Zhi Yanga,
Yuhong Huangb,
Fei Ma*ac,
Yaping Miaoac,
Hongwei Baoa,
Kewei Xu*ad and
Paul K. Chu*c
aState Key Laboratory for Mechanical Behavior of Materials, Xi'an Jiaotong University, Xi'an 710049, Shaanxi, China. E-mail: mafei@mail.xjtu.edu.cn; kwxu@mail.xjtu.edu.cn
bCollege of Physics and Information Technology, Shaanxi Normal University, Xi'an 710062, Shaanxi, China
cDepartment of Physics and Materials Science, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, China. E-mail: paul.chu@cityu.edu.hk
dDepartment of Physics and Opt-electronic Engineering, Xi'an University of Arts and Science, Xi'an 710065, Shaanxi, China
First published on 16th June 2016
A molecular dynamics (MD) simulation illustrates that different from single-crystal graphene sheets, the loading and unloading stress–strain curves of nanocrystalline ones do not coincide with each other, indicating substantial energy dissipation due to irreversible structural changes in the grain boundaries. An energy dissipation coefficient is proposed to quantitatively describe the effects of the grain size, temperature and strain rate dependent irreversible breaking and reforming of bonds in GBs, realignment of grain orientation, lattice-shearing-induced phase transformation, and formation of Stone–Wales defects and vacancies near GBs. The energy dissipation coefficient increases as the grain size decreases, especially at high temperature and low strain rate, and consequently, the reversibility of nanocrystalline graphene sheets under mechanical loading deteriorates compared to single-crystal graphene.
The as-prepared graphene sheets commonly contain various defects,12–16 and the graphene sheets fabricated by chemical vapor deposition (CVD) are usually polycrystalline. The grain boundaries (GBs) play dominant roles in the mechanical, electrical and chemical properties.17 Nano-indentation illustrated a slight reduction in the mechanical strength due to GBs,18,19 but some found that nanocrystalline graphene sheets have almost constant fracture strain and stress, independent on the grain size.20 The complex deformation behavior of nanocrystalline graphene sheets may be related to the structural evolution in GBs. Similar to metals, the GBs in polycrystalline graphene sheets can be considered as an array of edge dislocations but are usually disrupted by defects, and the dislocation density increases with increasing misorientation angle between the GBs.21,22 A large misorientation angle induces overlapping of neighboring dislocations and the strain field is compensated.23 As a result, large-angle tilt boundaries that have a large density of defects are as strong as the pristine one and even stronger than those with low-angle boundaries having fewer defects.24,25 Using nonequilibrium molecular dynamics (NEMD), Cao and Bagri illustrated that the thermal conductivity and Kapitza conductance is misorientation angle dependent.26,27 The first-principles calculations and classical molecular dynamics simulations have revealed an enhanced defect reactivity induced by an inhomogeneous strain field at grain boundaries.28 For nanocrystalline graphene sheets with GBs disrupted by various kinds of defects, the situation is more complicated. In this paper, the energy dissipation during mechanical deformation of nanocrystalline graphene sheets is described and an energy dissipation coefficient (η) is proposed to quantitatively describe the stability of the GBs in nanocrystalline graphene. It shows that energy dissipation increases with the grain boundary density, temperature and tensile strain, but is dependent on the strain rate little.
MD simulation is carried out using the LAMMPS (Large-scale Atomic/Molecular Massively Parallel Simulator) package. The interactions between carbon atoms are described by the adaptive intermolecular reactive empirical bond order (AIREBO) potential which can accurately describe the interactions between carbon atoms as well as bond breaking and reforming.31 The cutoff parameter describing the short-range C–C interaction is selected to be 2.0 Å in order to avoid spuriously large bonding forces and nonphysical results at large deformation.32 The Nosé–Hoover thermostat is utilized to account for the thermal effect.33 Uniaxial tensile loading is applied along the x axis using the deformation-control method.34 The atoms are allowed to move freely along the y axis and periodic boundary conditions are applied along the two in-plane directions. The interlayer separation of graphite, 3.4 Å, is taken as the effective thickness of the monolayer graphene. On the atomic level, the stress is computed according to the virial theorem.35 Prior to uniaxial tensile loading, the nanocrystalline graphene sheets are fully relaxed to an equilibrium state in the isothermal–isobaric ensembles at 3000 K for 20 ps initially and then relaxed at a given temperature for 600 ps. A time step of 1 fs is employed in the MD simulation and a Poisson's ratio of 0.165 is used.36 The reversibility is evaluated by releasing the nanocrystalline graphene sheets from a tensile strain of 5–25% at a given temperature and strain rate.
According to the principle of work and energy, the mechanical work W in the loading process is transferred into the elastic strain energy in the lattice and chemical energy in the GBs. The elastic strain energy in the lattice is released in the unloading process, but the chemical energy due to bond breaking and reforming in the GBs cannot be released.37 The area enclosed by the stress–strain curves in the loading and unloading processes describes the dissipated energy dW and an energy dissipation coefficient (η = dW/W) is proposed to quantitatively describe the process. The loading and unloading stress–strain curves of nanocrystalline graphene sheets with different grain sizes are plotted in the inset of Fig. 2(b). Accordingly the energy dissipation coefficient is calculated and the results are shown in Fig. 2(b) as a function of grain boundary densities. As the grain size decreases, the density of GBs increases, and the irreversible structural evolution during the loading process becomes more prominent. Consequently, the energy dissipation coefficient increases. For instance, in the graphene sheet with 10 nm grains corresponding to a grain boundary density of 0.016 nm−1, the average energy dissipation coefficient is 0.152. It increases to 0.19 by 25% in the graphene sheet with 2 nm grains and grain boundary density of 0.235 nm−1. Similar energy dissipation in mechanical loading has also been observed from traditional metals but it is produced by motion of dislocations,38 domain wall movement,39 grain boundary sliding,40 deformation twinning,41 and phase transformation.42
Fig. 3(a) displays the evolution of the atomic configuration in the graphene sheets with a grain size of 4.0 nm during the loading and unloading processes. The map is colored according to the potential energy of each atom. The potential energy increases with tensile strain up to 15% gradually, particularly at GBs, and then decreases upon unloading. When the stress is reduced to zero, a residual strain about 2.1% remains. Fig. 3(b) presents the enlarged atomic configurations of a representative region at three stages: (1) at a strain of 0% before loading, (2) loaded up to a strain of 15%, and (3) unloaded down to 2.1%. The red, black, and blue arrows track the evolution of the atomic configurations in the vicinity of the three defects which are metastable and can be maintained at low temperature,29 but they may be changed to low-energy configurations at a high temperature or under mechanical loading.30 When the strain is increased up to 15%, a new bond is formed at the left of the large ring marked by the red arrow to promote the formation of stable pentagon rings, while bond breaking and reforming take place subsequently in the local region indicated by the blue arrow. This results in the transformation from a heptagon–octagon topological defect into a pentagon–heptagon one. In the region indicated by the black arrow, a distorted octagon ring changes to a regular one during tensile loading. Moreover, lattice-shearing-induced phase transformation from hexagonal structure to orthorhombic one occurs, accompanied with Stone–Wales defects and vacancies. Since the atoms in GBs have higher energy, and they tend to reconstruction, which might promotes the grain boundary sliding and realignment, similar to that in metals. Hence, the final atomic configurations in the local regions cannot recover to the original states upon unloading owing to: (1) breaking and reforming of bonds in GBs, (2) realignment of grain orientation, (3) lattice-shearing-induced phase transformation, (4) formation of Stone–Wales defects and vacancies near GBs. These result in a residual strain after unloading, but the residual strain is reduced in the second loading and unloading to some degree. Besides grain boundaries, impurities and voids have also been observed in polycrystalline graphene sheets fabricated by chemical vapor deposition, and so on. The movements of impurities and voids should also result in energy dissipation. However, in the work of this manuscript, we mainly focus on the influences of GBs on the plastic deformation as well as on the energy dissipation in nanocrystalline graphene sheets, and thus only GBs are produced in the initial simulation models. The effects of other defects induced in the deformation process cannot be completely avoided, but it is difficult to be evaluated.
![]() | ||
Fig. 3 (a) Snapshots of the atomic configurations in a loading–unloading cycle; (b) structural evolution in local regions during the loading and unloading processes. |
Bond breaking and reforming in GBs are sensitive to the temperature and strain rate in mechanical loading.43 Therefore, the energy dissipation coefficient should also be temperature and strain rate dependent. Taking the graphene sheets with grains in the sizes of 2.2 nm, 4.0 nm and 6.6 nm as examples, two sets of MD simulations are performed: (1) in the temperature range of 10–1500 K but at a given strain rate of 108 s−1 and (2) in the strain rate range of 108 to 1010 s−1 but at a given temperature of 300 K. Fig. 4(a) and (b) show the energy dissipation coefficients as a function of temperature and strain rate, respectively. As shown in Fig. 4(a), the energy dissipation coefficients increase with temperature almost linearly, for example, from 0.075 at 10 K to 0.38 at 1500 K by 407% for the graphene sheet with 4.0 nm grains. Kinetically, the atoms in the GBs oscillate more at higher temperature. The energy barrier can be overcome more easily to promote bond breaking, rotation, and reforming in the GBs in addition to the formation of polygonal holes44 and buckling along the GBs.45 As shown in Fig. 4(b), the energy dissipation coefficients decrease with increasing strain rate slightly. At lower strain rate, the atoms in the GBs have enough time to rearrange during tensile loading and the irreversible structural evolution is more substantial thereby leading to a larger energy dissipation coefficient.44 However, the dependence on the strain rate is not appreciable. Furthermore, the graphene sheets with smaller grains have larger grain boundary densities [Fig. 1(a)] and irreversible bond rearrangement is more likely in the GBs. Hence, the energy dissipation coefficient becomes more sensitive to the temperature and strain rate in the graphene sheets with smaller grains.
![]() | ||
Fig. 4 (a) Energy dissipation coefficients of nanocrystalline graphene sheets at different temperature; (b) logarithmic plots of the energy dissipation coefficients a function of strain rate. |
Large strain usually leads to more structural rearrangement in the GBs and hence, energy dissipation is more substantial. Unloading from different strain values is simulated in the temperature range of 10–1500 K. Fig. 5(a) shows the simulated stress–strain curves of graphene sheets with a grain size of 4.0 nm at 300 K. Elastic lattice distortion dominates at small strain, but bond rearrangement in the GBs becomes more significant with increasing strain. Consequently, the residual strain increases from 0.23% to 2.1% and the enclosed area by the loading and unloading stress–strain curves is substantially enlarged. The strain from where the graphene sheets are unloaded increases from 5% to 15% and the energy dissipation coefficients increase with the maximum unloading strain, particularly at a high temperature [Fig. 5(b)]. For example, the energy dissipation coefficient at 300 K increases by 62.8% when the maximum unloading strain increases from 5% to 15%, whereas it increases by 115% at 1500 K. Tight-binding MD simulation indicates that evaporation of carbon dimers and sequential Stone–Wales transformation of carbon bonds lead to rapid motion and annihilation of the GBs particularly those with large carbon rings.46 The larger energy dissipation coefficient can be attributed to enhanced reconstruction of the local atomic configuration in the GBs.
![]() | ||
Fig. 5 (a) The loading stress–strain curves and those unloaded from different maximum strains; (b) energy dissipation coefficient at different temperatures. |
This journal is © The Royal Society of Chemistry 2016 |