You are on page 1of 7

Physics Letters A 375 (2011) 40344040

Contents lists available at SciVerse ScienceDirect

Physics Letters A
www.elsevier.com/locate/pla

Multiscale modeling of graphene- and nanotube-based reinforced polymer nanocomposites


A. Montazeri a , H. Rai-Tabar a,b,
a b

Computational Physical Sciences Research Laboratory, School of Nano-Science, Institute for Research in Fundamental Sciences (IPM), Tehran, Iran Department of Medical Physics and Biomedical Engineering, and Research Centre for Medical Nanotechnology and Tissue Engineering, Shahid Beheshti University of Medical Sciences, Evin, Tehran, Iran

a r t i c l e

i n f o

a b s t r a c t
A combination of molecular dynamics, molecular structural mechanics, and nite element method is employed to compute the elastic constants of a polymeric nanocomposite embedded with graphene sheets, and carbon nanotubes. The model is rst applied to study the effect of inclusion of graphene sheets on the Young modulus of the composite. To explore the signicance of the nanoller geometry, the elastic constants of nanotube-based and graphene-based polymer composites are computed under identical conditions. The reinforcement role of these nanollers is also investigated in transverse directions. Moreover, the dependence of the nanocomposites axial Young modulus on the presence of ripples on the surface of the embedded graphene sheets, due to thermal uctuations, is examined via MD simulations. Finally, we have also studied the effect of sliding motion of graphene layers on the elastic constants of the nanocomposite. 2011 Elsevier B.V. All rights reserved.

Article history: Received 1 July 2011 Received in revised form 30 August 2011 Accepted 31 August 2011 Available online 22 September 2011 Communicated by R. Wu Keywords: Graphene-based nanocomposite Multiscale modeling Molecular dynamics simulation Youngs modulus Temperature-based ripple effect

1. Introduction There is both scientic and technical interest in reinforced polymer nanocomposites due to their enhanced physical, mechanical and electrical properties as compared with neat polymers. In recent years, carbon nanotubes (CNTs) have been used in various types of composite materials. However, the high cost and long production time of puried CNTs have been barriers to the development of CNT-reinforced composites. Furthermore, CNTs form bundles that tend to agglomerate and produce weak spots within polymers [1]. In addition, due to their curved geometry, CNTs cannot be easily wetted by the resin, and this adversely affects the resulting composite with a poor interfacial contact. Novoselov et al. [2] were able to experimentally produce individual graphene sheets (GS), and Lee et al. [3] used AFM nanoindentation technique to measure the elastic properties and intrinsic strength of a suspended graphene sheet over a set of open holes on a Si substrate. Their measurements showed a critical stress of 130 GPa, and a strain of about 25% for a graphene membrane, while the Young modulus was obtained to be roughly 1 TPa. Moreover, the elastic and mechanical properties of GSs have also been computed via different computational and theoretical ap-

Corresponding author at: Computational Physical Sciences Research Laboratory, School of Nano-Science, Institute for Research in Fundamental Sciences (IPM), Tehran, Iran. Tel.: +98 2122835061; fax: +98 2122835058. E-mail address: rai-tabar@nano.ipm.ac.ir (H. Rai-Tabar). 0375-9601/$ see front matter 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.physleta.2011.08.073

proaches, such as the density functional theory (DFT) [4], quantum mechanical-based methods [5], molecular dynamics (MD) simulation [6], and continuum mechanics-based methods [7]. Results similar to the experimental ones have been obtained in these studies. The advanced properties of graphene together with its high specic surface area and its strong nanoller-matrix adhesion have made this material an obvious candidate for use in highperformance polymer-based composites. Graphene-based nanocomposites form a new class of lightweight superstrong functional materials in structural applications, energy storage, molecular sensors, and biomedical eld, to name but a few. Recently, it has been demonstrated that polymer-based nanocomposites, with exfoliated graphene sheets as reinforcing elements, show dramatic improvements in mechanical properties. Koratkar et al. [8] have shown that by adding graphene, equal to 0.1% of the weight of the composite, the strength and the stiffness of the resulting matrix were enhanced by the same degree as that of adding CNTs equal to 1% of the weight of the composite. This order of magnitude gain highlights the potential benets of graphene as a reinforcement material. These authors have also reported a signicant increase (up to 52%) in the critical buckling load of the nanocomposite with only 0.1% weight fraction of GSs introduced into the epoxy matrix [9]. The graphene llers also improve the resistance of the composite to fatigue crack propagation by nearly two orders of magnitude, compared with the baseline epoxy material [10].

A. Montazeri, H. Rai-Tabar / Physics Letters A 375 (2011) 40344040

4035

Limited theoretical and numerical simulation studies have been devoted to the investigation of the effect of presence of GSs on the mechanical properties of polymers. Cho et al. [11] studied the mechanical properties of an epoxy matrix reinforced with randomly distributed graphene sheets, using the MoriTanaka approach [12] in conjunction with molecular mechanics (MM). Elastic constants of graphene sheets, which constituted the inclusion phase in the micromechanical model, were computed via molecular force eld method. The calculations conrmed that the modulus of the nanocomposite was strongly dependent on the aspect ratios of the GSs, but not on their sizes. Employing MD simulation, Awasthi et al. [13] studied the nanoscale load-transfer between polyethylene and a graphene sheet. In this Letter we aim to provide a complete investigation of the mechanical properties of graphene-based nanocomposites. To our knowledge, no multiscale modeling of the mechanical behavior of such composites has been reported so far. To achieve this, we employ a molecular structural mechanics/nite-element (MSM/FE) multiscale modeling approach to investigate the effect of GS inclusions on the Young modulus of the polymer matrix. Furthermore, to explore the inuence of the ller geometry on these properties, the elastic constants of CNT-reinforced polymer composite are also computed within the same method and are compared with those of the GS-reinforced case. Our results show some discrepancies with the experimental results. We surmise that these could be due to the existence of ripples on the surface of GSs, and also their sliding motion. Previous MD simulations [14] have ascertained that thermal uctuations can lead to the appearance of ripples over the GS surface, affecting their mechanical behavior. Furthermore, spring-based niteelement models [15] have shown that GSs can slide over each other due to the non-bonded nature of interaction between the GS layers. Such a motion could lead to a lowering of the stiffness of these nanocomposites. Within the traditional theoretical framework, macroscopic mechanical properties of composites are studied by constructing a representative volume element (RVE) composed of a cubic or a cylindrical portion of the matrix containing the reinforcing agent. In a similar manner, in this work, the mechanical behavior of GS-polymer composites was modeled by an MSM/FE multiscale modeling of the RVE wherein the graphene sheet was resolved on atomistic scales, via the MSM method, and the polymer matrix was resolved on macroscopic scales via the continuum-based FE method. Also, MD simulations were performed to produce the relaxed rippled structure of graphene sheets at different temperatures. These rippled GSs were then employed in the MSM/FE model to explore the effect of thermal uctuations on the elastic constants of the nanocomposites. This Letter is organized as follows. In Section 2 we introduce our computational model. Numerical results for the Young modulus of nanocomposites, as well as the examination of the effects of nanoller geometry, temperature, and the sliding motion of graphene layers are presented in Section 3 and are compared with the available experimental data. Section 4 contains the concluding remarks. 2. Modeling 2.1. The MSM/FE multiscale model In the MSM method [16], an SWCNT, or a graphene sheet, is viewed as a space frame wherein the covalent bonds are modeled as connecting beams and the carbon atoms as their joint nodes. Uniaxial beam elements, characterized by their cross-sectional area, moment of inertia, and material properties, are used to represent the covalent bonds, and their parameters are determined

Fig. 1. Schematic illustration of computational model for GS/polymer composite: (a) side view and (b) end view.

by establishing equivalence between the computational chemistry and structural mechanics. This method has been successfully used to model the static, dynamic, and thermal properties of carbon nanostructures and their composites [1720]. The present MSM model of the GS can be incorporated into the FE-based model for prediction of the mechanical properties of the proposed composites. In our computations, the GS-polymer interactions were described via the LennardJones (LJ) potential, wherein spring elements with a nonlinear behavior were employed to couple the GS and the polymer phase. The spring elements used were characterized by two nodes and a spring constant. To construct these elements, the distances between the nodes on the graphene sheet and the nodes on the inner surface of the polymer matrix were computed, and a spring element was then inserted between any two nodes whose distance was smaller than the cutoff radius of the LJ potential, as depicted in Fig. 1. The spring stiffness of this element was determined by the second derivative of the LJ potential. The details of this multiscale modeling approach are given in Refs. [18,19]. 2.2. Molecular dynamics simulation MD simulations were employed to produce rippled GSs. Here, in line with previous studies of the mechanical properties of GSs [14,21,22], the second-generation Brenner many-body interatomic potential [23] was used to model the energetics and dynamics of the covalent bonds between carbon atoms in the graphene sheet. Briey, the Brenner potential has the form [23]:

Eb =
i j (>i )

V R (r i j ) b i j V A (r i j )

(1)

4036

A. Montazeri, H. Rai-Tabar / Physics Letters A 375 (2011) 40344040

where V R (r i j ) and V A (r i j ) denote the repulsive and attractive pairwise interactions, respectively, and r i j is the distance between pairs of nearest-neighbor atoms i and j. The term b i j represents the reactive empirical bond order depending on the local bonding environment. Detailed explanation about the terms of this potential can be found in Ref. [23]. Initially, all carbon atoms were considered to lie on a at surface of a honeycomb lattice with nearest-neighbor distance of 0.142 nm. The velocity Verlet algorithm [24] was used to integrate the equations of motion with a time-step of dt = 1 fs. This timestep provides a good balance between accuracy and computational costs. The NosHoover thermostat [25] was used to maintain the system at the desired simulation temperature. After reaching the equilibrium at different temperatures, the relaxed rippled structures of the GSs were stored and used in the MSM/FE model to explore the effect of thermal uctuations on the mechanical properties of the composites. 3. Results and discussion We implemented the three-dimensional MSM/FE multiscale model to study the effect of inclusion of graphene sheets on the elastic constants of nanocomposites. Furthermore, we also investigated the reinforcement role of SWCNTs in the axial and transverse directions of the nanocomposite embedded with this nanostructure. A combination of the MD simulation and continuum-based modeling was used to examine the temperature dependence of the results. Additional computations were also performed to ascertain the effect of GS sliding motion on the elastic constants when several sheets of graphene were present in the nanocomposite. The polymer matrix selected in all our simulations was PMMA (poly methyl methacrylate). The Young modulus of this isotropic amorphous polymer was set at 2.5 GPa, which is within the experimental range of 2.243.8 GPa [26], and its Poissons ratio was taken to be 0.35. 3.1. Effect of inclusion geometry The mechanical properties of the PMMA nanocomposite embedded with graphene sheets and SWCNTs were computed. A relevant problem here to investigate was the inuence of the geometry of the carbon nanostructure, i.e., which nanoller provided a more suitable inclusion for the polymer matrix? To obtain an estimate, we computed the Young modulus of the PMMA nanocomposite at a xed volume fraction of 5% GS, and 5% SWCNT. For the SWCNT, a (10, 10) armchair nanotube with diameter and length of 1.356 and 21.398 nm respectively was selected. Since the spatial dimensions of both of these nanollers should be the same, a GS with the same length as the SWCNT and width of 4.118 nm ( dSWCNT ) was selected as the second ller. The only unknown parameters, necessary for the modeling of the cubic computational RVEs, were the dimensions of the cross sections of the unit cells. For the SWCNT-polymer composite, this parameter depends on the SWCNT volume fraction ( f CNT ) present within the RVE, which is dened by
Fig. 2. Finite element macroscopic models of RVEs with different nanollers at a volume fraction of 5%: (a) GS-reinforced composite and (b) SWCNT-reinforced composite.

second RVE, representing the GS-polymer composite, and its crosssectional dimensions were determined to have a width = 8.060 nm and a height = 4 nm. It is to be noted that the RVEs used here were continuous. Thus, the lengths of the nanollers and the polymer matrix were the same in this work. After constructing the model, the macroscopic behavior of the RVEs could be evaluated using the FEM. In this case, the RVEs of our FEM models are schematically shown in Fig. 2. To investigate the effect of different nanollers on the elastic constants of the nanocomposites quantitatively, a tensile load was applied via a prescribed strain of 5% on the two RVEs. The elastic modulus of the nanocomposite (E) was then calculated via

f CNT =

( R CNT + hvdw )2
A Cell

(2)

E=

F / A Cell = H /H0

(3)

where hvdw is the equilibrium van der Waals separation distance between the SWCNT and the matrix, and A Cell is the crosssectional area of the unit cell transverse to the nanotube axis. The parameter hvdw depends on the nature of the SWCNT/polymer interfacial interactions and was taken to be 0.18 nm in accordance with an earlier work [26]. Inserting 5% for SWCNT in Eq. (2), we obtained the width of the cross-sectional area of the cubic matrix to be 6.801 nm. Also, the same procedure was adopted for the

where F (in units of nN) is the total force acting at the end of the RVE, H 0 is the RVEs initial length (in units of nm) and H is its elongation. The values of the axial Young modulus were obtained to be 59.536 GPa and 34.618 GPa for the graphenebased and nanotube-based composites, respectively, showing that graphene presents a more superior ller than the SWCNT, or any other known nanoller, for improving the mechanical properties in nanocomposite. Previous experimental studies also conrm this

A. Montazeri, H. Rai-Tabar / Physics Letters A 375 (2011) 40344040

4037

nding. Thus, Raee et al. [8] have compared the mechanical properties of epoxy nanocomposites embedded with graphene platelets and SWCNTs at a nanoller weight fraction of 0.1%. It was observed that whereas there was only a 4% increase in the Young modulus of the SWCNT-based composite compared with that of the neat epoxy, in the GS-based composite, the Young modulus showed an approximately 31% increase, from 2.85 GPa to 3.74 GPa. Their results indicated that graphene platelets signicantly out-performed the carbon nanotubes. Such a large difference in the reinforcement capabilities of these nanollers was also reported by Ramanathan et al. [27], wherein an increase of 52% and 80% in the axial Young modulus of the nanocomposite was observed by adding 1% wt SWCNT and functionalized graphene sheets, respectively. Similar experimental results concerning these two composites pertinent to other mechanical properties, such as tensile strength and fracture toughness [10] and the critical buckling load [9], have also been reported. Although the experimental results conrm our MSM/FE multiscale model regarding the reinforcement capabilities of graphene inclusions, vis--vis the CNT inclusions, there were, however, signicant discrepancies between our computational results and the experimental data. We observe that the experimental results are much lower than our computed results, where our model indicates a 24-fold increase in Youngs modulus when 5% volume fraction of GSs was added. The discrepancy may have been due to several factors; (1) presence of randomly aligned bres in the experiments, but uni-directionally aligned bres in our numerical simulations; (2) presence of actual defects in graphene sheets resulting in lower values of mechanical properties; (3) the wrinkled (wavy) structure of GSs within the polymeric matrix at different temperatures which is different from the at rectangular shape of the GSs assumed in our model; (4) the sliding of different layers of GSs over each other during the imposition of loads, which was not considered in our modeling. To examine the possible contribution of these factors, we considered points (3) and (4). Our results, given in the next sections, indicate that by accommodating these points, the computed results are improved signicantly and approach the experimental data. A further issue that has not been considered in the earlier works is the effect of these nanollers on the elastic moduli of the nanocomposites along directions other than the axial one. In the majority of experimental and modeling-based studies, the reinforcement role of nanotubes and graphene layers, with the load applied along the axial direction, has been explored. However, when these llers are embedded in nanocomposites, they can be subject to various types of loads along different directions. To this end, we computed the elastic moduli of the RVEs in the transverse directions using the same method as that used for the axial direction. The transverse Young modulus obtained was approximately 3.056 GPa for the case of SWCNT-reinforced RVE, which is much lower than its axial elastic constant of 34.618 GPa. Using a combination of MD and continuum-based elasticity theory, we have shown in our previous publication [28] that the transverse elastic modulus of SWCNTs is negligible in comparison with its axial one, which also conrms the results reported in this Letter. Also, the in-plane transverse Young modulus of the GS-reinforced composite was computed to be 3.859 GPa which implies a 9.8% increase in comparison with that of the SWCNT, again showing that GSs have greater effect on improving the mechanical properties of polymers than CNTs. Moreover, it was observed that the out-of-plane elastic constant for this case was about 2.951 GPa. It comes from the fact that the in-plane elastic constant of graphene layers is larger than the corresponding values in normal directions. The in-plane stiffness of graphite has been determined experimentally to be 1.06 TPa, whereas its stiffness in the direction perpendicular to the layers is only 0.036 TPa [29].

3.2. Effect of temperature Comparison of the results obtained from our multiscale modeling and the experimental data reveals that there exist rather signicant differences between them. We suggested four possible reasons to account for this discrepancy, among them, the effect of thermal uctuations of GS, which lead to the appearance of a wrinkled (rippled) topology in the graphene at the nanoscale, and the sliding of graphene layers over each other. We rst considered the ripples on the GS structure and investigated their effect on the elastic constants. It is known that graphene sheets are found not to be at, and that they can deform into a structure with nearly periodic ripples, which can be produced in a controlled way by means of simple thermal manipulation [30]. Using both analytical and MD-based numerical simulations, Abedpour et al. [31] studied the possibility that the appearance of ripples was due to thermal uctuations. Also, Monte Carlo-based simulations of equilibrium structures of GSs at nite temperatures have reported that ripples spontaneously form with a characteristic wavelength of about 8 nm and that the ripple amplitude is comparable with the carboncarbon interatomic distance ( 0.142 nm) even in very large graphene sheets [32]. It is concluded that ripples constitute an intrinsic feature of graphene sheets and are expected to strongly affect their mechanical and electronic properties [33]. Liu and Zhang [14] performed MD simulations to study how thermal uctuations affect the bending rigidity of graphene, and found that the bending rigidity decreased exponentially with increasing temperature. They proposed that a tendency of inducing a warping shape on edges of nanoribbons due to thermal uctuations would initiate a decrease in the bending rigidity. In the present work, we have focused on the effect of thermal uctuations on the reinforcement role of a graphene layer when loaded in the axial direction. To do so, MD simulations were performed to obtain the equilibrium structure of the graphene layer at the desired temperatures in the range 0.01 K to 400 K. The GS in each simulation was equilibrated for 50 ps (i.e., 50 000 time steps) during which, the four edges of the graphene sheet were clamped. Fig. 3 shows the rippled structure of GSs at equilibrium at different temperatures from which we extracted the coordinates of the GS atoms and used them in the RVE of our MSM/FE multiscale model. The results listed in Table 1 show that compared with the case of at graphene sheet, the presence of thermal uctuations causes a decrease in the axial Young modulus of the GS-polymer composites. The numbers in brackets in Table 1 give the percentage of reduction. As expected, by increasing the temperature, the elastic constant is reduced, but the rate of decrease is more pronounced in the rst stages of applying temperature uctuations (Fig. 4). This phenomenon originates from the fact that the strength of the GS is due to the sp2 hybridization of its carbon atoms and the application of temperature causes deterioration in the CC bonds whereby these bonds protrude out of the graphene plane and hence do not show their intrinsic mechanical strength in the axial direction of the RVE. The results listed in Table 1 reveal that the presence of ripples in the morphology of the GS causes a rapid decrease in the mechanical properties of the nanocomposite, and that above 50 K, the temperature dose not have any signicant extra effect and the results reach a plateau. These data conrm the suggestion that thermally-induced ripples in the GS play a signicant role in reducing the mechanical properties of the corresponding nanocomposites. 3.3. Effect of graphene layers sliding So far we have considered the reinforcement role of a single GS, and the effect of presence of ripples in the GS on the mechanical properties of the nanocomposites. We now consider the

4038

A. Montazeri, H. Rai-Tabar / Physics Letters A 375 (2011) 40344040

Fig. 3. MSM model of GSs relaxed at different temperatures via MD simulation: (a) 5 K; (b) 25 K; (c) 100 K and (d) 300 K. Table 1 Axial Youngs modulus of nanocomposites reinforced with wavy structured GSs at different temperatures compared with the case containing at graphene sheet (E = 59.536 GPa). Case Temperature (K) 5 10 25 50 100 200 300 400 Axial Youngs modulus of graphene-nanocomposite (GPa) 24.248 18.889 16.079 12.598 11.112 10.291 10.059 9.799 Reduction in the elastic constant due to thermal uctuations (%) 59.27 68.27 72.99 78.83 81.33 82.71 83.12 83.54

1 2 3 4 5 6 7 8

inuence of the sliding motion of graphene layers on the mechanical properties. In realistic modeling, multiple GSs are present in the nanocomposite and, hence, the interplay between these layers should be considered. We note that adjacent graphene sheets are held together by weak van der Waals forces and, hence, they can easily slide over each other. The mechanical performance of graphene-based nanocomposites may be greatly affected by this sliding motion. Debelak and Lafdi [1] showed that an increase of graphite content above a critical value causes a decrease in the exural strength of the host polymer. They suggested that the shear and glide of [001] plane of graphene layers may have contributed strongly to the low strength properties of the overall exfoliated lled polymer. To examine this phenomenon in our MSM/FE multiscale model, new RVEs containing a bilayer of graphene sheets were con-

Fig. 4. Axial Youngs modulus of graphene-based nanocomposites reinforced with GSs relaxed at different temperatures.

structed. Bilayer graphene is composed of two GSs that interact with each other via van der Waals forces, which can be expressed using the LennardJones potential energy function. To model these interactions the spring elements, introduced earlier, were used. To account for the effect of temperature uctuations and the sliding of the layers simultaneously three RVEs were studied containing double-layered graphene sheets (DGSs) at different temperatures of 0.01 K, 50 K and 300 K, as shown in Fig. 5. To model the bilayer graphene sheets, the GSs in the previous simulations were

A. Montazeri, H. Rai-Tabar / Physics Letters A 375 (2011) 40344040

4039

Fig. 5. Atomic networks of double-layered graphene sheets at different temperatures: (a) 0.01 K; (b) 50 K and (c) 300 K. Table 2 Effect of graphene layers sliding over each other on the axial Young modulus of nanocomposites. Case Temperature (K) Axial Youngs modulus (GPa) of graphene-nanocomposite reinforced with: GS 1 2 3 0.01 50 300 59.536 12.598 10.059 DGS 44.641 10.161 7.048 Reduction in the elastic constant due to graphene layers sliding (%) 25.02 19.34 29.93

that aggregation of graphene sheets may develop in the nanocomposites during production, since a good deal of GSs can rmly interlock with each other [34]. These aggregates may also result from the inadequate intercalation of monomers or polymer in solution into the space between graphene sheets. It has been shown that these may result in poor mechanical properties of the polymer systems [35]. 4. Conclusions We have employed a molecular structural mechanics (MSM)/ nite element (FE)-based multiscale modeling approach to compute the reinforcement effects of carbon nanostructures embedded in polymers. We have shown that at a low nanoller content, graphene platelets perform signicantly better than carbon nanotubes in terms of enhancing the Young modulus of the nanocomposite. There was a noticeable gap between the computed results and the available experimental data. Several possible factors can contribute to this, among them the contributions of ripples on the surface of the GSs, and the sliding of graphene layers over each other were numerically computed in this study. MD-based simulations to reveal the wavy structure of GSs at different temperatures were performed. We concluded that the presence of such ripples causes a decrease in the axial Young modulus of GS-polymer composites in comparison with the case of at embedded GS, with the decrease being much more noticeable during the early stages of thermal growth (below 50 K). Finally, the variation of Youngs modulus in the presence of double-layer sliding was investigated, and it was found that the elastic constant was considerably affected by the shear

utilized and were placed apart from each other at a distance of 0.34 nm. Also, the dimensions of new RVEs were obtained by considering 5% volume fraction of DGSs. Table 2 shows the comparison of the present results with those pertinent to the single GS-based composites. The table shows that the Young modulus reveals a considerable reduction due to the sliding that occurs between the graphene layers at different temperatures, with an average deviation of 20%. It is worth noting that the axial Young modulus of the nanocomposite at the room temperature (300 K) drops to only 7.048 GPa from the overestimated value of 59.536 GPa obtained when the ripples in the GSs and sliding between adjacent layers were not considered. The new results are now far more comparable with the available existing experimental data, conrming the validity of our combined atomistic (MD and MSM)/continuum (FE) multiscale model. The remaining discrepancy between the experimental and computational results is mainly due to the presence of defects in the structure of GSs and the random spatial arrangement of these nanollers in the polymer, as discussed before. Also, it is plausible

4040

A. Montazeri, H. Rai-Tabar / Physics Letters A 375 (2011) 40344040

and glide of graphene layers. The nal results from this enhanced model were now in good agreement with experimental data. Acknowledgements H.R.-T. would like to acknowledge the support of Iran National Science Foundation for a Research Chair in Nanotechnology. References
[1] B. Debelak, K. Lafdi, Carbon 45 (2007) 1727. [2] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos, I.V. Grigorieva, A.A. Firsov, Science 306 (2004) 666. [3] C. Lee, X. Wei, J.W. Kysar, J. Hone, Science 321 (2008) 385. [4] F. Liu, P. Ming, J. Li, Phys. Rev. B 76 (2007) 064120 (7 pp.). [5] Y.G. Yanovsky, E.A. Nikitina, Y.N. Karnet, S.M. Nikitin, Phys. Mesomech. 12 (2009) 254. [6] Z. Ni, H. Bu, M. Zou, H. Yi, K. Bi, Y. Chen, Phys. B 405 (2010) 1301. [7] M. Arroyo, T. Belytschko, Phys. Rev. B 69 (2004) 115415 (11 pp.). [8] M.A. Raee, J. Raee, I. Srivastava, Z. Wang, H. Song, Z.-Z. Yu, N. Koratkar, ACS Nano 3 (2009) 3884. [9] M.A. Raee, J. Raee, Z.-Z. Yu, N. Koratkar, Appl. Phys. Lett. 95 (2009) 223103. [10] M.A. Raee, J. Raee, I. Srivastava, Z. Wang, H. Song, Z.-Z. Yu, N. Koratkar, Small 6 (2010) 179. [11] J. Cho, J.J. Luo, I.M. Daniel, Compos. Sci. Technol. 67 (2007) 2399. [12] T. Mori, K. Tanaka, Acta Metall. 21 (1973) 571.

[13] A.P. Awasthi1, D.C. Lagoudas, D.C. Hammerand, Model. Simul. Mater. Sci. Eng. 17 (2009) 015002 (37 pp.). [14] P. Liu, Y.W. Zhang, Appl. Phys. Lett. 94 (2009) 231912 (4 pp.). [15] S.K. Georgantzinos, G.I. Giannopoulos, N.K. Anifantis, Mater. Des. 31 (2010) 4646. [16] C. Li, T.-W. Chou, Int. J. Solids Struct. 40 (2003) 2487. [17] I.C. Yeh, G. Hummer, Proc. Natl. Acad. Sci. USA 101 (2004) 12177. [18] C. Li, T.-W. Chou, Compos. Sci. Technol. 66 (2006) 2409. [19] A. Montazeri, R. Naghdabadi, Int. J. Multiscale Com. 7 (2009) 431. [20] A. Sakhaee-Pour, M.T. Ahmadian, R. Naghdabadi, Nanotechnology 19 (2008) 085702. [21] M. Neek-Amal, F.M. Peeters, Appl. Phys. Lett. 97 (2010) 153118 (3 pp.). [22] M. Neek-Amal, F.M. Peeters, Phys. Rev. B 81 (2010) 235437 (6 pp.). [23] D.W. Brenner, O.A. Shenderova, J.A. Harrison, S.J. Stuart, B. Ni, S.B. Sinnott, J. Phys. Condens. Mat. 14 (2002) 783. [24] M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids, Oxford University Press, New York, 1986. [25] W.G. Hoover, Phys. Rev. A 31 (1985) 1695. [26] Y. Han, J. Elliott, Comp. Mater. Sci. 39 (2007) 315. [27] T. Ramanathan, Nat. Nanotechnol. 3 (2008) 327. [28] A. Montazeri, M. Sadeghi, R. Naghdabadi, H. Rai-Tabar, Phys. Lett. A 375 (2011) 1588. [29] J.P. Lu, Phys. Rev. Lett. 79 (1997) 1297. [30] W. Bao, F. Miao, Z. Chen, H. Zhang, W. Jang, C. Dames, C.N. Lau, Nat. Nanotechnol. 4 (2009) 562. [31] N. Abedpour, M. Neek-Amal, R. Asgari, F. Shahbazi, N. Nafari, M.R.R. Tabar, Phys. Rev. B 76 (2007) 195407 (5 pp.). [32] A. Fasolino, J.H. Los, M.I. Katsnelson, Nat. Mater. 6 (2007) 858. [33] Q. Wang, Phys. Lett. A 374 (2010) 1180. [34] G. Chen, W. Weng, D. Wu, C. Wu, Eur. Polym. J. 39 (2003) 2329. [35] G. Chen, C. Wu, W. Weng, D. Wu, W. Yan, Polymer 44 (2003) 1781.

You might also like