is
increased. In the case of a perfectly balanced bridge, the measurement is ideal and no
stability problems arise at all. If this is not the case, asymptotic stability is obtained if
the following inequality is satis ed (Masarati [31]): R1 CP > R2 C . Because a perfect and
permanent balancing is impossible, the actual resistors must be chosen to obtain the best
balance as possible provided they satisfy this stability condition. This consideration holds
in the case of a direct feedback.
The use of a direct feedback on the bridge output, that is a velocity feedback, failed,
possibly due to the same reasons that led to bad results when a simple current to voltage
conversion has been applied to a piezoelectric sensor output. The technique ltering de-
scribed by Fanson [32] has been used: the sensor output is conditioned by a second order
low-pass lter with high resonance factor so that the signal is ampli ed at design frequency
while the higher harmonic content is ltered. The implementation of more sophisticated
adaptive "sensoriactuator" can be envisaged, i.e. the one proposed by Cole and Clark [17].
By using these devices a number of problems, e.g. uncertainities of material properties
and their dependence on environmental parameters, seem to be solved. But no dramatic
problems have been encountered in balancing the bridges, so the implementation of such
a techniques has been delayed.

3. EXPERIMENTAL SET-UP
The structure under test is a at panel (600x400x2mm), made of 2027 Aluminium Alloy,
clamped on each side to a steel frame (Fig. 2). The main structure is elastically sus-
pended to a support frame and it is excited by an electro-mechanical shaker (B&K Type
4810). To limit the dynamic response of the frame torsional mode (
= 450Hz), the shaker

9
has been placed on a symmetry plane (Fig. 2). Eight pairs of piezeceramic patches have
been used, made of PZ21A (See Tab.1), produced by Ferroperm (DK), the dimensions of
which are 50x50x0:5mm. The elements of each couple are bonded, by means of epoxy
resin, in corresponding locations on the two surfaces of the panel. These locations have
been chosen to control selected low frequency modes with an odd number of half waves
along the short side, due to the previously mentioned reason. To obtain better perfor-
mances the piezoelectrics have been intuitively placed as far as possible from areas where
the modal curvature of the selected modes vanishes, as depicted in Fig. 3. In Table 2 a
comparison is made between experimental and Finite Element natural frequencies of the
panel equipped with piezoelectric devices. No adjustements were made to improve this
comparison because the goal of the test was also to verify the robustness of the adopted
design techniques against model uncertainities.

3.1 Strain feedback control


The actual implementation of the strain feedback control is presented in Fig. 4. It is
based on a simple voltage follower that allows to measure a voltage proportional to the
charge generated by the direct piezoelectric e ect: it is built around a FAT operational
ampli er. The polarisation resistance has been chosen to give 0.5Hz corner frequency.
In Fig. 5 the sensor transfer function is depicted showing that, along the whole range of
interesting frequencies, it actually behaves as a proportional sensor.
The control power is supplied by a power stage based on a properly compensated APEX
PA85 operational ampli er, capable of supplying up to 100V . Its measured bandwidth
is greather then 4kHz when driving a 150F capacitive load. Due to the comparatively
high voltage sensor output the gain regulation has been obtained by simply partitioning
the measurement before entering the high voltage ampli er.

3.2 Modal feedback control with self-sensing actuators


In Fig. 6 the sketch of the control system, based on the use of a self-sensing actuator,
is presented (the second bridge, depicted by a dashed line, is used only in the case de-
scribed as "double self-sensing actuator"). The di erence of the voltages across the bridge
branches is carried out by means of an operational ampli er working in di erential mode
(Burr Brown INA118). The bridge is equipped with two trimmers to make its balancing

10
easier. The power stage and the previously described gains regulations have been used.
The modal lter implementation allows the signal to be ampli ed by a 5dB factor at
the corner frequency together with a 180deg change in phase across this frequency. It is
important to note that while the bridge allows to sense a strain rate, the 90deg change
in phase, produced by the modal lter at the corner frequency, transforms it in a strain
measurement.

4. NUMERICAL AND EXPERIMENTAL RESULTS


In this section the results related to the previously described control modes are presented.
Numerical and measured response will be compared.
4.1 Strain feedback with co-located sensors and actuators
These results are related to a direct strain feedback control. Figures 7 and 8 report the
comparison between the frequency response functions of the panel for the experimental
and numerical simulations, respectively with the control system on and o . They refer
to the measurement of an accelerometer placed close a corner of the panel. The actively
controlled response is black and solid, while the uncontrolled one is grey and dashed.
Results for active control with one and with three piezoelectric devices are presented.
The gains have been heuristically tuned by adding a single device at once and bringing
the system near the stability limit. Fig. 7 presents the data provided by the test system
while Fig. 8 is related to simulations carried out with MATLAB using 30 modes to
describe the panel dynamics; a total of 60 equations has been used when sensors and
actuators dynamics have been added. Data in Table 2 are instead referred to the active
control e ectiveness under harmonic excitations using devices A and B, working one at a
time. The maximum output of a reference accelerometer is presented when the structure is
harmonically forced close to some natural modes. The amount of the response attenuation
can be appreciated and a reasonable con dence in numerical predictions is envisaged. In
Figure 9 only the acceleration reduction is depicted, outlining that the correlation has
not the same quality in relation to the response reduction of di erent modes. It can be
noted that the e ect on the control e ect second mode is overestimated by the simulation.

11
4.2 The single self-sensing actuator
The use of a single piezoelectric device as a self-sensing actuator, combined with the modal
lter, gave satisfactory results. Also in this test case the results shown are related to a
xed frequency excitation and measurements are related to the output of an accelerome-
ter. The lters have been tuned by setting the working frequency to match speci c modes.
Simulations have been performed by including the power ampli er transfer function and
the bridge electric model. A 5% unbalance has been introduced and experimental gains
have been used. In fact this set of testing activities was carried out before the simulations,
because of a frequent operative overlapping, and simulations were in some case driven by
the experimental observations, e.g. in relation to the adopted control gains. In partic-
ular, during the tests the maximum stable gain has been achieved by a trial and error
procedure and these values have been used in simulations afterward. In Tables 4 and 5
the dynamic response, both in controlled and uncontrolled case are presented for several
excitation frequencies. The e ectiveness of the control is apparent also in this case, but
the correspondence between tests and numerical predictions is not satisfactory, as it can
be veri ed in Figure 10 and Table 6, where some of these experimental and numerical
results are compared in terms of acceleration reduction. In Figure 10 the rst two bars
refer to the performances measured during the tests and predicted by the simulation: it
is possible to appreciate that a lower authority is often predicted by the simulation.
Based on these data we noted that the control e ectiveness was often underestimated
by the numerical predictions. Furthermore we found that the experimental gains do not
mark the stability limit of the numerical model. So the same procedure has been applied
to the numerical simulations and the largest possible gains have been assessed: by this
approach the correlation improved signi cantly. The results of this procedure are shown
in Table 6, and depicted by the third bar in Figure 10: an overestimate of the e ectiveness
on the second mode is present also in this case, con rming the remarks of the previous
section, but the behaviour of the other modes is satisfactorily predicted as well.
4.3 Double self-sensing actuator
The basic exploitation of the availability of a self-sensing actuator consists in using a
couple of opposite elements for actuation purposes, so that a better performance, in re-
lation to the previous cases, is expected. Actually the bridge balancing resulted more
dicult in this case and lower gains were found to be usable at some frequencies. This

12
can be justi ed with the considerations previously presented: due to the system layout,
the capacitance driven by the power ampli er is doubled (two patches are controlled), in
this way reducing its dynamic range and exacerbating the sensitivity problems on higest
modes.
Also in this case the results of harmonic tests at several frequencies are presented. Fig-
ure 11 and Table 7 compare experimental measurements and simulations. Again a satis-
factory agreement has been found once simulation gains have been scaled to the maximum
allowable to avoid instabilities. The use of a pair of self-sensing actuators led in general
to better performances but at the same time both simulations and experiments showed
lower maximum gains can be used in some cases.
4.4 Acoustic
5. CONCLUDING REMARKS
The research activities demonstrated once again the possibilities o ered by piezoelectric
patches in suppressing vibrations, even by using very simple control laws, e.g. a direct
strain feedback.
The inecient behaviour of velocity sensors should be further investigated: in fact this
kind of feedback shows several typical advantages but due to the speci city of the co-
located piezoelectric control it could lose part of its appeal. Nonetheless the availability
of references successfully exploiting it cannot be disregarded, even if on this subject the
literature is contradictory.
The use of a self-sensing actuator, together with a modal lter of the input signal leads
to results comparable to those obtained using separate sensor and actuator devices, but
with the saving of a piezo patch. The use of a pair of self-sensing actuators improved,
but not dramatically, the e ectiveness of the active control, due to some diculties in
balancing the bridge and because lower maximum gains can be used at some frequencies.
The formulation adopted to simulate the dynamic behaviour of the coupled system and
to design the active control shows a satisfactory agreement with experimental results, so
that the related performances can be reasonably predicted. The availability of validated
models, conditioning circuits and power ampli ers allows one to continue with the design
and implementation of a digital control system.
Acknowledgements
The author is grateful to Mr.Mariofelice Zanardi for his helpful contribution to the work.
This research was supported in part by ENEL-CRA under Grant No.R23TC0009 and by
the National Council of Researches - CNR under Grant No.97.00891.PF34.

References
[1] Crawley E.F. and deLuis J, 1987 Use of piezoelectric actuators as elements of intel-
ligent structures, AIAA J., V.25, 1373-1385
[2] Eringen A.C. and G.A.Maugin, 1989, "Electrodynamics of Continua", Vol. 1 and 2,
Springer- Verlag.

14
[3] Tzou H.S. and Tseng C.T., 1990, Distributed piezoelectric sensor/actuator design
for dynamic measurement control of distributed parameter systems: a piezoelectric
nite element approach, J. of Sound and Vibration, V.138, 17-34
[4] Hwang W.S. and Park H.C., 1993, Finite element modeling of piezoelectric sensors
and actuators, AIAA J., V.31 ,No.5, May 1993, 930-937
[5] Wang B.T. and Rogers C.A., 1991, Modeling of nite-length spatially-distributed
induced strain actuators for laminate beams and plates, J.of Intell. Mater. Syst. &
Struct., V.2, 38-58
[6] Chaudhry Z. and Rogers C.A., 1994, The Pin-Force Model Revisitated, J.of Intell.
Mater. Syst. & Struct., Vol.5, May, pp.347-354
[7] Crawley E.F. and Anderson E.H., 1990 Detail models of piezoeletric actuation of
beams, J.of Intell. Mater. Syst. & Struct., V.XXX, 4-25
[8] Gaudenzi P., R. Barboni, R.Carbonaro and S. Accetalla "Direct position and velocity
feedback control on an active beam with PZT sensors and actuators", Proceedings
of the International Forum on Aeroelasticity and Structural Dynamics , 17-20 June
1997, Rome
[9] Yang S.M. and Lee Y.J. 1993 Vibration suppression with optimal sensor/actuator
location and feedback gain Smart Mater. & Struct. Vol.2 232-239
[10] Yang S.M. and Chiu J.W. 1993 Smart structures - vibration of composites with
piezoelectric materials Composite Structures, 25, 381-386
[11] Clark L.R. and Fuller C.R., 1992 Experiments on Active Control of Structurally
Radiated Sound Using Multiple Piezoceramic Actuator J.of Acoustical Society of
America, 91(6), 3313-3320
[12] Dimitradis E.K., Fuller C.R. and Rogers C.A., 1991 Piezoelectric actuators for dis-
tributed vibration excitation of thin plates, J.of Vib.Acoust., V.113, 100-107
[13] Dosch J.J., D.J. Inman, E. Garcia, 1992 "A self-sensing piezoelectric actuator for
collocated control", Journal of Intelligent Material System and Structures, Vol.3,
pp.659-667, Jan.

15
[14] Saunders W.R., D.G. Cole and H.H. Robertshaw "The impact of piezoelectric senso-
riactuators on active structural acoustic control", Proceedings of the SPIE, Vol.1917,
No.1,p.578-586
[15] Brusa E., S. Carabelli, A. Tonoli, 1996 " Self-Sensing collocated structures with dis-
tribuited piezoeletric transducers", Mechatronics Laboratory Politecnico di Torino,
ICAST 1996, pp.84-94, Roma, Italy
[16] Akishita S., Y. Mitani and H. Miyaguchi, 1994 "Sound transmission control through
rectangular plate by using piezoelectric ceramics as actuators and sensors", Journal
of Intelligent Material System and Structures, Vol 5, May
[17] Cole D.G., Clark R.L., 1994 "Adaptive compensation of piezoelectric sensoriactua-
tors", Journal of Intelligent Material System and Structures, Vol 5, pp.665-672
[18] Anderson E.H., Hagood N.W., 1992 "Self-sensing piezoelectric actuation: analysis
and application to controlled structures", AIAA/ASME/ASCE/AHS/ASC Struc-
tures, Structural Dynamics and Material Conf., Dallas TX, pp.2141-2155
[19] Tsou H.S., Hollkamp J.J, 1997, "Collocated independent modal control with self-
sensing orthogonal piezoelectric actuators", AIAA-94-1737-CP
[20] Ko B., Tongue B.H., 1995, "Acoustic control using a self-sensing actuator", J. of
Sound and Vibration, Vol.187(1),pp.145-165
[21] Chang-qing C., W.Xiao-ming, S.Ya-peng, 1996, "Finite element approach of vibration
control using self-sensing piezoelectric actuators", Computers & Structures, Vol.60,
No.3, pp.505-512.
[22] Ghiringhelli G.L., M. Lanz, P. Mantegazza, 1993. " Numerical modelling and exper-
imental testing of distribuited piezoeletric actuators", Int. Forum of Aeroelasticity
and Structure Dynamics, Vol.1, pp.277-292, Strasbourg
[23] ANSI/IEEE Standard notation for piezoelectricity, 1988
[24] Ghiringhelli G.L., " Numerical modelling and experimental testing of distribuited
piezoeletric actuators", l'Aerotecnica Missili e Spazio, 1995

16
[25] Hong S.Y., V.V. Varadan and V.K.Varadan, 1991 "Comparison of analog and digital
strategies for automatic vibration control of lightweight space structures", SPIE, Vol.
1489 Structure Sensing and Control, pp.75-83
[26] Aubrun J.N., "Theory of the control of structures by low-authority controllers",
Journal of Guidance and control, Vol.3, No.5, 1980, 444-451
[27] Chen C.L., "Direct output feedback control for large space structures", Dynamics
Laboratory Report DYNL-82-1, California Institute of Technology, 1982
[28] Sullivan J.M., S.E. Burke and J.E> Hubbard, 1995 "Experimental demonstration of
active broadband vibration suppression of a rectangular plate using gain-weighted,
shaped distributed transducers", AIAA-95-1119-CP, American Institute of Aeronau-
tics ans Astronautica, pp. 3379-3389
[29] Denoyer K.K., Kwak M.K. "Dynamic Modelling and Vibration Suppression of a
Slewing Structure utilizing Piezoelectric Sensors and Actuators", Journal of Sound
and Vibration, Vol.189, N.1, pp.13-31, 1996
[30] Hagood N.W., Anderson E.H., 1991, " Simultaneous Sensing and Actuation using
piezoelectric materials", SPIE Active and Adaptive Optical Components
[31] Masarati P., 1995 "Travi piezoelettiche: modellazione ed analisi", Politecnico di Mi-
lano, Thesis
[32] Fanson J.L. and T.K.Caughey, 1987 "Positive position feedback control for large
space structures", AIAA Paper 87-0902

17
Table 1: PZ21A piezoceramic properties
 Kg=m3 7915
TC C 500
s11 m =N 16:94 10,12
E 2

sE22 m2 =N 24:45 10,12


 :34
,d31 C=N 275 10,12
d33 C=N 660 10,12
T 3603
 S 1561

Table 2: Frequencies comparison FEM Vs. Experimental


Mode Shape FEM Test 
# [Hz] [Hz] %
1 1-1 88.07 92.30 -4.8
2 2-1 137.94 145.1 -5.2
3 1-2 221.98 214.4 3.4
4 3-1 227.58 223.2 1.9
5 2-2 270.31 266.2 1.5
6 4-1 344.69 322.5 6.4
7 3-2 355.88 334.2 6.1
8 1-3 429.22 397.1 7.5
9 2-3 465.48 436.9 6.1
10 4-2 477.26 454.2 4.8

18
Table 3: Comparison of acceleration reduction
Experiment Simulation
MODE ANC [g] AC [g]  % ANC [g] AC [g] %
Pair A
I 18.3 1.5 91.8 12.4 0.98 92.0
IV 34.6 13.9 59.9 45.1 3.02 91.1
VIII 33.3 17.1 48.6 22.2 12.1 45.5
Pair B
I 18.25 2.35 87.1 12.4 1.5 87.7
II 43.7 18.9 56.9 46.3 3.4 91.4
IX 9.3 1.3 86.3 11.9 1.34 88.1

Table 4: Measurements of modal control eciency


Pair f [Hz] ANC [g] AC [g]  % G
A 91 3.18 0.93 70.75 680
A 145 7.11 5.11 28.13 196
B 145 10.53 4.81 54.32 460
C 91 4.87 1.75 64.07 840
C 145 7.13 3 57.92 408
C 300 8.15 3.65 55.21 1360
C 416 11.63 3.8 67.33 840
D 268 6.97 2.66 61.84 108
D 300 7.79 6.41 17.72 144
D 397 6.68 3.77 43.56 96
D 416 7.31 6.08 16.83 80

19
Table 5: Numerical prediction of modal control eciency
Pair f [Hz] ANC [g] AC [g]  % G
A 87.44 1.24 0.61 50.54 680
A 87.44 1.24 0.38 68.91 900
B 137.86 4.63 1.97 57.39 460
B 137.86 4.63 1.46 68.51 550
C 87.44 1.24 0.54 55.99 840
C 87.44 1.24 0.43 65.21 950
C 137.86 4.63 2.96 36.09 408
C 137.86 4.63 1.62 64.92 600
D 467.21 7.15 3.9 45.5 155

Table 6: Comparison of acceleration reduction: Numerical Vs. Experimental


Experiment Simulation Scaled sim.
Piezo # Mode G % G % G %
A I 680 70.8 680 50.5 900 68.9
B II 460 54.3 460 57.4 550 68.5
C I 840 64.1 840 56.0 950 65.2
C II 408 57.9 408 36.1 600 64.9
D VIII 80 16.8 80 5.1 200 35.0

20
Table 7: Comparison of acceleration reduction: Numerical Vs. Experimental (Double self-
sensing actuator)
Experiment Simulation
Pair Mode G ANC [g] AC [g]  % G ANC [g] AC [g]  %
A I 1120 6.33 1.21 80.8 800 1.24 0.33 73.5
B I 1060 6.50 3.59 44.8 950 1.24 0.45 63.6
B II 260 8.76 4.16 52.57 260 4.63 2.41 47.9
C I 700 4.83 1.24 74.25 850 1.24 0.35 71.7
C II 600 12.10 4.45 63.22 450 4.63 1.73 62.6
C VIII 60 11.65 6.17 47.01 100 7.12 4.88 31.40
D VIII 104 10.00 3.59 64.08 110 7.15 4.03 43.7

Table 8: SPL measurements (single active device)


Load [N] 1.11 2.79 5.6 6.1
Frequency [Hz] 88.0 128.0 192.0 276.0
SPL [dB]
Uncontrolled 93.40.002 79.30.013 92.80.012 96.80.013
Pair A 77.80.005 70.40.147 Not Controllable 72.70.093
Di erence -15.6 -8.9 -24.1
Pair B 74.20.007 73.60.062 82.80.007 85.10.013
Di erence -19.2 -5.7 -10.0 -11.7
Pair C 75.40.057 71.80.037 Not Controllable 78.80.057
Di erence -18.0 -7.5 -18.0

Table 9: SPL measurements (three active devices)


Load [N] 0.92 3.72 6.7 6.34
Frequency [Hz] 88 128 196 284
SPL [dB]
Uncontrolled 96.00.057 76.40.003 91.00.037 96.90.037
Controlled 81.40.003 62.20.032 83.70.007 77.50.023
Di erence -14.7 -14.3 -7.3 -19.3

21
C V1

VC Z1

Z2

CP V2

Figure 1: The self-sensing actuator bridge circuit

Figure 2: Test experimental set-up

22
1th Mode 2th Mode

4th Mode 6th Mode

Figure 3: Piezoelectric location Vs mode null curvature

Voltage
Follower PA-85

LM-362
Sensor R pol

Actuator Polarisation
direction

Figure 4: The strain sensor layout

23
A [dB]

Phase

Frequency [Hz]

Figure 5: Strain sensor transfer function

Vcontrol C
V1 V2 LM-362
Filter
Piezoelectric
Vcontrol

INA118
Vcontrol

Double self-sensing actuator

Figure 6: The self-sensing strain-rate sensor layout

24
Single active device (A)

T
Pair A
100
Test
80
Sim.
% 60

40

20
0
I II VIII

Pair B
100

80

% 60

40

20
0
I II IX

Mode

Figure 9: Comparison of acceleration reduction

80
70 Test
60 Gspe
50 Gmax
40
%
30
20
10
0
I,A II,B I,C II,C VIII,D
Mode & Piezo #

Figure 10: Acceleration reduction (Single self-sensing actuator)

26
90
80 Experiment
70 Simulation
%
60
50
40
30
20
10
0
I I II I II VIII VIII
(A) (B) (B) (C) (C) (C) (D)
Mode & Piezo #

Figure 11: Acceleration reduction (Double self-sensing actuator)

27

Anda mungkin juga menyukai