You are on page 1of 16

Downloaded from http://rspa.royalsocietypublishing.

org/ on July 6, 2015

Proc. R. Soc. A (2010) 466, 23252340


doi:10.1098/rspa.2009.0609
Published online 10 March 2010

Thermal vibration of carbon nanotubes


predicted by beam models and molecular
dynamics
BY LIFENG WANG1, *, HAIYAN HU1
1 MOE

AND

WANLIN GUO2

Key Laboratory of Mechanics and Control of Aerospace Structures, and


of Nanoscience, Nanjing University of Aeronautics and Astronautics,
210016 Nanjing, Peoples Republic of China

2 Institute

The paper presents a detailed study on the thermal vibration of a single-walled carbon
nanotube by using different beam models of continuum mechanics, together with the law
of energy equipartition, and the molecular dynamics simulations. The basic nding of the
study is the relation, derived by using the Timoshenko beam model and the law of energy
equipartition, between the temperature and the root-of-mean-squared (RMS) amplitude
of thermal vibration at any cross section of the carbon nanotube. The molecular dynamics
simulations show that both the Euler beam model and the Timoshenko beam model
can roughly predict the thermal vibration of lower order modes for a relatively long
carbon nanotube. However, the Timoshenko beam model, compared with the Euler beam
model, offers a much better prediction of the RMS amplitude of the thermal vibration
near the xed end of the carbon nanotube. For the thermal vibration of a relatively
short carbon nanotube or higher order models of a relatively long carbon nanotube, the
difference between the Timoshenko beam and the Euler beam in dynamic prediction
becomes obvious, and the Timoshenko beam model works much better than the Euler
beam model.
Keywords: carbon nanotube; thermal vibration; Timoshenko beam; molecular dynamics

1. Introduction
Carbon nanotubes have been attracting continuous interest for over a decade
since their discovery (Iijima 1991). Carbon nanotubes exhibit superior mechanical
and electronic properties over any known materials and hold substantial promise
for new super-strong composite materials, among others (Qian et al. 2002).
Treacy et al. (1996) estimated Youngs modulus of isolated carbon nanotubes
by measuring, in the transmission electron microscope, the amplitude of their
intrinsic thermal vibrations. Krishnan et al. (1998) presented the relationship
between Youngs modulus, the size and the standard deviation of the vibration
amplitude at the tip of a carbon nanotube at a given temperature. Hsieh
et al. (2006) investigated the intrinsic thermal vibrations of a single-walled
carbon nanotube via molecular dynamics simulations. The recently developed
*Author for correspondence (walfe@nuaa.edu.cn).
Received 16 November 2009
Accepted 4 February 2010

2325

This journal is 2010 The Royal Society

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2326

L. F. Wang et al.

nanotechnology enables one to construct small devices working at a molecular


level and activated by the thermal uctuation. One example is the so-called
Brownian ratchet, where a particle moves unidirectionally over a switchable
anisotropic potential (Astumian 1997). Xu et al. (2006) studied thermally driven
large-amplitude uctuations in carbon nanotube-based devices by using molecular
dynamics simulations.
Recent years have witnessed numerous applications of the Timoshenko beam
model to analyse the vibration and wave propagation of carbon nanotubes.
For example, Yoon et al. (2004, 2005) studied the exural wave propagation
and vibration of a multi-walled carbon nanotube based on the model of the
Timoshenko beam with the rotary inertia and the shear deformation taken
into account. Wang & Hu (2005) found that the non-local elastic model of
the Timoshenko beam, compared with other beam models, can offer a better
prediction of the dispersion of the exural wave in carbon nanotubes. To the
best of our knowledge, however, neither experimental studies nor numerical
simulations have been available for the validation of the Timoshenko beam model
in studying the thermal vibration of any carbon nanotubes.
The primary objective of this study is to check the validity of the Timoshenko
beam model together with the law of energy equipartition in studying the
thermal vibration, simulated by the molecular dynamics, of a single-walled carbon
nanotube in argon atmosphere. For this purpose, 2 presents the root-of-meansquared (RMS) amplitude of the stochastically driven vibration of a Timoshenko
beam, which will be used to model single-walled carbon nanotubes. Then, 3
gives the molecular dynamics simulation for the thermal vibration of a carbon
nanotube under an argon atmosphere. Section 4 outlines a comparison between
the analytical results and the numerical results. Finally, the paper ends with
concluding remarks in 5.

2. Timoshenko beam model


This section starts with the dynamic equation of a Timoshenko beam with
uniform cross section placed along direction x in the frame of coordinates (x, y, z),
w(x, t) being the displacement of section x of the beam in direction y at the
moment t (Thomson 1972)


v4 v2 w
v2 w
2
=0
(2.1a)
rA 2 + bAG
vt
vx
vx
and
v2 4
rI 2 + bAG
vt



vw
4
vx




EI


v2 4
= 0,
vx 2

(2.1b)

where E is Youngs modulus, I is the moment of inertia for the cross section, r is
the mass density, A is the cross-section area of the beam, G is the shear modulus,
4 is the slope of the deection curve when the shearing force is neglected, b is the
form factor of shear depending on the shape of the cross section and b = 0.5 holds
for the circular tube of the thin wall (Timoshenko & Gere 1972). The boundary
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

Stochastically driven vibration of CNT

2327

conditions of a cantilever beam clamped at x = 0 are


v2 4(L, t)
v2 w(L, t)
=
0
and
= 0.
(2.2)
vx 2
vx 2
To study the free vibration of a Timoshenko beam, let the dynamic deection
and slope be given by
w(0, t) = 0,

4(0, t) = 0,

w = we
jut

and 4 = 4e
jut ,

(2.3)

where w represents the deection amplitude of the beam, and 4 the slope
amplitude of the beam due to bending deformation alone. Let
x
x= .
(2.4)
L
Substituting equation (2.3) into equation (2.1), one obtains
v4
v2 w
+ b 2 s 2 w = 0
L
2
vx
vx
and
s2

(2.5a)

v2 4
1 vw
(1 b 2 r 2 s 2 )4 = 0,
+
vx2
L vx

(2.5b)

where
b2 =

rAL4 u2
,
EI

r2 =

I
AL2

and

s2 =

EI
.
bAGL2

(2.6)

In the case of [(r 2 s 2 )2 + 4/b 2 ]1/2 > (r 2 + s 2 ), the solutions of equations (2.5a)
and (2.5b) read (Huang 1961)
w = C1 cosh ba1 x + C2 sinh ba1 x + C3 cos ba2 x + C4 sin ba2 x

(2.7a)

4 = C1 sinh ba1 x + C2 cosh ba1 x + C3 sin ba2 x + C4 cos ba2 x,

(2.7b)

and

where


 1/2

1
4 1/2
a1
2
2
2
2
= (r + s ) + (r s ) + 2
.
a2
b
2

(2.8)

In the case of [(r 2 s 2 )2 + 4/b 2 ]1/2 < (r 2 + s 2 ), then equations (2.7a) and (2.7b)
should be replaced by (Huang 1961)
w = C1 cos ba1 x + jC2 sin ba1 x + C3 cos ba2 x + C4 sin ba2 x

(2.9a)

4 = jC1 sin ba1 x + C2 cos ba1 x + C3 sin ba2 x + C4 cos ba2 x,

(2.9b)

and
where
a1
Proc. R. Soc. A (2010)


 1/2

1
4 1/2
2
2
2
2
= (r + s ) (r s ) + 2
.
b
2

(2.10)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2328

L. F. Wang et al.

In equations (2.7a), (2.7b) and (2.9a), (2.9b), only one-half of the constants are
independent since they are related by equations (2.5a) and (2.5b) as follows:
C1 =

b a21 + s 2
C1 ,
L a1

(2.11a)

C2 =

b a21 + s 2
C2 ,
L a1

(2.11b)

C3 =
C4 =

and

b a22 s 2
C3
L a2

b a22 s 2
C4 .
L a2

(2.11c)
(2.11d)

For the case of [(r 2 s 2 )2 + 4/b 2 ]1/2 > (r 2 + s 2 ), the natural frequency of the
cantilever beam yields the following equation (Huang 1961):
b(r 2 + s 2 )
sinh ba1 sin ba2 = 0.
(1 b 2 r 2 s 2 )1/2
(2.12)
Solving equation (2.12) for b gives an innite 
number of bn , n = 1, 2, . . ., which
determines the nth natural frequency un = EIbn2 /rAL4 of the Timoshenko
beam in this case. The nth normal mode (w n , 4 n ) for the cantilever beam is
(Huang 1961)
2 + [b 2 (r 2 s 2 )2 + 2] cosh ba1 cos ba2

w n = Dn fwn (x) = Dn [cosh bn an1 x ln zn dn sinh bn an1 x cos bn an2 x + d sin bn an2 x]
(2.13a)
and

4 n = Dn f4n (x) = Hn


qn
cosh bn an1 x +
sinh bn an1 x cos bn an2 x + qn sin bn an2 x ,
ln z n
(2.13b)

where
bn a2n1 + s 2 ln zn
Dn ,
L an1
qn
(1/ln ) sinh bn an1 sin bn an2
dn =
,
zn cosh bn an1 + cos bn an2
ln sinh bn an1 + sin bn an2
qn =
(1/zn ) cosh bn an1 + cos bn an2

 1/2

1
4 1/2
an1
2
2
2
2
= (r + s ) + (r s ) + 2
.
an2
bn
2
Hn =

and

Proc. R. Soc. A (2010)

(2.14a)
(2.14b)
(2.14c)

(2.14d)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

Stochastically driven vibration of CNT

2329

For the case of [(r 2 s 2 )2 + 4/b 2 ]1/2 < (r 2 + s 2 ), the nth natural frequency un of
the Timoshenko beam can be determined from (Huang 1961)
2 + [b 2 (r 2 s 2 )2 + 2] cos ba1 cos ba2

b(r 2 + s 2 )
sin ba1 sin ba2 = 0. (2.15)
(b 2 r 2 s 2 1)1/2

The nth normal mode (w n , 4 n ) for the cantilever beam is


w n = Dn fwn (x) = Dn [cos bn an1 x + ln zn hn sin bn an1 x cos bn an2 x + hn sin bn an2 x]
(2.16a)
and

4 n = Dn f4n (x) = Hn

cos bn an1 x


mn


sin bn an1 x cos bn an2 x + mn sin bn an2 x ,
ln zn
(2.16b)

where
Hn =
hn =

bn a2n1 + s 2 ln zn
Dn ,
L
an1
mn

(1/ln ) sin bn an1 sin bn an2


,
zn cos bn an1 + cos bn an2

ln sin bn an1 sin bn an2


(1/zn ) cos bn an1 + cos bn an2


1/2 1/2
1
4
an1 = (r 2 + s 2 ) (r 2 s 2 ) + 2
.
bn
2
mn =

and

(2.17a)
(2.17b)
(2.17c)

(2.17d)

In equations (2.12)(2.17ad), one has


zn =

(a2n1 + r 2 ) (a2n2 s 2 )
=
(a2n1 + s 2 ) (a2n2 r 2 )

(a2n1 + r 2 ) (a2n2 s 2 )
=
(a2n2 r 2 ) (a2n1 + s 2 )

(2.18a)

an1
= jln .
an2

(2.18b)

and
ln =

Now, the study turns to the stochastic vibration of the Timoshenko beam
from the viewpoint of energy analysis. The total energy En contained in the
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2330

L. F. Wang et al.

nth vibration mode can be found by calculating the elastic energy at the instant
of maximal deection when the cantilever is momentarily stationary, i.e. ejun t = 1,






L
EI v4n 2
vwn 2
elastic
=
+ bAG 4n
dx
En
vx
vx
0 2
t=4/un






2
2
L
vfwn
EI vf4n
2
= Dn
+ bAG f4n
dx
.
(2.19)
vx
vx
0 2
t=4/un

From the law of energy equipartition, there is an average energy of kT /2 per


degree of freedom for all of the relevant lateral vibration modes. There are both
elastic and kinetic energy degrees of freedom in a vibration mode then, on average,
En  = kT for each vibration mode, with En yielding the Boltzmann distribution
(Krishnan et al. 1998). From equation (2.19), it is easy to get
Dn2 = 

L
0

= 
L
0

Enelastic
(EI /2)(vf4n /vx)2 + bAG(f4n (vfwn /vx))2 dx
kT
(EI /2)(vf4n /vx)2 + bAG(f4n (vfwn /vx))2 dx


t=4/un

(2.20)

t=4/un

Then, the RMS amplitude of the nth mode at x is


wn (x) = Dn fwn (x).

(2.21)

As the vibration modes are mutually independent, the vibration prole for the
combined modes is also a Gaussian distribution, with standard deviation given
by the sum of the variances. The RMS amplitude of the carbon nanotube at x is



w 2 (x).
w(x) = 
n

(2.22)

n=0

3. Molecular dynamics model


In order to check the applicability of the above Timoshenko beam model to
studying the thermal vibration of a single-walled carbon nanotube, this section
presents the model of molecular dynamics simulations of a carbon nanotube under
an argon atmosphere as shown in gure 1.
In the corresponding molecular dynamics models, the interatomic interactions
are described by the TersoffBrenner potential (Brenner 1990), which has been
proved applicable to the description of mechanical properties of single-walled
carbon nanotubes. The structure of the TersoffBrenner potential is as follows:
 
[VR (rij ) B ij VA (rij )],
(3.1)
V (rij )
i

Proc. R. Soc. A (2010)

j(>i)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2331

Stochastically driven vibration of CNT


120

100

80

60

40

20

0
60

40

20

20

40

60

Figure 1. Molecular dynamics model of a single-walled carbon nanotube, under an argon


atmosphere, with black-lled atoms xed.

where rij is the distance from atom i to atom j; VR (rij ) and VA (rij ) are the
repulsive and attractive terms given by

 
Dij
exp 2Sij bij (rij r0 )
VR (rij ) fij (rij )
Sij 1

(3.2a)

and


Sij Dij
2
VA (rij ) fij (rij )
exp
bij (rij r0 ) .
Sij 1
Sij

(3.2b)

Here, Sij = 1.29, Dij = 6.325 eV, bij = 15 nm1 , r0 = 0.1315 nm, fij , Dij , Sij , bij
are scalars; fij (rij ) is a switch function used to conne the pair potential in a
neighbourhood with radius of r2 as follows:

1,
rij < r1 ,



1 
p(rij r1 )
fij (rij )
1 + cos
, r1 rij r2 ,

2
r2 r 1

0,
rij > r2 ,
Proc. R. Soc. A (2010)

(3.2c)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2332

L. F. Wang et al.

where r1 = 1.7 and r2 = 2.0 . In equation (3.1), B ij reads


B ij 12 (bij + bji ),

d

G(qjik )fik (rik )
bij 1 +

and

k
=i,j


and G(qijk ) a0 1 +

bji 1 +

(3.2d)
d
G(qijk )fjk (rjk )

k
=i,j

c02
c02

,
d02 d02 + (1 + cos qijk )2

(3.2e)
(3.2f )

where qijk is the angle between bonds ij and ik, d = 0.80469, a0 = 0.011304,
c0 = 19 and d0 = 2.5. In addition, the CC bond length in the model
is 0.142 nm.
The van der Waals interaction either between two argon atoms or between a
carbon atom and an argon atom is described by the LennardJones pair potential
 
 6

s 12
s

,
(3.3)
V (rij ) = 43
rij
rij
where 3CC = 3.19 103 eV, sCC = 3.345 (Razadeh 1974), 3ArAr = 1.032

102 eV, sArAr = 3.822 (Cleary & Mayne 2006), 3CAr = 3CC 3ArAr and
sCAr = (sCC + sArAr )/2.
The Verlet algorithm in the velocity form with time step 1 fs is used to simulate
the thermal vibration of carbon nanotubes

R(t + dt) = R(t) + dtV (t) + 12 dt 2 a(t)
(3.4)
and
V (t + dt) = V (t) + 12 dt[a(t) + a(t + d)],
where R represents the position of atoms, V the velocity of atoms, a the
acceleration of atoms and dt the time step. In the simulation, the periodic
boundary conditions are applied to argon atmospheres. The displacements of
nanotube of each section are observed every 1 ps, the total simulation time is
Ttat ps and the RMS amplitude of the cross section of carbon nanotube at x is

Ttot 2
i=1 wi (x)
,
(3.5)
w(x)

=
Ttot
where wi (x) is the amplitude of the cross section of carbon nanotube at x.

4. Comparison between analytical and numerical results


To predict the thermal vibration of a single-walled carbon nanotube from the
analytical results in 2, it is necessary to know Youngs modulus E and the
shear modulus G or Poissons ratio y in advance. The previous studies based
on the TersoffBrenner potential gave a great variety of Youngs moduli of
single-walled carbon nanotubes from the simulated tests of axial tension and
compression (Guo et al. 2006). When the thickness of the wall was chosen as
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2333

(a)

0.4

RMS amplitude (nm)

Stochastically driven vibration of CNT

0.3

(b) 0.06

0.04
0.2
0.02
0.1

(c) 0.020

(d ) 0.012

RMS amplitude (nm)

0.010
0.015
0.008
0.010

0.006
0.004

0.005
0.002
0

10
15
x coordinate (nm)

20

25

10
15
x coordinate (nm)

20

25

Figure 2. The RMS amplitude of thermal vibration of a (10, 10) armchair carbon nanotube of
24.6 nm long at 300 K. (a) The rst mode, (b) the second mode, (c) the third mode and (d) the
fourth mode. Solid line, T ; dashed line, E.

0.34 nm, for example, 1.07 TPa was reported by Yakobson et al. (1996), 0.8 TPa
by Cornwell & Wille (1997) and 0.440.50 TPa by Halicioglu (1998). Meanwhile,
Youngs modulus determined by Zhang et al. (2002) on the basis of the nanoscale
continuum mechanics was only 0.475 TPa when the rst set of parameters in the
TersoffBrenner potential was used. Wang et al. (2006) reported that Youngs
modulus was about 0.46 TPa, based on the higher order CauchyBorn rule. Hence,
it becomes necessary to compute Youngs modulus and Poissons ratio again
from the above molecular dynamics model for the single-walled carbon nanotubes
under the static loading.
For the same thickness of wall, Youngs modulus computed by using the
rst set of parameters in the TersoffBrenner potential was 0.46 TPa for the
armchair (5, 5) carbon nanotube and 0.47 TPa for the armchair (10, 10) carbon
nanotube for the axial tension. Furthermore, the simulated test of pure bending
gave the product of effective Youngs modulus E = 0.39 TPa and Poissons ratio
y = 0.22 for the armchair (5, 5) carbon nanotube and E = 0.45 TPa and y = 0.20
for the armchair (10, 10) carbon nanotube (Wang & Hu 2005). In the numerical
simulations of this section, Youngs moduli and Poissons ratios obtained from
the simulated test of pure bending for those two carbon nanotubes were used.
For the single-walled carbon nanotubes, the wall thickness is h = 0.34 nm and the
mass density of the carbon nanotubes is r = 2237 kg m3 .
Figure 2 shows the RMS amplitude of the thermal vibration of the rst four
modes of a (10, 10) armchair carbon nanotube of 24.6 nm long at 300 K. Here,
the symbol E represents the results predicted by using the Euler beam model
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2334
0.4

(b) 0.008

0.3

0.006

0.2

0.004

0.1

0.002

(c) 0.004

(d ) 0.020

RMS amplitude (nm)

(a)
RMS amplitude (nm)

L. F. Wang et al.

0.003

0.015

0.002

0.010

0.001

0.005

3
2
x coordinate (nm)

2
3
x coordinate (nm)

Figure 3. The RMS amplitude of thermal vibration of a (10, 10) armchair carbon nanotube of
4.9 nm long at 300 K. (a) The rst mode, (b) the second mode, (c) the third mode and (d) the
fourth mode. Solid line, T ; dashed line, E.

in appendix B, and the symbol T represents results predicted by using the


Timoshenko beam model. In gure 2, no difference between the results of the
Euler beam model and the Timoshenko beam model can be found for the rst
mode of vibration, and little difference can be observed for the second mode of
vibration. The differences between the results of the two beam models become
more and more obvious with an increase in the mode order. The RMS amplitude
predicted by the Timoshenko beam model is larger than that predicted by the
Euler beam model.
Figure 3 presents the RMS amplitudes of the thermal vibration of the rst
four modes of a (10, 10) armchair carbon nanotube of 4.9 nm long at 300 K.
Here, the symbols E and T represent the same models as in gure 2. Figure 3a
indicates that the difference between the results of the Euler beam model and
the Timoshenko beam model looks obvious even for the rst mode of vibration.
Figure 3bd shows again that the difference for the beam models becomes more
and more obvious with an increase in the mode order. The RMS amplitude
predicted by the Timoshenko beam model is also larger than that predicted by
the Euler beam model.
Figure 4 illustrates the RMS amplitudes of thermal vibration of armchair
carbon nanotubes of different sizes at 300 K, where gure 4a is for a (10, 10)
armchair carbon nanotube of 14.5 nm long, gure 4c is for a (10, 10) armchair
carbon nanotube of 9.6 nm long, gure 4e is for a (5, 5) armchair carbon nanotube
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2335

Stochastically driven vibration of CNT

RMS amplitude (nm)

(a) 0.20

(b) 0.020

0.15

0.015

0.10

0.010

0.05

0.005

10

15

RMS amplitude (nm)

(c) 0.10

(d )

0.03

0.08
0.02
0.06
0.04

0.01

0.02

(e) 0.25
RMS amplitude (nm)

10

(f)

0.20

0.08
0.06

0.15
0.04
0.10
0.02

0.05

4
6
x coordinate (nm)

10

2
x coordinate (nm)

Figure 4. The RMS amplitude of thermal vibration carbon nanotubes at 300 K. (a) A (10, 10)
armchair carbon nanotube of 14.5 nm long; (b) zoom view of (a). (c) A (10, 10) armchair carbon
nanotube of 9.6 nm long; (d) zoom view of (c). (e) A (5, 5) armchair carbon nanotube of 9.6 nm
long; (f ) zoom view of (e). Solid line, T ; dashed line, E; dashed dotted grey lines, MD1, MD2.

of 9.6 nm long and gure 4b,d,f are the zoom views of gure 4a,c,e, respectively.
Here, the symbols E and T have the same meaning as before, and MD1 and MD2
are the molecular dynamics simulations in x and y directions, respectively. The
parameters of the Euler beam model and the Timoshenko beam model are the
same as those given previously. In the molecular dynamics simulation, different
densities of argon were used, but the difference among those results was not
obvious. It can be seen that both the Euler beam model and the Timoshenko beam
model can roughly predict the thermal vibration of carbon nanotubes in gure 4.
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2336

L. F. Wang et al.

In the zoomed view of the xed end of gure 4b,d,f , however, the results show
that the Timoshenko beam model, compared with the Euler beam model, offers a
much better prediction of the RMS amplitude of thermal vibration simulated by
using molecular dynamics. For shorter carbon nanotubes, the difference between
the results of the Timoshenko beam model and the Euler beam model is obvious.
And the Timoshenko beam model can give a better prediction of the thermal
vibration than the Euler beam model for shorter carbon nanotubes.

5. Concluding remarks
The Timoshenko beam model, together with the law of energy equipartition,
enables one to establish the analytical relation between the temperature and the
RMS amplitude of thermal vibration at any cross section of the carbon nanotube.
The molecular dynamics simulations of both (5, 5) and (10, 10) armchair carbon
nanotubes show that the Timoshenko beam model, compared with the Euler
beam model, is able to offer a much better prediction for the thermal vibration
of those carbon nanotubes when the vibration modes are higher and the carbon
nanotubes are shorter, especially at the xed end of a carbon nanotube.
This work was supported in part by the National Natural Science Foundation of China under grant
nos. 10702026, 60910007.

Appendix A. Vibration prole of the tip of a carbon nanotube


(Krishnan et al. 1998)
This appendix presents the calculation of the vibration prole of the tip of a
carbon nanotube. For a classical simple oscillator of amplitude Dn , the oscillator
position y at the moment t is given by
y = Dn sin(ut),

(A 1)

where, as before, Dn is the amplitude that depends on the energy of the oscillator,
and u = 2p/f , with f being the frequency. In the interval between y and y + dy,
the oscillator spends a time dt, which is found by taking the derivative of
equation (A 1)
dy = Dn u cos(ut) dt,
(A 2)
or
dy
,
dt = 
u Dn2 y 2

Dn  y  Dn .

(A 3)

The probability P(un , y) dy of nding the oscillator between y and y + dy when


the amplitude is Dn is proportional to the time spent in this interval dt. After
normalization, it is found that

1 D 2 y 2 , |y|  Dn ,
n
(A 4)
P(Dn , y) = p

0,
|y| > Dn .
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

Stochastically driven vibration of CNT

2337

P(Dn , y) reaches a peak at the extrema y = Dn and has a minimum at y = 0.


However, the energy of the system, and hence the amplitude Dn , is changing in a
stochastic manner with an increase in time. Therefore, one needs also to average
over all the possible values of Dn that the system can be adopted. Furthermore,
one must also average over all of the activated modes.
The carbon nanotube vibration is essentially relaxed phonons that are in an
equilibrium at an ambient temperature T . The probability that the system is in
the state m of energy En = m hu
n is given by the Boltzmann factor




hu
m h un
exp(m hu
n
n /kT )
1 exp
, (A 5)
= exp
W (m) = 
kT
kT
n /kT )
p=0 exp(p hu
where h = 1.054 1034 J s is the Planck constant. The frequency un of a
vibrating nanotube of density r is given by

EI
.
(A 6)
un = a2n
rA
The energy in the nth mode is therefore quantized in units of hu
n . For typical
/kT

1,
for
the
lower
modes,
which
is important for
nanotubes in this paper hu
n
the RMS amplitude of thermal induced vibration of a carbon nanotube. Thus, to
a very good approximation, one has


m hu
hu
n
n
exp
.
(A 7)
W (m) =
kT
kT
If one sets En = m hu
n and dEn hu
n , then W (En ) dEn is the continuum limit
of W (m) in equation (A 7), and W (En ) dEn is the probability between En and
En + dEn of energy in the nth mode,


En
1
exp
dEn .
(A 8)
W (En ) dEn =
kT
kT
This result is expected when the thermal average number of phonons, kT /hu
n,
is high. The average energy is En = kT , the half of which comes from the kinetic
energy degree of freedom, and the other half from the elastic energy degree of
freedom.
The stochastically averaged probability amplitude is therefore



P(Dn , y)W (En ) dEn ,
(A 9)
p(y) =
0

which, from equations (A 4) and (A 9), reads





exp(En /kT )
1
dEn ,
p(y) =

pkT 0
Dn2 y 2

y 2  Dn2 .

(A 10)

From equation (B 4), one has


Dn2 =
Proc. R. Soc. A (2010)

8En
En
=2 ,
4
EILan
cn

(A 11)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2338

L. F. Wang et al.

where cn = EILa4n /4. Therefore, one reaches





exp(En /kT )
1
dEn .
p(y) =

pkT cn y 2 /2 2En /cn y 2

(A 12)

The substitution
cn y 2 (1 + x 2 )
and dEn = cn y 2 x dx
2
ensures that the condition En  cn y 2 /2 holds. Thus, one has



 


cn
cn y 2 x 2
cn y 2
dx.
p(y) =
x exp
exp
2kT
pkT
2kT
0
En =

The integral above can be easily worked out to give the Gaussian form





cn y 2
cn
exp
.
p(y) =
pkT
2kT

(A 13)

(A 14)

(A 15)

Using equation (A 11), one has the standard deviation


sn2 =

kT
.
cn

(A 16)

Appendix B. Stochastic vibration of a cantilever Euler beam


(Krishnan et al. 1998)
This appendix starts with the dynamic equation of an Euler beam of uniform
cross section placed along direction x in the frame of coordinates (x, y, z), with
w(x, t) being the displacement of section x of the beam in direction y at the
moment t. Upon the assumption of small amplitudes, the motion of a vibrating
Euler beam is governed by the following partial equation of the fourth order
(Thomson 1972)
v4 w(x, t)
v2 w(x, t)
+
EI
= 0,
(B 1)
rA
vt 2
vx 4
where E is Youngs modulus, I is the moment of inertia for the cross section,
r is the mass density and A is the cross-section area of the beam. The boundary
conditions for a cantilever Euler beam are
v3 w(L, t)
v2 w(L, t)
vw(0, t)
=
0
and
= 0.
(B 2)
= 0,
w(0, t) = 0,
vx
vx 2
vx 3
The solution for the nth mode reads (Thomson 1972)
wn (x, t) = w n (x)ejun t = Dn fn (x)ejun t


Dn
sin an L sinh an L
(sin an x sinh an x) ejun t ,
=
cos an x cosh an x +
2
cos an L + cosh an L
(B 3)

def
where j 1, a4n = rA
u2 , an L = 1.8751, 4.6941, 7.8548, 10.9955, 14.1372, . . . ,
EI n
and Dn is a constant decided from initial conditions.
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

Stochastically driven vibration of CNT

2339

The total energy En corresponding to the nth mode can be found by calculating
the elastic energy at the instant of maximal deection when the cantilever is
momentarily stationary for ejun t = 1

2

L 
vw
rA
EILD2n a4n
n
.
(B 4)
dx
=
Enelastic =
2 0
vt
8
t=0

From the law of energy equipartition, there is an averaged energy of kT /2 per


degree of freedom for all of the relevant lateral vibration modes. Because there are
both elastic and kinetic energy degrees of freedom for all of the relevant lateral
vibration modes, then, on average, En  = kT holds for each vibration mode, with
En yielding the Boltzmann distribution. It is straightforward to show that each
mode of a stochastically driven oscillator has a Gaussian probability prole and
the standard deviation of the vibration amplitude of the free tip of the Euler
beam is given by (Krishnan et al. 1998)
sn2 [L] = Dn2 =

8kT
.
EILa4n

(B 5)

Then, the RMS amplitude of the nth mode at x reads


s n (x) = Dn fn (x).

(B 6)

As all vibration modes are independent, their contributions can be added


incoherently. To average in coherently over all the vibration modes, one simply
adds the variances sn2 (x) to get another Gaussian distribution with the standard
deviation given by



s 2n (x).
(B 7)
s(x)

=
n=0

So, the RMS amplitude of the stochastic vibration of the Euler beam at x can be
obtained.

References
Astumian, R. D. 1997 Thermodynamics and kinetics of a Brownian motor. Science 276, 917922.
(doi:10.1126/science.276.5314.917)
Brenner, D. W. 1990 Empirical potential for hydrocarbons for use in simulating the chemical vapor
deposition of diamond lms. Phys. Rev. B 42, 94589471. (doi:10.1103/PhysRevB.42.9458)
Cleary, S. M. & Mayne, H. R. 2006 High-symmetry global minimum geometries for small mixed
Ar/Xe Lennard-Jones clusters. Chem. Phys. Lett. 418, 7983. (doi:10.1016/j.cplett.2005.10.070)
Cornwell, C. F. & Wille, L. T. 1997 Elastic properties of single-walled carbon nanotubes in
compression. Solid State Commun. 101, 555558. (doi:10.1016/S0038-1098(96)00742-9)
Guo, X., Wang, J. B. & Zhang, H. W. 2006 Mechanical properties of single-walled carbon nanotubes
based on higher order CauchyBorn rule. Int. J. Solids Struct. 43, 12761290. (doi:10.1016/
j.ijsolstr.2005.05.049)
Halicioglu, T. 1998 Stress calculations for carbon nanotubes. Thin Solid Films 312, 1114.
(doi:10.1016/S0040-6090(97)00369-6)
Hsieh, J. Y., Lu, J. M., Huang, M. Y. & Hwang, C. C. 2006 Theoretical variations in the Youngs
modulus of single-walled carbon nanotubes with tube radius and temperature: a molecular
dynamics study. Nanotechnology 17, 39203924. (doi:10.1088/0957-4484/17/15/051)
Proc. R. Soc. A (2010)

Downloaded from http://rspa.royalsocietypublishing.org/ on July 6, 2015

2340

L. F. Wang et al.

Huang, T. C. 1961 The effect of rotatory inertia and of shear deformation on frequency and normal
mode equations of uniform beams with simple end conditions. J. Appl. Mech. 28, 579584.
Iijima, S. 1991 Helical microtubules of graphitic carbon. Nature 354, 5658. (doi:10.1038/354056a0)
Krishnan, A., Dujardin, E., Ebbesen, T. W., Yianilos, P. N. & Treacy, M. M. J. 1998
Youngs modulus of single-walled nanotubes. Phys. Rev. B 58, 14 01314 019. (doi:10.1103/
PhysRevB.58.14013)
Qian, D., Wagner, G. J., Liu, W. K., Yu, M. F. & Ruoff, R. S. 2002 Mechanics of carbon nanotubes.
Appl. Mech. Rev. 55, 495533. (doi:10.1115/1.1490129)
Razadeh, H. A. 1974 An analytical-potential approach to the lattice dynamics of graphite. Physica
74, 135150. (doi:10.1016/0031-8914(74)90189-X)
Thomson, W. T. 1972 Theory of vibration with applications. Englewood Cliffs, NJ: Prentice-Hall,
Inc.
Timoshenko, S. & Gere, J. 1972 Mechanics of materials. New York, NY: Van Nostrand Reinhold
Company.
Treacy, M. M. J., Ebbesen, T. W. & Gibson, J. M. 1996 Exceptionally high Youngs modulus
observed for individual. Nature 381, 678680. (doi:10.1038/381678a0)
Wang, L. F. & Hu, H. Y. 2005 Flexural wave propagation in single-walled carbon nanotubes.
Phys. Rev. B 71, 195 412. (doi:10.1103/PhysRevB.71.195412)
Wang, J. B., Guo, X., Zhang, H. W., Wang, L. & Liao, J. B. 2006 Energy and mechanical
properties of single-walled carbon nanotubes predicted using the higher order CauchyBorn
rule. Phys. Rev. B 73, 115 428. (doi:10.1103/PhysRevB.73.115428)
Xu, Z. P., Zheng, Q.-S. & Chen, G. H. 2006 Thermally driven large-amplitude uctuations in
carbon-nanotube-based devices: molecular dynamics simulations. Phys. Rev. B 74, 195 445.
Yakobson, B. I., Brabec, C. J. & Bernholc, J. 1996 Nanomechanics of carbon tubes: instabilities
beyond linear response. Phys. Rev. Lett. 76, 25112514. (doi:10.1103/PhysRevLett.76.2511)
Yoon, J., Ru, C. Q. & Mioduchowski, A. 2004 Timoshenko-beam effects on transverse wave
propagation in carbon nanotubes. Compos. B Eng. 35, 8793. (doi:10.1016/j.compositesb.
2003.09.002)
Yoon, J., Ru, C. Q. & Mioduchowski, A. 2005 Terahertz vibration of short carbon nanotubes
modeled as Timoshenko beams. J. Appl. Mech. 72, 1017. (doi:10.1115/1.1795814)
Zhang, P., Huang, Y., Geubelle, P. H., Klein, P. A. & Hwang, K. C. 2002 The elastic modulus
of single-wall carbon nanotubes: a continuum analysis incorporating interatomic potentials.
Int. J. Solids Struct. 39, 38933906. (doi:10.1016/S0020-7683(02)00186-5)

Proc. R. Soc. A (2010)

You might also like