Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/225401642

Dynamic pull-in phenomenon in MEMS resonators

Article  in  Nonlinear Dynamics · April 2007


DOI: 10.1007/s11071-006-9079-z

CITATIONS READS

272 558

3 authors:

Ali H. Nayfeh Mohammad Ibrahim Younis


Virginia Polytechnic Institute and State University King Abdullah University of Science and Technology
798 PUBLICATIONS   26,562 CITATIONS    230 PUBLICATIONS   4,172 CITATIONS   

SEE PROFILE SEE PROFILE

Eihab M. Abdel-Rahman
University of Waterloo
184 PUBLICATIONS   3,400 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

MEMS Bifurcation Gas Sensors View project

Metal-Organic Frameworks coated resonators for chemical detection View project

All content following this page was uploaded by Mohammad Ibrahim Younis on 06 January 2016.

The user has requested enhancement of the downloaded file.


Nonlinear Dyn (2007) 48:153–163
DOI 10.1007/s11071-006-9079-z

ORIGINAL ARTICLE

Dynamic pull-in phenomenon in MEMS resonators


Ali H. Nayfeh · Mohammad I. Younis ·
Eihab M. Abdel-Rahman

Received: 29 April 2005 / Accepted: 18 April 2006 / Published online: 29 September 2006

C Springer Science + Business Media B.V. 2006

Abstract We study the pull-in instability in microelec- 1 Introduction


tromechanical (MEMS) resonators and find that char-
acteristics of the pull-in phenomenon in the presence Advances in microfabrication technology have enabled
of AC loads differ from those under purely DC loads. the design and fabrication of RF microelectromechani-
We analyze this phenomenon, dubbed dynamic pull-in, cal (MEMS) devices, which promise breakthrough de-
and formulate safety criteria for the design of MEMS velopments in telecommunications, radar systems, and
resonant sensors and filters excited near one of their personal mobiles. In this paper, we study the dynamics
natural frequencies. We also utilize this phenomenon of two important MEMS RF components: resonators
to design a low-voltage MEMS RF switch actuated with and switches.
a combined DC and AC loading. The new switch uses Resonant microbeams (resonators) have been
a voltage much lower than the traditionally used DC widely used as transducers in mechanical microsensors
voltage. Either the frequency or the amplitude of the since the mid-1980s [1]. As interest grew dramatically
AC loading can be adjusted to reduce the driving volt- in MEMS devices for wireless communications appli-
age and switching time. The new actuation method has cations and the demand for high-frequency and quality
the potential of solving the problem of high driving factor resonators increased rapidly, MEMS resonators
voltages of RF MEMS switches. were proposed in the mid-1990s [2, 3] as a feasible
alternative to conventional large-size resonators.
Keywords MEMS . Microbeams . Resonators . There are a number of actuation methods for MEMS
Pull-in . Primary resonance . Electric actuation devices. Electrostatic actuation is the most well estab-
lished of these actuation method because of its sim-
plicity and high efficiency [3]. In a microbeam-based
A. H. Nayfeh ()
Department of Engineering Science and Mechanics, MC resonator, the beam is deflected by a DC bias and then
0219, Virginia Polytechnic Institute and State University, driven to vibrate around its natural frequency by an
Blacksburg, VA 24061, USA AC harmonic load [2]. A key issue in the design of
e-mail: anayfeh@vt.edu
such a device is to tune the electric load away from the
M. I. Younis pull-in instability [4, 5], which leads to collapse of the
Department of Mechanical Engineering, State University of microbeam and hence the failure of the device.
New York at Binghamton, Binghamton, NY 13902, USA Many studies have [4–6] addressed the pull-in phe-
nomenon and presented tools to predict its location to
E. M. Abdel-Rahman
Department of Systems Design Engineering, University of enable designers to avoid it. However, these studies
Waterloo, Waterloo, Ontario, Canada N2L 3G1 investigate the stability of the static deflection of the

Springer
154 Nonlinear Dyn (2007) 48:153–163

microbeam rather than the stability of motions around altering either its amplitude or its frequency to mini-
this deflected position. Hence, they do not account for mize the driving voltage and pull-in time.
motions due to the AC loading or the transients due to We use the reduced-order model in [16, 17] to simu-
the DC loading. This is particularly important in light late the dynamic behavior of MEMS resonators and
of the fact that the behavior of these devices is non- switches excited near their fundamental natural fre-
linear [4, 5, 7–9]. Therefore, there is a possibility for a quencies. We apply a shooting technique [10] and long-
dynamic instability to trigger pull-in [10] (i.e., dynamic time integration of the equations of motion to generate
pull-in) below the statically predicted instability limit. periodic motions. This approach is valid for small and
The dynamic pull-in phenomenon was reported large motions and can be applied to a wide range of
and analyzed for switches actuated by a step voltage loadings and initial conditions, and hence it can be used
[11, 12] and various ramping rates [11]. Both studies to study the global dynamics of switches. We validate
[11, 12] indicate that the dynamic pull-in voltage can the numerical scheme using the analytical solution in
be as low as 91% of the static pull-in voltage. In the [7]. Then we use the global approach to simulate the
presence of squeeze-film damping, the dynamic pull-in response of the low-voltage switch.
voltage is shown to approach the static pull-in voltage
[12]. Seeger and Boser [13] also showed numerically
the presence of dynamic pull-in in the behavior of a gy- 2 Problem formulation
roscope mass driven by a harmonic parallel-plate elec-
tric load. We consider a clamped-clamped microbeam, Fig. 1,
Another MEMS device that has gained a great subject to viscous damping with a coefficient ĉ per
deal of attention in wireless communications is the unit length and actuated by an electric load VDC +
MEMS switch. MEMS switches overcome the limi- VAC cos( ˆ tˆ), where VDC is the DC polarization volt-
tations of conventional electromechanical and solid- age, and VAC and  ˆ are the amplitude and frequency of
state switches; they have many attractive features like the AC voltage. The equation of motion that governs
low-power consumption and high isolation. The ma- the transverse deflection ŵ(x̂, tˆ) is written as [4, 5]
jor drawbacks of these devices are their requirement
of high driving voltages and low reliability [3, 14]. It ∂ 4 ŵ ∂ 2 ŵ ∂ ŵ
is highly desirable to bring the actuation voltage to a EI + ρ A + ĉ
∂ x̂ 4 ∂ tˆ2 ∂ tˆ
level compatible or close to those of IC circuits and to    2 
increase their reliability. EA ∂ ŵ ∂ 2 ŵ
= d x̂ + N̂
In [4, 5], we presented a model that predicts the static 2 0 ∂ x̂ ∂ x̂ 2
pull-in phenomenon. In [7], we utilized perturbation
b[VDC + VAC cos (ˆ tˆ)]2
methods [15] to generate analytical expressions for the + (1)
dynamic behavior of resonators undergoing small mo- 2 (d − ŵ) 2

tions around their equilibria. In [16, 17], we developed a


reduced-order model to simulate the static and dynamic where x̂ is the position along the microbeam, A and
behaviors of resonators and switches undergoing small I are the area and moment of inertia of the cross-
and large motions. We utilized the model to simulate section, E is Young’s modulus, N̂ the applied tensile
the transient behavior of a MEMS switch and calculate axial force, tˆ is time, ρ is the material density, h the
its pull-in time. In this work, we study the dynamics microbeam thickness, d the capacitor gap width, and
of MEMS resonators around their natural frequencies  the dielectric constant of the gap medium. The last
and track their motions up to the pull-in instability. We term in Equation (1) represents an approximation of the
analyze the dynamic pull-in phenomenon and present
criteria for the design of MEMS resonators. We also use
this phenomenon to design a novel RF MEMS switch
actuated by a voltage much less than the static pull-in
voltage. The switch is actuated using a combined dc
and AC loading, similar to the actuation method used
in MEMS resonators. The AC loading can be tuned by Fig. 1 A schematic of an electrically actuated microbeam

Springer
Nonlinear Dyn (2007) 48:153–163 155

parallel-plate electric force assuming complete overlap deflection as


of the area of the microbeam and the stationary elec-
trode. The boundary conditions are
M
w(x, t) = u i (t)φi (x) (8)
i=1
∂ ŵ(0, tˆ) ∂ ŵ(, tˆ)
ŵ(0, tˆ) = ŵ(, tˆ) = 0, = = 0 (2)
∂ x̂ ∂ x̂
where u i (t) is the ith generalized coordinate and φi (x)
For convenience, we introduce the nondimensional is the ith linear undamped mode
shape of the straight
1
variables microbeam, normalized such that 0 φi φ j d x = δi j and
governed by

ŵ x̂ EI
w= , x= , t = tˆ , φiiv = N φi + ωi2 φi (9)
d  ρbh4
 φi = 0 and φi =0 at x =0 and x =1
ˆ ρbh
4
= (3) (10)
EI

Substituting Equation (3) into Equations (1) and (2), Here, ωi is the ith natural frequency of the microbeam.
we obtain We multiply Equation (4) by φn (x)(1 − w)2 , substitute
Equation (8) into the resulting equation, use Equation
∂ 4w ∂ 2w ∂w ∂ 2w (9) to eliminate φiiv , integrate the outcome from x = 0
+ + c = [α 1 (w, w) + N ]
∂x4 ∂t 2 ∂t ∂x2 to 1, and obtain
[VDC + VAC cos (t)]2
+ α2 
(1 − w)2
M 1
ü n − 2 ü i u j φi φ j φn d x
(4) i, j=1 0

w(0, t) = w(1, t) = 0,
M  1

∂w ∂w + ü i u j u k φi φ j φk φn d x + cu̇ n + ωn2 u n
(0, t) = (1, t) = 0 (5) i, j,k=1 0
∂x ∂x
 1

where the functional  is given by = α2 [VDC + VAC cos(t)] 2


φn d x
0


 M 1
1
∂ f1 ∂ f2 +2 ωi2 u i u j φi φ j φn d x
 ( f 1 (x, t), f 2 (x, t)) = dx (6)
0 ∂x ∂x i, j=1 0


M  1
The parameters appearing in Equation (4) are − ωi2 u i u j u k φi φ j φk φn d x
i, j,k,=1 0
 2 
ĉ4 d
M 1
c= , α1 = 6 , +2 u̇ i u j cφi φ j φn d x
EIT h 0
i, j=1
N̂ 2 64 
N = , α2 = (7)
M 1
EI Eh 3 d 3 − u̇ i u j u k cφi φ j φk φn d x
i, j,k=1 0
Next, we generate a reduced-order model [16, 17] by

M  1
discretizing Equations (4) and (5) into a finite-degree- + α1 u i u j u k (φi , φ j ) φn φk d x
of-freedom system consisting of ordinary-differential i, j,k=1 0
equations in time. We use the undamped linear mode

M  1
shapes of the straight microbeam as basis functions in − 2α1 u i u j u k u l (φi , φ j ) φk φl φn d x
the Galerkin procedure. To this end, we express the i, j,k,l=1 0

Springer
156 Nonlinear Dyn (2007) 48:153–163


M
+ α1 u i u j u k u l u m (φi , φ j )
i, j,k,l,m=1
 1
× φk φl φm φn d x for n = 1, 2, . . . , M (11)
0

Equation (11) represents a discretized system of M cou-


pled nonlinear ordinary-differential equations describ-
ing the dynamic behavior of an electrically actuated
microbeam. A single-mode approximation yields the
following equation:
  
1 1
u¨1 + cu˙1 + ω12 u 1 1 − 2u 1 φ13 d x + u 21 φ14 d x
0 0
 1
= α2 [VDC + VAC cos(t)]2 φ1 d x
0
Fig. 2 Equilibria of an electrostatically actuated resonator cal-
+ α1 u 31 (φ1 , φ1 ) culated using three (blue), four (green), and five (red) modes;
  solid lines: stable, dashed lines: unstable
1 1
× φ1 φ1 d x − 2u 1 φ12 φ1 d x
0 0

  with the DC voltage. While all three models converge


1
+ u 21 φ13 φ1 d x (12) to the lower branch of solutions, they diverge as the
0
voltage approaches zero on the upper branch. The sta-
ble (lower) branch and the unstable (upper) branch of
Using three or more modes in Equation (11) was shown equilibrium points collide in a saddle-node bifurcation
[16, 17] to guarantee convergence for the stable equi- at the static pull-in instability VDC ≈ 4.8 V, resulting
libria. In the following analysis, we use three or more in the destruction of both branches. This static analysis
modes to generate the results. suggests that MEMS resonators should be designed to
Equation (11) constitute a non-explicit system operate below 4.8 V to ensure stability, while MEMS
of second-order nonlinear ordinary-differential equa- switches should be actuated by a DC voltage larger than
tions. To calculate solutions of these equations, one can 4.8 V to trigger pull-in.
use an implicit scheme, such as a predictor–corrector For the dynamic analysis, we use two DC voltages:
scheme, or transform them into an explicit system in 3.5 and 2 V. It follows from Fig. 2 that four modes are
ü n by multiplying Equation (11) with the inverse of the sufficient to produce both of the stable and unstable
mass matrix. In the present calculations, we used the equilibria and hence four modes are adequate to repre-
latter approach. sent the global dynamics at these DC voltages.
Next, we study the dynamic behavior of the mi-
crobeam under VDC = 2 V and an AC harmonic excita-
3 Results tion near its fundamental natural frequency ω1 = 23.9.
We assume a quality factor Q = 1000, which is related
We utilize three, four, and five modes in the reduced- to the damping coefficient by c = ω1 /Q. We use a com-
order model, Equation (11), to calculate the equi- bination of a two-point boundary-value problem and a
libria of a microbeam under DC loading when  = shooting method [10] to calculate periodic solutions of
510 μm, h = 1.5 μm, b = 100 μm, d = 1.18 μm, and Equation (11) and study their stability using Floquet
the nondimensional axial load N = 8.7. Details of the theory [10]. In Fig. 3, we show variation of wMax the
calculations are presented in [16, 17]. Figure 2 shows maximum displacement of the microbeam mid-point
variation of the microbeam mid-point deflection wMax with the frequency of excitation when VAC = 0.01 V.

Springer
Nonlinear Dyn (2007) 48:153–163 157

Fig. 3 Frequency–response curve below the dynamic pull-in


instability: VDC = 2 V and VAC = 0.01 V

Fig. 5 Frequency–response curve showing the onset of dynamic


pull-in: VDC = 2 V and VAC = 0.1 V

Fig. 4 Variation of the normalized nonlinear resonance fre-


quency with the amplitude of AC loading; solid curve: perturba-
tion approach, stars: global approach
Fig. 6 Long-time integration results for a set of initial conditions
The solid lines denote stable branches and the dashed close to the orbit of bifurcation point B of Fig. 5
line denotes an unstable branch. The nonlinear res-
onance frequency in Fig. 3 is approximately 24.337, and the rest of the specifications are the same as those
where one of the Floquet multipliers approaches unity. used in Fig. 2. Figure 4 shows that the model is in
There is no indication of pull-in in the figure. good agreement with the analytical solution within the
To validate the reduced-order model, we compare approximation limits (small but finite motions) of the
in Fig. 4 the nonlinear resonance frequency obtained analytical solution.
using the model (stars) to that obtained using an ap- Figure 5 shows variation of wMax with  for the
proximate analytical solution [7] (the solid line) found resonator of Fig. 3 when the AC voltage is increased
by directly attacking the distributed-parameter system, to VAC = 0.1 V. In this case, the frequency–response
Equations (4) and (5). The figure shows the variation curve does not close within the figure. The dotted-
of the nonlinear resonance frequency normalized with dashed line represents a branch of saddles (unstable
respect to the natural frequency ω1 = 26.9523 of a mi- equilibria corresponding to the upper branch in Fig. 2).
crobeam of  = 310 μm and VDC = 2 V with the ex- There are three cyclic-fold bifurcations within the fre-
citation amplitude VAC . The quality factor is Q = 151 quency range 23–25 Hz where a stable branch collides

Springer
158 Nonlinear Dyn (2007) 48:153–163

Fig. 7 Long-time integration results for two sets of initial conditions close to the orbit of bifurcation point A of Fig. 5

with an unstable branch of solutions. Each bifurcation


point is characterized by the slope of the curve ap-
proaching infinity and a Floquet multiplier approaching
unity. One bifurcation point B is on the left side at
 = 24.47 and two points A and C are on the right
side at  = 24.31 and  = 24.80. Sweeping the exci-
tation frequency beyond any of these points will result
in the following qualitative changes in the response:
r A down-sweep past point A will result in a jump up
(to the left stable branch or the upper stable branch)
or in pull-in.
r An up-sweep past point B will result in a jump up (to
the upper stable branch) or down (to the lower right
stable branch) or in pull-in.
r An up-sweep past point C will result in a jump down
(to the lower right stable branch) or in pull-in.
Any of these abrupt changes in the response are un- Fig. 8 Frequency–response curves for VDC = 2 V and VAC =
desirable in devices like resonant microsensors. The 0.1 V (red), 0.2 V (green), and 0.5 V (blue)
present global analysis can predict the location of the
regions where jumps and hysteresis exist.
As an example of these qualitative changes, we show ity of the disturbance, the number of stable attractors
in Fig. 6 the results of long-time integration of Equation available at the operating point, and the size of the basin
(11) for a set of initial conditions near the bifurcation of attraction of each of the attractors. Once the system
point B and a forcing frequency  = 24.49 that slightly is disturbed away from its operating point, it might
exceeds the bifurcation threshold. The time is nondi- settle back to the same point, settle down to another
mensionalized with respect to T = 7.1195 × 10−5 s. stable attractor, or go to pull-in, depending on interac-
The system response stays for a while around the pre- tions among these factors. To illustrate this point, we
vious stable solution (its “ghost” [10]) and then leaves show in Fig. 7a and b time-history evolutions gener-
to the pull-in instability. In interpreting the results in ated from long-time integration of Equation (11) for
Fig. 6, note that w = 1 corresponds to collapse of the two sets of initial conditions close to the orbit of bifur-
microbeam into the fixed electrode. cation point A of Fig. 5. The forcing frequency is set to
The response of the system to external disturbances  ≈ 24.29, u̇ 1 = 0.2 in Fig. 7a, u̇ 1 = 0.23 in Fig. 7b,
depends on the system transient dynamics, the sever- and the rest of the initial conditions are identical. We

Springer
Nonlinear Dyn (2007) 48:153–163 159

tion points to higher values of wMax . At higher values of


VAC , an interval opens up around ω1 where the only pos-
sible solution is the upper stable branch. The basin of
attraction of this branch is small and resonators excited
in this frequency range can go to pull-in due to small
external disturbances.
Figure 9 shows variation of wMax with VAC when
VDC = 2 V,  = 24.16, and Q = 1000. The upper
branch has been truncated at wMax = 0.8 because the
continuation of the unstable branch beyond this point
is of little interest and the accuracy of the predic-
tions beyond wMax = 0.9 is dubious, as indicated in
Fig. 2. Three cyclic-fold bifurcations appear in the
figure where the slope of the curve approaches infinity
Fig. 9 Force–response curve showing the onset of dynamic pull- and a Floquet multiplier approaches unity. Unlike the
in when Q = 1000
frequency-sweep case in Fig. 5, the highest bifurcation
point at VAC ≈ 0.282 V will always lead to pull-in. It
note from Fig. 7a that the system settles back to a is a sufficient condition for dynamic pull-in. The other
point on the lower right stable branch. On the other two bifurcation points at VAC = 0.0498 and 0.00738 V
hand, Fig. 7b shows that the system jumps to pull-in. can lead to phenomena similar to those discussed in
As mentioned previously, two more possibilities ex- Fig. 5. We illustrate two of these possibilities, namely a
ist in which the system settles down on either the sta- jump up to a larger orbit and a jump down to a smaller
ble branch of the left side or the upper stable branch. orbit in Fig. 10a and b, respectively. The initial con-
However, in the present case, the domains of attraction ditions for the time integration are taken from orbits
of these stable branches are small (because the orbits corresponding to the two bifurcation points, while the
they represent are quite large and approach the saddle), excitation amplitudes are VAC = 0.05 and 0.00732 V,
hence it is hard for the system to jump to these states respectively, slightly beyond the bifurcation thresholds.
physically. The ensuing motion experiences a transient behavior,
Increasing the amplitude of the AC voltage VAC after which the system settles onto a totally different
from 0.1 to 0.5 V, as shown in Fig. 8, moves the left orbit from that before the jump. A force-sweep over
and right parts of the frequency–response curve further an interval including these two bifurcation points will
away from each other and the location of the bifurca- lead to hysteresis.

Fig. 10 Long-time integration results for points near the two lower cyclic-fold bifurcations of Fig. 9

Springer
160 Nonlinear Dyn (2007) 48:153–163

Figure 11 shows variation of wMax with VAC for the


same parameters of Fig. 8 but with Q = 100. The upper
branch has been truncated at wMax = 0.8 because the
continuation of the unstable branch beyond this point
is of little interest and the accuracy of the predictions
beyond wMax = 0.9 is dubious, as indicated in Fig. 2.
Figure 11 shows that decreasing the quality factor by an
order of magnitude eliminates the unstable branch and
the two lower bifurcation points. On the other hand,
the highest bifurcation point undergoes a small shift
to VAC = 0.299. So, while the increase in Q elimi-
nates the undesirable nonlinear phenomena of jumps
and hysteresis, it has a minimal effect on the use of
dynamic pull-in for switching applications. In fact, the
switch packaging requirements can be relaxed signifi-
cantly, thus the quality factor, at the expense of a slight
increase in the actuation voltage used to trigger pull-in.
Fig. 11 Force–response curve for the parameters of Fig. 8 and
We have observed fractal behavior in the system re-
Q = 100 sponse. Figure 12a and b show a time-history evolution
and a phase portrait of the system response at a point
Figures 5 and 9 show that the dynamic pull-in in- on the stable branch in Fig. 9 when VAC = 0.09 V. As
stability occurs at a voltage much lower than the static indicated in Fig. 9, the motion is stable and periodic. In
pull-in voltage 4.8 V. The root mean squares of the volt- Fig. 13a and b, we show a time-history evolution and a
age loads leading to dynamic pull-in are 2.07 and 2.2 V, phase portrait of the system response when a single ini-
respectively. These results illustrate the importance of tial condition has been changed from u 1 (0) = −0.265
taking into account dynamic pull-in when designing to −0.264. There is a qualitative difference in the sys-
MEMS resonators to avoid device failure under dy- tem response to these two closely-spaced sets of initial
namic loads. They also suggest a novel dynamic actu- conditions. While one of them (in Fig. 12) produces
ation method for MEMS switches, in which the pull-in a stable periodic motion, the other (Fig. 13) triggers
voltage is much less than the static pull-in voltage. This a transient response that collides with the stable man-
actuation method is a promising approach to reduce the ifold of the saddle located at WMax ≈ 0.91 (compare
actuation voltage of RF MEMS switches.

Fig. 12 Long-time integration results for VAC = 0.09 V and u 1 (0) = −0.265

Springer
Nonlinear Dyn (2007) 48:153–163 161

Fig. 13 Long-time integration results for VAC = 0.09 V and u 1 (0) = −0.264

the location of the saddle in Fig. 2 with that in Fig. 13)


and goes to pull-in. Dynamic pull-in occurs here away
from any bifurcation point, demonstrating yet another
pull-in scenario. These results show response sensitiv-
ity to initial conditions, an indication of fractal behavior
[10].
The ‘effective’ nonlinearity of the resonator changes
its sign from positive, corresponding to a hardening
behavior, to negative, corresponding to a softening be-
havior near VDC = 3.25 V. This is a result of the fact
that the device is locally (in-well) hardening due to
the effect of the mechanical field and globally soften-
ing due to the effect of the electric field, see [7] for a
detailed analysis of this phenomenon. The frequency–
response curves in the softening domain are not merely
bent to the left rather than to the right, they also have
a different structure from those found in the hardening
domain. Figure 14 shows a frequency–response curve Fig. 14 A frequency–response curve in the softening domain:
for a resonator exhibiting a softening-type nonlinearity VDC = 3.5 V and VAC = 0.1 V
where VDC = 3.5 V and VAC = 0.1 V. The left part of
the curve is composed of a stable and an unstable branch
of solutions. The right part starts with a stable branch of Likewise, force–response curves in the softening do-
solutions that goes through a period-doubling bifurca- main have a different structure from those in the hard-
tion at PD to develop a branch of stable period-two pe- ening domain. Figure 15 shows a force–response curve
riodic solutions, which in turn loses stability through a of a resonator exhibiting a softening-type effective non-
symmetry-breaking bifurcation at SB. As the frequency linearity where VDC = 4 V and σ = −0.5. As in Fig. 9,
is decreased towards PD, we find that a Floquet multi- two stable branches of period-one periodic solutions
plier exited the unit circle along the real axis through meet an unstable branch at two cyclic-fold bifurca-
−1, indicating a period-doubling bifurcation. As the tions. On the other and, the upper cyclic-fold bifur-
frequency is decreased further towards SB, a Floquet cation of Fig. 9 does not exist here, rather the branch
multiplier exited the unit circle through +1, indicat- goes through a period-doubling bifurcation PD to gen-
ing a symmetry-breaking bifurcation. Similar behav- erate a stable period-two periodic solution followed by
iors were reported in the problem of capsizing of ships a symmetry-breaking bifurcation SB where the solu-
[10, 18]. tion loses stability.

Springer
162 Nonlinear Dyn (2007) 48:153–163

the quality factor and forcing level are high enough to


produce multi-stable responses (two or more coexist-
ing stable branches of solutions). When the response is
single-valued it fails to exist.
To explain the third mechanism, we summarize first
our conclusions on the dynamics of capacitively driven
microbeams. The application of a DC voltage VDC to
a microbeam creates the stable fixed point (center)
and the unstable fixed point (saddle) shown in Fig. 2.
In the phase plane representing the mid-point of the
beam and for the case of free vibration of the mi-
crobeam (no forcing or damping), a homoclinic or-
bit [10] starts and ends at the saddle and encircles
the center. This orbit is composed of a stable mani-
fold and an unstable manifold that intersect nontrans-
versely away from the saddle. Any initial conditions
(disturbances) inside the closed orbit result in a sta-
Fig. 15 A force–response curve in the softening domain: VDC =
ble oscillatory motion around the center and any initial
4 V and σ = −0.5 conditions outside it result in an unstable motion. In
the presence of damping, the center becomes a stable
focus, the saddle remains a saddle, and the homoclinic
4 Discussion and conclusions orbit is destroyed. Here, the stable and unstable man-
ifolds do not intersect. Exciting this system with a
We analyzed the dynamics of MEMS resonators, iden- small AC force leads to a stable periodic motion in
tified a new instability, dynamic pull-in, in capacitively- the stable regime near the focus. As the AC forcing
actuated MEMS, and showed the need to take into ac- and thus the amplitude of the motion increases, the sta-
count the static and dynamic instabilities in the design ble manifold approaches the unstable manifold. Even-
of MEMS. The dynamic pull-in phenomenon can be tually, both manifolds intersect transversally infinitely
used to advantage to solve a challenging dilemma in many times, resulting in a complex dynamic behav-
the design of RF MEMS switches, namely the high ior called homoclinic tangle [10]. One implication of
driving voltage requirement. We proposed a novel ap- this complex behavior is the sensitivity to initial con-
proach to the design of low-voltage MEMS switches. ditions or the unpredictability of motion. Increasing
Experimental work is planned to investigate the feasi- the amplitude of excitation further results in an ero-
bility of the new switch design. sion of the basin of attraction of bounded motions;
We found three distinct mechanisms leading to the and the possibility of finding a set of initial conditions
dynamic pull-in instability. The first mechanism is a that leads to a stable motion decreases. So, while a
bifurcation (a cyclic-fold in the hardening-domain or a stable periodic orbit exist, small transient deviations
symmetry-breaking in the softening-domain) that leads around it can lead to pull-in, as seen in Figs. 12 and
directly to pull-in. It is not sensitive to initial conditions 13.
or transient dynamics, as a result the stability limits it A minimum of three modes were employed in the
predicts represent an upper bound to stable motions. reduced-order model used to predict the transient and
In the other two mechanisms, the transient dynamics steady-state responses reported here. This corresponds
of the system determines whether it settles down, fol- to a six-dimensional phase space. As a result, generat-
lowing a disturbance, to a stable orbit or pulls-in. In ing the device stability maps is a nontrivial task. Fur-
one of these mechanisms, the change in the amplitude ther study of the dynamic pull-in phenomenon as well
or frequency of the AC signal drives the device be- as generation of the stability maps under various op-
yond a lower cyclic-fold bifurcation, leading to a jump erating conditions is the subject of forthcoming work.
up from a lower stable branch (Fig. 7b). This pull-in Meanwhile, we would like to point out that the dy-
mechanism is not always present. It exists only when namic pull-in instability is quite similar to the problem

Springer
Nonlinear Dyn (2007) 48:153–163 163

of capsizing of biased ships [18, 19] and the motion 10. Nayfeh, A.H., Balachandran, B.: Applied Nonlinear Dynam-
of the probe of an atomic force microscope within a ics. Wiley, New York (1995)
11. Ananthasuresh, G.K., Gupta, R.K., Senturia, S.D.: An ap-
van der Waals potential field [20, 21]. All of these phe-
proach to macromodeling of MEMS for nonlinear dynamic
nomena are manifestation of a characteristic of biased simulation. In: Proceedings of the ASME International Con-
dynamical systems, namely escape from an asymmet- ference of Mechanical Engineering Congress and Exposition
ric potential well. This process is discussed in further (MEMS). Atlanta, GA, pp. 401–407 (1996)
12. Krylov, S., Maimon, R.: Pull-in dynamics of an elastic beam
details in [10].
actuated by distributed electrostatic force. In: Proceedings
of the 19th Biennial Conference in Mechanical Vibration
and Noise (VIB), Chicago, IL, Paper DETC2003/VIB-48518
(2003)
References
13. Seeger, J.I., Boser, B.E.: Parallel-plate driven oscillations
and resonant pull-in. In: Proceedings of the Solid-State Sen-
1. Elwespoek, M., Wiegerink, R.: Mechanical Microsensors. sor, Actuator and Microsystems Workshop. Hilton Head Is-
Springer, Berlin (2001) land, SC pp. 313–316 (2002)
2. Clark, J.R., Bannon, F.D. III, Ark-Chew, W., Nguyen, C.T.- 14. Tilmans, H.A., Raedt, W.D., Beyne, E.: MEMS for wire-
C.: Parallel-resonator HF micromechanical bandpass filters. less communications: From RF-MEMS components to RF-
In: Proceedings of the Conference on Solid State Sensors MEMS-Sip. J. Micromech. Microeng. 13, 139–163 (2003)
and Actuators: TRANSDUCERS ’97, Chicago, IL, Vol. 2, 15. Nayfeh, A.H.: Introduction to Perturbation Techniques.
pp. 1161–1164 (1997) Wiley, New York (1981)
3. Varadan, V.M., Vinoy, K.J., Jose, K.A.: RF MEMS and Their 16. Younis, M.I., Abdel-Rahman, E.M., Nayfeh, A.H.: A
Applications. Wiley, New York (2003) reduced-order model for electrically actuated microbeam-
4. Younis, M.I., Abdel-Rahman, E.M., Nayfeh, A.H.: Static based MEMS. J. Microelectromech. Syst. 12, 672–680
and dynamic behavior of an electrically excited resonant (2003)
microbeam. In: Proceedings of the 43rd AIAA Structures, 17. Abdel-Rahman, E.M., Younis, M.I., Nayfeh, A.H.: A non-
Structural Dynamics, and Materials Conference, Denver, linear reduced-order model for electrostatic MEMS. In:
CO, AIAA Paper, 2002–1305 (2002) Proceedings of the 19th Biennial Conference in Me-
5. Abdel-Rahman, E.M., Younis, M.I., Nayfeh, A.H.: Charac- chanical Vibration and Noise (VIB), Chicago, IL, Paper
terization of the mechanical behavior of an electrically ac- DETC2003/VIB-48517 (2003)
tuated microbeam. J. Micromech. Microeng. 12, 795–766 18. Nayfeh, A.H., Sanchez, N.E.: Chaos and dynamic instabil-
(2002) ity in the rolling motion of ships. In: Proceedings of the
6. Tilmans, H.A., Legtenberg, R.: Electrostatically driven 17th Symposium on Naval Hydrodynamics. Hague, The
vacuum-encapsulated polysilicon resonators. Part II. Theory Netherlands pp. 617–631 (1988)
and performance. Sensors Actuators A 45, 67–84 (1994) 19. Soliman, M.S.: An analysis of ship stability based on tran-
7. Younis, M.I., Nayfeh, A.H.: A study of the nonlinear re- sient motions. In: Proceedings of the 4th International Con-
sponse of a resonant microbeam to an electric actuation. ference on the Stability of Ships and Ocean Vehicles, Naples,
Nonlinear Dyn. 31, 91–117 (2003) Italy (1990)
8. Abdel-Rahman, E.M., Nayfeh, A.H., Younis, M.I.: Dynam- 20. Ashhab, M., Salapaka, M.V., Dahleh, M., Mezić, I.: Dynam-
ics of an electrically actuated resonant microsensor. In: Pro- ical analysis and control of microcantilevers. Automatica 35,
ceedings of the International Conference of MEMS, NANO 1663–1670 (1999)
and Smart Systems, Banff, Canada, pp. 188–196, (2003) 21. Ashhab, M., Salapaka, M.V., Dahleh, M., Mezić, I.:
9. Gui, C., Legrenberg, R., Tilmans, H.A., Fluitman, J.H.J., Melnikov-based dynamical analysis of microcantilevers in
Elwenspoek, M.: Nonlinearity and hysteresis of resonant scanning probe microscopy. Nonlinear Dyn. 20, 197–220
strain gauges. J. Microelectromech. Syst. 7, 122–127 (1998) (1999)

Springer

View publication stats

You might also like