Sie sind auf Seite 1von 10

Page 1 of 9

Physical Chemistry Chemical Physics


View Article Online

Mechanical Degradation of Graphene by Epoxidation: Insight from


First-principles Calculations

Published on 24 June 2015. Downloaded by Rensselaer Polytechnic Institute on 24/06/2015 17:10:21.

Qing Peng,1,2 , Liang Han1 , Jie Lian1 , Xiaodong Wen


Koratkar1 , Suvranu De1

3,4 ,

Sheng Liu2 , Zhongfang Chen5 , Nikhil

Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX


First published on the web Xth XXXXXXXXXX 200X
DOI: 10.1039/C3RA41347K

Oxidation is a common cause for the degradation of materials including graphene, where epoxidation (forming C-O-C bond)
is very common. In addition, graphene oxide, in which epoxy groups are one of the two major functional groups (the other are
hydroxy), is an important precursor material used for the bulk synthesis of graphene sheets. Information about the mechanical
stabilities, non-linear elastic properties, and elastic limits under various strains is invaluable for applications of these nanomaterials. Here, we investigate the mechanical properties of the epoxided graphene in ordered graphene oxide, namely C6 O1 , C6 O2 ,
C6 O3 , representing the carbon:oxygen ratio of 6:1, 3:1, and 2:1, respectively, using first-principles calculations within the frame
work of density functional theory. We predict a reduction of the Youngs modulus of graphene by a factor of 20%, 23%, and
27% for C6 O1 , C6 O2 , and C6 O3 , respectively, indicating a monotonic degradation with respect to the epoxidation. However,
there is no clear trend for Poissons ratio, implying that the local atomic configurations are dominant over oxygen concentrations
in determining the Poisson ratio. Our computed high order elastic constants are good for the design of graphene oxides based
flexible transparent electronics.
1

2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

INTRODUCTION

Graphene is known for its fascinating mechanical properties


including high strength and flexibility 13 . Since oxygen is
widely available in the environments, oxidation will readily
occur, leading to the degradation of this important engineering material. Besides such unintended oxidation, the intended
oxidation of graphene, as in graphene oxide (GO), attracted
extensive interests due to its electronic properties as well as its
importance in massively synthesizing graphene 410 . In addition, GO has attracted increasing interest for drug delivery and
other biomedical applications due to its good biocompatibility,
ultrahigh drug loading capability, and the ease of surface functionalization 11 . Graphene oxide has demonstrated distinct advantages as a biosensing platform due to its excellent capabilities for coupling with biomolecules and processing in solution 12 . With unique properties due to the presence of oxygenCorresponding author Qing Peng: qpeng.org@gmail.com
1 Address:Department of Mechanical, Aerospace and Nuclear Engineering,
Rensselaer Polytechnic Institute, Troy, NY 12180, U.S.A.
2 Address:School of Power and Mechanical Engineering, Wuhan University,
Wuhan, China, 430072
3 Address:State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan, China, 030001
4 Address:Synfuels China Co. Ltd. Huairou, Beijing, China, 101407
5 Address:Department of Chemistry, Institute for Functional Nanomaterials,
University of Puerto Rico, Rio Piedras Campus, San Juan, PR 00931, U.S.A

containing functional groups, GO holds potentials for a wide


range of technological applications in many fields, such as microelectronic and chemical devices 1315 , energy storage 16,17 ,
and composite materials 18,19 . For example, the ultra-low density (1/400 times desnity of graphite) three-dimensional foams
were fabricated using GO layers reinforced with conformal
hexagonal boron nitride platelets. Despite its importance,
one interesting and fundamental question is how the oxygencontaining groups affect the mechanical properties of the perfect graphene sheet. Practically, knowledge of GOs mechanical properties is of key importance for its promising applications.
Extensive experimental 2028 and theoretical 2934 studies of
the mechanical properties of GO sheets show a wide range
of distributions of Youngs modulus and intrinsic strength.
Despite these studies, there is still only limited knowledge
about mechanical properties of GO, including the elastic limits, the non-linear elastic properties, and the mechanical instabilities. These information are critical in designing devices
with GO regarding their practical applications since GO is
vulnerable to large strains, with or without intent, because of
their monatomic thickness 3537 . For instance, strains can be
induced because of the mismatch of lattice constants to the
substrates. In addition, the nonlinear behaviors and strain tolerance are critical in the design of graphene oxides based flexible transparent electronics.
19 | 1

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26

Physical Chemistry Chemical Physics Accepted Manuscript

DOI: 10.1039/C5CP02986D

Physical Chemistry Chemical Physics

Page 2 of 9
View Article Online

Fig. 1 Geometry Geometry of graphene and ordered graphene oxide monolayers with the epoxy groups. Graphene plane (a),
armchair-side-view (b), zigzag-side-view (c). The counter parts are in (d-f) for C6 O1 ; (g-i) for C6 O2 ; and (j-l) for C6 O3 , respectively.

1
2
3
4
5
6
7
8
9
10
11

12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33

It is generally accepted that GO mostly contains epoxy (-O) and hydroxyl (-OH) groups on its basal plane 38,39 , while the
atomic structures of GOs are rather complicated and remain
a subject under debate 3843 . There is much experimental evidence indicating that GOs are amorphous 38,40,41 , whereas theoretical investigations suggest that the ordered structural models are energetically favorable 10,16,32,44,45 . The inconsistency
between the experiments and theoretical predictions about the
atomic structures of GOs might be attributed to the kinetic factor and configuration entropy. 34 Here we limited our study to
the three ordered models as illustrated in Fig. 1.

The goal of this paper is to study the degradation of


graphene caused by epoxidation. To that end we used the ordered graphene oxide. To account for the concentration of
oxygen in graphene oxide, we selected three carbon:oxygen
ratios as 6:1, 3:1, and 2:1. Even for a specific concentration,
there are numerous atomic configurations. Here we only consider the common high symmetrical configurations of C6 O1 ,
C6 O2 , and C6 O3 which are shown in Fig. 1. The mechanical
properties of graphene 46 and C6 O 47 were reported previously.
We included them here for a systematic study. We focus on the
nonlinear elastic properties and the mechanical instabilities of
these three ordered graphene oxide monolayers. Using density functional theory calculations, we studied its mechanical
behaviors under various strains. We found an accurate continuum description of the linear and non-linear elastic properties
of graphene oxides with high order elastic constants up to fifth
order. With computed high order elastic constants, one can
predict the pressure dependence properties, including sound
velocities, the second order elastic constants, in-plane stiffness, and Poisson ratio. Our continuum formulation could be
useful in finite element modeling of mechanical properties of
graphene oxide at the continuum level.
2|

19

2 COMPUTATIONAL METHOD

2.1 First-principles calculations

We consider a conventional unit cell which contains six carbon


atoms forming the basal plane of graphene oxide with periodic
boundary conditions in plane directions only (Fig. 1). Such
a conventional unitcell is chosen to capture the soft mode,
which is a particular normal mode exhibiting an anomalous
reduction in its characteristic frequency and leading to mechanical instability. This soft mode is a key factor in limiting
the strength of monolayer materials that can only be captured
in unit cells with hexagonal rings 48 . A previous study shows
that the larger super cells, including (4 2), (6 3), (8 4),
and (10 10) unit cells have almost identical results 49 .
The armchair direction is along the direction of the project
of a carbon-carbon bond onto the basal plane, and it was set to
be parallel to the x axis, which is denoted 1 in superscript of
strains and stresses. The zigzag direction is perpendicular to
the armchair direction and was set as y-axis, which is denoted
2 in superscript of strains and stresses.
First-principles calculations were carried out with the VASP
code 50 , which is based on the Kohn-Sham Density Functional Theory 51 with the generalized gradient approximations as parameterized by Perdew, Burke, and Ernzerhof for
exchange-correlation functions 52 . The projector augmented
wave pseudo-potential 53 was used where C 2s2 2p2 and O
2s2 2p4 were treated as valence electrons. A plane-wave basis set with a kinetic-energy cutoff of 600 eV was used in all
the calculations. Atomic relaxations are performed using a
Gamma-centered 24 24 1 k-mesh for sampling and atomic
positions are converged until the Hellmann-Feynman forces
There was a 20
on each atom are smaller than 0.001 eV/A.

A thick vacuum region to reduce the inter-layer interaction to


model the single layer system.
To eliminate the artificial effect of the out-of-plane thick-

3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34

Physical Chemistry Chemical Physics Accepted Manuscript

Published on 24 June 2015. Downloaded by Rensselaer Polytechnic Institute on 24/06/2015 17:10:21.

DOI: 10.1039/C5CP02986D

Page 3 of 9

Physical Chemistry Chemical Physics


View Article Online

ness of the simulation box on the stress, we used the second


Piola-Kirchhoff (P-K) stress 54 to express the two-dimensional
forces per length with units of N/m. The Lagrangian strain 55
is used in this study, defined as = + 1/2 2 , where is the
stretch ratio 54 .

2.2

1
2
3

Published on 24 June 2015. Downloaded by Rensselaer Polytechnic Institute on 24/06/2015 17:10:21.

Linkage continuum elastic theory with atomistic


model

18

For a general deformation state, the number of independent


components of the second, third, fourth, and fifth order elastic tensors is 21, 56, 126, and 252, respectively 54 . However,
only fourteen independent elastic constants need to be explicitly considered due to the symmetry. By a two-step leastsquares fit, the fourteen independent elastic constants are obtained from stress-strain relationships 54 . In the first step, we
use a least-squares fit to five stress-strain responses of armchair, zigzag, and biaxial deformations. The second step is for
a least-square solution to the over- and well-determined linear
equations.

19

8
9
10
11
12
13
14
15
16
17

20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48

RESULTS AND ANALYSIS

Before straining, we first optimize the equilibrium lattice constant for graphene oxides. When the strains are applied, all the
atoms are allowed full freedom of motion. A quasi-Newton algorithm 56 is used to relax all atoms into equilibrium positions
within the deformed unit cell that yields the minimum total
energy for the imposed strain state of the super cell.
We defined the degree of epoxidation as the ratio of the
number of oxygen atoms over the number of carbon atoms
in the unit cells to quantify the concentration of oxygen atoms
in the GO. The ultimate strengths u , ultimate strains u , and
Youngs modulus Y along the armchair, zigzag, and biaxial
direction as a function of the degree of epoxidation are illustrated in Fig. 2. We observed a general trend of reduction of
the Youngs modulus upon degree of epoxidation, indicating a
monotonic degradation with respect to the epoxidation. This
could be understood as when the degree of epoxidation increases, more C-C ( ) bonds become weaker sp3 bonds
of C-O-C, which introduce discontinuities in the graphene lattice. As the density of the bonds decrease, the strength
decreases. However, there is no clear trend for Poissons ratio, implying that the local atomic configuration is a dominant
factor over oxygen concentrations in determining the Poisson
ratio.
The stresses are the derivatives of the total energies with
respect to the strains obtained from DFT calculations. The
second P-K stress versus Lagrangian strain relationship for
uniaxial strains along the armchair and zigzag directions and
biaxial strains are shown in Fig. 3. The material behaves in an
asymmetric manner with respect to compressive and tensile

Fig. 2 Elastic limits Ultimate strengths u , ultimate strainsu , and


Youngs modulus Y along the armchair, zigzag, and biaxial direction
as a function of the degree of epoxidation.

19 | 3

Physical Chemistry Chemical Physics Accepted Manuscript

DOI: 10.1039/C5CP02986D

Physical Chemistry Chemical Physics

Page 4 of 9
View Article Online

Fig. 3 Stress-strain relationship Stress-strain responses of graphene (a), C6 O1 (b), C6 O2 (c), and C6 O3 (d), under armchair (top), zigzag
(middle), and biaxial (bottom) strains. 1 (2 ) denotes the x (y) component of stress. Cont stands for the fitting of DFT calculations
(DFT) to continuum elastic theory.

4|

19

Physical Chemistry Chemical Physics Accepted Manuscript

Published on 24 June 2015. Downloaded by Rensselaer Polytechnic Institute on 24/06/2015 17:10:21.

DOI: 10.1039/C5CP02986D

Page 5 of 9

Physical Chemistry Chemical Physics


View Article Online

1
2
3
4
5
6

Published on 24 June 2015. Downloaded by Rensselaer Polytechnic Institute on 24/06/2015 17:10:21.

7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52

strains. With increasing strains, the C-C bonds are stretched


and eventually rupture.
Taking C6 O1 as an example, when strain is applied in the
armchair direction (mode a), the bonds parallel to this direction are more severely stretched than those in other directions.
Under the deformation mode z (zigzag), in which the strain
is applied perpendicularly to the armchair direction, there is
no bond parallel to this direction. The two C-C bonds that
have an incline to the zigzag direction with an angle of about
30 are more severely stretched than those in the armchair direction. Under the ultimate strain in mode z, which is 0.26, the
two C-C bonds that are not along the armchair direction and
closer to oxygen atom are observed to rupture. Under the ultimate strain in mode b (biaxial), ub =0.12, the two C-C bonds
that are not along the zigzag direction with an angle of about
30 and further away from the oxygen atom are observed to
rupture.
The elastic constants are critical parameters in finite element analysis models for mechanical properties of materials.
The fourteen independent elastic constants of graphene oxide
are determined by a least-square fit to the stress-strain results
from DFT calculations in two steps, detailed in our previous
work 47,54 , which had been well used to explore the mechanical properties of 2D materials 5769 . A brief introduction is
that, in the first step, we use a least-square fit of five stressstrain responses. Five relationships between stress and strain
are necessary because there are five independent fifth-order
elastic constants (FFOEC). We obtain the stress-strain relationships by simulating the following deformation states: uniaxial strain in the zigzag direction (zigzag); uniaxial strain in
the armchair direction (armchair); and equibiaxial strain (biaxial). From the first step, the components of the secondorder elastic constants (SOEC), the third-order elastic constants (TOEC), and the fourth-order elastic constants (FOEC)
are over-determined (i.e, the number of linearly independent
variables are greater than the number of constraints), and the
fifth-order elastic constants are well-determined (the number
of linearly independent variables are equal to the number of
constraints). Under such circumstances, the second step is
needed: least-square solution to these over- and well- determined linear equations.
The second order elastic constants model the linear elastic
response. The higher (> 2) order elastic constants are important to characterize the nonlinear elastic response of C6 O using a continuum description. These can be obtained using a
least-square fit of the DFT data and are reported in Table 1.
Corresponding values for graphene are also shown.
Our results of in-plane stiffness of GO is Ys = 273.6, 264.1,
and 247.5 N/m for 6 O1 , 6 O2 , and 6 O3 , respectively. Taking the
, our result of the in-plane stiffness
interlayer distance of 7 A
is 390.9, 377.3, and 353.6 GPa, respectively, in good agreement with the experiment (250 150 GPa), 23 MD prediction

Table 1 Elastic constants Nonzero independent components for


the SOEC, TOEC, FOEC, and FFOEC tensor components (in units
of N/m), Poissons ration and in-plane stiffness Ys of graphene
(C6 ) and the ordered GO (C6 O1 ,C6 O2 ,C6 O3 ) monolayer from DFT
calculations.

a (A)
Ys (N/m)

C11
2nd
C12
C111
3rd
C112
C222
C1111
C1112
4th
C1122
C2222
C11111
C11112
5th C11122
C12222
C22222

C6
2.468
340.8
0.178
352.0
62.6
-3089.7
-453.8
-2928.1
21927
2731
3888
18779
-118791
-19173
-15863
-27463
-134752

C6 O1
2.53
273.6
0.195
284.5
55.6
-2437.2
-518.3
-2993.6
11416
-4368
3277
5822
-42938
28630
-112832
-335145
99159

C6 O2
2.568
264.1
0.139
266.5
37.0
-2352.8
-155.9
-2041.1
18514
1439
-2119
9797
-105098
-1355
15476
21277
-97941

C4 O2
2.592
247.5
0.084
249.3
21.2
-2585.7
102.9
-1946.0
23677
-1799
-605
9409
-143392
-7460
-34349
-63325
-116554

(325 GPa), 29 and DFT prediction (380470 GPa). 32 We predict a reduction of the Youngs modulus of graphene by a
factor of 20%, 23%, and 27% for C6 O1 , C6 O2 , and C6 O3 ,
respectively. The in-plane stiffnesses of GO is smaller than
graphene, which indicates that GO is softened by the epoxidation. This could be understood as follows, taking C6 O1 as an
example. The separation of the two carbon atoms bridged by
about 6 percent larger than
the oxygen atom in C6 O is 1.505 A,
The C-C bonds have been stretched
that of graphene (1.42 A).
in C6 O1 a priori by the introduction of oxygen atoms as compared with the pristine graphene (Fig. 1). These stretched C-C
bonds are weaker than those un-stretched, resulting in a reduction of the mechanical strength.
A recent experimental study demonstrated that graphene
can maintain a large fraction of its pristine strength and stiffness when sp3 -type defects exist in a high density 28 . Since
the defects as in epoxidated graphene are sp3 -type, our results
agree well with the experiment.
Higher order (> 2) elastic constants are important quantities 70 and can be determined by measuring the changes of
sound velocities under the application of hydrostatic and uniaxial stresses 71 . The higher order elastic constants can be
utilized to study the nonlinear elasticity, thermal expansion
(through the gruneisen parameter), temperature dependence
of elastic constants, harmonic generation, phonon-phonon interactions, photon-phonon interactions, lattice defects, phase
19 | 5

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26

Physical Chemistry Chemical Physics Accepted Manuscript

DOI: 10.1039/C5CP02986D

Physical Chemistry Chemical Physics

Page 6 of 9
View Article Online

1
2
3
4
5
6

Published on 24 June 2015. Downloaded by Rensselaer Polytechnic Institute on 24/06/2015 17:10:21.

7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52

transitions, echo phenomena, strain softening, and so on 46 .


Using the higher order elastic continuum description, one
can calculate the stress and deformation state under uniaxial
stress, rather than uniaxial strain. Explicitly, when pressure
is applied, the pressure dependent second-order elastic moduli can be obtained from the higher order elastic continuum
description 54 . The third-order elastic constants are important
in understanding the nonlinear elasticity of materials, such as
changes in acoustic velocities due to finite strain. As a consequence, nano devices (such as nano surface acoustic wave
sensors and nano waveguides) could be synthesized by introducing local strain 57,72 .
The hydrostatic terms (C11 , C22 , C111 , C222 , and so on)
of GO monolayers are smaller in magnitude than those of
graphene, consistent with the conclusion that the GO is
softer. Both GO and graphene monolayers exhibit instability under large tension. Stress-strain curves in the previous
section show that they might soften when the strain is larger
than the ultimate strain. From the view of electron bonding,
this is due to the bond weakening and breaking. This softening
behavior is determined by the TOECs and FFOECs in the continuum aspect. The negative values of TOECs and FFOECs
ensure the softening of GO monolayer under large strain.
A good way to check the importance of the high order elastic constants is to consider the cases when they are missing.
With the elastic constants, the stress-strain response can be
predicted from elastic theory 54 . When we only consider the
second order elasticity, the stress varies with strain linearly.
Taking the biaxial deformation as an example, as illustrated in
Fig. 4, the linear behaviors are only valid within a small strain
range, about 0.02.With the knowledge of the elastic constants
up to the third order, the stress-strain curve can be accurately
predicted within the range of 0.06. Using the elastic constants up to the fourth order, the mechanical behaviors can be
well treated up to a strain as large as 0.13. For the strains
beyond 0.13, the fifth order elastic constants are required for
accurate modeling. The analysis of the uniaxial deformations
provides similar results.
Our results of the mechanical properties of graphene epoxide are limited to zero temperature due to current DFT calculations. Once the finite temperature is considered, the thermal
expansions and dynamics will in general reduce the interactions between atoms. As a result, the longitudinal mode elastic constants will decrease with respect to the temperature of
the system. The variation of the shear mode elastic constants
should be more complex at the temperature. A thorough study
will be interesting, which is, however, beyond the scope of this
study.
Our results of the ordered GOs could be useful in estimating the amorphous GOs which are common in experiments.
The amorphous GOs could be considered as combinations of
these ordered GOs. The linear fitting of the in-plane Youngs
6|

19

Fig. 4 Order effect The predicted stress-strain responses of biaxial


deformation of ordered graphene epoxide C6 O2 monolayer from
different orders (second, third, fourth, and fifth order) compared to
that from the DFT calculations (circle line).

modulus Ys of graphene and three ordered GOs with respect


to the the degree of epoxidation x gives the proportional relationship Ys (x) = 173.76x + 324.94. Therefore, the amorphous GOs in-plane Youngs modulus can be estimated from
the proportional relationship. The GOs with hydroxyl group
might have similar behaviors, which is an interesting further
research topic.

4 CONCLUSIONS
In summary, we shed lights on the mechanical degradation of
graphene by epoxidation through comparing the mechanical
properties of graphene and three ordered graphene epoxide,
C6 O1 , C6 O2 , and C6 O3 , which represents the carbon:oxygen
ratio of 6:1, 3:1, and 2:1, respectively, under various strains
using first-principles calculations.It is observed that these GO
exhibits anisotropic nonlinear elastic deformations up to ultimate strains. There is a reduction of the Youngs modulus
of graphene by a factor of 20%, 23%, and 27% for C6 O1 ,
C6 O2 , and C6 O3 , respectively. Thus, the strength of GO decreases with the degree of the epoxidation. However, there
is no clear trend for Poissons ratio, implying that the local
atomic configurations is a dominant factor over oxygen concentrations in determining the Poisson ratio. Our computed
second, third, fourth, and fifth order elastic constants could
be used for modeling of the elastic response to a mechanical
loading of GO at continuum level, especially in the design of
graphene/graphene oxides based flexible electronics.

1
2
3
4
5
6
7

9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26

Physical Chemistry Chemical Physics Accepted Manuscript

DOI: 10.1039/C5CP02986D

Page 7 of 9

Physical Chemistry Chemical Physics


View Article Online

DOI: 10.1039/C5CP02986D

2
3
4
5

Published on 24 June 2015. Downloaded by Rensselaer Polytechnic Institute on 24/06/2015 17:10:21.

6
7
8
9

ACKNOWLEDGEMENTS
The authors would like to acknowledge the generous financial support from the Defense Threat Reduction Agency
(DTRA) Grant # HDTRA1-13-1-0025 and the Office of Naval
Research grants ONR Award # N00014-08-1-0462 and #
N00014-12-1-0527. Computational resources were provided
by Rensselaer Polytechnic Institute through AMOS, the IBM
Blue Gene/Q system at the Center for Computational Innovations.

19

20

21

10

11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46

References
1 C. Lee, X. Wei, J. W. Kysar and J. Hone, Science, 2008,
321, 385388.
2 X. Wei, B. Fragneaud, C. A. Marianetti and J. W. Kysar,
Phys. Rev. B, 2009, 80, 205407.
3 F. Liu, P. Ming and J. Li, Phys. Rev. B, 2007, 76, 064120.
4 S. Gilje, S. Han, M. Wang, K. L. Wang and R. B. Kaner,
Nano Lett., 2007, 7, 33943398.
5 H. C. Schniepp, J.-L. Li, M. J. McAllister, H. Sai,
M. Herrera-Alonso, D. H. Adamson, R. K. Prudhomme,
R. Car, D. A. Saville and I. A. Aksay, J. Phys. Chem. B,
2006, 110, 85358539.
6 D. R. Dreyer, S. Park, C. W. Bielawski and R. S. Ruoff,
Chem. Soc. Rev., 2010, 39, 228240.
7 D. Li, M. B. Mueller, S. Gilje, R. B. Kaner and G. G. Wallace, Nat. Nanotechnol., 2008, 3, 101105.
8 M. J. Allen, V. C. Tung and R. B. Kaner, Chem. Rev., 2010,
110, 132145.
9 S. Park and R. S. Ruoff, Nat. Nanotechnol., 2009, 4, 217
224.
10 J.-A. Yan and M. Y. Chou, Phys. Rev. B, 2010, 82, 125403.
11 M. Inagaki and F. Kang, J. Mater. Chem. A, 2014, 2,
1319313206.
12 E. Morales-Narvez and A. Merkoi, Adv. Mater., 2012, 24,
32983308.
13 G. Eda, G. Fanchini and M. Chhowalla, Nat. Nanotechnol., 2008, 3, 270274.
14 J. D. Fowler, M. J. Allen, V. C. Tung, Y. Yang, R. B. Kaner
and B. H. Weiller, ACS Nano, 2009, 3, 301306.
15 C. Gomez-Navarro, R. T. Weitz, A. M. Bittner, M. Scolari,
A. Mews, M. Burghard and K. Kern, Nano Lett., 2007, 7,
34993503.
16 L. Wang, K. Lee, Y.-Y. Sun, M. Lucking, Z. Chen, J. J.
Zhao and S. B. Zhang, ACS Nano, 2009, 3, 29953000.
17 J. Xu, K. Wang, S.-Z. Zu, B.-H. Han and Z. Wei, ACS
Nano, 2010, 4, 50195026.
18 S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M.

22

23
24
25

26
27
28

29
30
31
32
33
34
35

36
37
38
39

Kohlhaas, E. J. Zimney, E. A. Stach, R. D. Piner, S. T.


Nguyen and R. S. Ruoff, Nature, 2006, 442, 282286.
T. Ramanathan, A. A. Abdala, S. Stankovich, D. A. Dikin,
M. Herrera-Alonso, R. D. Piner, D. H. Adamson, H. C.
Schniepp, X. Chen, R. S. Ruoff, S. T. Nguyen, I. A. Aksay,
R. K. Prudhomme and L. C. Brinson, Nat. Nanotechnol.,
2008, 3, 327331.
D. A. Dikin, S. Stankovich, E. J. Zimney, R. D. Piner,
G. H. B. Dommett, G. Evmenenko, S. T. Nguyen and R. S.
Ruoff, Nature, 2007, 448, 457460.
H. Chen, M. B. Mueller, K. J. Gilmore, G. G. Wallace and
D. Li, Adv. Mater., 2008, 20, 35573561.
C. Chen, Q.-H. Yang, Y. Yang, W. Lv, Y. Wen, P.-X. Hou,
M. Wang and H.-M. Cheng, Adv. Mater., 2009, 21, 3007
3011.
C. Gomez-Navarro, M. Burghard and K. Kern, Nano Lett.,
2008, 8, 20452049.
Y. Gao, L.-Q. Liu, S.-Z. Zu, K. Peng, D. Zhou, B.-H. Han
and Z. Zhang, ACS Nano, 2011, 5, 21342141.
J. T. Robinson, M. Zalalutdinov, J. W. Baldwin, E. S.
Snow, Z. Wei, P. Sheehan and B. H. Houston, Nano Lett.,
2008, 8, 34413445.
J. W. Suk, R. D. Piner, J. An and R. S. Ruoff, ACS Nano,
2010, 4, 65576564.
S.-H. Kang, T.-H. Fang, Z.-H. Hong and C.-H. Chuang,
Diam. Relat. Mater., 2013, 38, 7378.
A. Zandiatashbar, G.-H. Lee, S. J. An, S. Lee, N. Mathew,
M. Terrones, T. Hayashi, C. R. Picu, J. Hone and N. Koratkar, Nat. Commun., 2014, 5, 3186.
J. T. Paci, T. Belytschko and G. C. Schatz, J. Phys. Chem.
C, 2007, 111, 1809918111.
Q. Zheng, Y. Geng, S. Wang, Z. Li and J.-K. Kim, Carbon,
2010, 48, 43154322.
Z. Xu and K. Xue, Nanotechnology, 2010, 21, 045704.
L. Liu, J. Zhang, J. Zhao and F. Liu, Nanoscale, 2012, 4,
59105916.
S. D. Dabhi, S. D. Gupta and P. K. Jha, J. Appl. Phys.,
2014, 115, 203517.
L. Liu, L. Wang, J. Gao, J. Zhao, X. Gao and Z. Chen,
Carbon, 2012, 50, 16901698.
Q. Peng, J. Crean, A. K. Dearden, X. Wen, C. Huang,
S. P. A. Bordas and S. De, Mod. Phys. Lett. B, 2013, 27,
1330017.
Q. Peng, A. K. Dearden, J. Crean, L. Han, S. Liu, X. Wen
and S. De, Nanotechn. Sci. Appl., 2014, 7, 129.
Q. Peng and S. De, Nanoscale, 2014, 6, 1207112079.
W. Gao, L. B. Alemany, L. Ci and P. M. Ajayan, Nat.
Chem., 2009, 1, 403408.
W. Cai, R. D. Piner, F. J. Stadermann, S. Park, M. A.
Shaibat, Y. Ishii, D. Yang, A. Velamakanni, S. J. An,

19 | 7

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50

Physical Chemistry Chemical Physics Accepted Manuscript

Das könnte Ihnen auch gefallen