You are on page 1of 10

Zhi-Bin Shen

e-mail: zb_shen@yeah.net

Bin Deng
e-mail: dbhj0@sina.com
College of Aerospace and Material Engineering,
National University of Defense Technology,
Changsha, Hunan 410073, China

Xian-Fang Li
School of Civil Engineering,
Central South University,
Changsha, Hunan 410075, China
e-mail: xfli00@hotmail.com

Guo-Jin Tang1
College of Aerospace and Material Engineering,
National University of Defense Technology,
Changsha, Hunan 410073, China
e-mail: tanggj@nudt.edu.cn

Vibration of Double-Walled
Carbon Nanotube-Based Mass
Sensor via Nonlocal Timoshenko
Beam Theory
The potential of double-walled carbon nanotubes (DWCNTs) as a micromass sensor is
explored. A nonlocal Timoshenko beam carrying a micromass at the free end of the inner
tube is used to analyze the vibration of DWCNT-based mass sensor. The length of the
outer tube is not equal to that of the inner tube, and the interaction between two tubes is
governed by van der Waals force (vdW). Using the transfer function method, the natural
frequencies of a nonlocal cantilever with a tip mass are computed. The effects of the
attached mass and the outer-to-inner tube length ratio on the natural frequencies are discussed. When the nonlocal parameter is neglected, the frequencies reduce to the classical
results, in agreement with those using the finite element method. The obtained results
show that increasing the attached micromass decreases the natural frequency but
increases frequency shift. The mass sensitivity improves for short DWCNTs used in mass
sensor. The nonlocal Timoshenko beam model is more adequate than the nonlocal EulerBernoulli beam model for short DWCNT sensors. Obtained results are helpful to the
design of DWCNT-based resonator as micromass sensor. [DOI: 10.1115/1.4005489]
Keywords: DWCNT, Timoshenko beam theory, mass sensor, nonlocal elasticity, transfer
function method, free vibration

Introduction

Since the discovery of carbon nanotubes (CNTs) by Iijima [1]


in 1991, they have demonstrated a significant potential for use in a
diverse range of new and evolving applications [2,3]. In particular, the experimental evidence shows that CNTs have low weight,
high aspect ratio, extremely high stiffness and are highly sensitive
to their environment change [4]. Such features of CNTs make
them promising candidates for atomic-resolution mass sensor [5].
Poncharal et al. [6] first proposed the idea of using individual
CNTs as high sensitivity nanobalances in 1999. In general, the
principle of mass detection using CNT-based sensor from a vibration analysis is based on the fact that the resonant frequency is
sensitive to the attached mass [7]. Many researchers have
explored the potential of using single-walled carbon nanotubes
(SWCNTs) as nanomechanical resonators in atomic scale [811].
Multiwalled carbon nanotubes (MWCNTs) resonators have different mechanical structures than SWCNT ones, due to the interaction between intertubes such as the van der Waals (vdW) force.
Mateiu et al. [12,13] proposed an approach for building a mass
sensor based on MWCNTs. Elishakoff et al. [14] investigated
vibration of double-walled carbon nanotubes (DWCNTs) cantilever with attached bacterium on the tip. Recently, synthesis technique has made DWCNTs with different wall lengths possible be
manufactured [15]. Kang et al. [16,17] examined frequency
change of ultrahigh frequency nanomechanical resonators based
on DWCNTs with different wall lengths using molecular dynamics simulations.
Continuum models have been widely used to study the vibrational behavior of CNT-based mass sensor to avoid the difficulties
encountered during the experimental characterization of nanotubes
as well as the time-consuming nature of computational atomistic
simulations. For example, the classical Euler-Bernoulli beam
1
Corresponding author.
Manuscript received July 14, 2011; final manuscript received August 25, 2011;
published online January 9, 2012. Editor: Boris Khusid.

theory (EBT) was utilized to model a nanomechanical resonator


[9,18,19]. However, the effects of shear deformation and rotary
inertia are neglected in these analyses. When these two factors are
taken into account simultaneously, the Timoshenko beam theory
(TBT) Ref. [20] is necessary and also more effective. Some theoretical analyses involving wave propagation and free vibration of
CNTs have also been presented [2123] using the TBT.
In addition, owing to the fact that the classical continuum
approach is scale-independent and then cannot describe the size
effects arising from small scale. Although the classical continuum
theory is sometimes applied to analyze the mechanical behavior
of nanostructures such as CNTs, it is found to be inadequate
because of ignoring the size effects. To improve this situation, the
nonlocal elasticity theory presented by Eringen [24,25] has been
developed to tackle scale-dependent problems. Along this line,
Lee et al. [26] used the nonlocal EBT to analyze the frequency
shift of CNT-based mass sensors. As mentioned above, the EBT
neglects the influences of shear deformation and rotary inertia. To
consider these two factors, application of the nonlocal TBT in
analyzing the frequency shift of CNT-based mass sensors is a prerequisite. To the best of the authors knowledge, this topic has not
been studied, in particular for the vibration of DWCNT based
mass sensor with inner and outer nanotubes of different lengths.
Such DWCNTs with different wall lengths have been reported to
have some specific practical applications [27].
In the present study, the dynamic behavior of a DWCNT based
mass sensor with different wall lengths is studied. The DWCNT is
modeled as two Timoshenko cantilevers and their interaction is
described by the vdW force, and the size effects are described by
the nonlocal elasticity theory. For the case of a micromass at the
free end of the inner tube, the transfer function method (TFM)
Ref. [28] is used to determine the natural frequencies of the
micromass sensor. The effects of the attached tip mass, nonlocal
parameter, length ratio, DWCNT length, and rotary inertia on the
fundamental frequencies are analyzed. The natural frequencies of
DWCNT-based mass sensors when using the nonlocal TBT are
compared with those when using nonlocal EBT and the classical

Journal of Nanotechnology in Engineering and Medicine


C 2011 by ASME
Copyright V

AUGUST 2011, Vol. 2 / 031003-1

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

TBT. Our analysis shows novel characteristics of the fundamental


frequency shift.

Basic Equations

2.1 Nonlocal TBT. As in the classical TBT, there are only


two independent variables, the transverse displacement w and the
rotation of cross-section h, both of which depend on the longitudinal coordinate x and time t. According to the Hamiltons principle
and Eringens nonlocal elasticity theory [24,25], governing equations of the nonlocal TBT can be expressed as [21]




@2 @2w
@h @ 2 w


jGA
qA 1  e0 a2 2
2
@x @x2
@x @t


2
@
(1)
1  e0 a2 2 px; t
@x




@2 @2h
@w
@2h
qI 1  e0 a2 2

EI

jGA
h

0 (2)
@x
@x2
@x @t2
where q is the mass density, A and I are the cross-sectional area
and its second moment, respectively, px; t is the distributed force
per unit length, E and G are Youngs modulus and shear modulus,
respectively. j is the shear correction factor, which is introduced
to account for the relaxation of the inconsistency of the usual
shear-free boundary condition at the beam surface and to compensate for the error in assuming a constant shear stress over the
whole cross-section of the beam. It depends on the material and
geometric parameters of the beam. e0 a is a small scale parameter
with length unit which can be used to describe the size effects, e0
being a nondimensional material constant that can be determined
by experiments or numerical simulations from molecular dynamics simulation [29] and a being the CC bond length. Here, the
nonlocal effects are assumed to be present for both normal and
shear stresses. When e0 a 0, the nonlocal elasticity reduces to
the classical TBT.
In addition, the nonlocal bending moment M and the nonlocal
shearing force Q can be expressed below, respectively, as


@h
@2w
@3h

px;
t
(3)
Mx; t EI e0 a2 qA 2 qI
@x
@t
@x@t2




@w
@3w
@px; t
Qx; t jGA h
e0 a2 qA
(4)

@x
@x@t2
@x
For a DWCNT, two nanotubes are interacted through the vdW
force, which for simplicity can be described by the following linear relation [23]
px; t cw1 x; t  w2 x; t

(5)

where c is the vdW interaction coefficient between the inner and


outer nanotubes. Here, we take [22]
320  2R1 erg=cm2
;
c
0:16a2





@ 2 @ 2 w1
@h1 @ 2 w1

jGA

qA1 1  e0 a2 2
1
@x  @t2
@x
@x2

2
@
1  e0 a2 2 cw2  w1 0  x  L2
@x




@ 2 @ 2 h1
@w1

jGA
h

qI1 1  e0 a2 2
1
1
@x @t2
@x
2
@ h1
 EI1 2 00  x  L2
@x




@ 2 @ 2 w2
@h2 @ 2 w2
qA2 1  e0 a2 2

jGA

2
2
@x @t
@x
@x2


2
2 @
cw2  w1 0  x  L2
 1  e0 a
@x2




@ 2 @ 2 h2
@w2
qI2 1  e0 a2 2

jGA
h

2
2
@x @t2
@x
2
@ h2
 EI2 2 00  x  L2
@x




@ 2 @ 2 w3
@h3 @ 2 w3
qA3 1  e0 a2 2
0

jGA

3
@x @t2
@x
@x2

a 0:142 nm

(7)

(8)

(9)

(10)

L2 < x  L1
(11)



@ 2 @ 2 h3
@w3
@ 2 h3
jGA3 h3
 EI3 2 0
qI3 1  e0 a2 2
@x @t2
@x
@x


L2 < x  L1

(12)

where the subscripts 1 and 2 denote the quantities associated with the
inner and outer nanotubes, respectively. It is particularly noted that
for convenience we have introduced subscript 3 that specifies the
corresponding quantities of the inner tube lying in L2 < x  L1 ,
viz. w3 w1 ; h3 h1 , I1 I3 ; A1 A3 as L2 < x  L1 . This is
equivalent to say that for 0  x  L2 , the DWCNT is essentially
composed of two tubes, while L2 < x  L1 it is in fact a single tube.
For a cantilever DWCNT carrying a micro mass where micro
mass specifies those such as a buckyball, a molecular, a bacterium
or a virus, etc., at the free end, the corresponding boundary conditions can be stated as
wj 0; t 0;

hj 0; t 0 j 1; 2

M2 L2 ; t 0; Q2 L2 ; t 0; M3 L1 ; t 0
(6)

where R1 is the inner radius of the DWCNT. For the present


DWCNT with R1 0.35 nm, Eq. (6) gives c 6:943  1011 erg=cm3
69:43 GPa.
Now, let us consider transverse vibration of a DWCNT carrying
an atomic-resolution mass at the free tip, as shown in Fig. 1. Without loss of generality, we assume that the inner and outer nanotubes have different lengths, Lj and dj representing the length and
diameter of the inner ( j 1) and outer ( j 2) tubes, respectively,
and the atomic-resolution mass is attached to the free end of the
inner tube. In the following, we consider the case where the inner
tube is longer than the outer tube, as shown in Fig. 1, the governing equations for the DWCNT carrying an atomic-resolution mass
read
031003-2 / Vol. 2, AUGUST 2011

Fig. 1 A cantilever DWCNT-based mass sensor with inner and


outer nanotubes of different lengths

Q3 L1 ; t m0

@ 2 w3 L1 ; t
0
@t2

(13)
(14)
(15)

where m0 is the attached mass at the free tip of the inner tube. In
addition, at the position x L2, the continuity conditions must be
satisfied, namely
w1 L2 w3 L2 ;
M1 L2 ; t M3 L2 ; t;

h1 L2 h3 L2

(16)

Q1 L2 ; t Q3 L2 ; t

(17)

Furthermore, initial conditions can be assumed as


wj x; 0

@wj x; 0
0 j 1; 2; 3
@t

(18)

Transactions of the ASME

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

hj x; 0

@hj x; 0
0; j 1; 2; 3
@t

(19)

2.2 Nonlocal EBT. The governing equations for the nonlocal


EBTcan also
be analytically derived from the above. Eliminating
@2 w
jGA @h
@x @x2 from Eqs. (1) and (2) leads to

 3

 2
2
2
@ h
@ w
@3h
2 @
2 @
qA 1  e0 a
 EI 3
qI 1  e0 a
2
2
2
2
@x
@x @x@t
@x @t


2
@
(20)
1  e0 a2 2 px
@x
In the Euler-Bernoulli hypothesis, the effects of shear deformation
and rotary inertia of the beam are neglected. The Euler-Bernoulli
hypothesis is actually equivalent to sufficiently large shear modulus
or shear rigidity, which impedes shear deformation in the beam.
Consequently, in the above formulation, let jGA=EI ! 1 and
neglect the influence of the rotary inertia, meaning qI@ 2 h=@t2 0.
Equation (2) reduces to
h

@w
@x

(21)

Bearing qI@ 2 h=@t2 0 in mind, we substitute Eq. (21) into


Eq. (20) and find that the governing equation of nonlocal EBT




2
@4w
@2 @2w
2 @
px (22)

1

e
a
EI 4 qA 1  e0 a2 2
0
@x
@x @t2
@x2
is recovered. In this case, instead of Eqs. (3) and (4), the nonlocal
bending moment M and the nonlocal shearing force Q are,
respectively,


@2w
@2w
Mx; t EI 2 e0 a2 qA 2  px
(23)
@x
@t


 t
@ Mx;
@3w
@3w
@px
Qx; t
EI 3 e0 a2 qA

@x
@x
@x@t2
@x
(24)
Thus, with the only deflection as an unknown in the nonlocal
EBT, the transverse vibration of the DWCNT with different tube
lengths (Fig. 1) are governed by the following coupled equations:

 2
2
@ 4 w1
@ w1
2 @

qA
1

e
a
EI1
1
0
4
2
@x
@x
@t2


2
2 @
cw2  w1
1  e0 a
(25)
@x2

 2
2
@ 4 w2
@ w2
2 @
EI2

qA
1

e
a
2
0
2
@x4
@x
@t2

2
@
 1  e0 a2 2 cw2  w1
(26)
@x

 2
2
@ 4 w3
@ w3
2 @
EI3

qA
1

e
a
0
(27)
3
0
@x4
@x2 @t2
The boundary conditions in Eqs. (13)(15) as well as the continuity conditions at the position x L2 in Eqs. (16) and (17) and initial conditions in Eqs. (18) and (19) are the same, where M and Q
in those equations should be replaced by Eqs. (23) and (24).

example, Adhikari et al. [30] presented a closed form solution for


a beam with nonlocal damping using the TFM for the distributed
parameter system. In this section, the TFM is employed to investigate the free vibration of the DWCNT-based mass sensor.
With the aid of the initial conditions, we can perform Laplace
transform for the governing Eqs. (7)(12). Furthermore, by introducing the following nondimensional parameters:
X

x
;
L1

qL21 s2
;
E

Wi

wi
;
L1

ai

Ii
;
Ai L21

cL21
;
kGAi

wi

E
;
jG

e0 a
i 1; 2; 3
L1

the governing Eqs. (7)(12) may be rewritten as


~1
~2
@2W
@2W
 w1 k 2
2
@X
@X2
~1
@
h
~1  w1 W
~2 
Cb w1 W
@X
2 ~
~1
@ W2
@2W
 w2 k 2
1 Cbk2 w2 k2
2
@X
@X2
~2
@
h
~2  w2 W
~1 
Cb w2 W
@X
~3
@2W
Cb
1
@ h~3
~3 

W
2
2 @X
@X2
1 Cbk
1 Cbk
~i
@ 2 h~i
1
@W
1 ai bC ~

hi i 1; 2; 3

@X2 ai b1 Ck2 @X ai b1 Ck2


1 Cbk2 w1 k2

(28)

(29)
(30)
(31)

Combining Eq. (28) with Eq. (29), @@XW21 and @@XW22 may be expressed
as




~1
@2W
Cb
w1 ~
1
w1 k2 @ h~1


W1 
@X2
D
@X
1 Cbk2 D
1 Cbk2
2 ~
w1 ~
w1 k @ h2
 W2 
(32)
D
D @X




~2
@2W
Cb
w ~
1
w k2 @ h~2

2 W
 2
2
2
2
2
@X
D
D
@X
1 Cbk
1 Cbk
w2 ~
w2 k2 @ h~1
 W1 
(33)
D
D @X
where D 1 Cbk2 1 Cbk2 w1 k2 w2 k2 .
Next, if we introduce a state vector as
"
#T
~
~
~
~1 X;s; @ W1 X;s ; h~1 X;s; @ h1 X;s ;:::; @ h3 X;s
gX;s W
@X
@X
@X

121

where the superscript T denotes matrix transpose, Eqs. (30)(33)


can be rewritten in a compact form in state space
dgX; s
UsgX; s gX; s
dX

(34)

where Us Uij 1212 is a 12  12 matrix with variable s, the


details of which are given in the Appendix A, and gX; s is related
to distributed loading and vanishes in the present study.
Next, after performing the Laplace transform to two sides of
the boundary conditions (13)(15) and the continuity conditions
(16) and (17), with the vanishing initial conditions, we get

3.1 Nonlocal TBT. In solving free vibration of Timoshenko


beams, as a powerful semi-analytical and seminumerical method,
the TFM is frequently applied to treat relevant problems. For

~j 0; s 0; h~j 0; s 0 j 1; 2
(35)
W

2
2

~
w2 k ~
k
w ~
2 @ h2 d; s
0
C 2 W
W1 d; s
2 d; s 1 Ck
@X
a2 b
a2
b
(36)

Journal of Nanotechnology in Engineering and Medicine

AUGUST 2011, Vol. 2 / 031003-3

Solution Method

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

~1 d; s
~2 d; s
@W
@W
 1 Cbk2 w2 k2
h~2 d; s 0
@X
@X
(37)
~
C 2~
@ h3 1; s
0
(38)
k W3 1; s 1 Ck2
a3
@X
~
~3 1; s 1 Cbk2 @ W3 1; s h~3 1; s 0
mbCW
(39)
@X
~1 d; s W
~3 d; s;
W
h~1 d; s h~3 d; s
(40)


2
2
~


k
w ~
w k ~
2 @ h1 d; s
 1 W
C 1 W
1 d; s 1 Ck
2 d; s
@X
a1
b
a1 b
Ck2 ~
@ h~3 d; s

0
(41)
W3 d; s  1 Ck2
@X
a1
~1 d; s
~2 d; s
@W
@W
1 bCk2 w1 k2
h~1 d; s  w1 k2
@X
@X
~
2 @ W3 d; s
~
 1 bCk
 h3 d; s 0
(42)
@X

x
X

f
2p 2p

 w2 k 2

where m m0 =qAL1 is the ratio of the attached micromass to the


inner tube mass, and d L2 =L1 is the outer-to-inner tube length
ratio, respectively. Furthermore, using the introduced vector g,
Eqs. (35)(42) become
M b g0; s N b g1; s Rc sgd; s cs

(43)

where Mb Mi;j 1212 , Nb Ni;j 1212 and Rc Ri;j 1212 are


12  12 matrixes, the details of them are given in the Appendix
B. cs in Eq. (43) is a vector and cs 0 in the present study.
According to the TFM, the solution of Eq. (34) can be
expressed as follows [31]:
h
i
M b N b eUs Rc seUsd gX; s
L
(44)
GX; n; sgn; sdn eUsX cs
0

where

GX; n; s

Mb eUsn
Xn
N b eUs1n X < n

For the free vibration of the DWCNT-based nanomechanical sensor, keeping gX; s 0 and cs 0 in mind, meaning that the
right-hand side of Eq. (44) clearly vanishes, the existence of a
nontrivial solution of the corresponding homogenous Eq. (44)
allows us to get that the determinant of the coefficient matrix must
vanish
h
i
(45)
det Mb N b eUs Rc seUsd 0
This is the characteristic equation we want to look for.
Considering the nondimensional natural frequency
r
p
qL2 s2
X C  1
E

(46)

the circular frequency x of free vibration is related by the following relation:


s
E
(47)
x Ims X
qL21
where Ims denotes the imaginary part of a complex s. After getting the solution of Eq. (45), the natural frequencies
031003-4 / Vol. 2, AUGUST 2011

s
E
qL21

(48)

can be directly obtained from Eq. (47). Also, the corresponding


modal shapes can be evaluated by
gx; sk eUsk X fk

(49)

where sk denotes the imaginary frequency of the nanomechanical


sensor for mode k, and fk is a nontrivial vector obtained by substituting sk into the corresponding homogeneous Eq. (44).
3.2 Nonlocal EBT. In a similar manner, the nondimensional
governing equations, boundary conditions and continuity conditions for the DWCNT sensor based on the nonlocal EBT can be
expressed as



~1
Ck2 w1 k2 @ 2 W
C
w1 ~

W1
a1 a1 b
a1
a1 b @X2
~ 2 w1
w k2 @ 2 W
~2
 1

W
a1 b @X2
a1 b
 2



~2
~2
@4W
Ck
w1 k 2 @ 2 W
C
w2 ~

W2
a2 a2 b
@X4
a2
a2 b @X2
~ 1 w2
w k2 @ 2 W
~1
 2

W
a2 b @X2
a2 b
~1
@4W

@X4

(50)

(51)

~3 Ck2 @ 2 W
~3 C
@4W
~3

 W
(52)
4
a1
@X
a3 @X2
~
~j 0; s 0; @ Wj 0; s 0 j 1; 2
W
(53)
@X
~3 1; s
C 2~
@2W
k W3 1; s 
0
(54)
a1
@X2
~3 1; s @ 3 W
~3 1; s
mC ~
Ck2 @ W

0
(55)
W3 1; s
a3
@X
@X3
a3
 2

~2 d; s
w2 k2 ~
Ck
w k2 ~
@2W
2
0
W1 d; s
W2 d; s 
@X2
a2 b
a2
a2 b
(56)
 2

2
2
3
~1 d; s
~2 d; s @ W
~2 d; s
w k @W
Ck
w k @W
 2


2
0
@X
@X
@X3
a2 b
a2
a2 b
(57)
~
~
~1 d; s W
~3 d; s; @ W1 d; s @ W3 d; s
W
(58)
@X
@X
 2

~1 d; s w1 k2
Ck
w k2 ~
@2W
~2 d; s
1

W
W1 d; s 
@X2
a1
a1 b
a1 b
~3 d; s
Ck2 ~
@2W

0
(59)
W3 d; s
@X2
a1
 2

~1 d; s @ 3 W
~2 d; s
~1 d; s w1 k2 @ W
Ck
w k2 @ W

1

@X
@X3
@X
a1
a1 b
a1 b
2
3
~3 d; s @ W
~3 d; s
Ck @ W


0
(60)
@X
@X3
a1
A completely analogous procedure can deal with the following
equations using the TFM with the introduction of the following
state vector:


2 ~
3 ~
3 ~ T
~
~1 ; @ W1 ; @ W1 ; @ W1 ; :::; @ W3
:
gX; s W
@X @X2 @X3
@X3 121
Transactions of the ASME

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

Table 1 Comparison of the natural frequencies in GHz obtained using the TFM with those using
the FEM software for SWCNT sensor with L1 5 14 nm for different attached masses
m

L2/L1

0.0

0.0
0.5
1.0

1.0

0.0
0.5
1.0

TFM
FEM
%error
TFM
FEM
%error
TFM
FEM
%error
TFM
FEM
%error
TFM
FEM
%error
TFM
FEM
%error

f1

f2

f3

f4

f5

15.9730
15.8037
1.0599
32.8253
33.9794
3.5159
26.0066
25.6823
1.2470
7.0853
6.9354
2.1156
13.0307
13.3516
2.4626
16.9716
16.6581
1.8472

98.0645
96.8190
1.2701
110.5275
110.4688
0.0531
153.4298
149.6798
2.4441
72.7607
71.8114
1.3047
92.0735
90.7892
1.3949
120.5118
119.4134
0.9114

266.1967
262.1844
1.5073
292.1970
304.7367
4.2915
394.6100
381.7038
3.2706
221.4480
218.2323
1.4521
232.3664
243.4285
4.7606
314.2986
324.6293
3.2869

500.4702
491.5864
1.7751
545.5307
528.3862
3.1427
695.4094
671.3197
3.4641
440.1099
432.6588
1.6930
505.1855
490.4210
2.9226
532.4463
586.6382
10.1779

788.5085
772.3475
2.0496
765.8425
820.3664
7.1195
1024.0282
994.6709
2.8668
715.6924
701.6287
1.9650
683.4943
722.3286
5.6817
782.2438
877.5734
12.1867

 i;j 

Note that in this case, Us U
1212 ; M b Mi;j 1212 ; N b


Ni;j 1212 ; Rc Ri;j 1212 are still 12  12 matrixes, the details
of which are given in the Appendix C.

Results and Discussion

In this section, the effects of the nonlocal parameter, tip micromass, and outer-to-inner tube length ratio on the natural frequencies of the cantilever DWCNT are analyzed. Consider a DWCNT
with inner diameter d1 2R1 0.7 nm and outer diameter d2
2R2 1.4 nm, where R1 is the radius of the inner tube centerline, while R2 is the radius of the outer tube centerline. It is
assumed that the inner and outer tubes have the same Youngs
modulus E 1 TPa, shear modulus G 0.4 Tpa (with Poisson
ratio  0:25), and the effective thickness of SWCNT t 0.3 nm.
In accordance with the definition of the effective thickness
and the Youngs modulus mentioned above, a mass density
q 2:3g=cm3 is adopted. For a CNT with hollow circular
cross-section, the shear correction factor j is taken as 0.8 [32].

Fig. 2 A FEM model for cantilever DWCNT-based mass sensor


with different tube lengths

whether a tip micromass is attached or not. The fundamental frequencies are seen to be very sensitive to the change in the attached
tip mass, in particular for larger tip mass. This is the basic principle of the DWCNT as micromass detection [7].

4.1 Result Validation. Prior to the presentation of numerical


results of the natural frequencies, let us examine the validity and
accuracy of the method suggested here. The DWCNT-based
micromass sensor is modeled as a microcantilever with a concentrated mass at the free end of the inner tube. The natural frequencies of the DWCNT-based micromass sensor can be computed
and numerical results of a DWCNT with L1 14 nm for different
values of the tip micromass and the outer-to-inner tube length ratio are listed in Table 1 when the nonlocal parameter e0a/L vanishes, i.e., neglecting the size effects. For comparison, we also
tabulate the corresponding numerical results in Table 1 using a
commercial FEM software MSC.NASTRAN. Note that when adopting
MSC.NASTRAN, the vdW force between two tubes is modeled with
linear spring, and the concentrated mass is joined with the inner
nanotube by interpolation constraint element RBE3. The diagram
using 3D solid finite element model for a DWCNT with inner and
outer tubes of different lengths and carrying a micromass at the
inner tube tip is shown in Fig. 2. From Table 1, it is seen that the
theoretical results and the FEM simulation results are in good
agreement. For the fundamental frequencies, the maximum relative error is 3.51%, occurring in the absence of the tip mass with
L2/L1 0.5. This comparison provides high confidence for use of
the TBT model in further investigation as a mass sensor. Obtained
results indicate that with the tip mass rising, the corresponding
frequencies decrease. Moreover, as the ratio L2/L1 varies, the natural frequency of the DWCNT is strongly affected, no matter

4.2 Effect of Tip Mass on the SWCNT Based Micromass


Sensor. Due to the significance of the fundamental frequencies of
DWCNT based mass sensor, in what follows we focus our attention
on the fundamental frequency shift induced by an attached micromass. For this purpose, we denote the frequency shift Df as the difference between the fundamental frequencies of the cantilever
DWCNT with and without tip micromass, which serves as an index
to assess quantitatively the mechanical behavior of the mass sensor.
Figure 3 shows the fundamental frequency shift of a cantilever
DWCNT as a function of the attached mass with different outer
tube lengths and L1 14 nm. As pointed before, an increase in the
attached tip mass decreases the fundamental frequencies. However, from Fig. 3, the frequency shift is viewed to rise as the tip
mass is raised, which agrees qualitatively with the results
described by Lee et al. [26] and Li et al. [7]. Moreover, the variations of frequency shift are apparent when the attached mass is
larger than 1021 g. In other words, the mass sensitivity of this
nanomechanical sensor can reach at least 1021 g, which has the
same order as mentioned in Refs. [7] and [33]. Furthermore, the
effect of nonlocal parameter on the frequency shift is significant.
Note that the impact of the small scale on the frequency shift is
different depending on different length ratios. For example, for
L2/L1 0.5, the nonlocal parameter makes the frequency shift
become smaller. This trend is, however, reversed for L2/L1 1.0.
From Fig. 3, it is clear that there is a critical mass mc related to the
nonlocal parameter e0a and the length of DWCNT. The frequency
shift of the DWCNT-based mass sensor is more sensitive for the
tip micromass m0 less than mc if L2/L1 0.5, and for m0 larger
than mc if L2/L1 1.0.

Journal of Nanotechnology in Engineering and Medicine

AUGUST 2011, Vol. 2 / 031003-5

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

Fig. 3 The fundamental frequency shift of a DWCNT sensor versus attached mass with
L1 5 14 nm

In addition, the length of the DWCNT, L1, can influence the


changes in the fundamental frequency shift of the DWCNT-based
mass sensor. Figure 4 illustrates the effect of length ratio, L2/L1,
for two different DWCNT lengths, L1 7 nm and 14 nm, on the

frequency shift of the mass sensor. It is seen that the frequency


shift is significantly affected by the length ratio. The frequency
shift becomes larger when the length of DWCNT is shorter, especially for larger attached mass. Thus, the mass sensitivity

Fig. 4 The effect of length ratio and DWCNT length on the fundamental frequency shift of the
SWCNT sensor versus attached mass with e0a/L 5 0.3

031003-6 / Vol. 2, AUGUST 2011

Transactions of the ASME

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

[3] Kim, P., and Lieber, C. M., 1999, Nanotube Nanotweezers, Science, 286, pp.
21482150.
[4] Wong, E. W., Sheehan, P. E., and Lieber, C. M., 1997, Nanobeam Mechanics:
Elasticity, Strength, and Toughness of Nanorods and Nanotubes, Science, 277,
pp. 19711975.
[5] Jensen, K., Kim, K., and Zettl, A., 2008, An Atomic-Resolution Nanomechanical Mass Sensor, Nat. Nanotechnol., 3, pp. 533537.
[6] Poncharal, P., Wang, Z. L., Ugarte, D., and Heer, W. A. D., 1999, Electrostatic
Deflections and Electro-Mechanical Resonances of Carbon Nanotubes, Science, 283, pp. 15131516.
[7] Li, C. Y., and Chou, T. W., 2004, Mass Detection Using Carbon NanotubeBased Nanomechanical Resonators, Appl. Phys. Lett. 84, pp. 52465248.
[8] Joshi, A. Y., Sharma, S. C., and Harsha, S. P., 2010, Dynamic Analysis of a
Clamped Wavy Single Walled Carbon Nanotube Based Nanomechanical
Sensors, J. Nanotechnol. Eng. Med., 1, p. 031007.
[9] Mehdipour, I., Barari, A., and Domairry, G., 2011, Application of a Cantilevered SWCNT With Mass at the Tip as a Nanomechanical Sensor, Comput.
Mater. Sci., 50, pp. 18301833.
[10] Wu, D. H., Chien, W. T., Chen, C. S., and Chen, H. H., 2006, Resonant Frequency Analysis of Fixed-Free Single-Walled Carbon Nanotube-Based Mass
Sensor, Sens. Actuators, A, 126, pp. 117121.
[11] Joshi, A. Y., Sharma, S. C., and Harsha, S. P., 2011, Zeptogram Scale Mass
Sensing Using Single Walled Carbon Nanotube Based Biosensors, Sens.
Actuators, A, 168, pp. 275280.
[12] Mateiu, R., Davis, Z. J., Madsen, D. N., Mlhave, K., Bggild, P., Rassmusen,
A. M., Brorson, M., Jacobsen, J. H. C., and Boisen, A., 2004, An Approach to
a Multi-Walled Carbon Nanotube Based Mass Sensor, Microelectron. Eng.,
73, pp. 670674.
[13] Mateiu, R., Kuhle, A., Marie, R., and Boisen, A., 2005, Building a MultiWalled Carbon Nanotube-Based Mass Sensor With the Atomic Force Microscope, Ultramicroscopy, 105, pp. 233237.
[14] Elishakoff, I., Versaci, C., and Muscolino, G., 2011, Clamped-Free DoubleWalled Carbon Nanotube-Based Mass Sensor, Acta Mech., 219(5), pp. 2943.
[15] Collins, P. G., Hersam, M., Arnold, M., Martel, R., and Avouris, P., 2001,
Current Saturation and Electrical Breakdown in Multiwalled Carbon Nanotubes, Phys. Rev. Lett., 86, pp. 31283131.
[16] Kang, J. W., Kwon, O. K., Hwang, H. J., and Jiang, Q., 2011, Resonance Frequency Distribution of Cantilevered (5,5)(10,10) Double-Walled Carbon Nanotube With Different Intertube Lengths, Mol. Simul., 37(1), pp. 1822.
[17] Kang, J. W., Kwon, O. K., Lee, J. H., Choi, Y. G., and Hwang, H. J., 2009,
Frequency Change by inter-Walled Length Difference of Double-Wall Carbon
Nanotube Resonator, Solid State Commun., 149(1), pp. 15741577.

031003-10 / Vol. 2, AUGUST 2011

[18] Elishakoff, I., and Pentaras, D., 2009, Fundamental Natural Frequencies of
Double-Walled Carbon Nanotubes, J. Sound Vib., 322, pp. 652664.
[19] Elishakoff, I., Versaci, C., Maugeri, N., and Muscolino, G., 2011, ClampedFree Single-Walled carbon Nanotube-Based Mass Sensor Treated as BernoulliEuler Beam, ASME J. Nanotechnol. Eng. Med., 2, p. 021001.
[20] Timoshenko, S., 1921, On the Correction for Shear of the Differential Equation for Transverse Vibrations of Prismatic Bars, Philos. Mag., 41, pp.
744746.
[21] Li, X. F., and Wang, B. L., 2009, Vibrational Modes of Timoshenko Beams at
Small Scales, Appl. Phys. Lett., 94, p. 101903.
[22] Wang, C. M., Tan, V. B. C., and Zhang, Y. Y., 2006, Timoshenko Beam
Model for Vibration Analysis of Multi-Walled Carbon Nanotubes, J Sound
Vib., 294, pp. 10601072.
[23] Yoon, J., Ru, C. Q., and Mioduchowski, A., 2004, Timoshenko-Beam Effects
on Transverse Wave Propagation in Carbon Nanotubes, Composites, Part B,
35, pp. 8793.
[24] Eringen, A. C., 1983, On Differential Equations of Nonlocal Elasticity and Solution of Screw Dislocation and Surface Waves, J. Appl. Phys., 54(9), pp.
47034710.
[25] Eringen, A. C., 2002, Nonlocal Continuum Field Theories, Springer, New York.
[26] Lee, H. L., Hsu, J. C., and Chang, W. J., 2010, Frequency Shift of CarbonNanotube-Based Mass Sensor Using Nonlocal Elasticity Theory, Nanoscale
Res. Lett., 5, pp. 17741778.
[27] Natsuki, T., Lei, X. W., Ni, Q. Q., and Endo, M., 2010, Vibrational Analysis
of Double-Walled Carbon Nanotubes With Inner and Outer Nanotubes of Different Lengths, Phys. Lett. A, 374, pp. 46844689.
[28] Yang, B., and Tan, C. A., 1992, Transfer Function of One-Dimension Distributed Parameter System, ASME J. Appl. Mech., 59(4), pp. 10091014.
[29] Duan, W. H., Wang, C. M., and Zhang, Y. Y., 2007, Calibration of Nonlocal
Scaling Effect Parameter for Free Vibration of Carbon Nanotubes by Molecular
Dynamics, J. Appl. Phys., 101, p. 024305.
[30] Adhikari, S., Friswell, M. I., and Lei, Y., 2007, Modal Analysis of Nonviscously Damped Beams, ASME J. Appl. Mech., 74, pp. 10261030.
[31] Zhou, J. P., and Yang, B., 1995, A distributed Transfer Function Method for
Analysis of Cylindrical Shells, AIAA J., 33(9), pp. 16981708.
[32] Timoshenko, S., 1974, Vibration Problems in Engineering, Wiley, New York.
[33] Joshi, A. Y., Harsha, S. P., and Sharma, S. C., 2010, Vibration Signature Analysis of Single Walled Carbon Nanotube Based Nanomechanical Sensors, Physica E, 42, pp. 21152123.
[34] Arash, B., Wang, Q., and Varadan, V. K., 2011, Carbon Nanotube-Based Sensors for Detection of Gas Atoms, ASME J. Nanotechnol. Eng. Med., 2,
021010.

Transactions of the ASME

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

Fig. 6 The effect of small scale and attached mass on the fundamental frequency of the DWCNT sensor versus length ratio with L1 5 14 nm (a) e0a/L 5 0.0
(b) e0a/L 5 0.3

fNT/fCT 0.87 and 0.71 at about L2/L1 0.1 and 0.48, respectively, for variable tip mass. The greatest drop among the fundamental frequencies occurs around L2/L1 0.2 regardless of tip
mass. The location at which the greatest drop among the
fundamental frequencies occurs in Figs. 5(a) and 5(b) is clearly
dependent on the nonlocal parameter.
Next, let examine the merit of a DWCNT with different wall
lengths as mass sensor compared with a SWCNT. Figure 6 depicts
the effect of attached mass on the fundamental frequency ratio
fDT/fST, where the subscripts DT and ST stand for the DWCNT
(0 < L2 =L1  1) and SWCNT (L2 =L1 0), respectively. It can be
seen from Fig. 6(a) that the frequency fDT > fST always holds if
e0a/L 0, whereas it is true for L2/L1 > 0.35 if e0a/L 0.3. This is
to say that in a general case, the response of the fundamental
frequencies for DWCNT due to an attached tip mass is stronger
than that for SWCNT. Therefore, a design for nanomechanical
031003-8 / Vol. 2, AUGUST 2011

resonators that operate at various frequencies can be realized by


controlling the length ratio of DWCNTs. Furthermore, due to their
larger diameter, DWCNTs are easily produced and can prevent
undesirable kinking and bucking. So DWCNT may be the more
promising materials for mass sensor than SWCNT.
Finally, we compare the difference between the results of the
natural frequencies when adopting the nonlocal EBT and TBT.
The former neglects shear deformation and rotary inertia. A comparison of the frequency ratio fNT/fNE for a DWCNT with e0a/
L 0.1 and m 10 is demonstrated in Fig. 7. Here, the subscripts NT and NE represent the nonlocal TBT and EBT,
respectively. From Fig. 7, the values of fNT/fNE are always
less than unity. This means that the frequencies based on the
nonlocal EBT are overestimated, as those in the classical
theory. In particular, we find that frequencies using the nonlocal EBT overestimated in the largest extent for L2 =L1 0:5
Transactions of the ASME

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

Fig. 7 Effects of transverse shear deformation and rotary inertia on vibration


frequencies for DWCNT based mass sensor with m 5 10, and e0a/L 5 0.1 and
(NTnonlocal Timoshenko beam model; NEnonlocal Euler beam model.)

than those for L2 =L1 0; 1. In other words, when shear deformation and rotary inertia are taken into account, the nonlocal
TBT is more adequate than the nonlocal EBT, especially for
DWCNT with inner and outer tubes of different lengths.
This mainly results from the essential drawback of the EulerBernoulli hypothesis, irrespective of the nonlocal and classical
models. The difference between TBT and EBT is especially
evident for short DWCNT-based micromass sensors. However,
for slender SWCNT-based micromass sensors, the frequency
from the nonlocal EBT still gives enough accurate results. As
a consequence, for a short DWCNT-based micromass sensor,
it is accurate to employ the TBT instead of the EBT.

1Cai b
1
a b1Ck
(i 1,2,3), and other elements in
2 , U4i;4i1
ai b1Ck2

i
Us vanish.

Appendix B
In Eq. (43), the corresponding elements for the nonlocal TBT
2
are M1;1 M2;3 M3;5 M4;7 1, N6;11 1, N5;9 Ck
a1 , N5;12
2
2
1 Ck , N6;9 mbC, N6;10 1 Cbk , R8;7 R9;1 R10;3
2

R12;3 R9;9 R10;11 R12;11 1, R7;1  wa22kb , R7;5 ka2


2

In this paper, the frequency response of DWCNT-based micromass sensor was investigated. It was modeled as a microcantilever
carrying a concentrated tip mass at the free end of the inner tube.
The interaction of two tubes is described by the vdW force. Using
the nonlocal TBT, the governing equations were derived and the
fundamental frequencies were determined by the TFM. Our
results were confirmed by the results via using the FEM. The conclusions are drawn as follows

C wb2 , R7;8 R11;4 R11;12 1 Ck2 , R8;2 w2 k2 , R8;6

Conclusions

w2
w2 k
Cb
1
U6;8 1bCk
U4i;4i2
U6;4  w2Dk , U6;5 1Cbk
2 D,
2 D ,

Increasing the tip micromass decrease the natural frequencies


but increases the frequency shift.
The mass sensitivity of DWCNT-based mass sensor can be
enhanced when short DWCNTs are used in mass sensors,
especially for larger attached mass.
The mass sensor based on a DWCNT with different wall
lengths has a noticeable advantage over that based on a
SWCNT.
The nonlocal TBT is more adequate than the nonlocal EBT
since shear deformation and rotary inertia are taken into
account.

1 bCk2 w2 k2 , R11;1 ka1 C wb1 , R11;5  wa11kb , R11;9


2

2
2
2
 Ck
a1 , R12;2 1 bCk w1 k , R12;6 w1 k , R12;10 1
2
Cbk and other elements in M b , N b , and Rc vanish.

Appendix C
In Eqs. (34) and (43), the corresponding elements for the nonlocal EBT in place of those in the Appendixes A and B are
 4;1  C  w1 , U
 4;3
 i;i1 1 i 1; 2; 3; 5; 6; 7; 9; 10; 11, U
U
a1 b
a1
2
2
2
2
 4;5 w1 , U
 4;7  w1 k , U
 8;1 w2 , U
 8;3  w2 k ,
Ck w1 k , U
a1

a1 b

a1 b

a1 b

a2 b

a2 b

 8;7 Ck2 w2 k , U
 12;9  C , U
 12;11 Ck2 ,
 8;5  C  w2 , U
U
a2 b
a2
a2 b
a1
a2
a1
 2;2 M
 3;5 M
 4;6 1, N5;11 N6;12 1, N5;9 N6;10
 1;1 M
M
2
mC 






Ck
a3 , N6;9 a3 , R7;7 R8:8 R9;9 R10;10 R11;3 R12;4 1,
2
w
R9;1 R10;2 R11;11 R12;12 1, R7;1 R8;2  2 k , R7;5
a2 b

2
2
w2 k
w1 k
R8;6 Ck
R11;1 R12;2 Ck
R11;5 R12;6
a2 a2 b ,
a1 a1 b ,
2
2
w1 k
Ck
 a1 b and R11;9 R12;10  a3 , other elements in Us, M b ,
N b , and Rc vanish.

References

Appendix A
In Eq. (34), the corresponding elements for the nonlocal TBT
Cb
are U1;2 U3;4 U5;6 U7;8 U9;10 U11;12 1, U2;1 1Cbk
2
2

w1 k
w1
w1 k
w2
1
wD1 , U2;4 1bCk
2 D , U2;5  D , U2;8  D , U6;1  D ,

Journal of Nanotechnology in Engineering and Medicine

[1] Iijima, S., 1991, Helical Microtubules of Graphitic Ccarbon, Nature


(London), 354, pp. 5658.
[2] Dai, H. J., Hafner, J. H., Rinzler, A. G., Colbert, D. T., and Smalley, R. E.,
1996, Nanotubes as Nanoprobes in Scanning Probe Microscopy, Nature
(London), 384, pp. 147150.

AUGUST 2011, Vol. 2 / 031003-9

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

[3] Kim, P., and Lieber, C. M., 1999, Nanotube Nanotweezers, Science, 286, pp.
21482150.
[4] Wong, E. W., Sheehan, P. E., and Lieber, C. M., 1997, Nanobeam Mechanics:
Elasticity, Strength, and Toughness of Nanorods and Nanotubes, Science, 277,
pp. 19711975.
[5] Jensen, K., Kim, K., and Zettl, A., 2008, An Atomic-Resolution Nanomechanical Mass Sensor, Nat. Nanotechnol., 3, pp. 533537.
[6] Poncharal, P., Wang, Z. L., Ugarte, D., and Heer, W. A. D., 1999, Electrostatic
Deflections and Electro-Mechanical Resonances of Carbon Nanotubes, Science, 283, pp. 15131516.
[7] Li, C. Y., and Chou, T. W., 2004, Mass Detection Using Carbon NanotubeBased Nanomechanical Resonators, Appl. Phys. Lett. 84, pp. 52465248.
[8] Joshi, A. Y., Sharma, S. C., and Harsha, S. P., 2010, Dynamic Analysis of a
Clamped Wavy Single Walled Carbon Nanotube Based Nanomechanical
Sensors, J. Nanotechnol. Eng. Med., 1, p. 031007.
[9] Mehdipour, I., Barari, A., and Domairry, G., 2011, Application of a Cantilevered SWCNT With Mass at the Tip as a Nanomechanical Sensor, Comput.
Mater. Sci., 50, pp. 18301833.
[10] Wu, D. H., Chien, W. T., Chen, C. S., and Chen, H. H., 2006, Resonant Frequency Analysis of Fixed-Free Single-Walled Carbon Nanotube-Based Mass
Sensor, Sens. Actuators, A, 126, pp. 117121.
[11] Joshi, A. Y., Sharma, S. C., and Harsha, S. P., 2011, Zeptogram Scale Mass
Sensing Using Single Walled Carbon Nanotube Based Biosensors, Sens.
Actuators, A, 168, pp. 275280.
[12] Mateiu, R., Davis, Z. J., Madsen, D. N., Mlhave, K., Bggild, P., Rassmusen,
A. M., Brorson, M., Jacobsen, J. H. C., and Boisen, A., 2004, An Approach to
a Multi-Walled Carbon Nanotube Based Mass Sensor, Microelectron. Eng.,
73, pp. 670674.
[13] Mateiu, R., Kuhle, A., Marie, R., and Boisen, A., 2005, Building a MultiWalled Carbon Nanotube-Based Mass Sensor With the Atomic Force Microscope, Ultramicroscopy, 105, pp. 233237.
[14] Elishakoff, I., Versaci, C., and Muscolino, G., 2011, Clamped-Free DoubleWalled Carbon Nanotube-Based Mass Sensor, Acta Mech., 219(5), pp. 2943.
[15] Collins, P. G., Hersam, M., Arnold, M., Martel, R., and Avouris, P., 2001,
Current Saturation and Electrical Breakdown in Multiwalled Carbon Nanotubes, Phys. Rev. Lett., 86, pp. 31283131.
[16] Kang, J. W., Kwon, O. K., Hwang, H. J., and Jiang, Q., 2011, Resonance Frequency Distribution of Cantilevered (5,5)(10,10) Double-Walled Carbon Nanotube With Different Intertube Lengths, Mol. Simul., 37(1), pp. 1822.
[17] Kang, J. W., Kwon, O. K., Lee, J. H., Choi, Y. G., and Hwang, H. J., 2009,
Frequency Change by inter-Walled Length Difference of Double-Wall Carbon
Nanotube Resonator, Solid State Commun., 149(1), pp. 15741577.

031003-10 / Vol. 2, AUGUST 2011

[18] Elishakoff, I., and Pentaras, D., 2009, Fundamental Natural Frequencies of
Double-Walled Carbon Nanotubes, J. Sound Vib., 322, pp. 652664.
[19] Elishakoff, I., Versaci, C., Maugeri, N., and Muscolino, G., 2011, ClampedFree Single-Walled carbon Nanotube-Based Mass Sensor Treated as BernoulliEuler Beam, ASME J. Nanotechnol. Eng. Med., 2, p. 021001.
[20] Timoshenko, S., 1921, On the Correction for Shear of the Differential Equation for Transverse Vibrations of Prismatic Bars, Philos. Mag., 41, pp.
744746.
[21] Li, X. F., and Wang, B. L., 2009, Vibrational Modes of Timoshenko Beams at
Small Scales, Appl. Phys. Lett., 94, p. 101903.
[22] Wang, C. M., Tan, V. B. C., and Zhang, Y. Y., 2006, Timoshenko Beam
Model for Vibration Analysis of Multi-Walled Carbon Nanotubes, J Sound
Vib., 294, pp. 10601072.
[23] Yoon, J., Ru, C. Q., and Mioduchowski, A., 2004, Timoshenko-Beam Effects
on Transverse Wave Propagation in Carbon Nanotubes, Composites, Part B,
35, pp. 8793.
[24] Eringen, A. C., 1983, On Differential Equations of Nonlocal Elasticity and Solution of Screw Dislocation and Surface Waves, J. Appl. Phys., 54(9), pp.
47034710.
[25] Eringen, A. C., 2002, Nonlocal Continuum Field Theories, Springer, New York.
[26] Lee, H. L., Hsu, J. C., and Chang, W. J., 2010, Frequency Shift of CarbonNanotube-Based Mass Sensor Using Nonlocal Elasticity Theory, Nanoscale
Res. Lett., 5, pp. 17741778.
[27] Natsuki, T., Lei, X. W., Ni, Q. Q., and Endo, M., 2010, Vibrational Analysis
of Double-Walled Carbon Nanotubes With Inner and Outer Nanotubes of Different Lengths, Phys. Lett. A, 374, pp. 46844689.
[28] Yang, B., and Tan, C. A., 1992, Transfer Function of One-Dimension Distributed Parameter System, ASME J. Appl. Mech., 59(4), pp. 10091014.
[29] Duan, W. H., Wang, C. M., and Zhang, Y. Y., 2007, Calibration of Nonlocal
Scaling Effect Parameter for Free Vibration of Carbon Nanotubes by Molecular
Dynamics, J. Appl. Phys., 101, p. 024305.
[30] Adhikari, S., Friswell, M. I., and Lei, Y., 2007, Modal Analysis of Nonviscously Damped Beams, ASME J. Appl. Mech., 74, pp. 10261030.
[31] Zhou, J. P., and Yang, B., 1995, A distributed Transfer Function Method for
Analysis of Cylindrical Shells, AIAA J., 33(9), pp. 16981708.
[32] Timoshenko, S., 1974, Vibration Problems in Engineering, Wiley, New York.
[33] Joshi, A. Y., Harsha, S. P., and Sharma, S. C., 2010, Vibration Signature Analysis of Single Walled Carbon Nanotube Based Nanomechanical Sensors, Physica E, 42, pp. 21152123.
[34] Arash, B., Wang, Q., and Varadan, V. K., 2011, Carbon Nanotube-Based Sensors for Detection of Gas Atoms, ASME J. Nanotechnol. Eng. Med., 2,
021010.

Transactions of the ASME

Downloaded From: http://nanoengineeringmedical.asmedigitalcollection.asme.org/ on 02/20/2014 Terms of Use: http://asme.org/terms

You might also like