Você está na página 1de 5

Materials and Design 31 (2010) 44454449

Contents lists available at ScienceDirect

Materials and Design


journal homepage: www.elsevier.com/locate/matdes

Short Communication

Temperature-dependent elastic properties of single layer graphene sheets


Le Shen a, Hui-Shen Shen a,b,*, Chen-Li Zhang a
a
b

Department of Engineering Mechanics, Shanghai Jiao Tong University, Shanghai 200030, Peoples Republic of China
State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, Shanghai 200030, Peoples Republic of China

a r t i c l e

i n f o

Article history:
Received 28 January 2010
Accepted 8 April 2010
Available online 14 April 2010

a b s t r a c t
Elastic properties of single layer graphene sheets (SLGSs) with different values of aspect ratio are presented by using molecular dynamics simulation. SLGSs subjected to uniaxial tension, shear load and
transverse uniform pressure are simulated under temperature varying from 300 K to 700 K. Based on
the classical plate theory, an individual orthotropic plate model is adopted for SLGSs. By direct measuring
the bending deflections, the effective thickness of SLGSs is determined uniquely. It is found that SLGSs
exhibit anisotropic, size-dependent and temperature-dependent properties. The results reveal that
Youngs modulus decreases with increasing in temperature, whereas the shear modulus depends weakly
on temperature change. The results also show that the effective thickness of zigzag sheets is larger than
that of armchair sheets.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Nanomaterials have attracted a great deal of attention of scientific community due to their novel mechanical, chemical and electronic properties [1]. Two-dimensional layer of sp2-bonded carbon
with one-atom-thick is dubbed a graphene sheet. The suspended
graphene sheets can be used in a variety of ways such as pressure
sensors or gas detectors [2] or in the building of mechanical resonators [3]. Owing to their exceptional thermal and mechanical
properties and high electrical conductivity, graphene sheets are
also of great interest to serve as new nanoscale building blocks
to create novel composite materials [4].
Most studies on single layer graphene sheets (SLGSs) have focused on their material properties [512]. In these studies the
material properties of SLGSs are assumed to be isotropic, except
for [12]. However, a large variation of Youngs modulus E and Poissons ratio m, as well as effective thickness h was obtained and reported in the open literature (see Table 1). The wide dispersion of
the mechanical properties of SLGSs can be attributed principally
to the uncertainty associated to the thickness of these nanostructures. For the majority of models used, the assumed thickness of
the graphene layer is 0.34 nm [6,8,10,11]. The 0.34 nm value provides in-plane Youngs modulus of the order of 1 TPa [13]. In contrast, Hemmasizadeh et al. [9] used a mixed molecular dynamics
(MD)continuum mechanics model to obtain E = 0.939 TPa and
m = 0.19 with h = 0.1317 nm. Huang et al. [7] used the second generation Brenner potentials to calculate the in-plane Youngs modulus,
Poissons ratio and thickness of SLGSs and obtained E = 2.99 TPa and
* Corresponding author at: Department of Engineering Mechanics, Shanghai Jiao
Tong University, Shanghai 200030, Peoples Republic of China.
E-mail address: hsshen@mail.sjtu.edu.cn (H.-S. Shen).
0261-3069/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.matdes.2010.04.016

m = 0.397 with h = 0.0811 nm. Their value of Youngs modulus is


found to be about three times as large as that of Hemmasizadeh
et al. [9]. It has been reported that [1416] the material properties
of single walled carbon nanotubes (SWCNTs) are anisotropic, chirality- and size-dependent and temperature-dependent. From
these studies we believe that the elastic properties of SLGSs, such
as Youngs moduli and shear modulus, are also anisotropic and temperature-dependent. Therefore, all material properties and effective
thickness of a SLGS need to be properly given. Otherwise, the
numerical results for bending or vibration analysis of SLGSs may
be incorrect. However, it is remarkably difficult to directly measure
the mechanical properties of individual SLGS in the experiment due
to their extremely small size.
2. Simulation method
Motivated by the previous study for SWCNTs [16], MD simulations are now performed on SLGSs to investigate their mechanical
properties. It is found that the effective thickness can be calculated
from direct deflection measurement by using a thin plate model
under bending. The elastic properties, including Youngs moduli,
shear modulus, and Poissons ratios are determined for SLGSs with
armchair and zigzag helicity. SLGSs subjected to uniaxial tension,
shear load and transverse uniform pressure are simulated under
temperature varying from 300 K to 700 K. In our MD simulations,
the large-scale atomic/molecular massively parallel simulator
(LAMMPS) is used. In the LAMMPS, the adaptive intermolecular
reactive empirical bond order potential (AIREBO) is adopted. The
AIREBO potential can be represented by a sum over pairwise inter), LJ
actions, including covalent bonding REBO interactions (EREBO
ij
TORS
terms (ELJ
ij ), and torsion interactions (Ekijl ) [17]

4446

L. Shen et al. / Materials and Design 31 (2010) 44454449

Table 1
Youngs modulus, Poissons ratio and effective thickness values reported by several
authors.
Sources

E (TPa)

h (nm)

Kudin et al. [5]


Arroy and Belytschko [6]

1.029
0.694
0.714

0.149
0.142
0.397

0.339
0.34
0.34

Huang et al. [7]

2.693.81
2.994.23

0.142
0.397

0.08740.0618
0.08110.0574

Reddy et al. [8]


Hemmasizadeh et al. [9]

0.6590.682
0.939

0.3670.416
0.19

0.34
0.1317

Reddy et al. [10]


Armchair sheets
Zigzag sheets

1.0951.125
1.1061.201

0.4450.498
0.4420.465

0.34
0.34

Sakhaee-Pour [11]
Armchair sheets
Zigzag sheets
Chiral sheets

1.042
1.040
0.992

1.285
1.441
1.129

0.34
0.34
0.34

"
#
X X TORS
1 X X REBO
LJ
E
Eij Eij
Ekijl
2 i ji
ki;j li;j;k

where e11 and e22 are the elastic strain in the X and Y directions,
respectively.
Similarly, when the uniaxial tensile force Ny = nf is applied on
the other two opposite edges of the sheet, the in-plane stretching
rigidity, referred to as C22, can be expressed by

C 22 E22 h

C 66 G12 h

3. Results and discussion


Six types of SLGSs with different values of aspect ratio are considered, i.e. an armchair sheet with aspect ratio b = 1.97, 1.44 and
1.01, and a zigzag sheet with aspect ratio b = 1.95, 1.45 and 0.99.
Hence, for example, there are totally 1886 atoms for the armchair
sheet with length Lx = 9.519 nm and width Ly = 4.844 nm, and 1896
atoms for the zigzag sheet length Lx = 9.469 nm and width
Ly = 4.877 nm used in the simulations. In the present work, based
on the classical plate theory, an individual orthotropic plate model
is adopted for SLGSs. Therefore, four independent material parameters are needed to completely describe the elastic behavior, which
are denoted by E11, E22, G12 and v12, respectively, where E11 and E22
are the Youngs moduli, G12 is shear modulus, and v12 is Poissons
ratio. In order to determine these material properties, three MD
simulation tests are carried out. Firstly, the uniaxial tensile force
Nx = nf is applied on the two opposite edges of the sheet (Fig. 1a),
from which the in-plane stretching rigidity, referred to as C11,
can be expressed by

C 11

e22
e11

nfLy
L x DL y

Secondly, the tangential force Nxy = ng is applied on the four edges


of the sheet (Fig. 1b), from which, the in-plane shear rigidity, referred to as C66, can be expressed by

The full expressions for these energy items are given in details in
Stuart et al. [17].
In the frame work of MD simulations, the SLGS can be considered as a congeries of individual atoms. The integration of Newtonian dynamics function is used to determine the variation of the
instantaneous location and velocity of each atom. To account for
the thermal effect, we use the NoseHoover thermostat [18] to
maintain the temperature of the system. This thermostat provides
good conservation of energy and lead to less fluctuation in temperature. The process in our program is run by four steps: (1) to determine the initial locations and velocities of all the atoms, (2) to
optimize and to relax full structure to ensure the equilibrium
geometry being local minimum on the potential energy surfaces,
(3) to apply external load iteratively with the appropriate constraints, and (4) to calculate deformations by using statistical physics techniques.

nfLx
E11 h
Ly DLx

m12

where f is the magnitude of the tensile force at each atom, n is the


number of the atoms that tensile force is applied, DLx is the elongation of Lx, and h is the effective thickness of the SLGS. The Poissons
radio can be measured by

ng

cLy

where g is the magnitude of the tangential force at each atom, c is


the shear strain. Note that now n is the number of the atoms that
tangential force is applied.
From our MD simulations the values of in-plane stretching
rigidities C11 and C22 and in-plane shear rigidity C66 for these six
types of SLGSs are obtained at temperature T = 300, 500 and
700 K and are listed in Table 2. The different trends of C11, C22
and C66 varying with temperature can be explained by the fact that
the different change in bond length and bond angle of SLGSs under
different loading cases. For uniaxial tension, the sheet response to
the external force is largely taken by the changing bond length.
When an SLGS is subjected to a tangential force, the bond length
remains almost unchanged but the change of bond angle is great.
The higher the temperature, the easier it is for SLGSs to undergo
a change in CC length, giving rise to a smaller stretching rigidity.
In contrast, the change in bond angle becomes more difficult at
higher temperature, giving rise to larger shear rigidity.
It is worthy to note that five unknowns (i.e. E11, E22, G12, v12 and
h) cannot be uniquely determined by four Eqs. (2)(5). From Eqs.
(2), (4), and (5) it can be seen that the predicted values of the elastic modulus depend dramatically on the crucial choice of the effective thickness h. However, the effective thickness cannot be
directly obtained from MD simulations since it is only given as a
continuum assumption. In order to determine the effective thickness of SLGSs, a bending test is then carried out at T = 300 K. In this
simulation the transverse uniform load q is applied on each atom
of the SLGS, and the boundary condition is assumed to be simply
supported with no in-plane displacements. As in the case of
SWCNTs, the bending rigidity cannot be read accurately from MD
simulations [19]. Fortunately, the central deflection W can be directly measured, e.g. for the armchair sheet with b = 1.97 we have
W = 0.135 nm when q = 30.4 MPa, and for the zigzag sheet with
b = 1.95 we have W = 0.133 nm when q = 30.6 MPa. From Shen
[20], the loaddeflection relationship for a rectangular plate can
be expressed by

 
 3
qL4x
1 W
3 W
AW
:::
AW
h
h
D11 h

where D11 is the flexural rigidity of the plate in the X direction and
D11 = E11h3/[12(1 m12 m21 )].
From Eq. (6), one has

121m12 m21 qL4x

C 11
4
1
Aw W

where
1

AW

p6

16

3
A3
w W

31=2
5

1 2c4 b2 c3 b4

L. Shen et al. / Materials and Design 31 (2010) 44454449

4447

Fig. 1. A single layer graphene sheet under: (a) uniaxial tension and (b) tangential forces.

and

Table 2
In-plane stretching and shear rigidities of SLGSs in thermal environments.
T (K)

C11 (TPa  nm)

C22 (TPa  nm)

C66 (TPa  nm)

b Lx =Ly ; c3 C 22 =C 11 ; c4 m21 2C 66 =C 11 1 m12 m21

Armchair sheets (Lx = 9.519 nm, Ly = 4.844 nm, b = 1.97)


300
0.314
0.319
500
0.308
0.310
700
0.298
0.301

0.134
0.134
0.139

Armchair sheets (Lx = 6.995 nm, Ly = 4.847 nm, b = 1.44)


300
0.308
0.310
500
0.305
0.306
700
0.302
0.303

0.132
0.134
0.137

Armchair sheets (Lx = 4.888 nm, Ly = 4.855 nm, b = 1.01)


300
0.304
0.306
500
0.303
0.304
700
0.300
0.302

0.132
0.134
0.134

Zigzag sheets (Lx = 9.496 nm, Ly = 4.877 nm, b = 1.95)


300
0.311
0.304
500
0.305
0.298
700
0.300
0.292

0.136
0.139
0.139

Zigzag sheets (Lx = 7.065 nm, Ly = 4.887 nm, b = 1.45)


300
0.308
0.306
500
0.304
0.302
700
0.300
0.299

0.136
0.138
0.139

Zigzag sheets (Lx = 4.855 nm, Ly = 4.888 nm, b = 0.99)


300
0.306
0.304
500
0.304
0.303
700
0.302
0.300

0.132
0.134
0.134

A3
w

g 13 1 18c4 b2 81c3 b4 ; g 13 81 18c4 b2 c3 b4 ;

!
"
3p6
1
c3 b4
1 2c4 b2 c3 b4

g 13
64
g 31
#



2c3 m221 1 c3 b4 4m21 c3 b2
1 m12 m21

c3 m221

10

By substituting MD data into Eq. (7), we find that the effective thickness for armchair sheets is h = 0.129, 0.143 and 0.156 nm, and for
zigzag sheets is h = 0.145, 0.149 and 0.154 nm, respectively. These
values are close to that of a circular graphene sheet of Hemmasizadeh et al. [9], and are assumed to be unchanged in thermal
environments.
Table 3 shows the variation of Youngs and shear moduli for
armchair and zigzag SLGSs with temperature ranging from 300 K
to 700 K. The Youngs modulus E11 for the armchair sheet with
b = 1.97 is computed as 2.43 TPa at T = 300 K, which agrees well
with the Youngs modulus of graphene [7]. The effects of temperature changes on Youngs modulus E11 and shear modulus G12 for
the armchair sheet with b = 1.97 and the zigzag sheet with
b = 1.95 are shown in Fig. 2. It can be found that the predicted values of elastic moduli are essentially influenced by temperature,
while the Youngs and shear moduli show different trends with
the temperature variation. The Youngs moduli E11 and E22 strictly
decrease with increases in temperature. The maximum reduction
of Youngs modulus is about 6% when the temperature ranges from
300 K to 700 K. In contrast, the shear modulus almost remains constant as temperature increases from 300 K to 700 K. As expected,
the temperature dependence of mechanical properties for SLGSs
is in consistent with the thermal behavior of SWCNTs [16]. It
should be mentioned that the Poissons ratio is assumed to be constant in thermal environments. It can also be found that Poissons
ratio depends dramatically on the SLGS chirality. The value of v12
for the armchair sheet is smaller than that of the zigzag sheet.
Similar trend has been predicted by different methods [11]. However, in [11] the values of Poissons ratio are greater than 1.1, and,

4448

L. Shen et al. / Materials and Design 31 (2010) 44454449

Table 3
Elastic properties of SLGSs in thermal environments.
G12 (TPa)

a11 (106/K)

a22 (106/K)

Armchair sheet (Lx = 9.519 nm, Ly = 4.844 nm, h = 0.129 nm, m12 = 0.197)
300
2.434
2.473
500
2.388
2.403
700
2.310
2.333

1.039
1.039
1.078

2.2
1.8
1.9

2.0
1.9
1.9

Armchair sheet (Lx = 6.995 nm, Ly = 4.847 nm, h = 0.143 nm, m12 = 0.202)
300
2.154
2.168
500
2.133
2.140
700
2.112
2.119

0.923
0.937
0.958

2.3
2.0
1.8

2.1
2.1
1.7

Armchair sheet (Lx = 4.888 nm, Ly = 4.855 nm, h = 0.156 nm, m12 = 0.201)
300
1.949
1.962
500
1.942
1.949
700
1.923
1.936

0.846
0.859
0.859

1.9
2.0
2.1

2.1
2.3
2.4

0.938
0.959
0.959

1.7
2.0
2.1

1.5
1.7
2.0

0.913
0.926
0.933

1.8
1.8
1.7

1.6
2.0
1.9

0.857
0.870
0.870

2.1
2.3
2.4

1.9
2.0
2.1

T (K)

E11 (TPa)

E22 (TPa)

Zigzag sheet (Lx = 9.496 nm, Ly = 4.877 nm, h = 0.145 nm,


300
2.145
500
2.103
700
2.069

m12 = 0.223)

Zigzag sheet (Lx = 7.065 nm, Ly = 4.887 nm, h = 0.149 nm,


300
2.067
500
2.040
700
2.013

m12 = 0.204)

Zigzag sheet (Lx = 4.855 nm, Ly = 4.888 nm, h = 0.154 nm,


300
1.987
500
1.974
700
1.961

m12 = 0.205)

2.097
2.055
2.014
2.054
2.027
2.007
1.974
1.968
1.948

3.0

Young's modulus E11 (TPa)

Young's modulus E11 (TPa)

T=500 K
2.5

2.0

1.5

1.0
200

armchair sheet ( = 1.97)


zigzag sheet ( = 1.95)
400

600

armchair sheet
zigzag sheet

0
0.5

800

1.0

T (K)

1.5

2.0

2.5

Lx/Ly

2.0

2.0

Shear modulus G12 (TPa)

Shear modulus G12 (TPa)

T=500 K
1.5

1.0

0.5

0.0
200

armchair sheet ( = 1.97)


zigzag sheet ( = 1.95)
400

600

800

T (K)

1.5

1.0

armchair sheet
zigzag sheet

0.5

0.0
0.5

1.0

1.5

2.0

2.5

Lx/Ly

Fig. 2. The effect of temperature change on: (a) Youngs modulus and (b) shear
modulus for armchair and zigzag sheets.

Fig. 3. The effect of aspect ratio on: (a) Youngs modulus and (b) shear modulus for
armchair and zigzag sheets at 500 K.

therefore, these values are thoroughly inappropriate to SLGSs. It is


worthy to note that since v12 is less than v21, the Youngs modulus
E22 should be greater than E11 for the armchair sheet, while for the

zigzag sheet we still have E11 > E22. The thermal expansion coefficients a11 and a22 are also listed in Table 3 which can readily be obtained from MD simulation tests. It is found that the thermal

L. Shen et al. / Materials and Design 31 (2010) 44454449

expansion coefficients are also anisotropic and temperaturedependent.


Fig. 3 presents the effect of aspect ratio b on the Youngs modulus E11 and shear modulus G12 of armchair and zigzag sheets at
T = 500 K. As can be seen from the results both Youngs modulus
E11 and shear modulus G12 increase as aspect ratio increases. The
effect of aspect ratio on the Youngs modulus E11 and shear modulus G12 of zigzag sheets is less than that of armchair sheets. Note
that the effective thickness h is decreased with increase in the aspect ratio of SLGSs. These results confirm that the material properties of SLGSs are also size-dependent.
4. Conclusions
In summary, on the basis of the MD simulations, we have investigated the elastic properties of six types of SLGSs over a wide temperature range from 300 K to 700 K. By direct measuring the
bending deflections, the effective thickness of SLGSs can be determined uniquely. The results reveal that the Youngs modulus depends dramatically on the temperature change, and it decreases
as the temperature increases. In contrast, the shear modulus depends weakly on temperature change. The results confirm that
the material properties of SLGSs are also size-dependent.
Acknowledgment
The support for this work, provided by the National Natural Science Foundation of China under Grant 10802050, is gratefully
acknowledged.
References
[1] Li C, Thostenson ET, Chou T-W. Sensors and actuators based on carbon
nanotubes and their composites: a review. Compos Sci Technol
2008;68:122749.

4449

[2] Schedin F, Geim AK, Morozov SV, Hill EW, Blake P, Katsnelson MI, et al.
Detection of individual gas molecules adsorbed on grapheme. Nat Mater
2007;6:6525.
[3] Bunch JS, van der Zande AM, Verbridge SS, Frank IW, Tanenbaum DM, Parpia
JM, et al. Electromechanical resonators from graphene sheets. Science
2007;315:4903.
[4] Stankovich S, Dikin DA, Dommett GHB, Kohlhaas KM, Zimney EJ, Stach EA,
Piner RD, Nguyen ST, Ruoff RS. Graphene-based composite materials. Nature
2006;442:2826.
[5] Kudin KN, Scuseria GE, Yakobson IB. C2F, BN, and C nanoshell elasticity from ab
initio computations. Phys Rev B 2001;64:235406.
[6] Arroyo M, Belytschko T. Finite crystal elasticity of carbon nanotubes based on
the exponential CauchyBorn rule. Phys Rev B 2004;69:115415.
[7] Huang Y, Wu J, Hwang KC. Thickness of graphene and single-wall carbon
nanotubes. Phys Rev B 2006;74:245413.
[8] Reddy CD, Rajendran S, Liew KM. Equilibrium configuration and continuum
elastic properties of finite sized graphene. Nanotechnology 2006;17:86470.
[9] Hemmasizadeh A, Mahzoon M, Hadi E, Khandan R. A method for developing
the equivalent continuum model of a single layer graphene sheet. Thin Solid
Films 2008;516:763640.
[10] Reddy CD, Rajendran S, Liew KM. Equilibrium continuum modeling of
graphene sheets. Int J Nanosci 2005;4:6316.
[11] Sakhaee-Pour A. Elastic properties of single-layered graphene sheet. Solid
State Commun 2009;149:915.
[12] Scarpa F, Adhikari S, Phani AS. Effective elastic mechanical properties of single
layer graphene sheets. Nanotechnology 2009;20:065709.
[13] Lee C, Wei X, Kysar JW, Hone J. Measurement of the elastic properties and
intrinsic strength of monolayer graphene. Science 2008;321:3858.
[14] Jin Y, Yuan FG. Simulation of elastic properties of single-walled carbon
nanotubes. Compos Sci Technol 2003;63:150715.
[15] Chang T, Geng J, Guo X. Chirality- and size-dependent elastic properties of
single-walled carbon nanotubes. Appl Phys Lett 2005;87:251929.
[16] Zhang C-L, Shen H-S. Temperature-dependent elastic properties of singlewalled carbon nanotubes: prediction from molecular dynamics simulation.
Appl Phys Lett 2006;89:081904.
[17] Stuart SJ, Tutein AB, Harrison JA. A reactive potential for hydrocarbons with
intermolecular interactions. J Chem Phys 2000;112:647286.
[18] Hoover WG. Canonical dynamics: equilibrium phase-space distributions. Phys
Rev A 1985;31:16957.
[19] Wang CY, Zhang LC. A critical assessment of the elastic properties and effective
wall thickness of single-walled carbon nanotubes. Nanotechnology
2008;19:075705.
[20] Shen H-S. Nonlinear analysis of composite laminated thin plates subjected to
lateral pressure and thermal loading and resting on elastic foundations.
Compos Struct 2000;49:11528.

Você também pode gostar