Sunteți pe pagina 1din 7

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 22, NO.

3, JUNE 2013

553

Effect of Graphene Layers on Static Pull-in


Behavior of Bilayer Graphene/Substrate
Electrostatic Microactuators
Hossein Rokni and Wei Lu

AbstractA closed-form solution is obtained for the pull-in instability of curved multilayer graphene/substrate microcantilever
electrostatic actuators. The first-order fringing-field correction
and the interlayer shear between neighboring graphene layers
(GLs) and between the graphene and the substrate are incorporated into the analytical model. In the solution procedure, the governing fourth-order differential equation of variable coefficients is
converted into a Fredholm integral equation. The resulting equation is solved for the static pull-in voltages by adopting the first
natural mode of the cantilever beam as a deflection shape function.
The influence of GLs on the pull-in voltages of the electrostatic
microactuators is investigated. It is found that laying 10, 30, and
60 GLs on top of the substrate results in increases of about 95%,
190%, and 295%, respectively, in the pull-in voltage of the straight
bilayer graphene/substrate electrostatic microactuators. It is also
observed that the classical EulerBernoulli beam theory fails to
predict the pull-in voltages of the multilayer graphene/substrate
electrostatic microactuators, showing that the pull-in voltage is
highly affected by the graphene interlayer shear.
[2012-0089]
Index TermsClosed-form solution, graphene, microactuator,
pull-in voltage, shear effect.

I. I NTRODUCTION

WING to advantages such as favorable scaling property, low energy consumption, low cost, low driving
power, large deflection capacity, and relative ease of fabrication, electrostatic microactuators are widely used in microelectromechanical systems (MEMS) devices, such as micromirrors,
micromachined inertia sensors, microresonators, capacitive
pressure sensors, comb drivers, micropumps, inkjet printer
heads, and RF switches. One of the main interesting applications of quasi-static electrostatic pull-in instability is to extract
material parameters of thin films, such as Youngs modulus and
residual stress [1][3]. Accordingly, determination of the pullin voltage is critical in the design of MEMS devices and also
challenging due to the electromechanical coupling effect and
the nonlinearity of electrostatic force.
Microcantilever beams are mostly fabricated by surface micromachining techniques, which create residual stress gradient
in the cross section of the microbeam causing them to curl.
Manuscript received April 11, 2012; revised September 18, 2012; accepted
November 16, 2012. Date of publication December 21, 2012; date of current
version May 29, 2013. This work was supported in part by the National Science
Foundation under Award CMMI 1232883. Subject Editor N. Aluru.
The authors are with the Department of Mechanical Engineering, University
of Michigan, Ann Arbor, MI 48109 USA (e-mail: rokni@umich.edu; weilu@
umich.edu).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/JMEMS.2012.2230315

In addition, multilayer microcantilever beams may bend up


or down due to the thermomechanical mismatch between the
layers. Although numerous numerical [1], [4][9], analytical
[10][14], and experimental [1], [11], [15] investigations have
been conducted on the pull-in instability behavior of straight
microbeams subjected to electrostatic loadings, few research
groups [16][20] considered the curled beam.
Graphene is a one-atom-thick planar structure existing in either a single or multilayer form. Owing to its extraordinary mechanical, electronic, optical, and thermal properties, graphene
has attracted significant attention from the scientific community. According to recent mechanical experiments, graphene
appears to be the strongest material tested so far, with a Young
modulus of 1.0 TPa and an intrinsic strength of 130 GPa [21].
Therefore, graphene could be a promising candidate for the
reinforcement of MEMS devices, as the design of actuators
with higher stiffness, for instance, is desirable for an optimum
application of electrostatic actuators [22].
On the other hand, due to weak interlayer van der Waals
binding in multilayered graphene nanostrips, they are prone
to slide over each other and/or over substrate layers provided
that the external force exceeds the forces existing between
two graphene layers (GLs) or between the graphene and the
substrate. Slipping between graphene and the substrate, such
as silicon nitride and gold, has been experimentally observed
[23]. Since the classical EulerBernoulli beam theory is not
able to interpret this shear interlayer effect, multibeam shear
model theory has recently been developed [24], where interlayer shear modulus of graphene is involved. They performed
several molecular dynamics simulations to validate their proposed analytical model.
The present work is an attempt to study the pull-in instability
of curved multilayer graphene/substrate microcantilever electrostatic actuators within the context of the multibeam shear
model theory. The first-order fringing-field correction is incorporated into the analytical model. In the solution procedure,
the governing differential equation of variable coefficients is
transformed to a Fredholm integral equation. Then, the resulting Fredholm integral equation is solved by adopting the
first natural mode in order to describe the deflection profile of
the microcantilever beams under uniform electrostatic pressure.
The high accuracy of the current novel solution is verified
by comparing the proposed pull-in voltages with measured
data available in the literature. Finally, the effect of the different numbers of GLs on the pull-in voltages of the bilayer
graphene/substrate electrostatic microactuators is studied.

1057-7157/$31.00 2012 IEEE

554

Fig. 1.

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 22, NO. 3, JUNE 2013

Schematic of the curved bilayer graphene/SiNx microcantilever actuator.

II. M ODELING
Fig. 1 shows a curved multilayer graphene/silicon nitride
microcantilever beam with an electrostatic load induced by
an electric potential v. The microactuator, modeled by a thin
beam of length L, width w, and uniform substrate thickness
hs , is assumed to be made of silicon nitride (SiNx ), whereas
its top surface is covered by various numbers of GLs n with
a thickness of nhg . The coordinate system is attached to the
neutral axis at the left end of the microbeam, where x and z
refer to the horizontal and vertical coordinates, respectively.
A. Radius of Curvature
During the microfabrication process, the bilayer graphene/
substrate cantilever beam bends toward the GL due to the thermomechanical mismatch of the GL and the substrate cantilever.
For a curved cantilever beam with an initial radius of curvature
, the initial gap between the graphene electrode and the ground
plane g(x), including the dielectric contribution from the SiNx
layer [25], can be expressed as


hs
x
+ 1 cos
g(x) = g0 +
(1)
SiNx

where g0 is the gap between the fixed end of the cantilever beam
and the ground plane and SiNx (= 7.5) [26] is the relative
permittivity of the dielectric for the SiNx .
Considering the effect of the film thickness for the relatively
thin substrate, the radius of curvature , which is related to the
strain mismatch between graphene (g ) and the substrate (s ),
can be given by [27]
1
6EH(1 + H)
=
(g s )

hs

(2)

= 1 + 4EH + 6EH 2 + 4EH 3 + E 2 H 4

(3)

where

E = E g /E s , and H = nhg /hs . Also, E g and E s are the effective Young moduli of the graphene and the substrate layers,

Fig. 2. Variation of the radius of curvature against the graphene strain and the
number of GLs based on (2).

respectively. For a wide beam (w 5h, h = hs + nhg ), the


effective modulus E g (or E s ) can be approximated by the
plate modulus Eg /(1 g2 ) [or Es /(1 s2 )]; otherwise, E g
(or E s ) is the Young modulus Eg (or Es ) [7]. g and s are the
Poissons ratios of the graphene and the substrate, respectively.
As it is seen from (2), the radius of curvature is a function of the
graphene strain, substrate strain, and number of GLs. From the
experimental data reported in [23], the following conclusions
can be deduced.
1) The graphene strain itself is dependent on the number
of GLs. For instance, the graphene strain g changes
from one GL (1.45 102 , averaged from four samples) to two (1.35 102 , from one sample) and three
( 0.7 102 , from two samples) GLs attached to the
silicon nitride while the graphene strain of one GL on
gold is estimated to be about 0.3 102 averaged from
nine samples [23]. It seems that the graphene strain g
reduces as the number of the GLs increases.
2) The substrate strain of gold and silicon nitride is small
(s < 4 104 ) in comparison with the graphene strain.
Therefore, in (2), the effect of s on the radius of curvature can be neglected (i.e., (g s ) g ).
3) The radii of curvature of monolayer/SiNx , bilayer/SiNx ,
and trilayer/SiNx cantilever beams have been measured to
be about 3.6, 6.6, and 5.0 103 m1 , respectively. Unlike
the graphene strain, no certain trend can be observed
for the radius of curvature as a function of the number
of GLs.
In view of (2), the variation of the radius of curvature as a
function of the graphene strain and the number of graphene
nanoribbons (GNRs) is shown in Fig. 2 for the following parameters used in this paper: Eg = 1 TPa, Es = 220 GPa, hg =
0.35 nm (for single GL), hs = 188 nm, g = 0.165, s = 0.22,
and s = 4 104 [23]. It is seen that the radius of curvature
increases monotonically with the increase of the graphene strain
and the number of GLs which is somewhat inconsistent with
the experimental results [23] (see conclusion 3). It might be
attributed to the dependence of the graphene strain itself on the
number of GLs (see conclusion 1).

ROKNI AND LU: EFFECT OF GRAPHENE LAYERS ON PULL-IN BEHAVIOR OF ELECTROSTATIC MICROACTUATORS

Fig. 3. Variation of the graphene strain against the number of GLs based
on (4).

Due to the lack of experimental data available in the literature and the complexity of the graphenegraphene and
graphenesubstrate interlayer shear mechanisms, it is not feasible to find a clear relationship between the graphene strain
and the number of GLs. Nevertheless, the following equation
for the graphene strain as a function of the number of the GLs
can be proposed to achieve the best fit to the experimental
data [23]:
g = (0.15 + 4 e

) 102 .

(4)

The variation of the graphene strain against the number of the


GLs is shown in Fig. 3, which indicates that the aforementioned
function fits the experimental results very well. The following
points should be made about (4).
1) As mentioned earlier, due to very limited existing test
data, it is not prescribed to exactly follow the aforementioned function; however, it appears to be a reliable
rule of thumb to use this function as the graphene strain
decreases with an increase in the number of the GLs.
2) Owing to the negative thermal expansion coefficient of
graphene, the bilayer graphene/SiNx and graphene/gold
cantilever beams bend toward the GLs during the microfabrication process, leading to a tensile strain in the
GLs. This behavior has been reflected in (4), where
the graphene strain is always positive regardless of the
number of the GLs.
3) Due to the thermal expansion mismatch between the
substrate layer and the graphitic layer, whether few-layer
graphene or many-layer graphene (i.e., graphite), it is
reasonable to assume that, to some extent, the strain
in the graphitic layer approaches a constant value, as
the number of GLs increases. Since the graphite strain
induced in bilayer graphite/SiNx structures has not been
measured in the literature so far, the value of the graphite
strain is set to be 0.15 102 .
After substituting (4) into (2), the variation of the radius of
curvature against the number of the GLs is shown in Fig. 4. It is
observed that the proposed equation for the radius of curvature
of the graphene/substrate cantilever beams provides a good fit

555

Fig. 4. Variation of the radius of curvature versus the number of GLs using
(2) and (4).

to the experimental data [23]. It should be emphasized that


more experiments need to be conducted to better characterize
the effect of the number of the GLs on the graphene strain and
the radius of curvature.
B. Multibeam Shear
It is known that the graphene is atomically smooth, and its interaction with the substrate can be of van der Waals force or/and
friction force. If the external force exceeds these forces existing
between neighboring GLs or between the graphene and the
substrate, the GL slides over the substrate (graphene) to release
its strain energy [23]. Therefore, using the multibeam shear
model theory [24], the interlayer shear is incorporated into
the following analytical model for a more accurate estimation
of the pull-in instabilities of the multilayer graphene/substrate
microcantilever beams.
Based on the multibeam shear model theory, the governing
nonlinear differential equation for the pull-in instability of a
curved microcantilever beam structure incorporating the firstorder fringing-field correction is given by
E e Ie

4 uz (x)
2 uz (x)

[(n

1)G
+
G
]
A
gg
gs
g
x4
x2
=

where

0 wv 2
0.650 v 2
2 + 2 [g(x) u (x)]
2 [g(x) uz (x)]
z

(5)


2 
hs
h3s
+ hs
yc
E e Ie = E s w
12
2


2 
nh3g
nhg
+ nhg
+ hs y c
+ Eg w
12
2
yc =

E s h2s + E g n2 h2g + 2E g nhg hs


2(E s hs + E g nhg )

(6)

(7)

Ggg and Ggs are the graphenegraphene and the graphene


substrate interlayer shear moduli, respectively, Ag (= whg )
is the cross-sectional area of the graphene, 0 = 8.854
1012 C2 N1 m2 is the permittivity of vacuum, and yc is

556

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 22, NO. 3, JUNE 2013

the position of the neutral axis relative to the lower side of the
substrate. The interfacial shear moduli of the graphene/silicon
nitride and the graphene/graphene interface at room temperature are estimated to be Ggs 1 GPa [23] and Ggg 0.25 GPa
[24], respectively. Based on the EulerBernoulli beam theory,
the effective flexural rigidity of the bilayer graphene/substrate
cantilever beams can be given by [25]


2 
hs
h3s
+ hs
yc
E e Ie = E s w
12
2


2 
n3 h3g
nhg
+ nhg
+ hs y c
+ Eg w
. (8)
12
2

second integral

By comparing (6) and (8), it is seen that the entire second


term of (6) in the second brackets remains unchanged as this
term originates from the parallel axis theorem which reflects
the geometrical properties of the problem, whereas the first
term in the second brackets changes according to the multibeam
shear model theory in which the GLs are assumed to behave
as independent cantilever beams with the same transverse displacement [24].
By introducing the following dimensionless parameters:

X
X
V2
(X s)3 (s)ds
Uz (X) Q (X s)Uz (s)ds =
6

uz
g0

X=

x
L

Q=

[(n1)Ggg +Ggs ]Ag L2


E e Ie

V 2=

Uz =

0 v 2 wL4
2E e Ie g03

G0 =

G(X) = 1+

0.65g0
[7]
w



hs

L
(9)
+
1cos X
g0 SiNx g0

2 Uz (X)
QUz (X) = V 2
X 2

X
(X s)(s)ds+A1 X +A2 (14)
0

third integral
Uz (X)
Q
X

X

V2
Uz (s)ds =
2

X
X2
(X s)2 (s)ds + A1
2

+ A2 X + A3

(15)

fourth integral

X3
X2
+ A2
+ A3 X + A4
6
2

+ A1

(16)

where integration constants Ai (i = 1, 2, 3, 4) can be determined by applying boundary conditions at the beam ends and s
is an independent variable in the integration.
The following equations are taken into account in order to
satisfy the clamped and free boundary conditions at the two
ends of the beam:

Uz (X) 
=0
(17a)
X X=0


Uz (X)
=0
(17b)
X=0

and substituting (9) into (5), one can get


Uz (X) 
=0
X 2 X=1
 3

Uz (X)
Uz (X) 
Q
= 0.

X 3
X
X=1
2

4 Uz (X)
2 Uz (X)
Q
4
X
X 2
V2
G0 V 2
2 + [G(X) U (X)] .
[G(X) Uz (X)]
z

(10)

For the sake of convenience and generality, the aforementioned equation is rewritten as
(11)

where
G0
1
.
+
G(X) Uz (X)
[G(X) Uz (X)]2

(12)

Both sides of (11) are integrated four times with respect to X


from zero to X in order to reach the Fredholm integral equation
[28] as follows:
first integral
3

A3 = 0

(18a)

A4 = 0.

(18b)

Uz (X)
Uz (X)
Q
= V 2 (X)
X 4
X 2

(X) =

(17d)

Inserting (17a) and (17b) (i.e., clamped boundary condition)


into (15) and (16) results in

C. Fredholm Integral Equation

(17c)

Uz (X)
Uz (X)
=V2
Q
X 3
X

Substituting (17c) and (17d) (i.e., free boundary condition)


into (13) and (14), using (17d) to find Uz (X = 1), and doing
some mathematical manipulation lead to
1
A1 = V

(s)ds

2
A2 =
V2
Q+2

1 


Q
Q
3
s (1 s) +
(s)ds
6
6

X

1
(s)ds + A1

(19a)

(13)

(1 s)Uz (s)ds

Q2
0

(19b)

ROKNI AND LU: EFFECT OF GRAPHENE LAYERS ON PULL-IN BEHAVIOR OF ELECTROSTATIC MICROACTUATORS

Finally, substitution of (18) and (19) into (16) yields the


Fredholm integral equation for the microcantilever beam as
follows:
X
X
V2
Uz (X) Q (X s)Uz (s)ds
(X s)3 (s)ds
6
0

V2
+
6

1
X
0

Q2
Q+2


Q 2
s (s) (s)ds
(X 3s)
Q+2

1
X 2 (1 s)Uz (s)ds = 0.

(20)

D. Displacement

1
a2 (C) =
6



1 1
X 2 (X 3s) Q s2 (s3) (X)dX
Q+2
0

1

(X s)3 (X)dX .

Uz (X) = C(X)

At the pull-in state, the velocity and the acceleration of the


beam must be zero (i.e., quasi-static conditions) [14], [17], [30].
That is, the derivative of (23) with respect to C must be zero as
well. Hence, the following equation should also be satisfied in
addition to (23):

(X) = (cosh X cos X)

sinh sin
(sinh X sin X)
cosh + cos

(22)

where = 1.87510407. It should be pointed out that very good


results with negligible error (1%) and lower computational
cost can also be obtained when (X) = X 4 4X 3 + 6X 2 .
E. Closed-Form Pull-in Voltage
In order to yield one linear algebraic equation in terms of
C and V , we initially substitute (21) and (22) into (20), then
multiply the resulting Fredholm integral equation by (X), and
finally integrate it from zero to one. Thus, we get
Ca1 + V 2 a2 (C) = 0

a2 (C)
= 0.
C

(23)

(25)

Eliminating V from (23) and (25) leads us to obtain C at the


static pull-in, denoted by Cp as follows:

(21)

where C denotes the scaling of the displacement function and


(X) is the assumed deflection shape function satisfying the
boundary conditions.
Natural modes essentially satisfy the boundary conditions
and the homogeneous part of the governing equation of a
dynamic system, because they are the exact solution to the
free vibration of structures. Hence, the natural modes form
the foundation for forced response calculations in structural
dynamics. Accordingly, the first natural mode is adopted to
describe the deflection profile of cantilever beams subjected to
a uniformly distributed electrostatic load as follows [29]:

(24b)

a1 + V 2

The normalized displacement of the microcantilever beam


Uz (X), satisfying all four mechanical boundary conditions
given by (17a)(17d), is given by the following function:

557

a2 (Cp ) Cp

a2 (Cp )
= 0.
Cp

(26)

Substitution of Cp into (21) describes the deflection profile of


the microcantilever beam at static pull-in. The normalized static
pull-in displacement Uzp can now be obtained as the maximum
deflection of the microcantilever beam occurring at the point of
static pull-in. This is expressed as
Uzp (X)|X=1 = CP (X)|X=1 = 2CP .

(27)

The critical applied voltage (dimensionless static pull-in


voltage Vp ) corresponding to Cp is obtained from either (23)
or (25), i.e.,

Vp =

Cp a1
.
a2 (Cp )

(28)

Equation (28) provides the readers with a closed-form expression for the pull-in voltages of the bilayer graphene/
substrate electrostatic actuators. It should be noted that the
value of Cp is first calculated from (26) and then substituted
back into (28) to find the value of Vp .
III. R ESULTS AND D ISCUSSION
A. Comparison Study (No GLs)

where
1
a1 =

1 X
2
(X)dX Q (X s)(s)ds (X)dX

Q2
Q+2

1 1
X 2 (1s)(s)ds (X)dX
0

(24a)

In order to confirm the reliability and the high accuracy of the


aforementioned formulation, a comparison study is conducted
in Fig. 5 between the proposed pull-in voltages and those of
Gupta [20], who measured the pull-in voltages of polysilicon
microcantilever beams of 40-m width and 2.1-m thickness
with various effective lengths ranging from 100 to 500 m.
Based on the test bed parameters reported by Gupta [20], the
initial gap g0 and the initial radius of curvature are 2.4 m

558

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 22, NO. 3, JUNE 2013

presence of the interlayer slip. From Fig. 6, the following


findings can be made.

Fig. 5. Theoretical and measured pull-in voltages of the microcantilever


beams with different beam lengths.

Fig. 6. Comparison of pull-in voltages of the electrostatic microactuator


calculated by the EulerBernoulli and multibeam shear models for different
layers.

1) For a larger number of GLs, the classical EulerBernoulli


beam theory anticipates the microcantilever beams much
stiffer than what the multibeam shear model theory does.
2) For MLGNR electrostatic microactuators with fewer
GLs, the pull-in voltages obtained by the multibeam
shear model are greater than those of the classical
EulerBernoulli beam theory. It is mainly attributed to
the contribution of the shear rigidity toward increasing
the equivalent stiffness of the MLGNR electrostatic microactuators.
3) The growth rate of the pull-in voltage obtained by the
multibeam shear model theory reduces as the GLs become greater. This phenomenon has also been observed
by Liu et al. [24], where the growth rate of the first and
the second resonance frequencies of graphene cantilever
nanostrips reduces with the increase of the GLs.
4) All the aforementioned findings hold for both the straight
and the curved MLGNR electrostatic microactuators.
5) Placing 10, 30, and 60 GLs onto the substrate enhances
almost two-, four- and sixfold, respectively, the pull-in
voltage of the straight bilayer graphene/SiNx electrostatic
microactuators.
Before closing this section, it should be emphasized that
there are no experimental and/or numerical results in the
literature to verify the multibeam shear model for the bilayer graphene/substrate beams. An extensive experimental
study is underway to fabricate and test a set of bilayer
graphene/substrate electrostatic actuators.

IV. C ONCLUSION
and 40 mm, respectively, while the polysilicon Young modulus
is set to be 153 GPa. Also, the permittivity of dielectric medium
is = 0 r = 10.66 1012 F/m, where r = 1.2046 is the
dielectric constant of the dielectric medium between the beam
and the ground.
It is evident from Fig. 5 that the agreement between the
proposed pull-in voltages of the curled microcantilever beams
( = 40 mm) and those measured by Gupta [20] is found to
be excellent, whereas this is not the case for straight cantilever
beams ( = ).
B. Effect of GLs
In order to show the effect of GLs on the pull-in voltages
of the multilayer-GNR (MLGNR) electrostatic microactuators,
the variation of the pull-in voltage in volts versus the number of
GLs is shown in Fig. 6 for the classical EulerBernoulli model
(Q = 0) and the multibeam shear model (Q = 0) when Eg =
1 TPa, Es = 220 GPa, g = 0.165, s = 0.22, L = 1 m,
w = 0.1 m, g0 = 50 nm, hg = 0.35 nm (for single GL),
hs = 25 nm, and s = 0. For comparison purposes, the pullin voltages of the straight MLGNR electrostatic microactuators
( = ) are shown in Fig. 6 both in the absence and in the

In this paper, a novel method has been proposed for the


first time to obtain static pull-in voltages in curved multilayer
graphene/substrate microcantilever electrostatic actuators. The
multibeam shear model containing the interlayer shear rigidity
parameter was adopted to effectively capture the interlayer
shear effect. In the solution procedure, the governing fourthorder differential equation of variable coefficients was converted into a Fredholm integral equation. By adopting the first
natural mode of the microcantilever beams as a deflection shape
function, the resulting equation was solved for the static pullin voltages. The accuracy of the present analytical closed-form
solution was verified through comparing with the experimentally measured data existing in the literature for the electrostatic
microactuators with no GL. From the results, it was observed
that the pull-in voltage of the straight bilayer graphene/SiNx
electrostatic microactuators increased up to about 95%, 190%,
and 295% by attaching 10, 30, and 60 GLs, respectively, to
the substrate. It was also found that the pull-in voltages were
considerably affected by the slippage between neighboring GLs
and also between the graphene and the substrate. Thus, the
classical EulerBernoulli beam theory was not able to predict
the pull-in voltages of the multilayer graphene/substrate electrostatic microactuators.

ROKNI AND LU: EFFECT OF GRAPHENE LAYERS ON PULL-IN BEHAVIOR OF ELECTROSTATIC MICROACTUATORS

R EFERENCES
[1] P. M. Osterberg and S. D. Senturia, M-TEST: A test chip for MEMS
material property measurement using electrostatically actuated test structures, J. Microelectromech. Syst., vol. 6, no. 2, pp. 107118, Jun. 1997.
[2] H. Sadeghian, C. K. Yang, J. F. L. Goosen, E. van der Drift, A. Bossche,
P. J. French, and F. van Keulen, Characterizing size-dependent effective elastic modulus of silicon nanocantilevers using electrostatic pull-in
instability, Appl. Phys. Lett., vol. 94, no. 22, pp. 221903-1221903-3,
Jun. 2009.
[3] R. H. Poelma, H. Sadeghian, S. P. M. Noijen, J. J. M. Zaal, and
G. Q. Zhang, A numerical experimental approach for characterizing
the elastic properties of thin films: Application of nanocantilevers,
J. Micromech. Microeng., vol. 21, no. 6, p. 065003, Jun. 2011.
[4] C. OMahony, M. Hill, R. Duane, and A. Mathewson, Analysis of
electromechanical boundary effects on the pull-in of micromachined
fixedfixed beams, J. Micromech. Microeng., vol. 13, no. 4, pp. S75
S80, Jul. 2003.
[5] H. Sadeghian, G. Rezazadeh, and P. M. Osterberg, Application of the
generalized differential quadrature method to the study of pull-in phenomena of MEMS switches, J. Microelectromech. Syst., vol. 16, no. 6,
pp. 13341340, Dec. 2007.
[6] S. Chaterjee and G. Pohit, A large deflection model for the pull-in analysis of electrostatically actuated microcantilever beams, J. Sound Vib.,
vol. 322, no. 4/5, pp. 969986, May 2009.
[7] H. Sadeghian and G. Rezazadeh, Comparison of generalized differential quadrature and Galerkin methods for the analysis of microelectro-mechanical coupled systems, Commun. Nonlin. Sci. Numer.
Simul., vol. 14, no. 6, pp. 28072816, Jun. 2009.
[8] D. T. Haluzan, D. M. Klymyshyn, S. Achenbach, and M. Brner, Reducing pull-in voltage by adjusting gap shape in electrostatically actuated
cantilever and fixedfixed beams, Micromachines, vol. 1, no. 2, pp. 68
81, Jul. 2010.
[9] B. Wang, S. Zhou, J. Zhao, and X. Chen, Size-dependent pull-in instability of electrostatically actuated microbeam-based MEMS, J. Micromech.
Microeng., vol. 21, no. 2, pp. 027001-1027001-6, Feb. 2011.
[10] S. Pamidighantam, R. Puers, K. Baert, and H. A. C. Tilmans, Pull-in voltage analysis of electrostatically actuated beam structures with fixedfixed
and fixedfree end conditions, J. Micromech. Microeng., vol. 12, no. 4,
pp. 458464, Jul. 2002.
[11] Y. C. Hu, C. M. Chang, and S. C. Huang, Some design considerations on
the electrostatically actuated microstructures, Sens. Actuators A, Phys.,
vol. 112, no. 1, pp. 155161, Dec. 2004.
[12] S. Chowdhury, M. Ahmadi, and W. C. Miller, A closed-form model
for the pull-in voltage of electrostatically actuated cantilever beams,
J. Micromech. Microeng., vol. 15, no. 4, pp. 756763, Apr. 2005.
[13] S. Chowdhury, M. Ahmadi, and W. C. Miller, Pull-in voltage study of
electrostatically actuated fixedfixed beams using a VLSI on-chip interconnect capacitance model, J. Microelectromech. Syst., vol. 15, no. 3,
pp. 639651, Jun. 2006.
[14] M. M. Joglekar and D. N. Pawaskar, Closed-form empirical relations to
predict the static pull-in parameters of electrostatically actuated microcantilevers having linear width variation, Microsyst. Technol., vol. 17, no. 1,
pp. 3545, Jan. 2011.
[15] A. Ballestra, E. Brusa, M. G. Munteanu, and A. Som, Experimental characterization of electrostatically actuated in-plane bending of
microcantilevers, Microsyst. Technol., vol. 14, no. 7, pp. 909918,
May 2008.
[16] L. C. Wei, A. B. Mohammad, and N. M. Kassim, Analytical modeling
for determination of pull-in voltage for an electrostatic actuated MEMS
cantilever beam, in Proc. IEEE ISCE, Dec. 2002, pp. 233238.
[17] Y. C. Hu, Closed form solutions for the pull-in voltage of micro curled
beams subjected to electrostatic loads, J. Micromech. Microeng., vol. 16,
no. 3, pp. 648655, Mar. 2006.
[18] Y. C. Hu and C. S. Wei, An analytical model considering the fringing
fields for calculating the pull-in voltage of micro curled cantilever beams,
J. Micromech. Microeng., vol. 17, no. 1, pp. 6167, Jan. 2007.
[19] W. C. Chuang, Y. C. Hu, C. Y. Lee, W. P. Shih, and P. Z. Chang, Electromechanical behavior of the curled cantilever beam, J. Micro/Nanolith.
MEMS MOEMS, vol. 8, no. 3, pp. 033020-1033020-8, Jul. 2009.
[20] R. K. Gupta, Electrostatic pull-in test structures design for in-situ
mechanical property measurement of microelectromechanical systems
(MEMS), Ph.D. dissertation, Massachusetts Inst. Technol., Cambridge,
MA, 1997.
[21] C. Lee, X. Wei, J. W. Kysar, and J. Hone, Measurement of the elastic properties and intrinsic strength of monolayer graphene, Science,
vol. 321, no. 5887, pp. 385388, Jul. 2008.

559

[22] B. Ashrafi, Theoretical and experimental investigations of the elastic properties of carbon nanotube-reinforced polymer thin films,
Ph.D. dissertation, McGill Univ., Sainte-Anne-de-Bellevue, QC, Canada,
2008.
[23] H. Conley, N. V. Lavrik, D. Prasai, and K. I. Bolotin, Graphene
bimetallic-like cantilevers: Probing graphene/substrate interactions,
Nano Lett., vol. 11, no. 11, pp. 47484752, Nov. 2011.
[24] Y. Liu, Z. Xu, and Q. Zheng, The interlayer shear effect on graphene
multilayer resonators, J. Mech. Phys. Solids, vol. 59, no. 8, pp. 1613
1622, Aug. 2011.
[25] S. M. C. Abdulla, H. Yagubizade, and G. J. M. Krijnen, Analysis
of resonance frequency and pull-in voltages of curled micro-bimorph
cantilevers, J. Micromech. Microeng., vol. 22, no. 3, p. 035014,
Mar. 2012.
[26] S. Iwamoto, S. Ishida, Y. Arakawa, M. Tokushima, A. Gomyo,
H. Yamada, A. Higo, H. Toshiyoshi, and H. Fujita, Observation of
micromechanically controlled tuning of photonic crystal line-defect
waveguide, Appl. Phys. Lett., vol. 88, no. 1, pp. 011104-1011104-3,
Jan. 2006.
[27] M. R. Begley, The impact of materials selection and geometry on multifunctional bilayer micro-sensors and actuators, J. Micromech. Microeng.,
vol. 15, no. 12, pp. 23792388, Dec. 2005.
[28] H. Rokni, R. J. Seethaler, A. S. Milani, S. Hosseini-Hashemi, and
X. F. Li, Exact closed-form solutions for size-dependent static
pull-in behavior in electrostatic microactuators via Fredholm integral
method, Sens. Actuators A, Phys., vol. 190, no. 1, pp. 3243, Feb. 2013.
[29] L. Meirovitch, Fundamentals of Vibrations. New York: McGraw-Hill,
2001.
[30] G. N. Nielson and G. Barbastathis, Dynamic pull-in of parallel-plate
and torsional electrostatic MEMS actuators, J. Microelectromech. Syst.,
vol. 15, no. 4, pp. 811821, Aug. 2006.

Hossein Rokni received the B.S. degree in mechanical engineering from the Iran University of Science
and Technology, Tehran, Iran, in 2005 and the M.S.
degree in mechanical engineering from The University of British Columbia, Kelowna, BC, Canada, in
2011. He is currently working toward the Ph.D. degree in the Laboratory of Nanostructures, University
of Michigan, Ann Arbor.
His research interests include the design, modeling, and fabrication of MEMS/NEMS devices with
particular interest in electrostatic graphene-based
MEMS/NEMS sensors and actuators.

Wei Lu received the B.S. degree from Tsinghua University, Beijing, China, and the Ph.D. degree from
Princeton University, Princeton, NJ.
He is currently an Associate Professor with the
Department of Mechanical Engineering, University
of Michigan, Ann Arbor, where he is also currently a
Codirector of the research center GM/UM Advanced
Battery Coalition for Drivetrains. His research interests include self-assembled nanostructures and their
applications in materials, mechanics, energy and biological systems, mechanics of nano-/microstructures,
biomechanics, and innovative diagnostics for global health. He has over
70 publications in peer-reviewed scientific journals and has given over
100 presentations and invited talks at international conferences, universities,
and national laboratories. He also has numerous publications in conference
proceedings, encyclopedias, and book chapters. He serves as an Associate
Editor for the Journal of Computational and Theoretical Nanoscience and on
the Editorial Boards of several other journals.
Prof. Lu is a Fellow of the American Society of Mechanical Engineers.
He has been the recipient of many awards, including the Faculty Recognition
Award, the U.S. Air Force Summer Faculty Fellowship, the Robert M. Caddell
Memorial Research Achievement Award, the CAREER Award from the U.S.
National Science Foundation, and the Robert J. McGrattan Award from the
American Society of Mechanical Engineers.

S-ar putea să vă placă și