Você está na página 1de 12

1

DYNAMIC MODELING OF AN ELECTROSTATIC ACTUATED


CANTILEVERED MICROMIRROR
JIANLIANG YOU
MUTHUKUMARAN PACKIRISAMY
ION STIHARU
Optical Microsystems Laboratory, Concave Research Center
Department of Mechanical & Industrial Engineering, Concordia University
1455 de Maisonneuve Blvd. West, Montreal, Quebec H3M 1G8, Canada
This paper investigates dynamic performance of a rectangular micromirror supported by
a straight cantilevered beam and actuated by either one of the two bottom electrodes. The
established analytical model is based on assumptions that bending motion and torsion
motion of the actuator are independent with each other, and bending motion of the
microplate is decomposed as a translation at the cantilever tip and a rotation about the
lateral axis of the point, as well as that the microplate is assumed rigid. The model has
demonstrated the independence between torsion and bending motions. Hence a uniform
beam cross-section that leads to stiff bending and soft torsion of the structure can be
used to form the hinge for a simple torsional micromirror. This new concept is then
verified through case study of both analytical model and the corresponding FEA
simulation. Dynamic performance of the actuator was analyzed through dynamic
responses of the structure under various applied voltages. Comparisons of the calculated
resonant frequencies and responses with simulated results also proved accuracy of the
analytical model.
1. Introduction
Micro electrostatic actuators that consist of a microplate and the corresponding
suspension have vast applications in optical cross-connects (OXC), scanners,
optical switches, digital displays, accelerometers, etc. Though electrostatic
actuation exhibits binary unstable deflection when the drive voltage reaches at
the pull-in voltage of the structure, it shows many advantages that have been
proved in recent years, such as, simple structure, easy to fabricate, out-of-plane
motion, quasi-linearity before pull-in, tunable working frequency, and
controllable static and dynamic performance. The electrostatic actuators
fabricated are usually designed to realize either a translation or a rotation with
one DOF (degree of freedom) of motion. Coupled motion and pull-in
phenomenon of a structure by electrostatic actuation with two degrees of
freedom has been modeled and validated through FEA simulations [1-4].
2
However, there are few articles discussing on the dynamic performance of
electrostatic actuators. As it is common when dealing with problems of large
deflection and oscillation application, characterization of dynamic performance
of electrostatic actuators has become an essential issue. This characterization
helps predict the possible response of the actuator under excitation [5].
There have appeared several electrostatic torsion structures in recent years.
A rectangular mirror suspended symmetrically by two straight beams is a typical
structure for torsional micromirrors [6]. Another structure for torsional actuators
is composed of a rectangular micromirror that is supported by the two straight
beam segments attached to an edge of the mirror [1]. In order to soften the
torsional stiffness of the torsional micromirror, the straight beam is replaced by
the planar torsional serpentine spring [7] thereby reducing the pull-in voltage of
the structure and realizing the compatibility of integration with CMOS circuits. It
has been proved many times that torsional spring constant is much lower than
bending stiffness for the same torsional hinges. Starting from this understanding,
a cantilever suspended plate is modeled for applications such as spatial optical
switching or steering. Although nonlinearity of electrostatics around pull-in
status deviates a bit from the linear assumption of the performance, for a wide
range of applied voltage the calculated static bending displacement, torsion and
rotation angles according to the analytical model are still within an acceptable
error compared with the FEA results. The investigation thus has verified the
applicability of the deduced expressions for the electrical load with assumption
of small deflection [8].
With only one support cantilever the proposed torsional micromirror based
actuator is shown in Fig. 1. Dynamic characterization of the actuator is presented
in this paper. The spring constants for the cantilevered based microplate are less
than half of the original symmetric beam-plate structure. Investigation of
dynamic responses of the actuator to the step voltage excitation is the focus of
this work. Characteristic analysis of the dynamic performance is performed
through the lumped model with three degrees of freedom. The straight cantilever
beam with a rectangular cross-section is used as a simple suspension to support
the microplate attached to the tip of the cantilever. This simple structure can be
easily fabricated by Micragem technology (Micralyne GEneralized MEMS:
An SOI-based MEMS fabrication process) [9]. Verification of the dynamic
performance is implemented by FEA simulation using ANSYS. The torsional
micromirror with very soft torsional stiffness and negligible bending effect can
thus be realized through deliberate selection of cantilever geometry parameters.
3
2. The Coupled Dynamic Model
The cantilever-plate actuator is composed of an assumed rigid rectangular plate
and a straight cantilever beam with uniform rectangular cross-section. The
suspended plate can rotate with respect to the cantilever axis as well as along the
lateral direction of the attached location of beam and plate when a bias voltage is
applied between the plate and one of the two electrodes underneath the plate.
The two isolated, equal sized electrodes are symmetrically arranged with respect
to the axis of the cantilever and have an isolated distance of a between them. Fig.
1(a) shows the 3D view of the actuator and Fig. 1(b) and (c) also indicate
geometry parameters of the actuator.
Due to highly uniform surface and relatively large size, the SOI (silicon on
insulator) fabricated microplate is assumed to be rigid and thus can be used to
reflect the incident light beam without divergence or diffusion. The thin
cantilever beam is deliberately designed so that the resulted slope angle at the
cantilever tip due to an applied voltage bias on the actuator can be negligible as
compared with that of the torsion angle. It has been noticed in [8] that for a short
straight cantilever beam the torsion angle is much larger than that of the bending
slope at the tip or the torsional spring constant of a cantilever is several orders
smaller than the bending stiffness. However, as is shown in Fig. 1 (d) the
equivalent force balance diagram for the actuator considers the rigid plate as a
lumped mass and the electrostatic loads as vertical bending force F
e
, torque T
e

and bending moment M
e
that correspond to deflections z, and at the
cantilever tip, respectively.
Though the ratios of vertical bending stiffness with other torsion and slope
spring constants of the beam determine the coupling effect among the assumed
three DOFs, it is common that a specifically designed cantilevered actuator can
be used as a custom OXC (optical cross connection) with simple control
algorithm when the structure is intended to have negligible bending deflection
and no coupling effect between torsion and bending. As a matter of fact two
DOF among the above assumed 3-DOF model are actually belonged to the same
one DOF, i.e. both vertical translation and bending slope at the tip together
constitute the bending deflection of the actuator. Nevertheless, the decoupled 3D
analytical model is adopted in this paper for convenience.
The applied electrostatic loads on the cantilever tip of the actuator can be
given by [8]



(
(
(

2
2
ln
2
2
ln
2
0
0
0
0
2
a
L z g
a
z g
W
L z g
W
z g
V
F
e
(1)
4




























respectively, where is the permittivity of the air, V the applied bias, g
0
the gap
between the plate and the bottom electrode, W the width, and L the length of the
plate, a the isolated distance between two bottom electrodes.
Distributed mass of the lengthy cantilever beam has to be considered when
dealing with dynamic performance. The equivalent mass of the cantilever beam
with respect to the tip motions is deduced through assumption of third-order
polynomial mode shapes of the beam and their integration along the total length
is given with the following formula:


(b)
(c)
(d)
Fig. 1 The schematic view, geometry parameters and the equivalent force diagram of the
actuator.
(a)
(
(
(

2
2
ln ) (
2
2
ln ) (
2
0
0
0
0
0
0 2
2
W
z g
a
z g
z g
W
L z g
a
L z g
L z g
V
T
e
(2)
(
(
(
(
(
(
(
(
(
(
(



+
+

=




2
2
ln )
2
ln( )
2
(
)
2
ln( )
2
(
)
2
ln( )
2
( )
2
ln( )
2
(
2
0
0
0 0
0 0
0 0 0 0
2
2
W
L z g
a
L z g
L
W
L z g
W
L z g
a
L z g
a
L z g
a
z g
a
z g
W
z g
W
z g
V
M
e
(3)

=
l
j i ij
dx m m
0
(4)
5
where m, l are the mass per unit length and length of the beam. The equivalent
mass is thus derived as



According to Lagranges equations,



where T and U are the systems kinetic and potential energies; q
i
is a deflection
variable, Q
i
is the external load. The matrix dynamic equation of the actuator
when including the effect from the cantilever beam can then be given as follows




where




and M is the mass of the plate.
The stiffness matrix in the equation corresponds to


where I is the cross-sectional area moment of inertia of the beam and J the cross-
sectional polar moment of inertia of the beam, which are given by



E is Youngs modulus, G is the shear modulus. The polar moment of inertia of
the beam can also be defined more precisely as



For tw, w and t should be replaced between each other in the formulae.
It is noticed that the second order differential equations of the actuator in (7)
are nonlinear and coupled among the three variables due to the nonlinear
(
(
(

=
2
2 0 11
0 70 0
11 0 78
210
l l
l
ml
m
e (5)
( ) ( )
i
i i
Q
q
U T
q
U T
dt
d
=

|
|

\
|


&
(6)

(
(
(

+
(
(
(

(
(
(

e
e
e
M
T
F z
k k
k
k k z
m m
m
m m

33 31
22
13 11
33 31
22
13 11
0
0 0
0
0
0 0
0
& &
& &
& &
(7)
2
13
210
11
2
ml
ML
m =
ml
MW
m
3
1
12
2
22
+ =
13 31
m m =
3
2
33
105
1
4
ml
ML
m + =
ml M m
35
13
11
+ =
3 11
12
l
EI
k =
2 13
6
l
EI
k =
l
GJ
k =
22
13 31
k k =
l
EI
k
4
33
=
(

|
|

\
|
=
4
4 3
12
1 36 . 3
3
16
16 w
t
w
t wt
J
12
3
wt
I = (tw) (8)
(

\
| +
+
=

=
t
w k
k w
t wt
J
k
2
) 1 2 (
tanh
) 1 2 (
1 192
1
3
0
5 5
3

(tw) (9)
6
electrostatic property. Resonant frequencies of the structure can be estimated by
solving the determinant equation:


It can be seen from the equation that the torsional motion of the actuator is
independent from other motions, that is, the natural frequency for torsional
motion is deduced from letting the first parenthesis term equal to zero as

whereas the resonant frequency of bending motion is resolved from the second
parenthesis term.
3. Simulations and Discussions
Though the static property of the actuator that was investigated in the previous
work [8] has shown the nonlinear electrostatic curve when the applied voltage is
close to the pull-in value, it is applicable and acceptable for the non-deformation
assumption to be applied in dynamic analysis for its large range electrostatic
characterization. Thus with the same rigid plate assumption, this section presents
dynamic performance of the established analytical model and the corresponding
FEA simulations with specific parameters and dimensions of the actuator.
3.1. Electrostatic Equivalent Stiffness Matrix
Generally eigen-frequencies of a structure mainly depend on the stiffness term
and mass term of its governing vibration equation. The nonlinear electrostatic
loads imposed on the actuator can be approximated through Taylor series
expansion. The electrostatic load has been decoupled with respect to the three
coordinates as defined in the previous section, which are, the vertical bending
force at the tip, the bending moment about the tip and the torsional moment
about the lateral direction of the tip, as shown in Fig. 1 (d). Developing
expressions for F
e
, T
e
and M
e
about an initial static point (z
0
,
0
,
0
) of the
actuator under an applied voltage and discarding terms of the second and higher
orders, yield the following expressions for equivalent spring constants:






( )( ) ( ) ( ) 0 ) (
2
2
13 13
2
33 33
2
11 11
2
22 22
= m k m k m k m k (10)
22 22 2
/ m k =
(11)
z
z F
k
e
e

=
) , , (
0 0 0
11

=
) , , (
0 0 0
12
z F
k
e
e

=
) , , (
0 0 0
13
z F
k
e
e
z
z T
k
e
e

=
) , , (
0 0 0
21

=
) , , (
0 0 0
22
z T
k
e
e

=
) , , (
0 0 0
23
z T
k
e
e
z
z M
k
e
e

=
) , , (
0 0 0
31

=
) , , (
0 0 0
32
z M
k
e
e

=
) , , (
0 0 0
33
z M
k
e
e
7
where k
e11
, k
e12
, k
e13
mean the equivalent translational bending, rotational
bending and torsional spring constants due to the vertical electrostatic load F
e
.
The other equivalent stiffness coefficients are defined similarly. Substitution of
these coefficients into the governing matrix equation (7) yields




The right matrix term represents the electrostatic force, torque and moment
that are resulted from the electrical load about an equilibrium point. It is clear
from the stiffness matrix that each individual member in the matrix is subtracted
by a positive value indicating the softening effect of the electrostatics on the
general stiffness of the structure. The three fundamental natural frequencies can
be calculated by solving for the eigenvalues of the second order differential
equations.
3.2. Frequency Spectrum
In order to compare the analytical model with FEA simulation results, it is
necessary to set up all required parameters of the actuator, which are shown in
Table 1.
Table 1 Parameters of An Actuator for Analysis
microplate
LW
(m
2
)
cantilever
lw (m
2
)
thickness
t (m)
gap
g0
(m)
isolated
distance
a (m)
Young
modulus
E (GPa)
Poisson
ratio
density

(kg/m
3
)
200400 60020 10 12 40 129.5 0.21 2320
Simulation and calculation for free vibration of the actuator without applied
voltage give a slight difference to the fundamental frequency. As shown in Table
2, the first bending mode solved from ANSYS model has a slight higher
frequency than that of analytical model with an error of 1.8%. There even exists
a small difference of the value when SOLID45 and SOLID95 element meshing
are involved in the simulation. Still, it should be noticed that there has not
appeared the first torsional mode frequency in ANSYS simulation, which has
seldom been mentioned and analyzed in the literature so far although there
apparently exists this mode due to the relatively large lateral dimension of the
plate and several tens order smaller torsional stiffness of the cantilever beam.
With different voltages applied to the actuator the solved resonant
frequencies are shown in Table 3 for comparison. The difference of frequencies
is calculated through subtraction with that of the analytical results from
MATLAB. Using different element types in ANSYS model for meshing of the

(
(
(




+
(
(
(

(
(
(

0
0
0
33 33 32 31 31
23 22 22 21
13 13 12 11 11
33 31
22
13 11
0
0 0
0
e
e
e
e e e
e e e
e e e
M
T
F z
k k k k k
k k k k
k k k k k z
m m
m
m m

& &
& &
& &
(12)
8
structure resulted in a small difference of the values when the applied voltage is
lower than pull-in value. As the voltage increases, nonlinearity of electrostatics
becomes more dominant and this difference becomes larger. The results show
that Solid95 meshing has much closer values than Solid45 meshing to calculated
results. However, for an actuator of this type used for oscillation purposes the
applied voltage is usually within an almost linear range of the electrostatics, thus
either element types can be used for simulation though preference should be
given to Solid95.
Table 2 The fundamental frequency without applied voltage
Frequency MATLAB ANSYS (solid45) ANSYS (solid95) error
1st bending 5029.8 Hz 5122.5 5120.8 1.8%
1st torsion 10.56 Hz Not shown Not shown ---
Table 3 also validates that a linear working range exists since the difference
between analysis and simulation is within an acceptable range. However, when
dealing with pull-in or near pull-in performance of the actuator, analytical results
again shows a large deviation from simulated results, which demonstrated that
deformation of the microplate must be considered in this range. The first bending
frequency versus applied voltage from the three models is shown in Fig. 2.
Table 3 Variation of the first bending mode frequency with applied voltage
Voltage
(V)
Analyzed
(MATLAB)
Simulated
(ANSYS, solid45)
Difference
(%)
Simulated
(ANSYS, solid95)
Difference
(%)
0 5029.8 5122.5 1.8 5120.8 1.8
10 5024.2 4970.0 1.1 5097.7 1.5
20 5006.1 4888.5 2.3 5023.4 0.3
30 4970.7 4725.1 4.9 4878.4 1.9
40 4902.6 4401.9 10.2 4607.7 6.0
It is interesting to notice that the first torsional mode frequency varies only
within a 1Hz range when the applied voltage varies from zero to a maximum
value close to pull-in status (Fig. 3). Moreover there is a huge difference in
frequency between torsion and bending modes (10Hz and 5000Hz, respectively).
This phenomenon reveals a fact that the cantilevered microplates can be applied
for torsional micromirrors through proper selection of the cantilever dimensions.
As long as the bending deflection is kept within an acceptable range of error, the
cantilevered micromirrors can replace the current torsional micromirrors in the
applications for precise optical switching.
By Runge-Kutta method frequency spectrum of the actuator corresponding
to a stepped applied voltage is solved and shown in Fig. 4. There are two motion
frequencies within 25 kHz range, one with 10 Hz frequency that has a much
larger power density and the other with around 5000 Hz frequency that has a
relatively low power of vibration. This result agrees very well with the resonance
analysis.
9






































0
1000
2000
3000
4000
5000
6000
0 10 20 30 40 50 60
Applied Voltage(V)
F
u
n
d
a
m
e
n
t
a
l

B
e
n
d
i
n
g

F
r
e
q
u
e
n
c
y
(
H
z
)
MATLAB soild45 solid95
Fig. 2 The first bending frequency versus the applied voltage.
10.15
10.2
10.25
10.3
10.35
10.4
10.45
10.5
10.55
10.6
0 10 20 30 40 50 60
Applied Voltage (V)
F
u
n
d
a
m
e
n
t
a
l

T
o
r
s
i
o
n

F
r
e
q
u
e
n
c
y
(
H
z
)
Fig. 3 The first torsion frequency versus applied voltage
Fig. 4 Frequency spectrums with (a) no voltage and (b) applied voltage of 30 V.
(a) (b)
10
3.3. Dynamic Responses
Dynamic response of the actuator in this work only considered deflection
responses to a suddenly applied constant voltage. Fig. 5 shows the torsional
responses of the actuator with three different applied voltages. A lower applied
voltage on the actuator leads to a smaller amplitude of the response and the
torsional angle is always kept at a range from zero to a positive value. The
average torsion angle or the amplitude of torsion motion can be also estimated to
be a positive value. The period or frequency of the torsion motion can be read
out from it, that is, approximately 0.95s or 10.5Hz, which is in agreement with
the previous calculation.
Comparatively the bending and rotational bending motion has shown much
higher frequency. Fig. 6 shows the responses of bending motion at cantilever tip
and rotational bending at the same point. The left column lists the bending
responses whereas the right column lists the rotational bending. The two curves
on the same row are resulted from under the same applied voltage. It is clear
from any pair of curves on the same row that the rotational bending and
translational bending share the same frequencies. This phenomenon agrees with
the actual structure since two of the 3 DOF are coupled and belong to one degree
of freedom. The actual structure has only two degrees of freedom: one bending
motion and the other torsion motion.














Amplitude of the first bending mode decreases as applied voltage increases
whereas amplitude of the second bending mode varies inversely. However the
amplitude for the second bending mode is always smaller than that of the first
mode. It can also be noticed from the figure that the bending motion oscillates
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
0
0.2
0.4
0.6
0.8
1
x 10
-3
Time (s)
T
o
r
s
i
o
n

A
n
g
l
e

(
r
a
d
i
a
n
)
V=30 V
V=40 V
V=20 V
Fig. 5 Response of torsional mode to the step voltage applied.
11
around the original balanced position when the applied voltage is small. In other
words, the cantilevered plate is deflected up and down periodically. Due to the
relatively large length of the cantilever beam, amplitude of the bending motion is
in the same order with that of torsion motion. It can be used in spatial steering
for optical beam switching if equally deflected angles are reached by proper
selection of the dimensions of the cantilever and the plate. Another application is
for torsional micromirrors if the negligible bending deflection is realized through
stiffening the bending motion and softening the toraional motion.

















4. Conclusions
The dynamic model for electrostatic actuated cantilevered micromirrors is
established. The accuracy of the analytical model for dynamic performance of
the actuator is verified through FEA simulations. The simulation models from
two different element settings have resulted in different eigen-frequencies and
deflections especially when the conditions are close to pull-in. Element
SOLID95 instead of SOLID45 is suggested for dynamic simulations for these
electrostatic actuators. Dynamic responses of the actuator to the step constant
voltage are solved and analyzed. Two promising applications of the actuators are
proposed as long as the dimensions of both the cantilever beam and the plate are
properly selected.
Fig. 6 Responses of bending motion (left column) and slope rotational motion (right
column) to the step voltage of 0, 20, and 40V.
12
References
1. Ofir B. Degani and Yael Nemirovsky, J. of Microelectromechanical systems
11, 20(2002).
2. J.-M. Huang, A.Q. Liu, et al, Sensors and Actuators A115, 36 (2004).
3. Ofir B. Degani, Yael Nemirosky, Sensors and Actuators, A97-98, 569
(2002).
4. Z. Xiao, W. Peng and K. R. Farmer, J. of Microelectromechanical systems
12, 929(2003).
5. Jian Ping Zhao, Hua Ling Chen, Microsystem Technologies 11, 1301
(2005).
6. Avinash K. Bhaskar, Muthukumaran Packirisamy, Rama B. Bhat,
Mechanism and Machine Theory 39, 1399 (2004).
7. Jianliang You, Muthukumaran Packirisamy and Ion Stiharu, The 3rd Int.
Conf. on Intelligent Sensing and Information Processing, Bangalore, India
(2005).
8. Jianliang You, Muthukumaran Packirisamy and Ion Stiharu, IEEE Int.
Symposium on Industrial Electronics (ISIE'06), Montreal, Canada (2006).
(Accepted)
9. www.micralyne.com, Micralyne Inc. 1911-94 Street, Edmonton, Alberta,
Canada, T6N 1E6.
10. N. Lobontiu, E. Garcia, Mechanics of Microelectromechanical Systems,
Kluwer Academic Publishers, 2005.
11. G. K. Ananthasuresh, Optical Synthesis Methods for MEMS, Kluwer
Academic Publisher, 2003.

Você também pode gostar