Effects of loading mode and orientation on deformation ...Molecular dynamics simulation is performed...

5
Effects of loading mode and orientation on deformation mechanism of graphene nano-ribbons Y. J. Sun, 1 F. Ma, 1,2,a) Y. H. Huang, 3 T. W. Hu, 1,2 K. W. Xu, 1,4,a) and Paul K. Chu 2,a) 1 State Key Laboratory for Mechanical Behavior of Materials, Xi’an Jiaotong University, Xi’an 710049, Shaanxi, China 2 Department of Physics and Materials Science, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, China 3 College of Physics and Information Technology, Shaanxi Normal University, Xi’an 710062, Shaanxi, China 4 Department of Physics and Opt-electronic Engineering, Xi’an University of Arts and Science, Xi’an 710065, Shaanxi, China (Received 8 August 2013; accepted 25 October 2013; published online 7 November 2013) Molecular dynamics simulation is performed to analyze the deformation mechanism of graphene nanoribbons. When the load is applied along the zigzag orientation, tensile stress yields brittle fracture and compressive stress results in lattice shearing and hexagonal-to-orthorhombic phase transformation. Along the armchair direction, tensile stress produces lattice shearing and phase transformation, but compressive stress leads to a large bonding force. The phase transformation induced by lattice shearing is reversible for 17% and 30% strain in compressive loading along the zigzag direction and tensile loading along the armchair direction. The energy dissipation is less than 10% and resulting pseudo-elasticity enhances the ductility. V C 2013 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4829480] Since the successful exfoliation of single-layer gra- phene in 2004, 1 two-dimensional materials have become one of the research focuses in physics, chemistry, and mate- rials science. 25 Thereafter, many approaches have been developed to fabricate graphene and graphene nanoribbons, including mechanical exfoliation, 6,7 reduction of graphite oxide, 8,9 chemical vapor deposition, decomposition of SiC, 10 and un-zipping of carbon nanotube (CNT). 11,12 Generally, residual stress exists in the as-prepared samples as well as those after substrate transfer, 1315 and it is closely related to the localized ripples, edges, grain boundaries, adatom adsorption, and Stone-Wales defects produced in the synthesis and post thermal processes. 1618 The residual stress might influence the structural stability of this ultra- thin system in practice. In addition, graphene possesses me- chanical strength of about 130 GPa, 19 which is almost 100 times larger than that of steels, and so graphene is believed to be an ideal reinforcing component in composite materi- als. 20,21 Graphene sheets are typically distributed randomly in the composite host and when the materials are subjected to an external load along different orientations, the gra- phene sheets may be activated differently. Hence, it is of scientific and technological interest to determine the de- pendence of their deformation behavior on the loading mode and loading orientation. It is generally accepted that the competition between bond rotation and bond rupture determines the deformation mechanism of graphene. By conducting molecular dynam- ics (MD) simulation, Grantab et al. 22 found that graphene sheets with large-angle tilt boundaries having a high density of defects were as strong as the pristine one and much stronger than those with low-angle boundaries having fewer defects. They ascribed the abnormal behavior to the ability of the large-angle tilt boundaries to better accommodate the strained rings via bond rotation. 22 Bond-rotation-related mechanical deformation has been observed from graphene with Stone-Wales (SW) defects. 2325 Similar to plane slips in bulk metals, bond rotation depends on the loading direc- tion and can be described by a physical parameter resem- bling the Schmidt factor. 26 In bulk metals, the deformation mechanism is determined by the loading modes, namely, compression and stretching and hence, both the loading direction and loading mode affect the mechanism of me- chanical deformation. In this work, MD simulation is per- formed to determine the effects of compressive/tensile loading along the zigzag and armchair directions on the de- formation mechanism. MD simulation is carried out on the large-scale atomic/ molecular massively parallel simulator (LAMMPS). 27 The interaction between carbon atoms is described by the adaptive intermolecular reactive bond order (AIREBO) potential, which can accurately capture the interactions between carbon atoms as well as bond breaking and re-forming. 28,29 The cut- off parameter describing the short-range C-C interactions is chosen to be 2.0 A ˚ in order to avoid spuriously high bonding force and nonphysical results at large deformation. 30 A lattice constant of 1.426 A ˚ is adopted as the initial value and the layer separation of graphite of 3.4 A ˚ is taken as the effective thick- ness of the mono-layer graphene. 31 A Poisson’s ratio of 0.165 is used. 32 Prior to the simulation, the graphene sheets with periodic conditions in the two in-plane directions are relaxed to an equilibrium state in the isothermal-isobaric (NPT) ensembles for 1 000 000 MD steps with a time step of 1 fs. Graphene nanoribbons (GNRs) along the zigzag/armchair ori- entations and with a size of 50A ˚ Â 120 A ˚ with 2240/2279 atoms are created by deleting atoms from the outside part of the nanoribbons and a vacuum region 15 A ˚ in width is added a) Authors to whom correspondence should be addressed. Electronic addresses: [email protected]; [email protected]; and [email protected] 0003-6951/2013/103(19)/191906/5/$30.00 V C 2013 AIP Publishing LLC 103, 191906-1 APPLIED PHYSICS LETTERS 103, 191906 (2013)

Transcript of Effects of loading mode and orientation on deformation ...Molecular dynamics simulation is performed...

  • Effects of loading mode and orientation on deformation mechanismof graphene nano-ribbons

    Y. J. Sun,1 F. Ma,1,2,a) Y. H. Huang,3 T. W. Hu,1,2 K. W. Xu,1,4,a) and Paul K. Chu2,a)1State Key Laboratory for Mechanical Behavior of Materials, Xi’an Jiaotong University, Xi’an 710049,Shaanxi, China2Department of Physics and Materials Science, City University of Hong Kong, Tat Chee Avenue, Kowloon,Hong Kong, China3College of Physics and Information Technology, Shaanxi Normal University, Xi’an 710062, Shaanxi, China4Department of Physics and Opt-electronic Engineering, Xi’an University of Arts and Science, Xi’an 710065,Shaanxi, China

    (Received 8 August 2013; accepted 25 October 2013; published online 7 November 2013)

    Molecular dynamics simulation is performed to analyze the deformation mechanism of graphene

    nanoribbons. When the load is applied along the zigzag orientation, tensile stress yields brittle

    fracture and compressive stress results in lattice shearing and hexagonal-to-orthorhombic phase

    transformation. Along the armchair direction, tensile stress produces lattice shearing and phase

    transformation, but compressive stress leads to a large bonding force. The phase transformation

    induced by lattice shearing is reversible for 17% and 30% strain in compressive loading along the

    zigzag direction and tensile loading along the armchair direction. The energy dissipation is less

    than 10% and resulting pseudo-elasticity enhances the ductility. VC 2013 AIP Publishing LLC.[http://dx.doi.org/10.1063/1.4829480]

    Since the successful exfoliation of single-layer gra-

    phene in 2004,1 two-dimensional materials have become

    one of the research focuses in physics, chemistry, and mate-

    rials science.2–5 Thereafter, many approaches have been

    developed to fabricate graphene and graphene nanoribbons,

    including mechanical exfoliation,6,7 reduction of graphite

    oxide,8,9 chemical vapor deposition, decomposition of

    SiC,10 and un-zipping of carbon nanotube (CNT).11,12

    Generally, residual stress exists in the as-prepared samples

    as well as those after substrate transfer,13–15 and it is closely

    related to the localized ripples, edges, grain boundaries,

    adatom adsorption, and Stone-Wales defects produced in

    the synthesis and post thermal processes.16–18 The residual

    stress might influence the structural stability of this ultra-

    thin system in practice. In addition, graphene possesses me-

    chanical strength of about 130 GPa,19 which is almost 100

    times larger than that of steels, and so graphene is believed

    to be an ideal reinforcing component in composite materi-

    als.20,21 Graphene sheets are typically distributed randomly

    in the composite host and when the materials are subjected

    to an external load along different orientations, the gra-

    phene sheets may be activated differently. Hence, it is of

    scientific and technological interest to determine the de-

    pendence of their deformation behavior on the loading

    mode and loading orientation.

    It is generally accepted that the competition between

    bond rotation and bond rupture determines the deformation

    mechanism of graphene. By conducting molecular dynam-

    ics (MD) simulation, Grantab et al.22 found that graphenesheets with large-angle tilt boundaries having a high density

    of defects were as strong as the pristine one and much

    stronger than those with low-angle boundaries having fewer

    defects. They ascribed the abnormal behavior to the ability

    of the large-angle tilt boundaries to better accommodate the

    strained rings via bond rotation.22 Bond-rotation-related

    mechanical deformation has been observed from graphene

    with Stone-Wales (SW) defects.23–25 Similar to plane slips

    in bulk metals, bond rotation depends on the loading direc-

    tion and can be described by a physical parameter resem-

    bling the Schmidt factor.26 In bulk metals, the deformation

    mechanism is determined by the loading modes, namely,

    compression and stretching and hence, both the loading

    direction and loading mode affect the mechanism of me-

    chanical deformation. In this work, MD simulation is per-

    formed to determine the effects of compressive/tensile

    loading along the zigzag and armchair directions on the de-

    formation mechanism.

    MD simulation is carried out on the large-scale atomic/

    molecular massively parallel simulator (LAMMPS).27 The

    interaction between carbon atoms is described by the adaptive

    intermolecular reactive bond order (AIREBO) potential,

    which can accurately capture the interactions between carbon

    atoms as well as bond breaking and re-forming.28,29 The cut-

    off parameter describing the short-range C-C interactions is

    chosen to be 2.0 Å in order to avoid spuriously high bonding

    force and nonphysical results at large deformation.30 A lattice

    constant of 1.426 Å is adopted as the initial value and the layer

    separation of graphite of 3.4 Å is taken as the effective thick-

    ness of the mono-layer graphene.31 A Poisson’s ratio of 0.165

    is used.32 Prior to the simulation, the graphene sheets with

    periodic conditions in the two in-plane directions are relaxed

    to an equilibrium state in the isothermal-isobaric (NPT)

    ensembles for 1 000 000 MD steps with a time step of 1 fs.

    Graphene nanoribbons (GNRs) along the zigzag/armchair ori-

    entations and with a size of 50 � 120 Šwith 2240/2279atoms are created by deleting atoms from the outside part of

    the nanoribbons and a vacuum region 15 Å in width is added

    a)Authors to whom correspondence should be addressed. Electronic addresses:

    [email protected]; [email protected]; and [email protected]

    0003-6951/2013/103(19)/191906/5/$30.00 VC 2013 AIP Publishing LLC103, 191906-1

    APPLIED PHYSICS LETTERS 103, 191906 (2013)

    http://dx.doi.org/10.1063/1.4829480http://dx.doi.org/10.1063/1.4829480http://dx.doi.org/10.1063/1.4829480mailto:[email protected]:[email protected]:[email protected]://crossmark.crossref.org/dialog/?doi=10.1063/1.4829480&domain=pdf&date_stamp=2013-11-07

  • perpendicular to the length direction and normal to graphene

    sheets, so that the atoms near one edge of the GNRs do not

    interact with those near the opposite edge because of the peri-

    odic boundary conditions. Simulation is performed under

    compressive/tensile loading along the width direction of the

    GNRs in the NVT ensemble using the deformation-control

    method. A strain increment of about 10�4 corresponding to a

    strain rate of 109 s�1 is applied in every time step to the load-

    ing and controlled unloading processes. Four combinations

    including two orientations (armchair and zigzag) and two

    loading modes (tensile and compressive) are considered, that

    is, compressive loading along the zigzag direction (CZZ), ten-

    sile loading along the zigzag direction (TZZ), compressive

    loading along the armchair direction (CAC), and tensile load-

    ing along the armchair direction (TAC). The temperature is

    kept at 5 K using the Nose-Hoover thermostat. Only in-plane

    deformation is allowed so that wrinkles can be avoided thus

    enabling the study on deformation behavior of an ideal

    two-dimensional system. The strain and stress values are cal-

    culated according to Ref. 30. The evolution of atomic configu-

    rations and statistical bond lengths are analyzed in different

    stages by ATOMEYE.33

    Fig. 1 shows the TZZ, TAC, CZZ, and CAC stress-

    strain curves of the graphene nanoribbons. The positive and

    negative strain values indicate tensile and compressive load-

    ing, respectively. During TZZ and CAC, the absolute stress

    increases with strain smoothly but abruptly becomes zero at

    strain of 30% and 10%, corresponding to the destructive

    structural changes in the GNRs. On the other hand, during

    CZZ and TAC, three deformation stages are involved. In the

    first stage denoted by I, the stress also increases smoothly

    with strain, but stress oscillation corresponding to periodic

    change in the structure is observed from the second stage

    denoted by II (strain of about 12.5% and 20%). In the third

    stage denoted by III, the stress increases sharply initially and

    then suddenly becomes zero at strain of about 19% and 47%.

    In comparison with TZZ and CAC, the ductility in CZZ and

    TAC is enhanced by about 111% and 56.7%. The fracture

    strain in TAC is the largest (about 47%) whereas that in

    CAC is the smallest (10%). In the two compressive loading

    processes of CZZ and CAC, the atomic separation is short-

    ened and atomic interaction increases thereby producing the

    positive second-order elastic constants (second-order deriva-

    tive of stress with respect to strain) in stage I. However, in

    the two tensile loading processes of TZZ and TAC, the grad-

    ually reduced atomic interaction yields negative second-

    order elastic constants. The stress terrace in the strain range

    between 20% and 45% in TAC resembles elongation of the

    yield point induced by Cottrell atmospheres around disloca-

    tions and grain boundaries in low-carbon steels.34 It is thus

    obvious that the deformation mechanism of graphene sheets

    is determined by the combined effect of the loading mode

    and orientation.

    In order to understand the stress oscillation as well as

    enhanced ductility during CZZ and TAC, we systemically

    analyze the evolution of the atomic configuration and the

    typical snapshots are depicted in Fig. 2. Fig. 3 displays the

    FIG. 1. Stress-strain curves of graphene nanoribbons subjected to a strain

    rate of 10�4 ps�1 along different orientations (armchair or zigzag) and underdifferent loading modes (tensile or compressive) at 5 K. CZZ, TZZ, CAC

    and TAC represent compressive loading along the zigzag direction, tensile

    loading along the zigzag direction, compressive loading along the armchair

    direction, and tensile loading along the armchair direction, respectively, and

    I, II and III denote the elastic, shearing, and hardening deformation stages,

    respectively.

    FIG. 2. Typical snapshots showing the

    local atomic structures of graphene

    nanoribbons during loading of differ-

    ent modes: (a) TZZ at strain of 0% (I),

    10% (II), 20% (III), and 30% (IV); (b)

    CAC at strain of 0% (I), 5% (II), 8%

    (III), and 9% (IV); (c) TAC at strain of

    0% (I), 10% (II), 20% (III), and 47%

    (IV); (d) CZZ at strain of 0% (I), 10%

    (II), 12.5% (III), and 19% (IV).

    191906-2 Sun et al. Appl. Phys. Lett. 103, 191906 (2013)

  • statistical bond length in local regions with time. During

    TZZ, the length of bond 1 and bond 2 increases with strain

    but at a strain of 30%, the bond length decreases sharply to

    that of the ideal crystal lattice [Fig. 3(a)] as a result of brittle

    fracture as illustrated in Fig. 2(a). During CAC, the length of

    the two types of bonds decreases with strain and the rate is

    larger for bond 1 than bond 2 [Fig. 3(b)]. At about 10%

    strain, the length of bond 1 is 1.2 Å and the system is sub-

    jected to spuriously high bonding force yielding nonphysical

    results as shown in Fig. 2(b). If out-of-plane movement is

    allowed, wrinkles should occur in this case. A large lattice

    change is also observed during TAC and CZZ. Lattice shear-

    ing is activated from the edges of the GNRs along certain

    directions at 20% and 12.5% strain and then extends inwards

    gradually [Figs. 2(c) and 2(d)]. During lattice shearing, the

    bond length in the sheared regions changes suddenly. For

    instance, in TAC, the length of bond 1 diminishes from 1.7 Å

    to 1.42 Å and it is just as long as the length of bond 2 in the

    unsheared region. That of bond 2 increases dramatically

    from 1.42 Å to 1.63 Å, which is close to the length of bond 1

    in the unsheared region [Fig. 3(b)]. Similar abrupt changes

    in the bond length are also observed from the sheared regions

    during CZZ. The lattice change corresponds to phase trans-

    formation from hexagonal to orthorhombic.26,35 Lattice

    shearing proceeds gradually and an energy barrier is over-

    come resembling plane slipping in metals. This leads to

    stress oscillation in Stage II as shown in Fig. 1. Bond rupture

    at local regions occurs after completion of the phase transfor-

    mation in the whole system and fracture of GNRs occurs

    finally. Our results indicate that the phase transformation

    induced by lattice shearing is the origin of stress oscillation

    and enhanced ductility during CZZ and TAC.

    The orientation-dependent deformation behavior can be

    understood by comparing to bulk metals. The Schmidt factor

    is suggested to bridge the applied normal stress r and theresolved shear stress s along a given shearing direction in this

    two-dimensional system. It can be expressed as sin u � cos u,in which u denotes the angle between the loading and shear-ing directions.26 Since there is the largest atomic line density

    and the largest separation between atomic arrays along arm-

    chair directions, they are believed to be the possible shearing

    directions. In the TAC process, the Schmidt factor has a large

    value of about 0.48 and the resolved shear stress exceeds the

    critical stress (42.2 GPa (Ref. 26)) required by lattice shear-

    ing. Lattice shearing occurs after elastic deformation and

    before destructive fracture, as schematically shown in Fig.

    4(a). However, in the TZZ process, the Schmidt factor has a

    value of 0.44 and decreases with strain gradually. Hence, the

    resolved shear stress along the closely packed direction can-

    not activate the lattice shearing. Instead, brittle fracture takes

    place immediately after elastic deformation, as schematically

    shown in Fig. 4(b). Because of the orthorhombic relationship

    between the zigzag and armchair directions, CZZ is equiva-

    lent to TAC as a result of Poisson’s effect, while CAC is

    equivalent to TZZ. Therefore, lattice-shearing-enhanced duc-

    tility is found from both the TAC and CZZ processes, but not

    in the CAC and TZZ processes as discussed above. Table I

    summarizes the main results.

    The reversibility of the phase transformation is illus-

    trated during both CZZ and TAC when the lattice-sheared

    GNRs are unloaded at a controlled strain rate of 109 s�1 from

    a strain smaller than the fracture strain. Figs. 5(a) and 5(b)

    depict the stress-strain curves of a loading-unloading cycle

    during CZZ and TAC for maximum loading strain of 17%

    and 30%, respectively. In the low strain range between 0%

    and 9% in CZZ and 0% to 12% in TAC, the stress-strain

    curves in the unloading process almost coincide with those

    in loading process suggesting elastic deformation. However,

    at large loading strain, lattice-shearing-induced phase trans-

    formation from hexagonal to orthorhombic occurs and

    Stone-Wales defects and vacancies are produced sometimes.

    In both cases, an energy barrier should be overcome

    FIG. 3. Evolution of bond length as a

    function of strain during loading along

    (a) zigzag and (b) armchair directions

    with the positive and negative strain val-

    ues indicating tensile and compressive

    loading, respectively. When lattice

    shearing appears locally, “Sheared”

    denotes the bond length evolution in

    these regions and “Un-Sheared” in other

    regions.

    FIG. 4. The schematic models depict-

    ing the lattice evolution and the defor-

    mation mechanisms in two typical

    loading modes: (a) TAC and (b) TZZ.

    191906-3 Sun et al. Appl. Phys. Lett. 103, 191906 (2013)

  • correspondingly and the mechanical energy from the outside

    is stored as chemical energy in graphene. Upon unloading,

    the stored chemical energy will be dissipated suddenly as

    thermal energy but not mechanical energy. This leads to the

    difference in the stress-strain curves in the loading and

    unloading processes, and stress-strain hysteretic loops are

    formed. The magnified loops in grey are shown in the insets

    of panels (a) and (b) and the integrated areas reflect the

    energy dissipation. The calculated energy dissipation is less

    than 10% and quite small. It can be ascribed to the extraordi-

    narily low energy barrier of about 50 meV for the inverse

    phase transformation.35

    Fig. 6 displays the typical snapshots of the atomic config-

    uration during loading and unloading and pseudo-color is

    employed to indicate the magnitude of the shear strain on

    each atom. Lattice shearing occurs mainly along the armchair

    directions which are believed to have the highest resolved

    shear stress.26 During CZZ, the phase transformation induced

    by lattice shearing is nearly completed at 17% strain and the

    perfect hexagonal lattice is recovered upon controlled unload-

    ing [Fig. 6(a)]. No defects are produced in the circle. If com-

    pressive loading is applied again, the phase transformation

    takes place. The repeatable behavior endows graphene with

    pseudo-elasticity similar to that possessed by metal

    nanowires.36–38 During TAC as shown in Fig. 6(b), the bond

    length is significantly increased and bond fracture and rotation

    in TAC become easier compared to CZZ. Therefore, some

    SW defects are produced locally as shown in the inset of Fig.

    6(b). The defects are maintained during controlled unloading

    and prevent subsequent phase transformation under tensile

    loading. Therefore, the pseudo-elasticity at large tensile strain

    is degenerated, although the ductility along the armchair

    direction is the best. At an elevated temperature, bond fracture

    and rotation become easier and so is formation of defects.

    Accordingly, the ductility of graphene at high temperature is

    not good and it is contrary to the behavior of conventional

    bulk metals. This is also the reason why lattice shearing is sel-

    dom observed from graphene sheets. Since graphene sheets

    are usually distributed randomly in the composite host,

    different loading modes along various orientations may be

    activated at the same time. The resulting ductility of graphene

    sheets can be effectively utilized to improve the mechanical

    strength of composites.

    In summary, MD simulation is conducted to study the

    mechanical behavior of two-dimensional graphene and the

    deformation mechanism is determined by both the loading

    orientation and loading mode. When the load is applied

    along the zigzag orientation, tensile stress gives rise to brittle

    fracture, whereas compressive stress results in lattice shear-

    ing and phase transformation from hexagonal to orthorhom-

    bic. However, when the load is applied along the armchair

    direction, tensile stress leads to lattice shearing and phase

    transformation, but compressive stress yields large bonding

    force. The phase transformation induced by lattice shearing

    is reversible even at large strain of 17% and 30% during

    CZZ and TAC and the energy dissipation is less than 10%.

    As a result, pseudo-elasticity exists along some orientations

    under certain loading modes and the ductility is enhanced.

    Graphene sheets can be utilized as strengthening materials in

    composites to accomplish high elasticity, large strength, and

    low energy dissipation.

    FIG. 6. Evolution of atomic structures during loading and unloading: (a)

    CZZ and (b) TAC. The inset in (b) illustrates nucleation and development of

    SW defects.

    TABLE I. Deformation mechanisms of GNRs in four loading modes.

    Zigzag Armchair

    Tensile Brittle fracture Shearing enhanced ductile

    Compressive Shearing enhanced ductile Non-physical result

    FIG. 5. Loading-unloading stress-strain

    curves of graphene nanoribbons at 5 K:

    (a) CZZ and (b) TAC. The dissipation

    energy can be evaluated from the mag-

    nified grey area in the inset.

    191906-4 Sun et al. Appl. Phys. Lett. 103, 191906 (2013)

  • This work was jointly supported by Key Project of

    Chinese National Programs for Fundamental Research and

    Development (Grant No. 2010CB631002), National Natural

    Science Foundation of China (Grant Nos. 51271139,

    51171145, and 51302162), New Century Excellent Talents

    in University (NCET-10-0679), Fundamental Research

    Funds for the Central Universities, and Hong Kong Research

    Grants Council (RGC) General Research Funds (GRF) No.

    CityU 112212, and City University of Hong Kong Applied

    Research Grants (ARG) 9667038 and 9667066.

    1K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V.

    Dubonos, I. V. Grigorieva, and A. A. Firsov, Science 306, 666 (2004).2A. K. Geim, Science 324, 1530 (2009).3A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K.

    Geim, Rev. Mod. Phys. 81, 109 (2009).4S. Park and R. S. Ruoff, Nat. Nanotechnol. 4, 217 (2009).5C. N. R. Rao, A. K. Sood, K. S. Subrahmanyam, and A. Govindaraj,

    Angew. Chem., Int. Ed. 48, 7752 (2009).6P. Kumar, RSC Adv. 3, 11987 (2013).7J. N. Coleman, M. Lotya, A. O’Neill, and S. D. Bergin, Science 331, 568(2011).

    8P. Kumar, K. S. Subrahmanyam, and C. N. R. Rao, Mater. Express 1, 252(2011).

    9P. Kumar, B. Das, B. Chitara, and K. S. Subrahmanyam, Macromol.

    Chem. Phys. 213, 1146–1163 (2012).10W. Choi, I. Lahiri, R. Seelaboyina, and Y. S. Kang, Crit. Rev. Solid State

    Mater. Sci. 35, 52 (2010.).11P. Kumar, L. S. Panchakarla, and C. N. R. Rao, Nanoscale 3, 2127 (2011).12N. Behabtu, J. R. Lomeda, M. J. Green, and A. L. Higginbotham, Nat.

    Nanotechnol. 5, 406 (2010).13Y. Liu and B. Yakobson, Nano Lett. 10, 2178 (2010).14S. J. Chae, F. Gǔnes, K. K. Kim, E. S. Kim, G. H. Han, S. M. Kim, H. J.

    Shin, S. M. Yoon, J. Y. Choi, M. H. Park, C. W. Yang, D. Pribat, Y. H.

    Lee, Adv. Mater. 21, 2328 (2009).

    15P. Y. Huang, C. S. Ruiz-Vargas, A. M. Van der Zande, W. S. Whitney, M.

    P. Levendorf, J. W. Kevek, S. Garg, J. S. Alden, C. J. Hustedt, Y. Zhu, J.

    Park, P. L. McEuen, and D. A. Muller, Nature 469, 389 (2011).16D. A. Schmidt, T. Oha, and T. E. Beechem, Phys. Rev. B 84, 235422

    (2011).17Q. Lu and R. Huang, Phys. Rev. B 81, 155410 (2010).18M. Zhou, A. Zhang, Z. Dai, Y. P. Feng, and C. Zhang, J. Phys. Chem. C

    114, 16541 (2010).19C. Lee, X. D. Wei, J. W. Kysar, and J. Hone, Science 321, 385 (2008).20X. Zhao, Q. H. Zhang, D. J. Chen, and P. Lu, Macromolecules 43, 2357

    (2010).21B. Das, K. E. Prasad, U. Ramamurty, and C. N. R. Rao, Nanotechnology

    20, 125705 (2009).22R. Grantab, B. S. Vivek, and S. R. Rodney, Science 330, 946 (2010).23Y Wei, J. Wu, H. Yin, X. Shi, R. Yang, and M. Dresselhaus, Nature

    Mater. 11, 759 (2012).24J. Zhang, J. Zhao, and J. Lu, ACS Nano 6, 2704 (2012).25M. C. Wang, C. Yan, L. Ma, N. Hu, and M. W. Chen, Comput. Mater. Sci.

    54, 236 (2012).26F. Ma, Y. J. Sun, D. Y. Ma, K. W. Xu, and P. K. Chu, Acta Mater. 59,

    6783 (2011).27S. Plimpton, J. Comput. Phys. 117, 1 (1995).28S. J. Stuart, A. B. Tutein, and J. A. Harrison, J. Chem. Phys. 112, 6472

    (2000).29D. W. Brenner, O. A. Shenderova, J. A. Harrison, S. J. Stuart, B. Ni, and

    S. B. Sinnott, J. Phys.: Condens. Matter 14, 783 (2002).30O. A. Shenderova and D. W. Brenner, Phys. Rev. B 61, 3877 (2000).31G. Cocco, E. Cadelano, and L. Colombo, Phys. Rev. B 81, 241412 (2010).32J. Zhou and R. Huang, J. Mech. Phys. Solids 56, 1609 (2008).33J. Li, Modell. Simul. Mater. Sci. Eng. 11, 173 (2003).34M. Calcagnotto, Y. Adachi, D. Ponge, and D. Raabe, Acta Mater. 59, 658

    (2011).35P. W. Chung, Phys. Rev. B 73, 075433 (2006).36Q. Cheng, H. A. Wu, Y. Wang, and X. X. Wang, Appl. Phys. Lett. 95,

    021911 (2009).37H. S. Park, K. Gall, and J. A. Zimmerman, Phys. Rev. Lett. 95, 255504

    (2005).38F. Ma, S. L. Ma, K. W. Xu, and P. K. Chu, Appl. Phys. Lett. 92, 123103

    (2008).

    191906-5 Sun et al. Appl. Phys. Lett. 103, 191906 (2013)

    http://dx.doi.org/10.1126/science.1102896http://dx.doi.org/10.1126/science.1158877http://dx.doi.org/10.1103/RevModPhys.81.109http://dx.doi.org/10.1038/nnano.2009.58http://dx.doi.org/10.1002/anie.200901678http://dx.doi.org/10.1039/c3ra41149dhttp://dx.doi.org/10.1126/science.1194975http://dx.doi.org/10.1166/mex.2011.1024http://dx.doi.org/10.1002/macp.201100451http://dx.doi.org/10.1002/macp.201100451http://dx.doi.org/10.1080/10408430903505036http://dx.doi.org/10.1080/10408430903505036http://dx.doi.org/10.1039/c1nr10137dhttp://dx.doi.org/10.1038/nnano.2010.86http://dx.doi.org/10.1038/nnano.2010.86http://dx.doi.org/10.1021/nl100988rhttp://dx.doi.org/10.1002/adma.200803016http://dx.doi.org/10.1038/nature09718http://dx.doi.org/10.1103/PhysRevB.84.235422http://dx.doi.org/10.1103/PhysRevB.81.155410http://dx.doi.org/10.1021/jp105368jhttp://dx.doi.org/10.1126/science.1157996http://dx.doi.org/10.1021/ma902862uhttp://dx.doi.org/10.1088/0957-4484/20/12/125705http://dx.doi.org/10.1126/science.1196893http://dx.doi.org/10.1038/nmat3370http://dx.doi.org/10.1038/nmat3370http://dx.doi.org/10.1021/nn3001356http://dx.doi.org/10.1016/j.commatsci.2011.10.032http://dx.doi.org/10.1016/j.actamat.2011.07.036http://dx.doi.org/10.1006/jcph.1995.1039http://dx.doi.org/10.1063/1.481208http://dx.doi.org/10.1088/0953-8984/14/4/312http://dx.doi.org/10.1103/PhysRevB.61.3877http://dx.doi.org/10.1103/PhysRevB.81.241412http://dx.doi.org/10.1016/j.jmps.2007.07.013http://dx.doi.org/10.1088/0965-0393/11/2/305http://dx.doi.org/10.1016/j.actamat.2010.10.002http://dx.doi.org/10.1103/PhysRevB.73.075433http://dx.doi.org/10.1063/1.3183584http://dx.doi.org/10.1103/PhysRevLett.95.255504http://dx.doi.org/10.1063/1.2902293