Você está na página 1de 7

REFLECTIVE FIBER OPTIC DISPLACEMENT SENSOR APPLIED TO

CHARACTERIZATION OF PIEZOELECTRIC FLEXTENSIONAL ACTUATORS


Joo Marcos Salvi Sakamoto, Alexandre Csar Rodrigues Silva, Ricardo Tokio Higuti, Cludio Kitano
Department of Electrical Engineering, Universidade Estadual Paulista UNESP
P.O. Box 31 15385 000, Ilha Solteira So Paulo, Brazil
kitano@dee.feis.unesp.br

Gilder Nader, Emlio Carlos Nelli Silva


Department of Mechatronics and Mechanical Systems Engineering - Escola Politcnica da Universidade de So Paulo
Av. Prof. Mello Moraes, 2231 - 05508-900 So Paulo, S.P., Brazil
ecnsilva@usp.br

Abstract The laser interferometer method is a wellestablished technique for the characterization of
piezoelectric actuators. However, this method requires
precise optical alignment and meticulous operation.
There is great interest in developing displacement and
vibration measurement systems using reflective fiber
optic displacement sensors (RFODS) because of their
inherent simplicity, small size, wide frequency range,
extremely low displacement detection limit, and ability to
perform measurements without contact or affecting the
vibrating system. In this work a piezoelectric
flextensional actuator (PFA), designed with the topology
optimization method, is experimentally characterized by
the measurement of its nanometric displacements using a
RFODS. The linearity and frequency response of the PFA
are evaluated up to 45 kHz, and the tracking error
phenomenon is discussed.
Index TermsFiber optic lever, nanometric displacement
measurement, piezoelectric flextensional actuator, reflective
fiber optic sensor

I. INTRODUCTION
In the precision engineering industry, research and
development is being undertaken towards the development of
piezoelectric actuators. Piezoelectric ceramics such as PZT
can convert electrical energy to mechanical form. Every time
the PZT drive voltage changes, the piezo element changes its
dimensions. Because the free strain or displacement (in
plane: d31; out of plane: d33) of the piezoceramics is small,
they generally cannot be used directly as actuators in their
raw form; rather, amplification is required. Piezoelectric
ceramic actuators often employ mechanical amplifiers to
convert the small-induced strain of the ceramic material to a
large output displacement, which can be used for practical
applications. Compliant mechanism is often used as
mechanical displacement amplifiers to prevent displacement
losses that can occur in conventional pin-jointed
mechanisms.
If used in a restraint, the piezo actuators can generate
forces, which is always coupled with a reduction in
displacement. At maximum force generation, displacement is
zero. This maximum force (blocked force) depends on the
piezo actuator stiffness and maximum displacement.
Piezoelectric actuators can produce extremely fine position
changes, down to the sub nanometer range. The small

changes in operation voltage are converted into smooth


movements. They offer fast response time down to
microsecond time constants, and shows no wear and tear
because their displacement is based on solid state dynamics
and have neither gears nor rotating shafts. The devices only
absorb electrical energy during movement in static
operation, even holding heavy loads, they do not consume
power.
According to Uchino, application fields of piezoelectric
actuators are classified into three main categories:
positioners, motors, and vibration suppressors [1]. Regarding
to the first one, Uchino reports applications to optical
instruments such as lasers and interferometers, and the
positioning accuracy for fabricating semiconductor chips,
which must be adjusted using solid-state actuators, whose
manufacturing precision is of the order of 100 nm. On the
other side, in scanning probe microscopy, a powerful tool
frequently used in biology (for example), various systems are
applied for its scanning stages. In this case, piezoelectric
actuators are often used, allowing scanning area of several
square microns [2]. In cell manipulation there is a number of
different micromanipulation tasks, whose setups usually
consist of linear positioners, which carry the appropriate
micromanipulation tools (needles or micro pipettes, etc).
Piezoceramics are very well suited to replace conventional
manual positioning elements as well as motorized systems,
increasing the resolution and reducing the overall size of the
systems [3]. Recently, it was described a piezoelectric micro
actuator to drive the head suspension assembly of a 3.5 inch
high-performance commercial hard disk drive [4]. The
system has been proposed as a possible solution for attaining
wider bandwidths, larger than those associated to
electromagnetic micro actuators.
Piezoelectric ultrasonic motors are characterized by low
speed and high torque, which is contrasted with high speed
and low torque of electromagnetic motors. Ultrasonic motors
whose efficiency is insensitive to size are superior in the
mini-motor area. Tiny motors smaller then 1 cm, often
required in factory automation equipment, and are produced
with sufficient energy efficiency. They have higher
precision, high power-to-weight ratio, simplicity (fewer
parts), lower profile (no iron cores required), and minimal
EMI/RFI noise [5]. Another example, called inchworm, is a
linear motor where the strain output from actuators is
increased by using the frequency performance of the

piezoelectric material to rapidly move the actuator in one


direction in a series of small steps [6].
Vibration suppression in space structures and military
vehicles using piezoelectric actuators is also a promising
technology. Active vibration control has been used in a
number of smart structure applications, for example to isolate
an automobile engine from the car chassis, a helicopter rotor
from the fuselage, and marine engines from a ships hull.
These can be done by using variable reluctance, hydraulic,
electrodynamic, magnetosctrictive, and PZT actuators. By
using the product of the blocked force and the free
displacement, as figure of merit for the actuators, Brennan et
al. ranked PZT actuators as the better ones [7].
A number of piezo actuator patents have been found in
several companies all over the world: USA (Burleigh
Instruments Inc., EDO Corporation, Polytec PI Inc., Morgan
Matroc Inc., Dynamic Structures and Materials LCC, etc),
Europe (Marco Systemanalyse und Entwicklung GmbH,
Cedrat, MIDE Technology Corporation, etc), Canada (Sensor
Technology Ltd., etc), and Japan (NEC, TOTO Corporation,
Matsushita Electric, Toyota Motors, Hitachi Metal, Toshiba,
etc) [1], [8].
For the interest of this work, piezo actuators are essentially
used to produce displacements, so the study will emphasize
positioner actuators. A very efficient d33-type actuator and
compliant mechanism amplifier, designed to be used as
positioners, is the so called Piezoelectric Flextensional
(Flexural - Extensional) Actuator (PFA), which was first
developed in the late 1960s. Le Letty et al. described various
applications of flextensional actuators for aerospace as piezo
stages, tilt mechanisms, 6-degrees-of-freedom mechanisms,
active damping of vibrations of structures, and proportional
valves [9]. Flextensional actuators are currently being
manufactured by Dynamic Structures and Materials LCC
(Franklin, TN), EDO Corporation (Salt Lake City, UT), and
Cedrat (Meylan, France) [8].
In the present paper a new PFA designed by the topology
optimization method in a previous work [10] is
experimentally characterized by the measurement of its
nanometric displacements using a fiber optic displacement
sensor.
Fiber optic displacement sensors are useful for many
industrial applications in which noncontact measurements of
position, vibration and thickness are desired. Optical fiber
interferometry is one of the most accurate methods to
measure very small displacements, although its high
resolution is not necessarily required for most applications of
a displacement sensor. Instead, what is desired here is
dynamic measurement with a large dynamic range.
Among the broad variety of optical fiber sensor
modulation techniques and solutions that have been
discussed in literature, amplitude modulated optical fiber
sensor represent one of the simplest and earliest
developments in the field, which are still being extensively
investigated. The reason for this is their inherent simplicity,
reliability, flexibility, and relatively low cost added to the
well-known advantages of such sensors which include small
size, immunity to interference from electrical and magnetic
fields, total safety in explosive environments, chemical
inertness and intrinsic galvanic isolation.

A unit capable of measuring dynamic displacements of


piezoelectric actuators is the reflective fiber optic
displacement sensor (RFODS), also called reflective
intensity modulated fiber optic sensor [11], fiber optical
reflective displacement micrometer [12], fiber optic lever
displacement transducer [13], or fotonic sensor [14].
The basic RFODS structure consists of an input
(transmitting) fiber and a receiving fiber, both kept close to a
reflecting surface, as shown in Fig. 1. Any motion of the
reflector modulates the light intensity of the reflected light
beam entering the receiving fiber, consequently generating
an electrical signal proportional to the light variations (after
falling on the photo detector).
As the displacement
increases, the reflected light spot overlaps the core of the
receiving fiber and the output goes through a maximum.
After this point, the output decays slowly. The sensor can
work in either the positive or the negative slope region of the
curve. Other variations of this kind of sensors include a
single fiber with a directional Y-coupler and the Y-shaped
fiber optic bundle [13] - [15].

Fig. 1. Schematic diagram of the RFODS and the principle of the


displacement measurement.

Several industrial applications using the RFODS have been


described in the literature: dynamic measurement of high
frequency micro inch displacements of a magnetostrictive
transducer system used in ultrasonic welding [16];
measurement of the velocity of the back surface of an
irradiated aluminum target exposed to an intense pulsed
electron beam [17]; applications in the design of sensitive,
low frequency hydrophones of the pressure and pressuregradient types [15]; measurement and description of electric
arc breaking and the associated contact movement in a
vacuum chamber circuit breaker [18]; measurement of a solid
fraction in concentrated transparent flow [19]; the fiber
microphone for dynamic measurement of acoustic pressure
aiming active sound control applications [20], etc.
This work is divided in five Sections: in the Section 2 a
RFDOS modeling is presented; in Section 3 the PFA
assembly is described; in Section 4 the reflective fiber optic
sensor is tested and in Section 5 the experimental results are
listed. Linearity and frequency response of the PFA

displacement are evaluated, and the tracking error


phenomenon is discussed. Conclusions are presented in
Section 6.
II. THE RFODS MODELLING
The theoretical
formulation of the reflective light
intensity function at the distal end plane developed by He &
Cuomo, is applied, assuming that a lossless, monomode stepindex fiber is illuminated in a manner such that each light ray
propagating through the optical fiber core at different angles
( ) with respect to the fiber axis carries the same power [21].
Referring to Fig. 2, k = q / x 0 is defined as a dimensionless
coordinate on the image plane, while k c = q c / x 0 gives the
boundary of the illuminated area, where x0 is the transmitting
fiber core radius.

indexes of core and cladding for the monomode transmitting


fiber, respectively.
Similarly, light rays exiting the distal end will form a cone
with a maximum angle c. The reflected intensity on the
receiving plane is equivalent to that on the image plane.
Using the uniangular beam concept, as defined by He &
Cuomo, and considering the receiving fiber as multimode, it
can be derived the reflected optical power (Po, subtended by
the receiving multimode fiber) to the total optical power
exiting the transmitting fiber (Pi) ratio as [21]

Po 2
=
Pi

b
m p

I k (k )
k dk
Io

(1)

where the parameters b, and m are defined as:


if k c m + p
k c
b=
m + p if k c > m + p

(2)

= cos 1

m2 + k 2 p 2

2mk

(3)

m = x o + c m + c mr + xor

(4)

and where cm and c mr = h.c m are the cladding thicknesses of


transmitting and receiving fibers, respectively, x 0 r = p.x 0 is
the radius of the receiving fiber (h and p are constants), and
I0 is the total light intensity exiting the distal end of the
transmitting fiber. The relation between the intensity
distribution on the image plane as function of fiber core sizes
(xo and xor, for transmitting and receiving fiber, respectively),
numerical aperture (NA), and distance (D) between the fiber
end and reflector is given by
x0 I 0
(k c 1) 2 [1 + A 2 (k 1) 2 ]
ln

2
2
2
16 D c (k 1) [1 + A (k c 1) ]

for k c > 2 , k > 2 , k c k 2

I k (k ) =
x I
2
2
2
0 0 ln (k + 1) [1 + A (k 1) ]
16 D c (k 1) 2 [1 + A 2 (k 1) 2 ]

for k c > 2, k > 2, k c k > 2

(4)
Fig. 2. Coordinates of a two-fiber RFODS.

A light source illuminating the transmitting fiber will


contain beams arriving from all angles and the rays limited
by c , where c is the cutoff angle, will couple into the
fiber. This cutoff angle is determined by the numerical
aperture (NA) of the fiber and the refractive index of the
surrounding medium (n) according to: c = sen 1 ( NA / n) , in
which NA = n12 n22 , where n1 and n2 are the refractive

and where A = x o / 2 D . This model will be applied in the


next sections.
III. PIEZOELECTRIC FLEXTENSIONAL ACTUATOR
ASSEMBLY
The PFA prototype is shown in Fig. 3. The metal endcaps
serve as mechanical transformers for converting and
amplifying the lateral motion of the ceramic into a large axial
displacement normal to the endcaps. Both the d31 and d33
coefficients of the piezoceramic contribute to the axial

displacement of the composite structure. Displacement


decreases always from the center of the endcap, where the
maximum displacement is obtained, to the edge, where
displacement is equal to that of the piezoceramic without the
endcap.
Mirror

(a)

there is no light coupling to the receiving fiber. By increasing


D, the output detected signal goes through a maximum for
925 m. For higher values of D the output signal decays
smoothly.

Analysis point

(b)

Fig. 3. Prototype of manufactured flextensional piezoactuator. a)


Top view. b) Lateral view.

The aluminum flexible structure was manufactured by


using wire EDM (Electrical Discharge Machining). The
piezoceramic (PZT 5A, American Piezoceramics, 30 mm x
14 mm x 3 mm in directions 1, 2 and 3, respectively) is
polarized in direction 3 and electrodes are deposited on the 12 plane. The endcaps are opened and the piezoceramic is
bonded to the flexible structure with epoxy resin, so its
extensional vibration mode is coupled to the aluminum
structure by shearing in the epoxy resin. Due to the
difficulties to polish the irregular piezoactuator surface to an
optical degree, a 200 m thickness mirror, obtained by
aluminum vaporization over a glass plate, was bonded to the
actuator surface with epoxy resin. As the stiffness is of the
order of 1010 N/m2, the effect of an additional oscillator,
constituted by the mirror and epoxy resin, is minimized [22].
The PFA is fixed to a holder by three points, perpendicular to
the displacement to be measured. As a result the actuator is
free to vibrate in directions 1 and 3.
A review of modeling of stroke-amplified piezoceramic
actuators and their applications to compliant mechanism
design is given in [10].

Fig. 4. Experimental setup for testing the RFODS.

Fig. 5. Microscopic view of the transmitting and receiving fiber


heads.

IV. RFODS EXPERIMENTAL SETUP


A low cost version of the reflective optical fiber sensor
was built and tested. The RFODS setup, used to measure the
PFA vibration amplitudes, is sketched in Fig. 4. The PIN
photodiode output signal is amplified, digitized by an
oscilloscope (Tektronix TDS2022) and transferred to a
microcomputer via RS232 port. The effect of the ambient
light is cancelled by the use of a synchronous detection
device (lock-in amplifier).
A microscopic view of the transmitting and receiving fiber
heads is shown in Fig. 5. The monomode optical fiber
parameters are: n1=1.465, n2=1.460, x0=4 m and cm=58.5
m. The relevant parameters of the multimode optical fiber
are: x0r=31.25 m and cmr=31.25 m.
Figure 6 shows the theoretical and experimental outputs of
the receiving fiber versus displacement using the model
developed in Section II. When the gap D (the distance
between the fiber and the reflector) is smaller than 400 m

Fig. 6. Response curves (theoretical and experimental) of the


RFODS.

The only obvious mismatch between the results is that the


experimental curves are shifted a little to the right toward a
larger displacement. This is probably due to nonuniform
intensity within the light spot image.

V. EXPERIMENTAL RESULTS
With the RFODS operating around the mid point (adjusted
with the aid of a micrometer-translation stage) in the positive
slope region shown in Fig.6, the PFA is driven with
sinusoidal signals. At frequencies of approximately 15 kHz
and 23 kHz with varying amplitudes applied to the
piezoactuator, the input-output relations are shown in Fig. 7.
These frequencies correspond to two PFA resonances. The
curves show a linear response in the displacement range
considered.

Fig. 7. Output detected voltage as a function of the input voltage


applied to the PFA, at the frequencies of 14.77 kHz and 23.4 kHz.
The least-squares fit to these plots is also shown.

terms of the output voltage amplitude normalized by the


input voltage.
The first resonance at 3.86 kHz (with a small amplitude),
is far away from the operating frequency of 0 10 Hz for a
typical application of the device. Resonance frequencies are
also found at 15, 23 and 32 kHz. From d.c. to approximately
3.8 kHz the PFA frequency response is approximately flat.
All these results are in accordance with those obtained by the
authors in a previous work, using a Michelson homodyne
interferometer [23].
For a PFA positioner applications, a rapid drive voltage
change must result in a rapid position change. This is
necessary in applications such as switching of
valves/shutters, generation of shock-waves, vibration
cancellation systems, etc. However, rapid actuation of
nanomechanisms can cause recoil-generated ringing of their
loads and any adjacent components. This ringing can take
hundreds of milliseconds to damp out. In the following
example, the PFA is optically tested at 818 and 1800 Hz
frequencies of triangular drive voltages. In Figs. 9 and 10 it
is illustrated their respective photo-detected voltages, which
are proportional to the actuator surface displacements.
Apparently, it seems that the signals are noisy, however, a
detailed analysis reveals that this behavior is mainly due to
the tracking error phenomenon. As the resonant frequencies
of the piezoactuator are not sufficiently high, unwanted
dynamics of the stage are often a problem, especially for
scanning motion with high frequency, but also if trajectories
have to be followed with high velocities. Under this mode of
operation there is no more direct relationship between the
control voltage and the PFA displacement.
According to the literature, a PZT actuator can reach its
nominal displacement in approximately 1/3 of the period of
the resonant frequency, albeit with significant overshoot.
Nevertheless, as described before, its settling time can take
various milliseconds. Next the PFA is driven by 1000 Hz
square waveforms (whose periods are smaller than the
settling time), as illustrated in Fig. 11. As it is well known,
from the step response of the system the natural frequencies
could be obtained. However, the square waveform is more
illustrative, permitting the comparison between the
oscillating response and its period.

Fig. 8. PFA frequency response in terms of output detected voltage


normalized by applied input voltage.

A dynamic test was performed on the compliant


mechanism actuator to determine the resonant frequencies of
the system. Sinusoidal driving voltages of different
frequencies were used, and the measured response is shown
in Fig. 8 (for frequencies between 1 kHz and 45 kHz) in

Fig. 9. Response of the PFA to a 31 Vpeak-to-peak , 817.8 Hz


triangular drive signal.

ACKNOWLEDGEMENT
Support for this project from the sponsor agency CAPES
is gratefully acknowledged.
REFERENCES
[1]

[2]

Fig. 10. Response of the PFA to a 31.6 Vpeak-to-peak , 1799 Hz


triangular drive signal.

As can be observed, there are 23 oscillation cycles per


period of the 1000 Hz square waveform in Fig. 11, which
result in 23 kHz frequency oscillations, the main PFA
resonance frequency as obtained before (see Fig. 8).

[3]

[4]

[5]

[6]

[7]
Fig. 11. PFA response to a 13.1 Vpeak-to-peak, 1000 Hz square
waveform voltage.

[8]

A potential real-time feed forward technology called


inputShaping can cancel out resonances before they start,
rather than waiting for them to damp out [3]. By applying
this technique it is possible to obtain the fastest possible
motion, with virtually instant settling. Currently the authors
are getting evolved in the study of this kind of problem.

[9]

VI. CONCLUSION

[10]

A new PFA, designed with the topology optimization


method, was characterized. Its dynamic performance was
investigated, emphasizing the tracking error problem.
Linearity and frequency response of the displacement
amplitudes were evaluated using a reflective optical fiber
sensor. All the resonance frequencies up to 45 kHz were
determined. The gain was fairly smooth from d.c. up to
approximately 3.8 kHz. The results obtained in this work
agreed with those measured with a Michelson interferometer
in a previous work. However, the RFODS system is simpler,
cheaper and easier to operate than the interferometer.

[11]

[12]

[13]

Uchino, K., Recent Trends of Piezoelectric Actuator


Developments, IEEE International Symposium on
Micromechatronics and Human Science, pp.3-9, 1999.
Wille, A., Rau, S., Gloess, R., Calibration Method
Enables Alignment Down to 1 m, Photonics Spectra
Magazine, September 2003 edition, www. photonics.
com/
spectra/
tech/
XQ/
ASP/
techid.
1588/QX/read.htm, accessed in Sep. 2005.
Bergender, A., Driesen, W., Varidel, T., and Breguet,
J.M., Development of Miniature Manipulators for
Applications in Biology and Nanotechnologies, In:
Proceedings of Workshop "Microrobotics for
Biomanipulation", IEEE IROS 2003, Las Vegas, USA,
pp. 11-35, Oct. 27-31, 2003.
Koganezawa, S., Hara, T., Development of ShearMode Piezoelectric Microactuator for Precise Head
Positioning, Fujitsu Sci. Tech. J., 37 (2), pp.212-219,
Dec. 2001.
EDO Model PDM130 Piezoelectric Motor
Application Notes, EDO Document Control Number
52350,
In:
www.
edoceramic.
com/
Assemblies_Products/
PEMot%20AppNotes.htm,
accessed in Sep. 2005.
Silva, E.C.N., Topology Optimization Applied to the
Design of Linear Piezoelectric Motors, Journal of
Intelligent Material Systems and Structures, vol. 14,
pp.309-322, April/May 2003.
Brenan. M.J., Bonito, J.G., Elliot, S.J., David, A.,
Pinnington, R.J., Experimental Investigation of
Different Actuator Technologies for Active Vibration
Control, Smart. Mater. Struct., 8, pp. 145-153, 1999.
Niezrecki, C., Brei, D., Balakrishnan, S., Moskalik, A.,
Piezoelectric Actuation: State of the Art, The Shock
and Vibration Digest, vol. 33 (4), pp.269-280, Jul.
2001.
Le Letty, R., Claeyssen, F., Barillot, F., Lhermet, N.,
Amplified Piezoelectric Actuators for Aeroespace
Applications, AMAS Workshop on Smart Materials
and Structures SMART03, Jadwisin, pp. 51-62, Sep.
2-5, 2003.
Silva, E.C.N., Kikuchi, N., Topology Optimization
Design of Flextensional Actuators, IEEE Transactions
on Ultrasonics, Ferroelectrics, and Frequency Control,
vol.47, no.3, pp.597-605, May 2000.
Zheng, J., Albin, S., Self-Referenced Reflective
Intensity Modulated Fiber-Optic Displacement
Sensor, Opt. Eng., vol.38 (2), pp.227-232, Feb. 1999.
Kleiza, V., Paukste, J., Verkelis, J., Modelling Light
Transmission in a Fiber-Optical Reflection System,
Nonlinear Analysis: Modelling and Control, vol. 9 (1),
pp.55-63, 2004.
Cook, R.O., Hamm, C.W., Fiber Optic Lever
Displacement Transducer, Applied Optics, vol. 18
(19), pp.3230-3241, Oct. 1979.

[14] Menadier, C., Kissinger, C., Adkins, H., The Fotonic


Sensor, Instrument & Control Systems, vol.40,
pp.114-120, Jun.1967.
[15] Cuomo, F., Pressure and Pressure Gradient FiberOptic Lever Hydrophones, J. Acoust. Soc. Am.,
vol.73, n.5, pp.1848-1857, May 1983.
[16] Crispi, F.J., Maling Jr., G.C., Rzant, A.W.,
Monitoring Microinch Displacement in Ultrasonic
Welding Equipment, IBM J. Res. Develop., pp.307312, May 1992.
[17] Salins, R.B., Plastic Optical Fiber Displacement
Sensor for Study of the Dynamic Response of a Solid
Exposed to an Intense Pulsed Electron Beam, Rev.
Sci. Instrum., vol. 46 (7), pp.879-882, Jul. 1975.
[18] Anghel, F., Pavelescu, D., Grattan, K.T.V., Palmer,
A.W., Contact Dynamic Recording and Analysis
System Using an Optical Fiber Sensor Approach,
Review of Scientific Instruments, vol. 68 (9), pp.35833589, Sep. 1997.
[19] Bergougnoux, L., Ripault, J.M., Firpo, J.L.,
Characterization of an Optical Fiber Bundle Sensor,
Review of Scientific Instruments, vol. 69, n.5,
pp.1985-1990, May 1998.

[20] Bucaro, J.A., Lagakos, N., Lightwave Fiber Optic


Microphones and Accelerometers,
Review of
Scientific Instruments, vol. 72, n.6, pp.2816-2821, Jun.
2001.
[21] He, G., Cuomo, F.W., A Light Intensity Function
Suitable for Multimode Fiber-Optic Sensors, Journal
of Ligthwave Technology, vol. 9, n.4, pp.545-551,
Apr. 2001.
[22] Zhang, Q.M., Pam, W.Y., Cross, L.E., Laser
Interferometer for the Study of Piezoelectric and
Electrostrictive Strains, Journal of Applied Physics,
vol. 63, pp. 2492-2496, 1988.
[23] Maral, L.A.P., Leo, J.V.F., Nader, G., Silva, E.C.N.,
Higuti, R.T., Kitano, C., Analysis of Linearity and
Frequency Response of a New Piezoelectric Actuator
Using a Homodyne Interferometer and the J1-J4
Method, Procceedings of the 2005 IEEE
Instrumentation and Measurement Technology
Conference, Ottawa, Ontario, Canada, pp.1048-1053,
May 2005.

Você também pode gostar