Você está na página 1de 34

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/226044957

Reduced-Order Models for MEMS Applications

Article in Nonlinear Dynamics · January 2005


DOI: 10.1007/s11071-005-2809-9

CITATIONS READS

163 306

3 authors:

Ali H. Nayfeh Mohammad Ibrahim Younis


Virginia Polytechnic Institute and State University King Abdullah University of Science and Techn…
799 PUBLICATIONS 25,975 CITATIONS 195 PUBLICATIONS 3,153 CITATIONS

SEE PROFILE SEE PROFILE

Eihab M. Abdel-Rahman
University of Waterloo
157 PUBLICATIONS 2,619 CITATIONS

SEE PROFILE

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

Metal-Organic Frameworks coated resonators for chemical detection View project

Modelind, Simulation and Testing of initially curved microplates. View project

All content following this page was uploaded by Eihab M. Abdel-Rahman on 07 September 2014.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Reduced-Order Models for MEMS Applications

Ali H. Nayfeha , Mohammad I. Younisb , Eihab M. Abdel-Rahmana

a
Department of Engineering Science and Mechanics, MC 0219, Virginia Polytechnic Institute and State
University, Blacksburg, VA 24061, USA, anayfeh@vt.edu
b
Mechanical Engineering Department, State University of New York at Binghamton, Binghamton, NY 13902,
USA

Abstract
We review the development of reduced-order models for MEMS devices. Based on their implementation pro-
cedures, we classify these reduced-order models into two broad categories: node and domain methods. Node
methods use lower-order approximations of the system matrices found by evaluating the system equations
at each node in the discretization mesh. Domain-based methods rely on modal analysis and the Galerkin
method to rewrite the system equations in terms of domain-wide modes (eigenfunctions). We summarize
the major contributions in the field and discuss the advantages and disadvantages of each implementation.
We then present reduced-order models for microbeams and rectangular and circular microplates. Finally, we
present reduced-order approaches to model squeeze-film and thermoelastic damping in MEMS and present
analytical expressions for the damping coefficients. We validate these models by comparing their results with
available theoretical and experimental results.
Keywords: Reduced-order models, MEMS, microbeams, microplates, squeeze-film damping, thermoelastic
damping.

1 State-of-the-Art
The dynamics of MEMS are represented by partial-differential equations (PDEs) and associated boundary
conditions. The most widely used method to treat these distributed-parameters problems is to reduce them
to ordinary-differential equations (ODEs) in time and then solve the reduced equations either numerically
or analytically. Three approaches are used in the reduction:

• Idealization of the device flexible structural elements as rigid bodies.

• Discretization using finite-element methods (FEM), boundary-element methods (BEM), or finite-


difference methods (FDM).

• Construction of reduced-order models (ROM).

1
The first and second approaches, while lying at opposite extremes of complexity, are currently the most
widely used. The pressure for better designs, less trial-and-error in the design process, and better device
performance demands better models than idealized rigid bodies. Numerous researchers compared the pull-in
voltage of electrostatically actuated cantilever [1] and clamped-clamped [2] microbeams obtained by solving
the distributed-parameter system to those obtained using a spring-mass model and found that the spring-
mass model underpredicts the pull-in voltage.
Although FEM/BEM and FDM simulations are adequate for the analysis of the static deflections (equi-
librium positions) of MEMS devices, they are inadequate for dynamic simulations because they require the
time integration of thousands of second-order ODEs (one for each degree of freedom in the model). This
is a very expensive process, making system-level simulation, device optimization, interactive design, and
evolutionary design almost impossible. As a result, reduced-order modeling of MEMS is gaining attention
as a way to balance the need for enough fidelity in the model against the numerical efficiency necessary to
make the model of practical use in MEMS design.
There are two main classes of methods used to create ROMs: node and domain methods. All of these
methods aim to create a transformation from the physical coordinates of the device to a set of q generalized
coordinates associated with the eigenfunctions corresponding to the q lowest eigenvalues. The differences
among these methods lie in the techniques used to obtain this transformation.

1.1 Node Methods


This class of methods eliminates the spatial dependence in the PDEs by evaluating them at the nodes of
a discretization of the device structures (and cavities). The variation over time of the displacement (and
coupled parameters) at these nodes constitutes the state vector {x(t)}n of the discretized system of ODEs;
that is,
{ẋ}n = {f (x)}n + [B]nm {u}m (1)

where {f (x)} is a nonlinear function of the state vector and {u(t)} is the input to the system. Linear algebra
techniques are then used to create a transformation matrix [T ] to reduce the size of the state vector from n
to q, according to
{x}n = [T ]nq {y}q

This transformation is then used to reduce system (1) to

{ẏ}q = {fˆ(y)}q + [B̂]qm {u}m (2)

To this end, the Guyan method [3] uses the following procedure:

• In the mesh, designate q states as master states {xq }, representing the motions of the dominant modes
in the system, and designate the rest of the states {xn−q } as slave states.

2
• Set the time derivatives of the slave states in equation (1) equal to zero.

• Linearizes {f (x)} around an equilibrium point {xo } in state space, set the input equal to zero, and
retain only the last n − q equations of (1) to get
( )
h i x
q
K(n−q)q K(n−q)(n−q) =0
xn−q

• Use this equation to write the slave states {xn−q } in terms of the master states {xq } and obtain the
transformation " #
Iqq
{x} = −1 {xq }
−K(n−q)(n−q) K(n−q)q

• Insert this transformation into the linearized form of system (1) to obtain the reduced system

{ẏ}q = [A]qq {y}q + [B̂]qm {u}m (3)

where {y} = {xq }.

The reduced-order model (3) can predict small motions around an equilibrium position in the neighborhood
of the point {xo } in state space. However, the model performs poorly in predicting transients, large motions,
and motions around equilibrium positions away from {xo }[4]. The “substructuring technique” in ANSYS
5.6 and 5.7 employs the Guyan method [5] to create reduced-order models of MEMS.
Krylov subspace methods use the Arnoldi or Lanczos procedures to create an orthonormal basis for the
subspace Kq {[ ∂f
∂x ], [B]} of system (1), which is then used to construct [T ]. Both procedures are equivalent
to a moment-matching procedure in the s-space between a Taylor series expansion around s = so (usually
taken equal to zero) of the Laplace transform of the original model, linearized at a point {xo } in state space,
and the Laplace transform of the ROM. The accuracy of the ROM depends on the order of the moment
matching (number of vectors q in the basis set). The resulting ROM can predict transient and steady-state
responses in the neighborhood of {xo } as long as their frequency content is in the neighborhood of so .
The truncated balanced realization (TBR) method uses the controllability and observability Grammians of
system (1), linearized around a point {xo }, to construct a transformation matrix [T ] out of the eigenfunctions
corresponding to the highest q Hankel singular values of the linearized model.
Bechtold et al. [6] used each of the Arnoldi procedure and the TBR to produce a ROM describing the
thermo-electric behavior of a micro-ignition unit. The original PDE was linear, as a result they were able
to apply directly both the TBR and a classical Krylov subspace method. They found that, while both
reduction methods reproduce the full model behavior accurately, the transient response of a Krylov-based
model was less accurate than that of a TBR-based model of the same order q. Also, the steady-state response
of a Krylov-based model diverged from that of the full-scale model at a lower excitation frequency than a
TBR-based model of the same order q. On the other hand, the computational cost of a Krylov-based model

3
is of order O(n2 ), while that of a TBR-based model is of order O(n3 ). As a result, it is not practical to
use the TBR for systems where n is more than a few hundred. Bechtold et al. [4] compared the use of the
Guyan and Krylov subspace methods to produce reduced-order models for the device of Bechtold et al. [6].
They found that the performance of the Guyan method is inferior to that of a Krylov subspace method of
the same order q in predicting the device transient response.
Wang and White [7] used the Arnoldi procedure to produce a ROM from a linearized full model of the
micro-switch of Hung et al. [8]. They found that the reduced model predictions are accurate for small
motions but deviate significantly from the original model results for large motions. Bai et al. [9] used
the Lanczos algorithm to implement a reduced-order modeling capability in the MEMS analysis software
SUGAR 2.0 and 3.0. The routine also starts with a linearized full model except for the electrostatic force,
which was represented using the closed-form parallel-plate formula in both the original and reduced models.
They used the routine to generate ROMs for a gap-closing actuator [10] and a torsional micro-mirror [11].
In both devices, the nonlinear mechanical restoring forces were minimal. The transient and steady-state
predictions of the ROMs were in good agreement with the full model predictions. They also found that the
ROMs, where q is on the order of O(10) [11], are able to produce accurate results for excitation frequencies
up to the order of a few tens of KHz. Srinivasan et al. [12] also used the Lanczos algorithm to generate
a ROM of a comb-drive microresonator based on a linear FEM model. They found the system responses
predicted by the ROM around the first natural frequency of the resonator in good agreement with those
produced using the FEM model, analytical formulae, and the software package NODAS.
Chen and White [13] extended the Krylov subspace methods to quadratic systems by applying a modified
Arnoldi procedure to a second-order Taylor series expansion of the original model, equation (1), around a
point in state space {xo }. They used this procedure to generate a ROM for a nonlinear model of a capacitor-
resistor circuit. On the other hand, Ramaswamy and White [14] and Chen and Kang [15, 16, 17] extended
the Krylov subspace methods to weakly nonlinear systems where the electrostatic force, rather than the
full system model, is expanded in a higher-order Taylor series. They used the Arnoldi procedure to produce
ROMs linear on the mechanical side and quadratic or cubic on the electric force side. Ramaswamy and White
[14] generated ROMs for a capacitively-driven cantilever beam and a rectangular micro-mirror, whereas Chen
and Kang [15, 16, 17] generated progressively refined ROMs for the micro-switch of Hung et al. [8]. They
found that (a) the quadratic ROM predictions match those of the original model over a larger range of
motion than a linear ROM of the same order q and (b) the cubic ROM is able to match the original model
predictions over an even larger range than a quadratic ROM of the same order q. They used the same
procedure [18] to extract a quadratic ROM of a micro-mirror. They found that the model is able to match
both of the transient and steady-state responses predicted by the original model away from pull-in (the
snap-down angle).
While Krylov subspace methods have been used to create linear, quadratic, and cubic ROMs, the memory
and computational costs of the process grow exponentially [18, 19, 20] with the order of the ROM. As a result,
it is limited in practice to cubic models beyond which it becomes too expensive. This is a significant drawback

4
since the electrostatic forcing represents a strong (high-order) nonlinearity. The traditional approach used
to treat it is to expand the forcing term in a Taylor series. Truncation of the higher-order terms introduces
significant errors for large device motions and leads to an overestimation of the pull-in voltage.
To overcome this limitation, Rewieński and White [19] proposed an alternate implementation of Krylov
subspace methods: a trajectory piecewise-linear (TPWL) approach. First a training signal is used in a full
(or approximate) model to generate the trajectory of the device response to the signal, then linear expansions
of the model are obtained around a series of equi-distant states {xi } along the trajectory. The matrix [T ] is
constituted from the union of the Krylov basis at the initial state {x0 } and the vector {x0 } orthonormalized
with respect to the basis set. The transformation matrix is then used to reduce the size of the TPWL
model from n to q. A ROM constructed using this method for the micro-switch of Hung et al. [8] was able
to predict responses to various input signals as long as the response trajectory was close to the training
trajectory. They found that the ROM predictions were more accurate than the linear model of Wang and
White [7] and the quadratic model of Chin and White [13] of the same order q. The model, however, failed to
predict the pull-in behavior when the training signal used to generate the ROM did not trigger pull-in. As a
result, the application of TPWL is limited to responses qualitatively similar to those lying on the trajectory
used to produce the model.
Rewieński and White [20, 21] extended the basis set used to generate [T ] by incorporating the Krylov
basis sets and the vectors {xi } at each of the linearization states {xi } along the training trajectory. They
also used the reduced-order model produced at point {xi−1 } as an approximation of the full model to step in
time to point {xi } where another linearized model is produced. At each linearization point, they performed
singular value decomposition (SVD) on [T ] to eliminate redundant vectors and vectors with singular values
less than a set accuracy limit. They found that the accuracy of a ROM based on the extended basis set was
better than a ROM produced using the simple set of Rewieński and White [19] even where the extended set
order q was smaller than the simple set order.
Vasilyev et al. [22] compared the implementation of the Krylov subspace method of Rewieński and White
[19] to the use of TBR and the use of a Krylov subspace method to perform an intermediate order reduction
followed by TBR, all applied at the initial state {x0 }, to create a TPWL reduced-order model. They found
that the ROMs produced using TBR and Krylov-TBR have better accuracy than the ROM based on a
pure Krylov reduction for the same model order q. The Krylov-TBR approach has the added advantage of
numerical efficiency, allowing it to be applied to systems larger than those where a pure TBR is typically
useful. On the other hand, they found that the TBR and Krylov-TBR based ROMs are only stable and
accurate where the model order q is even; models with odd q are unstable.

1.2 Domain Methods


Methods in this class eliminate the spatial dependence in the PDEs using the Galerkin method. The
displacement (and coupled parameters) are expressed as a linear combination of a complete set of linearly

5
independent basis functions φi (x, y, z) in the form

X
w(x, y, z, t) = ui (t)φi (x, y, z)
i=1

where ui (t) is the generalized coordinate associated with basis function φi (x, y, z). Truncating the summation
series to a finite number n,
n
X
w̃(x, y, z, t) = ui (t)φi (x, y, z) (4)
i=1

substituting Equation (4) into the PDEs, and requiring the residue to be orthogonal to every basis function,
we obtain n second-order ODEs in time in terms of the generalized coordinates ui (t).
The basis set can be chosen arbitrarily, as long as its element satisfy all of the boundary conditions
and are sufficiently differentiable. To enhance convergence, the basis set has to be chosen to resemble the
behavior of the device. Two ways have been used to generate the basis set:

• Conducting experiments or solving the PDEs using FEM or FDM to generate snapshots describing
the variation of the states over time (motion of the device) under a training signal, then applying
a modal analysis method, one of the variations of the proper orthogonal decomposition method [23]
(singular value decomposition, SVD; Karhunen-Loève decomposition, KLD; and principal component
analysis, PCA) to the time series to extract the mode shapes of the device structural elements (and
corresponding eigenfunction of the coupled domains).

• Solving the linear undamped eigenvalue problem (EVP) of the device to obtain the mode shapes of
the structural elements (and corresponding eigenfunctions of the coupled domains).

1.2.1 Basis Set from Time Series

Hung et al. [8] and Hung and Senturia [24] simulated the dynamics of a capacitive micro-switch made of a
clamped-clamped microbeam and represented by two coupled PDEs accounting for the microbeam motions
and the pressure of the air trapped underneath it. They generated a basis set for the beam deflection and
another for the air pressure by applying SVD to a time series produced from a few runs of a fully meshed
finite-difference solution of the PDEs. They indicated that both basis sets are similar to the linear mode
shapes and eigenfunctions of the undeflected microbeam. Chen and Kang [15] used KLD to generate basis
sets for the same model. The results of the ROMs generated using both approaches converged to those of the
original model as the order q of the ROM was increased. There was no clear difference in the convergence
speed between SVD and KLD.
Chen and Kang [18] used KLD to generate a basis set for the air pressure under a rigid micro-mirror
and extract a ROM for the device. They found that the model was able to match both of the transient
and steady-state responses predicted by the full model even for large motions close to the snap-down angle.

6
They also used KLD [25], modified to use snapshots taken over constant distances along the Jacobian of
the system rather than time, to generate two basis sets to represent the diaphragm deflection and the air
pressure of a capacitive pressure sensor. They found that, for the same ROM order q, the standard KLD
smoothed out fast variations in the original model response, while the modified KLD was able to reproduce
those variations.
Liang et al. [23, 26] simulated the dynamics of a clamped-clamped microbeam using a generalized
Hebbian algorithm to perform PCA on a noise-free [26] and a noisy [23] time series from a finite-difference
solution of the PDEs describing the microbeam motion and its interaction with the air underneath it. The
snapshots were obtained from the device response to two step voltages larger than the pull-in voltage and
a basis set was generated for each of the beam deflection and air pressure. They found that the basis sets
obtained using PCA and KLD were similar for noise-free data. When noise was injected into the data, KLD
produced distorted basis sets, while a robust PCA algorithm smoothed out the noise from the basis sets.
Lin et al. [27] produced the first model of a complex device made of more than one primitive structural
element (beam, plate, or desk). They used finite differences to generate a time series for a micro-mirror
made of a plate suspended from two beams over an air gap and actuated by a step voltage beyond the pull-in
voltage. Applying KLD to the time series, they found the local basis sets for each of the structural elements
and the trapped air, then they used component mode synthesis to create a ROM for the overall structure.
They compared predictions of this ROM and those of another ROM generated using a global basis set for
all three structures to the finite-difference results and found that the ROM generated using local basis sets
had an accuracy superior to that of a ROM generated using a global basis set.
Qiao and Aluru [28] modeled electroosmotic transport in straight micro-channels using the Poisson-
Boltzman and Naiver-Stokes PDEs. They used the finite-cloud method to solve the PDEs and obtained time
series representing the velocity profile of the flow for a given training signal, then they used SVD to extract a
basis set for the flow velocity. The resulting ROM was valid only for actuating voltages in the neighborhood
of the training signal. To obtain a model valid over a significant range of voltages, they used a time series
composed of snapshots taken over the whole voltage range. The new model was valid over the whole range
and had an extra mode not present in the original basis set.
De and Aluru [29] modeled a capacitive micro-switch made of a clamped-clamped microbeam using a
PDE describing the plate deflections and an integral equation describing the electric potential across the
capacitor. They used the finite-cloud and boundary-cloud methods to solve the equations and obtain time
series for the deflection and electric charge distribution for a given training signal, then they used SVD
to extract a basis set for each of them. They employed the basis sets in conjunction with the collocation
method, rather than the Galerkin method, to produce a ROM for the device. The model was valid only for
small motions.
We conclude that ROMs generated using this approach can only predict motions qualitatively similar
to those of the original time series. For example, a ROM cannot predict the pull-in dynamics using a time
series representing the response to actuation voltages below the pull-in voltage. Also, the accuracy of the

7
model degrades as the magnitude of the applied voltage deviates from that used as a training signal(s). In
other words, these models represent local approximations of the original PDEs.

1.2.2 Basis Set from EVP

Anathasuresh et al. [30] and Grtétillat et al. [31] used the linear undamped mode shapes of a straight
beam as a basis set to produce a ROM describing the dynamics of a micro-switch made of an asymmetric
clamped-clamped microbeam. They found a discrepancy between the results obtained using the model
and experimental data. The model underestimated the pull-in time (switching time) because it neglected
mid-plane stretching and residual stresses in the beam.
Gabbay et al. [32] developed an automated procedure to generate a ROM from FEM/BEM simulations
of a device response to a few training signals spanning the operation domain. The procedure is limited to
conservative systems and small motions around a statically deflected position of the structural element. They
used the linear mode shapes of the deflected structural element as a basis set to represent the inertia and
linear stiffness terms in the ODEs. The simulation results were used to express the electrostatic energy as a
fitted rational polynomial in terms of the generalized coordinates. The electrostatic force was then introduced
into the ODES as the derivative of the electrostatic energy with respect to the generalized coordinates. They
reported that this approach overestimates the structural stiffness and fails to correctly predict the dynamics
of a clamped-clamped microbeam at a DC voltage beyond 30% of the pull-in voltage. Varghese et al. [33]
used this approach to produce a ROM of the beam response to a Lorentz force generated by a magnetic
field.
Mehner et al. [34] modified this procedure to address problems involving mid-plane stretching and large
displacements. They modified the constraints on the nodes of the FEM code to stretch and contract as
they bend to avoid stiffness overestimation and used the modified code to extract the mode shapes of the
deflected structural element, which were adopted as a basis set for the system. Simulations of the device
under various training signals were used to write the strain energy and the electrostatic energy as fitted
rational polynomials in terms of the generalized coordinates associated with the mode shapes. They found
that the modified mode shapes were close to, but different from, the linear mode shapes of Gabbay et al.
[32].
Bennini et al. [35] modified the procedure of Gabbay et al. [32] to express both of the strain energy and
electrostatic energy as fitted regular polynomials in terms of the generalized coordinates and to allow for
constant modal damping. Mehner et al. [36] extended this procedure to account for the squeeze-film effects
using a few FEM runs to create polynomials, in terms of the generalized coordinates, representing linear
stiffness and damping coefficients corresponding to each basis function. The ROM140 of ANSYS 7.0 element
uses this process to represent the effect of squeeze-film damping, while the ROM144 element of ANSYS 7.0
uses the procedure of Bennini et al. [35] to produce ROMs for MEMS devices.
Westby and Fjeldly [37] and Xie et al. [38] used the linear mode shapes of the structural element as a

8
l
b x
Microbeam
z
v(t) h
y Stationary Electrode

Figure 1: A schematic of an electrically actuated microbeam.

basis set to express the motion of undamped MEMS exhibiting quadratic and cubic nonlinearities in terms
of a few ODEs. They used the center manifold method to reduce the number of these ODEs further to those
describing the temporal variation of one [38] or two [37] dominant nonlinear normal modes. The ODEs were
then solved analytically for the device response.
In the following sections, we present reduced-order models for electrically actuated microbeams and
rectangular and circular microplates using the linear undamped mode shapes of the unactuated structure
as a basis set in the Galerkin procedure. We present results showing the efficiency and accuracy of these
models.

2 Microbeams
We consider a clamped-clamped microbeam, Figure 1, subject to viscous damping with a coefficient ĉ per
unit length and actuated by an electric load v(t̂) = VDC + VAC cos(Ωt̂), where VDC is the DC polarization
voltage and VAC and Ω are the amplitude and frequency of the AC voltage. The nondimensional equation
of motion and boundary conditions that govern the transverse deflection of the microbeam are written as
[39, 40]

∂4w ∂2w ∂w ∂2w α2 v(t)2


+ + c = [α 1 Γ(w, w) + N ] + 2 (5)
∂x4 ∂t2 ∂t ∂x2 (1 − w)
∂w ∂w
w(0, t) = w(1, t) = 0, (0, t) = (1, t) = 0 (6)
∂x ∂x
where x, t, and w, are the nondimensional position, time, and transverse deflection; respectively. They are
related to the dimensional variables (denoted by hats) by

ŵ x̂ t̂
w= , x= , t= (7)
d ℓ T
q
4
where d is the capacitor gap width, ℓ is the length of the beam, and T = ρAℓEI . The parameters appearing
in Equation (5) are
 2
d 6ǫℓ4 ĉℓ4 N̂ ℓ2
α1 = 6 , α2 = , c = , N = (8)
h Eh3 d3 EIT EI

9
where A and I are the area and moment of inertia of the cross section, ρ is the material density, E is Young’s
modulus, h is the microbeam thickness, ǫ is the dielectric constant of the gap medium, and N̂ is an applied
tensile axial force. The functional Γ is given by
Z 1
∂f1 ∂f2
Γ (f1 (x, t), f2 (x, t)) = dx (9)
0 ∂x ∂x

We generate a ROM [41, 42] by discretizing Equations (5) and (6) into a finite-degree-of-freedom system
consisting of ordinary-differential equations in time. We use the linear undamped mode shapes of the straight
microbeam (VDC = 0) as basis functions in the Galerkin procedure. To this end, we express the deflection
as
M
X
w(x, t) = ui (t)φi (x) (10)
i=1

We multiply Equation (5) by (1 − w)2 , substitute Equations (10) into the resulting equation, use the
linear undamped mode shape equation [41, 42] to eliminate φiv
i , multiply by φn (x), integrate the outcome
from x = 0 to 1, and obtain
M
X M
X
ün − 2 Λijn uj üi + Λijkn uj uk üi − cu̇n − ωn2 un = α2 Λn v(t)2
i,j=1 i,j,k=1
M
X M
X M
X M
X
+2 ωi2 Λijn ui uj − ωi2 Λijkn ui uj uk + 2c Λijn uj u̇i − c Λijkn uj uk u̇i (11)
i,j=1 i,j,k,=1 i,j=1 i,j,k=1
M
X Z 1 M
X Z 1
+ α1 ui uj uk Γ(φi , φj ) φn φ′′k dx − 2α1 ui uj uk ul Γ(φi , φj ) φk φ′′l φn dx
i,j,k=1 0 i,j,k,l=1 0

M
X Z 1
+ α1 ui uj uk ul um Γ(φi , φj ) φl φm φn φ′′k dx for n = 1, 2, . . . , M
i,j,k,l,m=1 0

where the prime denotes differentiation with respect to space x, the overdot denotes differentiation with
respect to the time t, ωi is the ith natural frequency of the microbeam, and the functional Λ is defined by
Z 1 Z 1 Z 1
Λn = φn dx , Λin = φi φn dx , Λijn = φi φj φn dx , . . .
0 0 0

Using three or more modes in Equation (11) was shown [41, 42] to give good convergence for the stable
equilibria.
In Figure 2, we compare the microbeam mid-point deflection WMax calculated using the ROM and
employing the first three and five symmetric modes with results obtained by solving the static boundary-
value problem using a shooting method [39, 40]. The five-mode solution is in excellent agreement with the
results of the shooting method for both of the upper and lower branches.
We use the ROM to calculate the natural frequencies of a resonant microsensor. For a given voltage v,
we substitute the static solution corresponding to the lower branch into the Jacobian matrix of Equations

10
1

0.8

0.6
WMax

0.4

0.2

0 2 4 6 8 10
v (V)

Figure 2: Variation of WMax calculated using the ROM for two cases: three symmetric modes (dashed
line) and five symmetric modes (solid line). The discrete points are results obtained by solving the static
boundary-value problem using a shooting method [39, 40].

(11) and find the corresponding eigenvalues. Then by taking the square root of the magnitudes of the indi-
vidual eigenvalues, we obtain the natural frequencies of the device. In Figure 3, we compare the normalized
fundamental natural frequency calculated using the ROM and employing five symmetric modes in the dis-
cretization (solid line) with results obtained by solving the eigenvalue problem of the distributed-parameter
system (triangles) using a shooting method [39, 40] and the experimental results (circles) obtained by Tilmans
and Legtenberg [43] for a resonator with the specifications l = 210µm, h = 1.5µm, b = 100µm, d = 1.18µm,
E = 166GP a, and N̂ = 0.0009 Newton. There is an excellent agreement among the results. The ROM shows
robustness in predicting the natural frequency over the whole range even as the microbeam approaches its
stability limit where the frequency approaches zero.
To demonstrate the ROM ability to predict the dynamic behavior of microbeam-based MEMS, we cal-
culate the pull-in time of a pressure sensor. We plug the φi and ωi corresponding to the first M symmetric
modes into Equations (11) and integrate them in time for the ui (t). To obtain the deflection variation with
time, we use Equation (10) with the calculated φi and ui (t). We find the pull-in time by monitoring the
beam response over time for a sudden rise in the displacement, at that point we report the time as the

11
1

0.8

0.6

w 1/w 0

0.4

0.2

0 2 4 6 8
v (V)

Figure 3: A comparison of the normalized fundamental natural frequency calculated using the ROM and
employing five symmetric modes in the discretization (solid line) with results obtained by Abdel-Rahman
et al. [39] and Younis et al. [40] (triangles) and the experimental results (circles) obtained by Tilmans and
Legtenberg [43].

pull-in time. Figure 4 shows the evolution of u1 , the dominant coefficient, with the nondimensional time
obtained by integrating Equations (11) using the first five symmetric modes. The nondimensional pull-in
time is approximately t = 3.4, where a sudden rise in u1 occurs.

12
1

0.8

0.6
u1(t)

0.4

0.2

0 1 2 3 4
t

Figure 4: Evolution of u1 with the nondimensional time demonstrating the onset of pull-in.

3 Rectangular Microplates
We model a capacitively actuated microplate by using the dynamic analog of the von Kármán equations to
account for moderately large deflections [44, 45]; that is,
 2
∂ 2 v̂ ∂ ŵ ∂ ŵ2 ∂ ŵ ∂ 2 ŵ
  
1 ∂ û 1
(1 − ν) + + (1 − ν) +
2 ∂ ŷ 2 ∂ x̂∂ ŷ 2 ∂ x̂ ∂ ŷ 2 ∂ ŷ ∂ x̂∂ ŷ
2 2 2
∂ û ∂v̂ ∂ ŵ ∂ ŵ ∂ ŵ ∂ ŵ
+ 2 +ν + +ν =0 (12)
∂ x̂ ∂ x̂∂ ŷ ∂ x̂ ∂ x̂2 ∂ ŷ ∂ x̂∂ ŷ
 2
∂ 2 v̂ ∂ ŵ ∂ 2 ŵ ∂ ŵ ∂ 2 ŵ
  
1 ∂ û 1
(1 − ν) + 2 + (1 − ν) +
2 ∂ x̂∂ ŷ ∂ x̂ 2 ∂ x̂ ∂ x̂∂ ŷ ∂ ŷ ∂ x̂2
∂ 2 v̂ ∂ 2 û ∂ ŵ ∂ 2 ŵ ∂ ŵ ∂ 2 ŵ
+ 2 +ν + +ν =0 (13)
∂ ŷ ∂ x̂∂ ŷ ∂ ŷ ∂ x̂∂ ŷ ∂ x̂ ∂ x̂∂ ŷ
1 2 ∂ 4 ŵ ∂ 4 ŵ ∂ 4 ŵ ρ(1 − ν 2 ) ∂ 2 ŵ ǫ(1 − ν 2 )v(t̂)2
 
h 4
+2 2 2 + 4
+ −
12 ∂ x̂ ∂ x̂ ∂ ŷ ∂ ŷ E ∂ t̂2 2Eh(d − ŵ)2
N̂xx ∂ 2 ŵ N̂xy ∂ 2 ŵ N̂yy ∂ 2 ŵ ∂ û ∂ 2 ŵ ∂v̂ ∂ 2 ŵ ∂v̂ ∂ 2 ŵ
= + 2 + + + + ν
Eh ∂ x̂2 Eh ∂ x̂∂ ŷ Eh ∂ ŷ 2 ∂ x̂ ∂ x̂2 ∂ ŷ ∂ ŷ 2 ∂ ŷ ∂ x̂2
2 2 2
∂ û ∂ 2 ŵ 1 ∂ ŵ ∂ 2 ŵ 1 ∂ ŵ ∂ 2 ŵ 1 ∂ 2 ŵ
     
∂ ŵ
+ν 2
+ 2
+ 2
+ ν (14)
∂ x̂ ∂ ŷ 2 ∂ x̂ ∂ x̂ 2 ∂ ŷ ∂ ŷ 2 ∂ ŷ ∂ x̂2
 2 2   2
1 ∂ ŵ ∂ ŵ ∂ û ∂v̂ ∂ ŵ ∂ ŵ ∂ ŵ
+ ν + (1 − ν) 13 + +
2 ∂ x̂ ∂ ŷ 2 ∂ ŷ ∂ x̂ ∂ x̂ ∂ ŷ ∂ x̂∂ ŷ
where û(x̂, ŷ, t̂), v̂(x̂, ŷ, t̂), and ŵ(x̂, ŷ, t̂) are the displacements in the x̂, ŷ, and ẑ-directions, N̂ij is the applied
force on the i-edge in the j-direction, and ν is Poisson’s ratio. For convenience, we introduce the following
nondimensional variables [46]:
aû av̂ ŵ N̂ij
u= , v= , w= , Nij = ,
2d2 2d2 d Eh
2x̂ 2ŷ 2ht̂
x= − 1, y= − 1, t= p
a b 3(1 − ν 2 )ρa4 /E
This choice of x and y shifts the center of the plate to the point (x = 0, y = 0). The nondimensional
counterparts of Equations (12)–(14) are
 2
∂2v ∂w ∂ 2 w ∂w ∂ 2 w
  
1 ∂ u 1
(1 − ν) +α + (1 − ν) +
2 ∂y 2 ∂x∂y 2 ∂x ∂y 2 ∂y ∂x∂y
2 2 2
∂ u ∂ v ∂w ∂ w ∂w ∂ 2 w
+ α2 2 + να + α2 + ν =0 (15)
∂x ∂x∂y ∂x ∂x2 ∂y ∂x∂y
∂2u 2
∂w ∂ 2 w ∂w ∂ 2 w
   
1 2∂ v 1
(1 − ν) α +α + (1 − ν) α +α
2 ∂x∂y ∂x2 2 ∂x ∂x∂y ∂y ∂x2
2 2 2 2
∂ v ∂ u 1 ∂w ∂ w ∂w ∂ w
+ 2 + να + + να =0 (16)
∂y ∂x∂y α ∂y ∂y 2 ∂x ∂x∂y
∂4w 2 ∂4w 1 ∂4w ∂2w v(t)2
4
+ 2 2 2+ 4 4
+ 2 − α2
∂x α ∂x ∂y α ∂y ∂t (1 − w)2
∂2w ∂2w ∂2w
 
2 2 1
= 3α0 Nxx 2 + Nxy + Nyy 2
∂x α ∂x∂y α2 ∂y
 2 2
1 ∂v ∂ 2 w
  
2 ∂u ν ∂v ∂ w α1 ∂u
+ 12α1 + + 12 2 ν + (17)
∂x α ∂y ∂x2 α ∂x α ∂y ∂y 2
 2
α21 ∂w ∂w ∂ 2 w

1 ∂u 1 ∂v ∂ w
+ 12α21 (1 − ν) + + 12 (1 − ν)
α2 ∂y α ∂x ∂x∂y α2 ∂x ∂y ∂x∂y
2 2 ! 2 2 2 ! 2
α21
   
2 ∂w ν ∂w ∂ w 1 ∂w ∂w ∂ w
+ 6α1 + 2 +6 2 +ν
∂x α ∂y ∂x2 α α2 ∂y ∂x ∂y 2

The parameters appearing in Equations (15)–(17) are


b a d 3 1 − ν2 4
α= , α0 = , α1 = , α2 = ǫa (18)
a h h 8 Eh3 d3
Zhao et al. [46] solved the linear undamped eigenvalue problem using the hierarchical finite-element method
(HFEM) to obtain the microplate eigenfunctions φi (x, y) and write the transverse displacement field as
N
X
w(x, y, t) = qi (t)φi (x, y) (19)
i=1

Substituting Equation (19) into Equations (15) and (16) and considering the associated in-plane boundary
conditions yields a set of boundary-value problems for u and v. Using the HFEM, Zhao et al. [46] solved for

14
u and v in terms of qi (t); that is,

u = u(x, y, qi (t)) and v = v(x, y, qi (t)) (20)


2
Multiplying both sides of Equation (17) by (1 − w) , substituting Equations (19) and (20) into the outcome,
and applying the Galerkin procedure, then obtained a set of nonlinearly coupled ODEs, which is the ROM
for the microplate.
Francais and Dufour [47] measured the center deflection of a fully clamped square microplate under
various electrostatic actuations. In Figure 5, we compare the deflection wmax at the center of the plate
calculated using the ROM with the experimental results of Francais and Dufour. The ROM shows good
agreement and robustness, being able to predict deflections up to pull-in. The dots correspond to unstable
equilibrium solutions and the solid line corresponds to stable equilibrium solutions calculated using the ROM.

1 +

0.8

0.6
w max

0.4
+ +
0.2 + +
+
+
+
2 4 6 8 10 12
α2Vp2

Figure 5: Comparison of wmax calculated using the ROM (solid and dotted curves) with the experimental
results (+) of Francais and Dufour [47].

When the microplate is deflected, the linear mode shapes and natural frequencies change correspondingly.
Figure 6 shows variation of the fundamental natural frequency ω1 of the deflected plate, normalized with
respect to the natural frequency ω10 of the flat plate, with the electrostatic load α2 Vp2 . For low values of
α1 = d/h, the fundamental natural frequency decreases as the electrostatic force increases and approaches
zero as pull-in develops. As α1 increases, the fundamental natural frequency increases for the same level
of electrostatic forcing. At high values of α1 , the fundamental natural frequency first increases with the
electrostatic force, then decreases, and eventually approaches zero.

15
1

0.8

0.6
ω1
0.4 α1 =0.5 α1 =1.5
0.2

0 α1 =1.0 α1 =2.0
50 100 150 200 250 300
α2Vp2

Figure 6: Variation of the normalized fundamental natural frequency ω1 with the electrostatic load α2 Vp2
for various values of α1 .

4 Circular Microplates
We consider a circular plate with radius R fully clamped above a parallel electrode. The plate is sub-
ject to viscous damping with a coefficient ĉ per unit length. The nondimensional equations governing the
axisymmetric transverse deflection of the plate ŵ can be written as [48]
" #
∂2w v(t)2
   
4 1 ∂ ∂w ∂Φ σ ∂ ∂w ∂w
+ ∇ w = β + r − 2c + F (r, t) + 2 (21)
∂t2 r ∂r ∂r ∂r r ∂r ∂r ∂t (wmax − w)
1 ∂ 2 w ∂w
∇4 Φ = − (22)
r ∂r2 ∂r
where ∇4 is the polar biharmonic operator, σ is the residual stress, F (r, t) is an additional axisymmetric
pressure, and Φ(r, t) is the stress function. The nondimensional variables and parameters appearing in
Equations (21) and (22) are related to the dimensional quantities (denoted by hats) according to the following
relations:
1/2
h2 24(1 − ν 2 )

2 ρh 1/2
r̂ = Rr , t̂ = R t,
w , ĉ =ŵ = ρh5 D c,
D R R4
12(1 − ν 2 )Dh4 2 2Dh6 2 Eh4 Eh5
F̂ = F , v̂ (t) = v (t) , σ̂ = σ , Φ̂ = Φ, (23)
R7 ǫR7 R4 R2
dR 12(1 − ν 2 )h2
wmax = 2 , and β =
h R2
Eh3
where D = 12(1−ν 2 ) is the plate flexural rigidity. The boundary conditions are

∂w(1, t)
w(1, t) = 0 , =0, and w(0, t) is bounded (24)
∂r

16
To generate the ROM, Vogl and Nayfeh [48] let
N
X
w(r, t) = ηm (t)φm (r) (25)
m=1
N
X
Φ(r, t) = ηm (t)ηn (t)ψmn (r) (26)
m,n=1

where φm (r) is the mth axisymmetric linear undamped mode shape of the flat plate and the ψmn (r) are
unknown axisymmetric functions to be determined in the course of the analysis. Substituting Equations (25)
and (26) into Equations (21), (22), and (24) and following the Galerkin procedure, they obtained
N
X
2
η̈q + 2cη̇q + ωq2 ηq − 2wmax Λijq ηi η̈j + 2cη̇j + ωj2 ηj
 
wmax
i,j=1
N N
"
X X Z 1
ωk2 ηk 22

+ Λijkq ηi ηj η̈k + 2cη̇k + = Λq v(t) + β − ωmax ηi ηj ηk φ′i φ′q ψjk

dr
i,j,k=1 i,j,k=1 0

N
X Z 1 N
X Z 1
+ 2ωmax ηi ηj ηk ηl (φi φq )′ φ′j ψkl

dr − ηi ηj ηk ηl ηm (φi φj φq )′ φ′k ψlm

dr (27)
i,j,k,l=1 0 i,j,k,l,m=1 0

N
X Z 1 N
X Z 1 N
X Z 1
2 ′ ′
− σωmax ηi rφ′i φ′q dr + 2σωmax ηi ηj rφ′j (φi φq ) dr − σ ηi ηj ηk rφ′k (φi φj φq ) dr
i=1 0 i,j=1 0 i,j,k=1 0

N N
#
Z 1 X Z 1 X Z 1
2
+ ωmax F rφq dr − 2ωmax ηi F rφi φq dr + ηi ηj F rφi φj φq dr q = 1, 2, . . . , N
0 i=1 0 i,j=1 0
r φ′i φ′j r 1 φ′ φ′ 1
r 1 r r1+ν
Z Z Z Z
′ i j
ψij (r) = − dξ + ξφ′i φ′j dξ + dξ + ξφ′i φ′j dξ i, j = 1, 2, . . . , N (28)
4 0 ξ 4r 0 4 0 ξ 41−ν 0

where the functional Λ is defined by


Z 1 Z 1 Z 1
Λq = rφq dx , Λiq = rφi φq dx , Λijq = rφi φj φq dx , . . .
0 0 0

Vogl and Nayfeh [48] validated the ROM with experimental data. Osterberg [49] measured the pull-
in voltage vpi for multiple radii R of clamped circular microplates made of silicon with the specifications
h ≈ 3µm and d ≈ 1µm. Osterberg developed a statistics-based model to approximate vpi and solved for the
optimal statistical coefficients by fitting his model to the experimental data. Vogl and Nayfeh [48] fit the
physics-based model, Equations (27) and (28), to the experimental data by solving for the values of E, σ, ν, d,
and h that minimize the objective function
14  model 2
X v (E, σ, ν, d, h) − v exp
i i
W = (29)
i=1
δi

17
Figure 7: pull-in voltage versus plate radius.

where the δi , vimodel , and viexp are, respectively, the experimental standard deviations, the model pull-in
values, and the experimental pull-in values for the 14 different experimental radii. The objective function W
is a weighted sum of the square of the deviations between the ROM and experimental values. They found
out a local minimum of W for d = 1.014 µm, h = 3.01 µm, E = 150.6 GPa, ν = 0.0436, and σ̂ = 7.82 MPa,
which seems to be the global minimum. The pull-in voltages from this optimum model are displayed in
Figure 7 along with the experimental data. Standard deviation bars for the experimental data are also
shown in the figure.

5 Reduced-Order Approaches to Model Damping in MEMS


In this section, we show how to reduce the computational cost in the simulation of MEMS devices in noncon-
servative systems. The reduction can be achieved by solving analytically the equation governing the energy
dissipation. For examples, in the case of squeeze-film damping, the Reynolds equation is solved to obtain
a relation between the pressure distribution and the microplate mode shape. In thermoelastic damping,
the heat equation is solved for the temperature variation in terms of the elastic strains. These analytical
solutions decouple the coupled physical domains (fluidic, thermal, and structural) and reduce the global
number of variables to be solved. Moreover, they transform the design problem from a 3D problem, where
for instance the gas in the capacitor gap has to be meshed or the temperature variation across the thickness
of a structure has to be determined, to a 2D problem on the plate domain only. Another computational
reduction can be achieved by extracting explicit analytical expressions for the damping coefficients. These
coefficients can be implemented in reduced-order models to account for energy losses.
There are several mechanisms of energy dissipation in MEMS devices [50]. The most common include

18
losses into the surrounding fluid due to acoustic radiation and viscous damping, losses into the structure
mounts, and intrinsic damping caused by losses inside the material of the mechanical structure. Among all
of the damping sources, viscous damping is the most significant source of energy loss in MEMS. For typical
MEMS devices employing a parallel-plate capacitor, viscous damping corresponds to squeeze-film damping.
The majority of models for squeeze-film damping are derived for rigid plates. Thus, these models are
inaccurate for flexible structures. Starr [51] modeled the behavior of a capacitive parallel-plate accelerometer
using the linearized Reynolds equation, assuming small deflection and pressure variation and incompressible
fluid. He derived an exact expressions for the damping force of a circular disk and an approximate expression
for that of a rectangular plate. To account for large displacements and compressibility effect, he provided
correction factors for the analytical expressions.
Blech [52] solved analytically the linearized compressible Reynolds equation for the pressure in the case
of oscillating rigid plates of rectangular and circular shapes with trivial pressure boundary conditions and
derived analytical expressions for the spring and damping coefficients due to squeeze-film damping. Darling
[53] extended the Blech model [52] to arbitrary venting conditions in the case of rigid plates. Their solution
scheme is based on the Green’s function.
A group of researchers used statistical thermodynamics [54]. Kàdàr et al. [55] and Li et al. [56]
modified the Christian model [57], which determines the effect of a moving rigid body on changing the linear
momentum of the gas molecules, by improving the distribution function of the velocity molecules to reflect
more the physics of the problem. They compared their theoretical results to the theory and experimental
data of Zook et al. [58] and found that their theory reduced the discrepancy; however the discrepancy was
still significant.
Bao et al. [59] used an energy-transfer model to study the effect of a moving structure on changing the
kinetic energy of the gas molecules. They derived an expression for the quality factor, similar to that of
Christian [57], but modified by a correction factor, which is proportional to the gap width and the inverse
of the plate length. They compared their theoretical results to the theoretical results and experimental data
of Zook et al. [58] and found out that their theory improved the agreement with the experimental data;
however there was still significant discrepancy.
Extrinsic damping mechanisms can be minimized by a proper design of devices and their operating
conditions. For example, squeeze-film damping can be minimized by increasing the distance between the
capacitor electrodes and encapsulating the device at a very low pressure. Intrinsic mechanisms on the other
hand are more difficult to control because they depend primarily on the material and geometric properties of
the structures. There are many mechanisms that contribute to intrinsic damping. Recent studies [60] have
shown that thermoelastic damping can be a dominant source of intrinsic damping in MEMS. Thermoelastic
damping results from the irreversible heat flow generated by the compression and decompression of an
oscillating structure.
The first to analyze thermoelastic damping rigorously is Zener [61], who gave an analytical approximation
for the quality factor of metallic beams due to thermoelastic damping. In a recent work, Lifshitz and Roukes

19
l
x

z b
Microplate
Vp+v(t)
h

y Stationary Electrode d

Squeezed Fluid

Figure 8: Schematic of a MEMS device.

[62] solved the problem of thermoelastic damping of beams and derived an analytical expression for the
quality factors. They calculated the quality factors of various microbeams and found that their model
yields results close to that of the Zener [61] model. Nayfeh and Younis [63, 64] analyzed thermoelastic and
squeeze-film damping in microplates. They extracted analytical expressions for the temperature, pressure,
and quality factor of a microplate due to these dissipation mechanisms. Next, we briefly give an account of
these expression.
We consider a microplate (Figure 8) actuated by an electrostatic load of magnitude Vp and subject to
a net pressure force P̂ (x̂, ŷ, t̂) per unit area due to the gas in the gap. Assuming small displacements (i.e.
small electric loading), we obtain the following linear equation of motion governing the transverse deflection
of the microplate including the effect of thermoelastic damping [65, 66, 67]:

ǫVp2
Z h/2 Z h/2
4 ∂ 2 ŵ ∂ 2 ŵ Eαt 2 Eαt
D∇ ŵ − N̂1 2 + ρh 2 = 3 ŵ − P̂ − ∇ ŵ (T̂ − T0 )dẑ − ẑ∇2 (T̂ − T0 )dẑ (30)
∂ x̂ ∂ t̂ d 1−ν −h/2 1 − ν −h/2

where N̂1 is the axial force per unit length in the x̂ direction, T̂ (x̂, ŷ, ẑ, t̂) is the temperature distribution, T0
is the stress-free temperature, and αt is the coefficient of thermal expansion. The temperature distribution
is governed by the linearized heat conduction equation [63]

∂ T̂ Eαt T0 ∂
k∇2 T̂ = ρCp − (ẑ∇2 ŵ) (31)
∂ t̂ 1 − ν ∂ t̂
where Cp is the heat capacity coefficient at constant pressure. The pressure is governed by the Reynolds
equation. Assuming small variations around the static pressure Pa in the air gap, we obtain the following
linearized equation governing the pressure distribution underneath the microplate
!
∂ 2 P̂ ∂ 2 P̂ ηef f ∂ P̂ ∂ ŵ
+ = 12 d − Pa (32)
∂ x̂2 ∂ ŷ 2 Pa d3 ∂ t̂ ∂ t̂

where ηef f is the effective viscosity of the fluid in the gap [68].
Equations (30)-(32), along with appropriate sets of boundary conditions, is a distributed-parameter
system. Numerical solution of the free damped vibration of the device is computationally cumbersome.

20
Instead, we follow Nayfeh and Younis [63, 64] and show how to reduce the computational cost of the
simulation by using perturbation techniques.

5.1 Thermoelastic Damping


We start by driving an analytical expression for thermoelastic damping in microplates. We assume that
there is no squeeze-film damping (P̂ = 0), in which case thermoelastic damping is an important source
of dissipation. By noting that the temperature variation across the plate thickness is much larger than its
variation across the plane of the plate, Nayfeh and Younis [63] reduced the linear damped eigenvalue problem
to
ǫVp2
Z h/2 Z h/2
4 ∂ 2 φn Eαt 2 Eαt 2
∇ φn − N̂1 + ∇ φ n θ n dẑ + ẑ∇ θ n dẑ − φn = ωn2 φn (33)
∂ x̂2 1−ν −h/2 1 − ν −h/2 d3
∂ 2 θn Eαt T0
k 2
= iωn ρCp θn − iωn ẑ∇2 φn (34)
∂ ẑ 1−ν
where φn (x̂, ŷ) and θn (x̂, ŷ, ẑ) are the nth complex mode shapes of the plate and the associated temperature
variation, respectively, and ωn is the nth complex eigenvalue. The temperature boundary conditions are
assumed to be zero-heat flux from the plate to the ambient environment; that is,
∂θn 1 1
=0 at ẑ = h and − h (35)
∂ ẑ 2 2
The solution of Equations (34) and (35) can be expressed as
 
Eαt T0 sin(Kp ẑ)
θn = ∇2 φn ẑ − (36)
(1 − ν)ρCp Kp cos( 21 Kp h)

where r
ωn ρCp
Kp = (1 − i)
2k
Substituting Equation (36) into Equation (33), carrying out the integrations, and retaining the linear terms,
we obtain
∂ 2 φn ǫVp2
DT ∇4 φn − N̂1 − φn = ωn2 φn (37)
∂ x̂2 d3
where DT = D + Dt and Dt is given by

E 2 α2t T0 h3
 
h 2 tan(Kp h/2)
Dt = + − (38)
(1 − ν)2 ρCp 12 Kp2 Kp3

Equations (37) and (38) can be used, along with any appropriate set of structural boundary conditions, to
simulate the behavior of an electrostatically actuated microplate accounting for thermoelastic damping.
Next, we derive an analytical expression for the quality factors of microplates due to thermoelastic
damping. Because Dt << D, we apply the method of strained parameters [69, 70] to Equation (37). The

21
analysis is applicable to any boundary conditions. To this end, we seek a first-order solution to Equation
(37) and associated boundary conditions in the form

φn ≈ φn0 + ε1 φn1 (39)


ωn ≈ ωn0 + ε1 ωn1 (40)
T
D ≈ D + ε1 Dt (41)

where ε1 is a small nondimensional bookkeeping parameter. Substituting Equations (39-41) into Equation
(37) and using the method of strained parameters, we obtain

Dt h 1 b/ℓ
∂ 2 φn0 ǫVp2
Z Z i
2
ωn1 = N̂1 φn0 dx̂dŷ + + ω n (42)
2ωn0 D 0 0 ∂ x̂2 d3 0

R 1 R b/ℓ
where φn0 is normalized such that ˆ = 1. This result is applicable to plates with general
φ2n0 dx̂dy
0 0
shapes and boundary conditions and low levels of DC voltages.
For the special case of N̂1 = 0 and Vp = 0, the analytical expression for the quality factor reduces to the
following simple expression:

h3 ρCp (1 − ν)
Q=   (43)
3 2 tan(Kp h/2)
12(1 + ν)Eα2t T0 Imag h12 + Kh2 − Kp3
p

where Imag denotes the imaginary part.


Nayfeh and Younis [63] set ν = 0 in Equation (43), calculated Q for a microbeam (Figure 1) with ℓ = 10b
and b = 10h for various T0 , and obtained results in full agreement with those obtained using the model of
Lifshitz and Roukes [62] as shown in Figure 9. They also calculated the variation of Q for a fully clamped
plate oscillating in the first mode at T0 = 300K with the plate thickness h as shown in Figure 10. The plate
specifications are ℓ = 200µm, b = 100µm, and ν = 0.25.

5.2 Squeeze-Film Damping


To determine an approximation to squeeze-film damping, Nayfeh and Younis [64] neglected the effect of
thermoelastic damping in Equations (30)-(32). Because the analysis depends on the structural and acoustical
boundary conditions, we confine the analysis to the case of Figure 8 and then give an overview of the procedure
for other cases. The pressure boundary conditions for this case are zero flux at the clamped edges of the plate
and trivial pressure at the open edges. Following Nayfeh and Younis [64], we introduce the nondimensional
variables
x̂ ŷ t̂ ŵ P̄ˆ
x= , y= , t= , w= , P̄ = (44)
ℓ ℓ T d Pa

22
5
x 10
16

14

12

10
Q

2
100 150 200 250 300 350 400
T (K)
0

Figure 9: Comparison of the quality factors calculated using the model of Nayfeh and Younis [63] (solid line)
to that calculated using the model of Lifshitz and Roukes [62] (stars).

q
4
where T = ρhℓ D . Substituting Equations (44) into Equations (30) and (32), we obtain the following system
of equations:

∂ 4w ∂4w ∂4w ∂2w ∂2w


+ 2 + − N + 2 = −Pnon P̄ + αVp2 w + αVp ve (45)
∂x4 ∂x2 ∂y 2 ∂y 4 ∂x2 ∂t
∂ 2 P̄ ∂ 2 P̄
 
∂ P̄ ∂w
+ =σ − (46)
∂x2 ∂y 2 ∂t ∂t

where the nondimensional parameters appearing in Equations (45) and (46) are

N̂1 ℓ2 Pa ℓ4 12ηef f ℓ2 ǫℓ4


N1 = , Pnon = , σ= , α= (47)
D dD d2 Pa T d3 D
The nondimensional boundary conditions for the case of Figure 8 are
At y = 0 and y = b/ℓ

∂2w ∂2w
+ ν =0 (48)
∂y 2 ∂x2
∂3w ∂ 3w
+ (2 − ν) =0 (49)
∂x3 ∂x2 ∂y
P̄ = 0 (50)

23
4
x 10
7

4
Q

0
2 3 4 5 6 7 8 9 10 11
h (µ m)

Figure 10: Variation of Q of the first mode of a fully clamped plate with h [63].

At x = 0 and x = 1

w=0 (51)
∂w
=0 (52)
∂x
∂ P̄
=0 (53)
∂x
The linear damped eigenvalue problem is governed by [64]
∂ 4 φn ∂ 4 φn ∂ 4 φn ∂ 2 φn
4
+2 2 2 + 4
−N + Pnon ψn − αVp2 φn = ωn2 φn (54)
∂x ∂x ∂y ∂y ∂x2
∂ 2 ψn ∂ 2 ψn
2
+ = iωn σ (ψn − φn ) (55)
∂x ∂y 2

At y = 0 and y = b/ℓ

∂ 2 φn ∂ 2 φn
+ ν =0 (56)
∂y 2 ∂x2
∂ 3 φn ∂ 3 φn
+ (2 − ν) =0 (57)
∂x3 ∂x2 ∂y
ψn = 0 (58)

24
At x = 0 and x = 1

φn = 0 (59)
∂φn
=0 (60)
∂x
∂ψn
=0 (61)
∂x
where φn (x, y) and ψn (x, y) are the nth complex mode shapes of the plate and pressure, respectively, and ωn
is the nth complex nondimensional eigenvalue. Equations (54-61) can be solved numerically for the complex
eigenvalues and mode shapes.
Because typically |ωn | σ >> 1, the boundary-value problem represented by Equations (55), (58), and
(61) is a singular-perturbation problem [70]. Such a case occurs when the gradient of a dependent variable
undergoes rapid changes over a very narrow region. In this problem, the pressure changes sharply near the
free edges. Applying the method of matched asymptotic expansion to Equations (55), (58), and (61) yields
[64]
b/ℓ−y y
− 1+i
√ √ − 1+i
√ √
ψn (x, y) = φn (x, y) − φn (x, b/ℓ)e 2 ε2
− φn (x, 0)e 2 ε2
+ ··· (62)

1
where ε2 = σ | ωn 0 |
. Equation (62) gives an approximate analytical expression for the nth complex pressure
mode shape in terms of the nth complex plate mode shape and eigenvalue.
We substitute Equation (62) into Equation (54) and obtain an equation, which along with the boundary
condition (57), (58), (60), and (61), represent a linear distributed-parameter system for the dynamic behavior
of the microplate under the coupled effect of squeeze-film damping, structural forces, and linear electrostatic
forces. This system is solved for the nth complex mode shape and eigenvalue. The real part of the complex
eigenvalue yields the frequency of the microplate, whereas the ratio between its real part and twice its
imaginary part yields the quality factor.
Nayfeh and Younis [64] calculated the quality factors of an electrically actuated microplate of length
310µm [71], employed as a transducer in a resonant sensor, under various gas pressures. Figure 11 shows a
comparison of the calculated quality factor Q (stars) to the experimental data (triangles) of Legtenberg and
Tilmans [71]. The agreement is excellent, thereby validating the model.
1
The perturbation approach depends primarily on the fact that (ε2 = σ | ωn 0 |
<< 1). Fully clamped
plates also have very high natural frequencies. The pressure boundary conditions for a fully clamped plate
demand zero flux at all edges; hence there are no boundary layers to first order in this case because the
pressure boundary conditions are similar to the structural boundary conditions. For this case, the pressure
distribution is the same as the plate mode shape (the last two terms in Equation (62) are zero). Hence, to the
first approximation, the pressure has a pure spring-force effect. To derive an expression for the quality factor,
Nayfeh and Younis [64] applied the method of strained parameters [70] and sought a first-order solution to

25
5
10

4
10

3
Q

10

2
10

1
10
−6 −5 −4 −3 −2 −1 0 1
10 10 10 10 10 10 10 10
Pa (mbar)

Figure 11: Comparison of the calculated quality factors (stars) to the experimental data (triangles) of
Legtenberg and Tilmans [71].

Equations (54) and (55) and their boundary conditions [64] in the form

φn ≈ φn0 + ε2 φn1 (63)


ψn ≈ ψn0 + ε2 ψn1 (64)
ωn ≈ ωn0 + ε2 ωn1 (65)

Their final result gives the following imaginary part of the complex natural frequency:
 2
∂2 φ
R 1 R b/ℓ 
∂ φ
−iPnon 0 0 φn0 ∂xn2 0 + ∂yn2 0 dxdy
ωn1 = R 1 R b/ℓ (66)
2ωn0 0 0 φ2n0 dxdy

which yields the damping of the system.


For the case of a clamped annular microplate of outer radius R1 and inner radius R2 exhibiting axisym-
metric bending, following a procedure similar to that used for fully clamped rectangular plates, Nayfeh and
Younis [64] obtained the damping coefficient
h  i
RR ∂ ∂φ
−iPnon R21/R1 φn0 ∂r r ∂rn0 dr
ωn1 = RR (67)
2ωn0 R21/R1 rφ2n0 dr

The pressure distribution is also the same as the structural mode shape. For clamped circular plates, R2

26
is set equal to zero in Equation (67). Also, in calculating ωn0 and φn0 , the boundary conditions at R2 are
replaced with conditions that require finite values of w and P̄ at r = 0.
Equations (66) and (67) can be used in reduced-order models of microplates to account for squeeze-
film damping (without the need to include the Reynolds equation or the Navier-Stokes equations in the
simulation).

6 Summary and Conclusions


We presented a review of ROMs for MEMS devices. We classified the ROMs into node and domain methods.
We concluded that node methods perform poorly in predicting transients, large motions, and nonlinear
behavior compared to domain methods. We distinguished between two categories of domain methods. The
first category requires the use of the time series resulting from a finite-element or finite-difference simulation
of the full nonlinear model or experiments to extract a basis set to be used in the Galerkin procedure.
We concluded that such methods are limited to operation conditions near that used to generate the time
series. More importantly, these models cannot predict any qualitative device behavior, which is not present
in the time series. The second category employs the mode shapes of the device. This approach can capture
qualitative and quantitative changes in the device behavior and has better performance in the presence of
nonlinearities provided that enough modes are retained in the approximation.
We used the second category and developed ROMs for microbeams and rectangular and circular mi-
croplates and validated them with experimental and full simulation results. We presented ROMs of squeeze-
film damping and thermoelastic damping in MEMS and validated them with available theoretical and exper-
imental results. We conclude that these ROMs need to be extended to a broader class of devices, employing
structural components of different geometries and boundary conditions.

References
[1] Zavracky, P. M., Majumder, S., and McGruer, N. E., ‘Micromechanical switches fabricated using nickel
surface micromachining’, Journal of Microelectromechanical Systems 6, 1997, 3–9.

[2] Choi, B. and Lovell, E. G., ‘Improved analysis of microbeams under mechanical and electrostatic loads,’
Journal of Microelectromechanical Systems 7 1997, 24–9.

[3] Guyan, R. J., ‘Reduction of stiffness and mass matrices’, AIAA Journal 3, 1965, 380.

[4] Bechtold, T., Rudnyi, E. B., and Korvink, J.G., ‘Automatic order reduction of thermo-electric models
for MEMS: Arnoldi versus Guyan’, in Proceedings of the Fourth International Conference on Advanced
Semiconductor Devices and Microsystems, Smolenice, Slovakia, 2002, pp. 333–336.

27
[5] Ostergaard, D.F. and Gyimesi, M., ‘Finite element based reduced order modeling of Micro Electro Me-
chanical Systems (MEMS)’, in Proceedings of the International Conference on Modeling and Simulation
of Microsystems, San Diego, CA, 2000, pp. 684-687.

[6] Bechtold, T., Rudnyi, E. B., and Korvink, J.G., ‘Automatic order reduction of thermo-electric model
for micro-ignition unit’, in Proceedings of the International Conference on Simulation of Semiconductor
Processes and Devices, Kobe, Japan, 2002, pp. 131–134.

[7] Wang, F. and White, J., ‘Automatic model order reduction of a microdevice using the Arnoldi approach’,
in Proceedings of the International Mechanical Engineering Congress and Exposition, Anaheim, CA,
1998, pp. 527–530.

[8] Hung, E. S., Yang, Y.-J., and Senturia, S. D., ‘Low-order models for fast dynamical simulation of MEMS
microstructures’, in Proceedings of the International Conference on Solid-State Sensors and Actuators:
Transducers 1997, Chicago, IL, Vol. 2, 1997, pp. 1101-1104.

[9] Bai, Z. , Bindel, D. , Clark, J. V., Demmel, J., Pister, K. S. J., and Zhou, N., ‘New numerical techniques
and tools in Sugar for 3D MEMS simulation’, in Proceedings of the International Conference on Modeling
and Simulation of Microsystems, Hilton Head Island, SC, 2001, pp. 31–34.

[10] Clark, J. V., Bindel, D., Zhou, N., Bhave, S., Bai, Z., Demmel, J., and Pister, K. S. J., ‘Sugar:
Advancements in a 3D multi-domain simulation package for MEMS’, in Proceedings of the Microscale
Systems: Mechanics and Measurements Symposium, Portland, OR, 2001.

[11] Clark, J. V., Bindel, D., Kao, W., Zhu, E., Kuo, A., Zhou, N., Nie, J., Demmel, J., Bai, Z., Govindjee,
S., Pister, K. S. J., Gu, M., and Agogino, A., ‘Addressing the needs of complex MEMS design’, in
Proceedings of the International Conference on Micro Electro Mechanical Systems, Las Vegas, NV,
2002, pp. 204–209.

[12] Srinivasan, V., Jog, A., and Fair, R. B., ‘Scalable macromodels for microelectromechanical systems’, in
Proceedings of the International Conference on Modeling and Simulation of Microsystems, Hilton Head
Island, SC, 2001, pp. 72-75.

[13] Chen, Y. and White, J., ‘A quadratic method for nonlinear model order reduction’, in Proceedings of
the International Conference on Modeling and Simulation of Microsystems, San Diego, CA, 2000, pp.
477-480.

[14] Ramaswamy, D., White, J., ‘Automatic generation of small-signal dynamic macromodels from 3-D
simulation’, in Proceedings of the International Conference on Modeling and Simulation of Microsystems,
Hilton Head Island, SC, 2001, pp. 2730.

28
[15] Chen, J. and Kang, S.-M., ‘Techniques for coupled circuit and micromechanical simulation’, in Pro-
ceedings of the International Conference on Modeling and Simulation of Microsystems, San Diego, CA,
2000, pp. 213-216.

[16] Chen, J. and Kang, S.-M., ‘An algorithm for automatic model-order reduction of nonlinear MEMS
devices’, in Proceedings of the International Symposium on Circuits and Systems, Geneva, Switzerland,
Vol. 2, 2000, pp. 445-448.

[17] Chen, J. and Kang, S.-M., ‘Model-order reduction of weakly nonlinear MEMS devices with Taylor
series expansion and Arnoldi process’, in Proceedings of the IEEE Midwest Symposium on Circuits and
Systems, Lansing, MI, 2000, pp. 248-251.

[18] Chen, J. and Kang, S.-M., ‘Dynamic macromodeling of MEMS mirror devices’, in Technical Digest of
the International Electron Devices Meeting, Washington, DC, 2001 , pp. 925–928.

[19] Rewieński, M. and White, J. , ‘A trajectory piecewise-linear approach to model order reduction and
fast simulation of nonlinear circuits and micromachined devices’, in Proceedings of the IEEE/ACM
International Conference on Computer-Aided Design, San Jose, CA, Vol. 1, 2001, pp. 252-257.

[20] Rewieński, M. and White, J. , ‘A Trajectory piecewise-linear approach to model order reduction and
fast simulation of nonlinear circuits and micromachined devices’, IEEE Transactions on Computer-Aided
Design of Integrated Circuits and Systems 22, 2003, 155–170.

[21] Rewieński, M. and White, J. , ‘Improving trajectory piecewise-linear approach to nonlinear model
order reduction for micromachined devices using an aggregated projection basis’, in Proceedings of the
International Conference on Modeling and Simulation of Microsystems, San Juan, PR, Vol. 1, 2002, pp.
128-131.

[22] Vasilyev, D., Rewieński, M., and White, J. , ‘A TBR-based trajectory piecewise-linear algorithm for
generating accurate low-order models for nonlinear analog circuits and MEMS’, in Proceedings of the
Conference on Design Automation, Anaheim, CA, 2003, pp. 490–495.

[23] Liang, Y. C., Lin, W. Z., Lee, H. P., Lim, S. P., Lee, K. H., and Sun, H., ‘Proper orthogonal decompo-
sition and its applications - part II: Model reduction for MEMS dynamical analysis’, Journal of Sound
and Vibration 256(3), 2002, 515-532.

[24] Hung, E. S. and Senturia, S. D., ‘Generating efficient dynamical models for microelectromechanical
systems from a few finite-element simulations runs’, Journal of Microelectromechanical Systems 8, 1999,
280–289.

29
[25] Chen, J. and Kang, S.-M., ‘Model-order reduction of nonlinear MEMS devices through arclength-
based Karhunen-Loeve decomposition’, in Proceedings of the International Symposium on Circuits and
Systems, Sydney, Australia, Vol. 3, 2001 , pp. 457–460.

[26] Liang, Y. C., Lin, W. Z., Lee, H. P., Lim, S. P., Lee, K. H., and Feng, D. P., ‘A neural-network-based
method of model reduction for the dynamic simulation of MEMS’, Journal of Micromechanics and
Microengineering 11, 2001, 226-233.

[27] Lin, W. Z., Lee, K. H., Lim, S. P., and Liang, Y. C., ‘Proper orthogonal decomposition and component
mode synthesis in macromodel generation for the dynamic simulation of a complex MEMS device’,
Journal of Micromechanics and Microengineering 13, 2003, 646-654.

[28] Qiao, R. and Aluru, N. R., ‘Mixed-domain and reduced-order modeling of electroosmotic transport in
Bio-MEMS’, in Proceedings of the International Workshop on Behavioral Modeling and Simulation,
Orlando FL, 2000, pp. 51–56.

[29] De, S. K. and Aluru, N. R., ‘Physical and reduced-order dynamic analysis of MEMS’, in Proceedings of
the International Conference on Computer Aided Design, San Jose, CA, 2003, pp. 270–273.

[30] Anathasuresh, G. K., Gupta, R. K., and Senturia, S. D., ‘An approach to macromodelling of MEMS
for nonlinear dynamic simulation’, in Proceedings of the Symposium on Mechanics in Microelectrome-
chanical Systems (MEMS), Atlanta, GA, 1996, ASME-DSC, pp. 401-407.

[31] Grtétillat, M. A., Yang, Y. J., Hung, E. S., Rabinovich, V., Ananthasuesh, G. K., Rooij, N. F., and
Senturia, S. D., ‘Nonlinear electromechanical behavior of an electrostatic microrelay’, in Proceedings of
the Iternational Conference on Solid State Sensors and Actuators: TRANSDUCERS 1997, Chicago, IL,
Vol. 2, 1997, pp. 1141–1144.

[32] Gabbay, L. D., Mehner, J. E., and Senturia, S. D., ‘Computer-aided generation of nonlinear reduced-
order dynamic macromodels I: Non-stress-stiffened case’, Journal of Microelectromechanical Systems 9,
2000, 262–269.

[33] Varghese, M., Rabinovich, V. L., and Senturia, S. D., ‘Reduced-order modeling of Lorentz force actuation
with modal basis functions’, in Proceedings of the International Conference on Modeling and Simulation
of Microsystems, San Juan, PR, 1999, pp. 155–158.

[34] Mehner, J. E., Gabbay, L. D., and Senturia, S. D., ‘Computer-aided generation of nonlinear reduced-
order dynamic macromodels II: Stress-stiffened case’, Journal of Microelectromechanical Systems 9,
2000, 270–278.

30
[35] Bennini, F., Mehner, J., and Dötzel, W., ‘Computational methods for reduced order modeling of cou-
pled domain simulations’, in Proceedings of the International Conference on Solid-State Sensors and
Actuators: TRANSDUCERS ’01, Munich, Germany, 2001, pp. 260–263.

[36] Mehner, J. E., Doetzel, W., Schauwecker, B., and Ostergaard, D., ‘Reduced order modeling of fluid
structural interactions in mems based on modal projection techniques’, in Proceedings of the Iternational
Conference on Solid-State Sensors, Actuators and Microsystems: TRANSDUCERS ’03, Vol. 2, 2003,
pp. 1840–1843.

[37] Westby, E. R. and Fjeldly, T. A., ‘Nonlinear analytical reduced-order models for MEMS’, in Proceedings
of the International Conference on Modeling and Simulation of Microsystems, San Juan, PR, Vol. 1,
2002, pp. 150-153.

[38] Xie, W. C., Lee, H. P., and Lim, S. P., ‘Nonlinear dynamic analysis of MEMS switches by nonlinear
modal analysis’, Nonlinear Dynamics 31, 2003, 243–256.

[39] Abdel-Rahman, E. M., Younis, M. I., and Nayfeh, A. H., ‘Characterization of the mechanical behavior of
an electrostatically actuated microbeam’, Journal of Micromechanics and Microengineering 12, 2002,
759–766.

[40] Younis, M. I., Abdel-Rahman, E. M., and Nayfeh, A. H., ‘Static and dynamic behavior of an electri-
cally excited resonant microbeam,” in Proceeding of the AIAA Structures, Structural Dynamics, and
Materials Conference, Denver, Colorado, 2002, AIAA Paper 2002-1305.

[41] Abdel-Rahman, E. M., Younis, M. I., and Nayfeh, A. H., ‘A nonlinear reduced-order model for electro-
static MEMS,” in Proceedings of the Biennial ASME Conference on Mechanical Vibration and Noise,
Chicago, IL, 2003, DETC2003/VIB-48517.

[42] Younis, M. I., Abdel-Rahman, E. M., and Nayfeh, A. H., ‘A Reduced-Order Model for Electrically
Actuated Microbeam-Based MEMS’, Journal of Microelectromechanical Systems 12, 2003, 672–680

[43] Tilmans, H. A. and Legtenberg, R., ‘Electrostatically driven vacuum-encapsulated polysilicon res-
onators. Part II. Theory and performance’, Sensors and Actuators A 45, 1994, 67–84.

[44] Nayfeh, A. H. and Mook, D. T., Nonlinear Oscillations, Wiley, New York, 1979.

[45] Nayfeh, A. H., Nonlinear Interactions, Wiley, New York, 2000.

[46] Zhao, X., Abdel-Rahman, E. M., and Nayfeh, A. H., ‘Mechanical behavior of an electrically actuated
microplate’, in Proceedings of the Biennial ASME Conference on Mechanical Vibration and Noise,
Chicago, IL, 2003, DETC2003/VIB-48531.

31
[47] Francais, O. and Dufour, I., ‘Normalized abacus for the global behavior of diaphragms: Pneumatic,
electrostatic, piezoelectric or electromagnetic actuation’, Journal of Modeling and Simulation of Mi-
crosystems 2, 1999, 149–160.

[48] Vogl, G. W. and Nayfeh, A. H., ‘A Reduced-order model for electrically actuated clamped circular
plates’, in Proceedings of the Biennial ASME Conference on Mechanical Vibration and Noise, Chicago,
IL, 2003, DETC2003/VIB-48530.

[49] Osterberg, P. M., Electrostatically Actuated Microelectromechanical Test Structures for Material Prop-
erty Measurement, Ph.D. Dissertation, Massachusetts Institute Technology, Cambridge, MA, 1995.

[50] Tilmans, H. A., Elwespoek, M., and Fluitman, J. H., ‘Micro resonant force gauges’, Sensors and Actu-
ators A 30, 1992, 35–53.

[51] Starr, J. B., ‘Squeeze-film damping in solid-state accelerometers’, in Proceeding of the IEEE Solid-State
Sensor and Actuator Workshop, Hilton Head Island, SC, 1990, pp. 44–47.

[52] Blech J. J., ‘On isothermal squeeze films’, Journal of Lubrication Technology A 105, 1983, 615–620.

[53] Darling, R. B., Hivick, C. and, Xu, J., ‘Compact analytical modeling of squeeze film damping with
arbitrary venting conditions using a Green’s function approach’, Sensors and Actuators A 70, 1998,
32–41.

[54] Gupta, M. C., Statistical Thermodynamics, Wiley, New Delhi, 1990.

[55] Kàdàr, Z., Kindt, W., Bossche, A., and Mollinger, J., ‘Quality factor of torsional resonators in the
low-pressure region’, Sensors and Actuators A 53, 1996, 299–303.

[56] Li, B., Wu, H., Zhu, C., and Liu, J., ‘The theoretical analysis on damping characteristics of resonant
microbeam in vacuum’, Sensors and Actuators A 77, 1999, 191–194.

[57] Christian, R. G., ‘The theory of oscillating-vane vacuum gauges’, Vacuum 16, 1966, 175–178.

[58] Zook, J. D., Burns, D. W., Guckel, H., Sniegowski, J. J., Engelstad, R. L., and Feng, Z., ‘Characteristics
of polysilicon resonant microbeams’, Sensors and Actuators A 35, 1992, 290–294.

[59] Bao, M., Yang, H., Yin, H., and Sun, Y., ‘Energy transfer model for squeeze-film air damping in low
vacuum’, Journal of Micromechanics and Microengineering 12, 2002, 341–346.

[60] Roszhart, T. V., ‘The effect of thermoelastic internal friction on the Q of micromachined silicon res-
onators’, in Technical Digest of Solid-State Sensors and Actuators Workshop, Hilton Head Island, SC,
1990, pp. 13–16.

32
[61] Zener, C., ‘Internal friction in solids: I. Theory of internal friction in reeds’, Physical Review 52, 1937,
230–235.

[62] Lifshitz, R. and Roukes, M. L., ‘Thermoelastic damping in micro- and nanomechanical systems’, Physical
Review B 61, 2000, 5600–5609.

[63] Nayfeh, A. H. and Younis, M. I., ‘Modeling and simulations of thermoelastic damping in microplates’,
Journal of Micromechanics and Microengineering 14, 2004, 1711–1717.

[64] Nayfeh, A. H. and Younis, M. I., ‘A new approach to the modeling and simulation of flexible microstruc-
tures under the effect of squeeze-film damping’, Journal of Micromechanics and Microengineering 14,
2004, 170–181.

[65] Boley, B. A. and Weiner, J. H., Theory of Thermal Stresses, Wiley, New York, 1960.

[66] Leissa, A. W., Vibration of Plates, NASA, Washington DC, 1969.

[67] Nayfeh, A. H. and Pai, F. P., Linear and Nonlinear Structural Mechanics, Wiley, New York, 2004.

[68] Veijola, T., Kuisma, H., Lahdenperä, J., and Ryhänen, T., ‘Equivalent-circuit model of the squeezed
gas film in a silicon accelerometer’, Sensors and Actuators A 48, 1995, 239–248.

[69] Nayfeh, A. H., Perturbation Methods, Wiley, New York, 1973.

[70] Nayfeh, A. H., Introduction to Perturbation Techniques, Wiley, New York, 1981.

[71] Legtenberg, R. and Tilmans, H. A., ‘Electrostatically driven vacuum-encapsulated polysilicon res-
onators. Part I. Design and fabrication’, Sensors and Actuators A 45, 1994, 57–66.

33

View publication stats

Você também pode gostar