Sunteți pe pagina 1din 9

Home

Search

Collections

Journals

About

Contact us

My IOPscience

Effects of temperature and strain rate on the mechanical properties of hexagonal boron nitride
nanosheets

This content has been downloaded from IOPscience. Please scroll down to see the full text.
2014 J. Phys. D: Appl. Phys. 47 025303
(http://iopscience.iop.org/0022-3727/47/2/025303)
View the table of contents for this issue, or go to the journal homepage for more

Download details:
IP Address: 103.37.201.23
This content was downloaded on 23/03/2015 at 11:41

Please note that terms and conditions apply.

Journal of Physics D: Applied Physics


J. Phys. D: Appl. Phys. 47 (2014) 025303 (8pp)

doi:10.1088/0022-3727/47/2/025303

Effects of temperature and strain rate on


the mechanical properties of hexagonal
boron nitride nanosheets
Tongwei Han, Ying Luo and Chengyuan Wang
School of Civil Engineering and Mechanics, Jiangsu University, Jiangsu Zhenjiang 212013,
Peoples Republic of China
E-mail: chengyuan.wang@gmail.com (C Wang)
Received 28 August 2013, revised 4 November 2013
Accepted for publication 7 November 2013
Published 11 December 2013
Abstract

The effect of temperature and strain rate on mechanical properties remains an open topic in
research of hexagonal boron nitride (h-BN) nanosheets. To examine these fundamental issues
we have performed molecular dynamics simulations to record the stressstrain curves in
tensile tests and measure Youngs modulus, fracture strength and fracture strain in armchair
and zigzag directions. Comparing the results obtained at different temperatures and strain rates
we have quantified the effects of the two factors on the tensile properties of the h-BN
nanosheets. The influence of crystal orientation is also examined in the present study. It is
found that the h-BN nanosheets are basically an anisotropic material whose tensile properties
vary substantially with temperature and strain rate. In particular, a yielding platform is
observed for the h-BN nanomaterial at relatively low temperature.
Keywords: hexagonal boron nitride, temperature, strain rate, mechanical properties,
molecular dynamics
(Some figures may appear in colour only in the online journal)
of 2D h-BN nanosheets has also started very recently. Bosak
et al [24] measured the elastic moduli (811 GPa) of h-BN
based on inelastic x-ray scattering measurements. Using
atomic force microscopy, Golbergs group [14] found that
the bending modulus of the h-BN sheets increases with
decreasing thickness and approaches an asymptotic value of
monolayer BN sheets when the thickness is below 50 nm. In
addition, notably improved thermal and mechanical properties
have been observed for the polymer composites reinforced
by BN nanosheets [19]. In addition to the experimental
work, theoretical studies have been conducted to simulate the
behaviour of h-BN nanosheets. Ohba et al [25] reported the
elastic constants and bulk modulus based on first-principles
calculations. The ab initio studies were also carried out by
several groups [2629], where Youngs modulus achieved for
h-BN nanosheets fell in the range of [809.1 GPa, 846.1 GPa].
More recently, Bohayra Mortazavi et al [30] investigated the
tensile response of single layer BN sheets using molecular
dynamics simulation (MDS). The obtained elastic moduli are
close to those of BN nanotubes, which vary between 0.8 and
0.85 TPa depending on the chirality. A Youngs modulus of
716.3 GPa was also obtained for h-BN nanosheets in Zhao

1. Introduction

Similar to graphene [1], other single-atom nanosheets, e.g.,


h-BN, WS2 , MoS2 and NbSe2 [24], have received considerable attention due to their unique mechanical, electrical and
chemical properties, and potential applications in nanotechnology [58]. Among them, h-BN nanosheets are one of the
most intriguing two-dimensional (2D) nanostructures, where
B and N atoms are organized in hexagonal rings. The boron
nitride nanotubes (BNNTs) have also been synthesized, which
can be imaged as a result of rolling up the h-BN nanosheets.
The most distinctive advantages of BN nanomaterials
over their carbon counterparts are that the BN systems
are electrically insulating [911] and more stable at high
temperature and various chemical environments [12]. In the
meantime, the two systems are comparable in terms of thermal
conductivity and mechanical robustness [1317]. These
unique features of BN nanomaterials promise great potential
for constructing nanodevices [18], functional nanocomposites
[19, 20], hydrogen accumulators, etc [21, 22]. As reviewed in
[23], a great deal of effort has already been made to quantify the
mechanical properties of BNNTs. The study of the mechanics
0022-3727/14/025303+08$33.00

2014 IOP Publishing Ltd Printed in the UK

J. Phys. D: Appl. Phys. 47 (2014) 025303

T Han et al

et als MDS [31]. So far, to the best of our knowledge,


the effects of charility, temperature and strain rate on the
mechanical properties of h-BN nanosheets have not yet been
studied, which, however, are shown to considerably alter
Youngs modulus, fracture strength and strain of their carbon
counterparts [3234]. It is thus of great interest to examine
further these fundamental issues for the h-BN nanosheets.
In the present paper, the tensile properties were measured
for monolayer h-BN nanosheets based on MDS at different
temperatures and strain rates. The stressstrain relation was
studied and the dependences of Youngs modulus, fracture
strength and strain on temperature and strain rate were
obtained for armchair and zigzag h-BN nanosheets. The
results are expected to provide important guidelines for the
practical applications of h-BN nanosheets in nanodevices and
nanoelectronics.

Table 1. Tersoff-like interatomic potential parameters for


BN [44, 45].

2. MDS on h-BN sheets

between boron and nitrogen atoms [4049]. For instance,


Sekkal et al [40] and Verma et al [41] slightly changed
the parameters corresponding to carbon systems [36, 37] in
order to fit c-BN and BNNTs, respectively. Here the boron
nitride was treated as a one-component system, where the
BN interactions were neglected. Matsunaga et al [42, 43]
considered c-BN as a real two-component system with BN
interactions. The Tersoff potential parameters of B were then
achieved and used to simulate cubic boron carbonitrides. Albe
et al [44, 45] developed the Tersoff-like potential for BN,
which is able to describe sp2 structures of BN polymorphs,
BN clusters and pure B and N bonding. These potential
parameters have been successfully employed for describing the
bonding interactions and investigating the structural, thermal
and mechanical properties of BNNTs [5054], h-BN nanosheet
[55] and BN nanobamboos [56]. Recently, based on the
force-matching method etc, the adjusted Tersoff-like potential
parameters were obtained for boron nitride [4649] by fitting
the obtained bond length and cohesive energy to experimental
data. In the present study, the Tersoff-like potential with
the parameters proposed by Albe et al [44, 45] was chosen
to describe the interactions between B and N atoms. These
parameters are shown in table 1, which were efficiently
used in studying BN nanostructures of various configurations
previously. Here different coefficients are identified in table 1
due to the distinct definitions of these coefficients given by
Albe et al [44, 45] and Tersoff [35].
To study the mechanical properties of h-BN nanosheets
under various temperatures and loading conditions, MDS
were implemented by using LAMMPS (large-scale atomic/
molecular massively parallel simulator) open-source package
[57], developed by Sandia National Laboratories. The
schematics of a rectangular h-BN nanosheet investigated in this
study is shown in figure 1. The ochre and blue atoms represent
boron and nitrogen atoms, respectively. The model consists
of 3720 atoms and its dimensions are around 100 100 .
The armchair and zigzag directions are oriented along the X
and Y axis, respectively. The tested h-BN nanosheet was first
optimized by using the conjugate gradient method to fully
relax and reach its equilibrium state. Subsequently, a uniaxial
tension was applied to stretch the optimized nanosheets along

3
c
d
h
n

2
B
R
D
1
A

Many empirical model potentials have been developed for


semiconductors, among which Tersoff [3537] and Tersoff
Brenner (Tersoff-like) bond order potentials [38, 39] are
most successful in describing the interaction between boron
and nitrogen atoms in boron nitride nanosystems. The
TersoffBrenner potential takes the same form as that of
the Tersoff potential, except for repulsive and attractive pair
potentials. In the Tersoff potential, the energy E, between two
neighbouring atoms i and j , has the form [35]

1
Ei =
Vij
E=
2 i=j
i
with
Vij = fC (rij )[fR (rij ) + bij fA (rij )],
where

fR (r) = Ae1 r
fA (r) = Be2 r

1,
fC (r) = 21 21 sin

0,

(rR)
,
2D

r <RD
RD <r <R+D
r > R + D.

Here fR (r) and fA (r) are the repulsive and attractive pair
potentials, respectively. The cutoff function fC (r) is defined to
limit the range of the potential and thus save the computational
resources required in the MDS. Normally the cutoff distance R
is chosen to include only the first-neighbour shell. In addition,
bij is the bond order function that determines the strength of
the attractive term. It takes the form below.
bij = (1 + n ijn ) 2n

m
m
ij =
fC (rij )g(ij k )e3 (rij rik )
1

k=i,j

g( ) = 1 +

c2
c2
.

d2
(d 2 + (h cos )2 )

In the literature, several sets of Tersoff and Tersoff-like


potential parameters have been developed for the interaction
2

BN-interaction

NN-interaction

BB-interaction

3
1
1.9925
1092.9287
12.38
0.5413
0.364 153 367
0.000 011 134
2.784 247 207
3613.431 337
2
0.1
2.998 355 817
4460.833 973

3
1
0
17.7959
5.9484
0
0.618 4432
0.019 251
2.627 272 104
2563.560 342
2
0.1
2.829 309 329
2978.952 79

3
1
0
0.526 29
0.001 587
0.5
3.992 9061
0.000 0016
2.077 498 242
1173.196 962
2
0.1
2.237 257 857
1404.475 204

J. Phys. D: Appl. Phys. 47 (2014) 025303

T Han et al

the chosen interatomic potential. The results are summarized


in table 2 in comparison with all available theoretical and
experimental data. As will be shown later, the h-BN
nanosheets exhibit different Youngs modulus and fracture
strength in armchair and zigzag directions. In spite of this,
some authors reported the values without giving the specific
direction [2529, 46, 62]. Thus, to compare with these data,
the Youngs modulus, fracture strength and strain obtained for
armchair and zigzag directions in the present study and [30, 31]
were averaged in table 2. First it is noted in the table that the
Youngs modulus of 881.1 GPa given by the present MDS at
very low strain rate and room temperature falls right in the
middle of the range of 716 to 950 GPa achieved in an inelastic
x-ray scattering measurement [24]. This value is also found
very close to that of boron nitride nanotubes measured using the
electric-field-induced resonance method inside a transmission
electron microscope [63]. Secondly, a fracture strength of
133.3 GPa is obtained in this work, which is comparable to
the 165 and 120.4 GPa calculated previously based on MDS
[30, 31]. Moreover, it is also noted in table 2 that the fracture
strain given by our MDS at room temperature reaches 0.332
(averaged 0.260 for armchair and 0.405 for zigzag) very close
to 0.302 [30] (the average of 0.282 and 0.321) and 0.280 [31]
calculated in previous MDS. From these comparison results,
it is followed that the present MDS is reliable and efficient
in calculating the tensile properties of the h-BN nanosheets
up to a large strain of more than 0.3. Here we would like to
point out that, in measuring the fracture strength and strain, a
direct comparison cannot be made between the present study
and experiments (or quantum mechanics simulations) due to
lack of data in the literature.

Figure 1. The simulation model of the h-BN nanosheets tested in


this study. The model size is 100 100 . Here boron and
nitrogen atoms are represented by ochre and blue balls, respectively.

armchair or zigzag directions. A periodic boundary condition


(PBC) is applied along the X and Y directions, which shows
that the simulations are representative of BN sheets with
infinite geometric size. The equations of motion were solved
using the velocity-Verlet algorithm [58] with a time step of
1 fs. In individual simulations, the temperature was kept
at the specified value by using the NoseHoover method
[59, 60]. The temperature effect was examined at the constant
engineering strain rate 1 109 s1 , while the study of the
strain rate effects was conducted at room temperature. To
obtain the stressstrain curves, the viral stress in the loading
direction was calculated at each strain level. Youngs modulus
was represented by the slope of the linear part of the curves,
and the fracture strength and strain were taken as the maximal
stress and strain before the onset of fracture. It should be
pointed out that the equivalent Youngs modulus of 2D h-BN
nanosheets scale with their effective thickness. In other words,
the specific value of Youngs modulus is associated with the
effective thickness selected. In the present study, the effective
thickness of h-BN nanosheets was taken as 0.333 nm, which
was used previously [41].

3.2. Thermal effect

In this section we studied the thermal effect on the tensile


properties of the h-BN nanosheets by performing tensile tests
for zigzag and armchair h-BN nanosheets. In doing this, the
temperature increased from 0 to 2000 K while the strain rate
was kept constant at 1 109 s1 .
In the tensile tests, the stressstrain
curves were recorded and shown in figures 2(a) and (b) for
armchair and zigzag h-BN nanosheets, respectively. One of
the interesting results is the yielding platform or material flow
observed at low temperature, e.g., T  400 K. In this case,
figure 2 shows that the stress first increases with rising strain at
a relatively low rate; when it reaches a critical value (yield ), the
stress remains constant while the strain increases substantially.
In other words, the material flows without raising the external
loading further. This phenomenon indicates yielding the h-BN
nanosheets. Subsequently, when the strain exceeds a certain
value the regular stressstrain relation resumes, where the
stress rises with growing strain at a rate even greater than that of
the first stage. In figures 2, the yielding strength yield does not
significantly change with temperature but varies considerably
with the chirality of the BN nanosheets. For example, yield is
around 75.5 GPa for the armchair sheets but increases to around
93.5 GPa for zigzag BN sheets. Here the length of the yielding

3.2.1. Tensile behaviour.

3. Results and discussion

Based on the MDS technique and the chosen interatomic


potentials, the tensile tests have been performed for zigzag and
armchair h-BN sheets at different temperatures and strain rates.
The fundamental mechanical properties, such as Youngs
modulus, fracture strength and strain, have been quantified
for the tested sample sheets and shown graphically against
temperature or strain rates. In particular, the MDSs used here
are compared with previously used methods to demonstrate
its efficiency and accuracy in studying the nanomechanics of
h-BN nanosheets.
3.1. Validation of the present MDS

To validate the present MDS we have calculated Youngs


modulus and the fracture strength of h-BN nanosheets based on
3

J. Phys. D: Appl. Phys. 47 (2014) 025303

T Han et al

Table 2. Comparison of the calculated Youngs modulus and strength of h-BN nanosheets with previous experiments and calculations.
Effective thickness of 0.333 nm of h-BN nanosheets is adopted when dealing with the published data.

Reference

Year

Method

Youngs modulus
(GPa)

Present study
Mortazavi et al [30]
Zhao et al [31]
Bosak et al [24]
Green et al [61]
Ohba et al [25]
Kudin et al [26]
Topsakal et al [27]
Peng et al [28]
Mirnezhad et al [29]
Eun-Suok Oh [46]
Boldrin et al [62]

2013
2012
2013
2006
1976
2001
2001
2010
2012
2013
2011
2011

Molecular dynamics
Molecular dynamics
Molecular dynamics
Inelastic x-ray scattering
Lennard-Jones and electrostatic potentials
First-principles calculations
ab initio
ab initio
ab initio
ab initio
Continuum lattice approach
Molecular mechanics(SH-UFF)

881.1
825
716.3
811
802.5
951.5
810
809.1
846.1
829
977.2
797.0

Strength (GPa)

Strain

133.2
165
120.4

0.332
0.302
0.280

Figure 3. A hexagonal (h) ring elongated along the armchair


direction. Solid lines represent the original h-ring and dashed lines
represent the deformed configuration.

On the other hand, when higher temperature is involved,


e.g., T > 400 K, the length of the yielding platform turns out to
be very small and the stress keeps increasing with rising strain
until the onset of fracture. Nevertheless, the conversion of the
shape of the stressstrain curves from convex to concave can
always be observed in figure 2, showing a softeninghardening
transition of the h-BN nanosheets at the vicinity of the point
(75 GPa, 0.16). As a matter of the fact, such a transition
zone is also observed for some macroscopic metallic materials
and explained by sliding and strengthening of dislocations.
This mechanism however cannot be validated in the present
simulations as the breakage or formation of chemical bonds
(i.e., the nucleation of defects) has not been observed in the
h-BN nanosheets before the onset of fracture. As illustrated
in figure 3, the present MDS showed that the stretching of the
nanosheets is implemented mainly via the elongations of the
bond lengths AB and ED, and the rotation of the bond angles
and . These bond lengths and angles can be considered as
equivalent nonlinear (line or angle) springs, whose behaviour
is entirely controlled by the interatomic potentials in h-BN
nanosheets. Thus the above-mentioned convexconcave
conversion in figure 2 reflects the softeningstrengthening
process of the equivalent springs (e.g., the increase and
decrease of the variable spring constants), which can only
be understood via calculation based on molecular mechanical
theory instead of morphological or structural analyses. Indeed
the physical mechanisms of this unique feature deserve to be
studied further in the near future.

Figure 2. The stressstrain curves obtained for (a) armchair and


(b) zigzag h-BN nanosheets in uniaxial tensile tests. The strain rate
is 1 109 s1 and temperature rises from 0 K to 2000 K.

platform can be measured by the associated increment of the


strain during the material yielding. At T = 0, this length
is found to increase from around 10% of the armchair sheet
(figure 2(a)) to around 20% of the zigzag sheet (figure 2(b)).

Based on the stressstrain curves


shown in figure 2 we have calculated some key material

3.2.2. Tensile properties.

J. Phys. D: Appl. Phys. 47 (2014) 025303

T Han et al

Figure 4. Temperature dependence of Youngs modulus obtained


for h-BN nanosheets. The strain rate is kept constant at 1 109 s1 .

Figure 6. Temperature dependence of the fracture strain obtained


for h-BN nanosheets. The strain rate is kept constant at 1 109 s1 .

carbon nanotubes and graphene sheets consisting of hexagonal


carbon rings but has to be taken into consideration for h-BN
sheets where different atoms (i.e., boron and nitride atoms) are
found.
The decrease of fracture strength with rising temperature is
also observed in figure 5. The reduction of the fracture strength
is around 100 GPa as the temperature rises from 0 to 2000 K.
This is true regardless of the chirality of the nanosheets. In
the meantime, as seen from figure 5, the armchair sheets
generally exhibit a fracture strength greater than that of the
zigzag sheets. This chirality effect on the fracture strength
is found to be substantial at T < 1500 K, and leads to the
maximum difference of 37 GPa in fracture strength between
the armchair and zigzag sheets. At higher temperature, i.e.,
T > 1500 K, the fracture strengths obtained for the two types
of nanosheets are close to each other. The result shows that
the fracture strength becomes less sensitive to the chirality at
high temperature.
A similar trend of fracture strain is shown in figure 6
but its behaviour is more complicated than those of Youngs
modulus (figure 4) and fracture strength (figure 5). As shown
in figure 6, the fracture strain of zigzag sheets is greater than
that of armchair sheets at low temperature. At the same
time, the greater rate of decrease with temperature is also
achieved for the fracture strain in the zigzag directions. These
finally lead to the distinct behaviour of the fracture strain,
i.e., at low temperature T < 750 K, the zigzag sheets exhibit
higher fracture strain whereas, at relatively high temperature
T > 750 K, the fracture strain of the armchair sheets becomes
even greater than that of the zigzag sheets. Considering
the uncertainty in the MDS, we believe that the sensitivity
of the ultimate strain to temperature does not significantly
change throughout the range of temperatures considered for
both armchair and zigzag h-BN sheets. But similar to
Youngs modulus (figure 4) and ultimate strength (figure 5),
the rate of change in the ultimate strain with temperature alters
considerably with the crystal directions, i.e., the armchair and
zigzag directions. This finally leads to the intersection of the
two curves at around 750 K in figure 6.

Figure 5. Temperature dependence of the fracture strength obtained


for h-BN nanosheets. The strain rate is kept constant at 1 109 s1 .

properties for the sample sheets at a temperature rising from 0


to 2000 K. The temperature dependence of Youngs modulus,
fracture strength and strain is plotted in figures 4, 5 and 6,
respectively, for armchair and zigzag h-BN nanosheets. To
confirm the reliability of the present MDS, the compression
tests in the elastic region were also performed on the h-BN
nanosheets at 0 K. The obtained Youngs modulus is shown in
the inset of figure 2, which has a magnitude nearly identical to
that obtained in tensile tests at the same temperature.
In figure 4 it is noted that Youngs modulus of the h-BN
nanosheets generally decreases with increasing temperature.
The average rate of change is about 121 MPa K1 almost
independent of the nanosheet chirality. In the meantime,
Youngs modulus of zigzag sheets is found to be greater than
that of armchair sheets. The difference is around 30 GPa at
all temperatures considered, showing the strong anisotropy
of the h-BN nanosheets. Indeed, the hexagonal structure
generally leads to anisotropic (overall) material properties
with the largest discrepancy found between the armchair and
zigzag directions. Such a discrepancy is usually neglected for
5

J. Phys. D: Appl. Phys. 47 (2014) 025303

T Han et al

Figure 9. Fracture strain of h-BN nanosheets as a function of strain


rate. The results are obtained at room temperature.

Figure 7. Youngs modulus of h-BN nanosheets as a function of

strain rate. The results are obtained at room temperature.

Specifically, in the range of high strain rate ( > 109 s1 ),


Youngs modulus of the zigzag sheets becomes greater than its
armchair counterpart.
Subsequently we examined the strain rate effect on the
fracture strength in figure 8. It is seen from the figure that,
in general, the fracture strength tends to grow with increasing
strain rate . Also consistent with figure 5, figure 8 shows that
the fracture strength of the armchair sheets is greater than that
of the zigzag sheets. The difference in the fracture strength
however decreases monotonically with increasing strain rate
and almost vanishes when the strain rate approaches 1011 s1 .
To give some idea, at = 5 107 s1 , the fracture strength of
the armchair sheets is around 17 GPa greater than the strength
of the zigzag sheets. But the difference reduces to less than
7 GPa when rises to 6 109 s1 .
Finally, we investigated the effect of strain rate on the
fracture strain in figure 9. The results indicate that the
fracture strain of the armchair and zigzag nanosheets increases
significantly with increasing strain rate. The fracture strain
of the zigzag sheets is larger than its armchair counterpart
and the difference due to the chirality effect remains nearly
unchanged throughout the length of the strain rate considered.
To reveal the possible coupling effect of temperature and strain
rate on the mechanical properties of h-BN nanosheets, as
shown in figure 10, Youngs modulus and fracture strength
were calculated as functions of strain rate at temperatures
T = 300 K, 900 K and 1500 K, respectively. The results
suggest that temperature change does not significantly alter
the tendency of the material properties to change with strain
rate. In the meantime, the figure also shows that the thermal
effect on the material properties remains qualitatively similar
at different strain rates. It is thus concluded that the coupling
effect between temperature and strain rate is trivial for the
tensile properties of h-BN nanosheets.

Figure 8. Fracture strength of h-BN nanosheets as a function of


strain rate. The results are obtained at room temperature.

3.3. Strain rate effect

It is well known that in a tensile test the strain rate can


substantially affect the values of the material properties
measured. It is thus expected that such an effect of strain rate
would also be significant for the h-BN nanosheets. To examine
this issue, the tensile tests were performed for the armchair
and zigzag nanosheets at room temperature and strain rate
increasing from 107 to 1011 s1 . The strain rate dependence
of Youngs modulus, fracture strength and fracture strain is
shown in figures 7, 8 and 9, respectively.
As shown in figure 7, at low strain rate i.e., < 108 s1 ,
Youngs modulus of the h-BN nanosheets is nearly a constant
(around 880 GPa) independent of strain rate as well as chirality.
It then declines slightly when rises from 108 to 109 s1 .
At a higher strain rate > 109 s1 , Youngs modulus of the
h-BN nanosheets is found to decrease rapidly with rising strain
rate. In the process when the strain rate rises from 109 to
1011 s1 , Youngs modulus of the zigzag sheets declines by
13.2% while that of the armchair sheets decreases by 16.8%.

4. Conclusions

The tensile tests have been performed based on MDS for


armchair and zigzag h-BN nanosheets at different temperatures
6

J. Phys. D: Appl. Phys. 47 (2014) 025303

T Han et al

decrease is negligible for Youngs modulus, small for the


fracture strength but more pronounced for the fracture
strain. In addition, a yielding platform is observed for
h-BN nanosheets at low temperature (e.g., T < 400 K)
but vanishes at relatively high temperature (T  400 K).
(3) The higher strain rate generally leads to lower Youngs
modulus but higher fracture strength and strain of the h-BN
nanosheets. This effect on Youngs modulus is negligible
at small strain rate (e.g., 108 s1 ) but becomes substantial
at relatively high strain rate. The effect on the fracture
strength and strain is always significant throughout the
length of the strain rate considered. In addition, the strain
rate effect varies significantly with crystal orientations for
Youngs modulus and fracture strength but remains nearly
unchanged for the fracture strain.
Acknowledgments

This work is financially supported by the Natural Science


Foundation of the Jiangsu Higher Education Institutions of
China (12KJB130001), Natural Science Foundation of Jiangsu
Province (BK2011490) and NSFC (21204031).
References
[1] Novoselov K S, Geim A K, Morozov S V, Jiang D, Zhang Y,
Dubonos S V, Grigorieva I V and Firsov A A 2004 Science
306 6669
[2] Novoselov K S, Jiang D, Schedin F, Booth T J,
Khotkevich V V, Morozov S V and Geim A K 2005 Proc.
Natl Acad. Sci. USA 102 104513
[3] Wang Y, Zhou C H, Wang W C and Zhao Y P 2013 Indust.
Eng. Chem. Res. 52 437982
[4] Matte H S S R, Gomathi A, Manna A K, Late D J, Datta R,
Pati S K and Rao C N R 2010 Angew. Chem. Int. Edn Engl.
49 405962
[5] Geim A K and Novoselov K S 2007 Nature Mater. 6 18391
[6] Nag A, Raidongia K, Hembram K P S S, Datta R,
Waghmare U V and Rao C N R 2010 Acs Nano 4 153944
[7] Zeng H B, Zhi C Y, Zhang Z H, Wei X L, Wang X B, Guo W L,
Bando Y and Golberg D 2010 Nano. Lett. 10 504955
[8] Hwang W S et al 2012 Appl. Phys. Lett. 101 013107
[9] Rubio A, Corkill J L and Cohen M L 1994 Phys. Rev. B
49 50814
[10] Blase X, Rubio A, Louie S G and Cohen M L 1994 Europhys.
Lett. 28 33540
[11] Zhang Z H, Guo W L and Dai Y T 2009 J. Appl. Phys.
105 084312
[12] Chen Y, Zou J, Campbell S J and Le Caer G 2004 Appl. Phys.
Lett. 84 24302
[13] Chopra N G and Zettl A 1998 Solid State Commun.
105 297300
[14] Li C, Bando Y, Zhi C Y, Huang Y and Golberg D 2009
Nanotechnology 20 385707
[15] Dumitrica T, Bettinger H F, Scuseria G E and Yakobson B I
2003 Phys. Rev. B 68 085412
[16] Golberg D, Bando Y, Huang Y, Terao T, Mitome M, Tang C C
and Zhi C Y 2010 Acs Nano 4 297993
[17] Arenal R, Wang M S, Xu Z, Loiseau A and Golberg D 2011
Nanotechnology 22 265704
[18] Dean C R et al 2010 Nature Nanotechnol. 5 7226
[19] Zhi C Y, Bando Y, Tang C C, Kuwahara H and Golberg D
2009 Adv. Mater. 21 288993

Figure 10. (a) Youngs modulus and (b) the fracture strength of the
h-BN nanosheets as a function of strain rate at temperatures
T = 300, 900 and 1500 K.

and strain rates. The tensile response has been studied and the
effects of temperature and strain rate have been examined for
the tensile properties of h-BN nanosheets. The new findings
obtained in the present simulations are as follows.
(1) The single-atom layer h-BN nanosheets are basically
anisotropic materials with Youngs modulus, fracture
(yielding) strength and fracture strain varying substantially with crystal orientations. In the armchair direction,
the h-BN nanosheets exhibit lower Youngs modulus but
higher fracture strength than those in the zigzag direction.
Lower fracture strain is achieved in the armchair direction
at low temperature but it turns out to be even greater than
that along the zigzag direction at high temperature.
(2) Raising temperature generally exerts a softening effect
on the h-BN nanosheets. As the temperature rises from
0 to 2000 K, Youngs modulus, fracture strength and
fracture strain decrease by around 26.5%, 62.1% and
68.2%, respectively. The chirality effect on the rate of
7

J. Phys. D: Appl. Phys. 47 (2014) 025303

T Han et al

[20] Gao Y W, Gu A J, Jiao Y C, Yang Y L, Liang G Z, Hu J T,


Yao W and Yuan L 2012 Polym. Adv. Technol. 23 91928
[21] Lu F S, Wang F, Cao L, Kong C Y and Huang X C 2012
Nanosci. Nanotechnol. Lett. 4 94961
[22] Pakdel A, Zhi C Y, Bando Y and Golberg D 2012 Mater. Today
15 25665
[23] Ghassemi H M and Yassar R S 2010 Appl. Mech. Rev.
63 020804
[24] Bosak A, Serrano J, Krisch M, Watanabe K, Taniguchi T and
Kanda H 2006 Phys. Rev. B 73 041402
[25] Ohba N, Miwa K, Nagasako N and Fukumoto A 2001 Phys.
Rev. B 63 115207
[26] Kudin K N, Scuseria G E and Yakobson B I 2001 Phys. Rev. B
64 235406
[27] Topsakal M, Cahangirov S and Ciraci S 2010 Appl. Phys. Lett.
96 091912
[28] Peng Q, Ji W and De S 2012 Comput. Mater. Sci. 56 117
[29] Mirnezhad M, Ansari R and Rouhi H 2013 Superlatt.
Microstruct. 53 22331
[30] Mortazavi B and Remond Y 2012 Physica E 44 184652
[31] Zhao S J and Xue J M 2013 J. Phys. D. Appl. Phys. 46 135303
[32] Wei C Y, Cho K J and Srivastava D 2003 Phys. Rev. B
67 115407
[33] Zhao H and Aluru N R 2010 J. Appl. Phys. 108 064321
[34] Yi L J, Yin Z N, Zhang Y Y and Chang T C 2013 Carbon
51 37380
[35] Tersoff J 1988 Phys. Rev. B 37 69917000
[36] Tersoff J 1989 Phys. Rev. B 39 55668
[37] Tersoff J 1988 Phys. Rev. Lett. 61 287982
[38] Brenner D W 1990 Phys. Rev. B 42 945871
[39] Brenner D W, Shenderova O A, Harrison J A, Stuart S J, Ni B
and Sinnott S B 2002 J. Phys.: Condens. Matter
14 783802
[40] Sekkal W, Bouhafs B, Aourag H and Certier M 1998 J. Phys.:
Condens. Matter 10 497584
[41] Verma V, Jindal V K and Dharamvir K 2007 Nanotechnology
18 435711

[42] Matsunaga K, Fisher C and Matsubara H 2000 Japan. J. Appl.


Phys. 39 L4851
[43] Matsunaga K and Iwamoto Y 2001 J. Am. Ceram. Soc.
84 22139
[44] Albe K, Moller W and Heinig K H 1997 Radiat. Eff. Defects
Solids 141 8597
[45] Albe K and Moller W 1998 Comput. Mater. Sci. 10 1115
[46] Oh E S 2010 Mater. Lett. 64 85962
[47] Liao M L, Wang Y C, Ju S P, Lien T W and Huang L F 2011
J. Appl. Phys. 110 054310
[48] Kinaci A, Haskins J B, Sevik C and Cagin T 2012 Phys. Rev. B
86 115410
[49] Sevik C, Kinaci A, Haskins J B and Cagin T 2011 Phys. Rev. B
84 085409
[50] Moon W H and Hwang H J 2004 Nanotechnology 15 4314
[51] Ebrahimi-Nejad S and Shokuhfar A 2013 Physica E
50 2936
[52] Song J, Wu J, Huang Y and Hwang K C 2008 Nanotechnology
19 445705
[53] Ali Shokuhfar A and Ebrahimi-Nejad S 2013 Physica E
48 5360
[54] Shen H 2009 Comput. Mater. Sci. 47 2204
[55] Slotman G J and Fasolino A 2013 J. Phys.: Condens. Matter
25 045009
[56] Tang D M, Ren C L, Wei X L, Wang M S, Liu C, Bando Y and
Golberg D 2011 Acs Nano 5 73628
[57] Plimpton S 1995 J. Comput. Phys. 117 119
[58] Swope W C, Anderson H C, Berens P H and Wilson K R 1982
J. Chem. Phys. 76 63749
[59] Hoover W G 1985 Phys. Rev. A 31 16957
[60] Nose S 1984 Mol. Phys. 52 25568
[61] Green J, Bolland T and Bollandt J 1976 J. Chem. Phys.
64 65662
[62] Boldrin L, Scarpa F, Chowdhury R and Adhikari S 2011
Nanotechnology 22 505702
[63] Suryavanshi A P, Yu M F, Wen J G, Tang C C and Bando Y
2004 Appl. Phys. Lett. 84 2527

S-ar putea să vă placă și