Characterization of the Melting Transition in Two Dimensions at Vanishing External Pressure Using Molecular Dynamics Simulations

Daniel Asenjo1,4, Fernando Lund1, Sim´onPoblete2, Rodrigo Soto1, and Marcos Sotomayor3 1Departamento de F´ısica and CIMAT, Facultad de Ciencias F´ısicas y Matem´aticas, Universidad de Chile, Santiago, Chile 2Max Planck Institute for Polymer Research, Ackermannweg 10, 55128 Mainz, Germany 3 Howard Hughes Medical Institute and Neurobiology Department, Harvard Medical School, Boston, MA, USA 4Department of Chemistry, University of Cambridge, Lensfield Road, Cambridge CB2 1EW, United Kingdom.

A molecular dynamics study of a two dimensional system of particles interacting through a Lennard-Jones pairwise potential is performed at fixed temperature and vanishing external pressure. As the temperature is increased, a solid-to-liquid transition occurs. When the melting temperature Tc is approached from below, there is a proliferation of pairs and the elastic constant approaches the value predicted by the KTHNY theory. In addition, as Tc is approached from above, the relaxation time increases, consistent with an approach to criticality. However, simulations fail to produce a stable using systems with up to 90,000 particles. A significant jump in enthalpy at Tc is observed, consistent with either a first order or a continuous transition. The role of external pressure is discussed.

I. INTRODUCTION On the numerical side, molecular dynamics and Monte- Carlo simulations19,20 of systems with a small number of particles (N) broadly detected a transition where the Melting of an infinite solid in two dimensions has number of proliferates, but failed to pro- been described as a process driven by a proliferation vide clear evidence for the nature of the observed tran- of thermally excited dislocation pairs in the Kosterlitz- sition. First order melting has been reported in the Thouless-Halperin-Nelson and Young (KTHNY) the- 19,21–23 1–4 literature, , while other calculations support a con- ory . The theory, formulated at vanishing external tinuous transition24,25. pressure, predicts the existence of a new, “hexatic”, inter- mediate thermodynamic phase. While solids are charac- The critical properties of the KTHNY transition are terized by long range translational and orientational or- a consequence of the effect that small der, liquids only present short range order. The predicted scale fluctuations have on large scale fluctuations. For hexatic phase presents long range orientational order but this mechanism to be operative, well separated length lacks long range translational order. The KTHNY theory scales must exist, suggesting a minimum size for numer- predicts a second order from the crys- ical simulations in two dimensions of 104. Indeed, Chen talline to the hexatic phase at which point there is a uni- et al. 26 performed molecular dynamics simulations of a versal jump of a normalized elastic constant from a finite Lennard-Jones system with a varying number N of par- value to zero. This first transition is followed by a second ticles. They found a metastable hexatic phase for sys- transition from the hexatic phase to the liquid phase at tems with N ≥ 36, 864, but not for N ≤ 16, 386. In all a higher temperature. The KTHNY theory continues to cases, simulations were performed at a significant exter- generate interest, specially because increased numerical nal pressure, a fact that alters the dislocation generation capabilities and new experimental techniques currently mechanism: the interaction between the components of a allow for new and more accurate testing of theoretical dislocation pair tends to close it down, while the external predictions. Indeed, there have been numerous attempts pressure, for some orientations, tends to open it up. The at the verification, both experimentally and numerically, whole process becomes one of thermal activation, much of the KTHNY theoretical predictions, with mixed out- like nucleation, and the likelihood of having isolated comes. dislocations—and an hexatic phase—increases. A subse- quent study in terms of inherent structure theory showed On the experimental side, studies with colloidal par- consistency with the KTHNY theory27. More recently, ticles have provided evidence of two stage melting with a molecular dynamics study28 carried out at constant an intermediate hexatic phase5–8, and of elasticity be- 9 volume, involving 36,000 particles interacting through a arXiv:1101.4337v2 [cond-mat.stat-mech] 18 Mar 2011 havior in agreement with the KTHNY predictions . Lennard-Jones potential also reported the presence of an The observed transition, however, appears to be first hexatic phase between the solid and liquid phases. How- order8,10. Similar results have been obtained with diblock ever, phase coexistence in NVT ensembles precludes un- copolymers11. Recently, melting in two steps with an in- ambiguous interpretation of these results. termediate hexatic phase has been observed in monolay- ers of polycristalline colloidal films, but not in thin or On a different vein, in a three dimensional continuum thick multilayer films12. Also recently, but in a differ- elastic solid, dislocation loops drive a mechanical instabil- ent context, dislocations have been directly observed in ity at a finite temperature29,30, at which point the shear graphene13, prompting a renewed interest on the role of modulus vanishes as a function of reduced temperature, defects in this two dimensional material14–18. following a power law with an exponent whose value is a 2 function that is independent of the microscopic details 0.005% per million iterations. We also verified momen- of the elastic solid. Numerical calculations of super- tum conservation and monitored pressure and tempera- heated Lennard-Jones crystals near the melting transi- ture throughout simulations. Thermostat and barostat 36 tion in three dimensions show the appearance of dislo- frequencies ωp and ωb were set to values between 80 cations as the temperature is raised31,32. However, it and 100, and 0.1 and 0.4, respectively. The cutoff radius is unclear whether these dislocations play a central role was set to 4 and a Verlet-neighbours list was used with during the phase transition or are just a by-product of a radius of 5.9. another transition-driving mechanism. Initial conditions were set by initial positions corre- The present work used molecular dynamics (MD) sim- sponding to a perfect triangular lattice, or a perfect tri- ulations in the NpT ensemble to study the melting tran- angular lattice plus a randomly oriented displacement sition in two dimensions. MD has been chosen in this of 0.05 LJ units, as well as initial velocities given by a occasion over the possible Monte Carlo alternative in Gaussian distribution corresponding to each temperature order to study the time evolution of the system, espe- (initial condition IC1). Otherwise, initial conditions were cially its relaxation behavior near criticality. We found assigned by equilibrium values of positions and velocities a single-step solid-to-liquid transition (as determined by obtained from a previous run with similar values of T the enthalpy change) when a vanishing external pressure and p (initial condition IC2). Most runs were carried out was applied to the system, in contrast to a multi-step with N = 36, 864, with up to 1.7 × 107 time iterations, transition at high pressure like the one presented in26 and a smaller number of runs with N = 90, 000. using the same number of particles. However, within a To characterize the state of the system throughout sim- narrow temperature interval defining the solid-to-liquid ulations we monitored the time evolution of the system’s transition, the monitored relaxation times, elastic con- enthalpy and computed the pair and orientational cor- 26 stants, and the evolution in number of dislocations were relation functions g(r) and g6(r), respectively . The all consistent with the KTHNY theory. We suggest that, Lam´eparameters were also computed, as a function of because of the necessary interplay between many length temperature. The number of dislocation pairs present in scales, a stable hexatic phase will be unambiguously ob- the crystal below the transition was examined counting served only in systems with at least ∼ 106 particles at the number of nearest neighbors for each particle, and zero external pressure. through visualization with the aid of a Voronoi construc- tion.

II. MOLECULAR DYNAMICS SIMULATIONS III. RESULTS Molecular dynamics (MD) simulations were carried out using a parallel MD code developed “in house” based on A. Simulations at different temperatures show the libraries presented in35. Simulation systems com- melting transition prised N identical particles of mass m in two dimensions interacting through a pairwise truncated and shifted Two sets of simulations were performed to study the Lennard-Jones (LJ) potential, melting behavior of a LJ system. The first set of simu- lations, aimed at reproducing the results presented in26,   σ 12 σ 6 were performed at p = 20 with N = 36, 864. Three sim- φLJ(r) = 4 − − C (1) r r ulations performed at temperatures T1 = 2.15, T2 = 2.16 and T3 = 2.17 were carried out using a perfect crys- for inter-particle distances r smaller than a cutoff ra- talline lattice as initial condition. Despite the use of a dius rc and 0 for r ≥ rc. The value of C was chosen different set of equations of motion that are modularly to ensure continuity of the potential. All simulations invariant, we obtained results that are consistent with were performed using periodic boundary conditions with those presented in26. The enthalpy h as a function of a fully-flexible cell in the NpT ensemble defined by the time remained stable during the simulation at T1, in- non-Hamiltonian equations of motion described by Mar- creased rapidly in an apparent single step to achieve a 36 tyna, Tobias, and Klein . The equations of motion were stable value during the simulation performed at T3, and integrated using the 5-value Gear predictor-corrector al- increased in two steps with a transient state in the sim- gorithm. ulation performed at temperature T2 (data not shown). All simulation parameters and monitored quantities The pair and orientational correlations functions, g(r) ∗ ∗p 2 are expressed in reduced units: x = x σ; t = t mσ /; and g6(r), were consistent with a solid phase for the ∗ T = T /kB, where kB is the Boltzmann constant; E = simulation performed at T1 and with a liquid phase at ∗ ∗ 3 E ; and p = p /σ (in the case of Argon,  = 0.0104 eV the end of simulations performed at T2 and T3. The ˚ and σ = 3.4 A, leading to a a unit of temperature T be- transient state observed at T2 exhibits long-range orien- ing 120.6 K and a unit of pressure p being 42 MPa). The tational but not translational order, consistent with an integration time step was set to 0.0005pmσ2/ (34 fs hexatic phase. Similar results were obtained with a simu- 26 for Argon) , ensuring extended-energy conservation to lation performed at T2 = 2.16 that used a thermalized (at 3

-65000 ble at low temperature (T ≤ 0.40725), but increased in an apparent single-step transition to an equilibrium -70000 value for high temperatures (T ≥ 0.4095). The nature of the phase was characterized using g(r) and g6(r) at the ending configurations of each simulation, confirm- -75000 ing a liquid phase for T ≥ 0.4095, and a solid phase for T ≤ 0.40725 (Figure 2). As opposed to the results

Enthalpy -80000 obtained at p = 20, no intermediate and transient state (possibly corresponding to a hexatic phase) was observed. It is possible that a hexatic phase with an algebraically -85000 decaying g6(r) and an exponentially decaying g(r) could be found within the interval of temperatures given by -90000 T− = 0.40725 < T < T+ = 0.4095. However, it was 0 1000 2000 3000 4000 5000 6000 7000 8000 not possible to reach equilibrium in between these two Time temperature values within the available simulation time -65000 scales (data not shown). Defining the relative change in enthalpy as ∆h ≡ 2(h+ − h−)/(h+ + h−) (where h ≡ h(T )) and the relative change in temperature as -70000 ± ± ∆T ≡ 2(T+ − T−)/(T+ + T−), we find ∆h = 0.1433 and ∆T = 0.0044. Within this narrow temperature interval, -75000 our data is consistent both with an abrupt jump from a low to a high enthalpy value at some intermediate tem- perature, as would be the case for a first-order transition Enthalpy -80000 (i.e., with latent heat), as well as with a two-step change of enthalpy as a function of temperature within the tem- -85000 perature interval, as would be the case for a continuous transition, without latent heat (Figure 3).

-90000 0 1000 2000 3000 4000 5000 6000 7000 8000 Time B. Relaxation times slightly above melting increase as the melting temperature is approached FIG. 1: Enthalpy as a function of time for several temper- atures. Upper panel: initial conditions are given by initial positions corresponding to a perfect triangular lattice plus a Near a critical point, relaxation times increase as crit- randomly oriented displacement of 0,05 in LJ units, and initial icality is approached. This is due to the increasing size velocities given by a Gaussian distribution corresponding to of fluctuations37, that reach macroscopic dimensions at each temperature (initial condition IC1). From the bottom- the critical point. To determine the nature of the phase right to the top-left curves the temperatures are T=0.4050, transition observed at p = 0, we monitored the relaxation 0.4095, 0.4130, 0.4160, 0.4200, and 0.4250.Lower panel: posi- tions and velocities are provided, at each temperature, by the time (tR) as a function of temperature in the simulations equilibrium values obtained by a previous run at T = 0.40725 mentioned at T > Tc (Figure 4). (initial condition IC2). From the bottom-right to the top- The relaxation time tR is defined, grossly, as the in- left curves the temperatures are T=0.40725, 0.41000, 0.41250, stant where the second derivative of enthalpy vs time 0.41500, 0.42000, 0.42500, and 0.44000. All simulations were vanishes, and more precisely as follows: the curve en- carried out at vanishing external pressure. thalpy vs time is interpolated by a smooth function whose second derivative is computed numerically, and the times tmax, where curvature is a maximum and tmin, where it T = 2.15) set of initial conditions. These results show the is a minimum, determined. The relaxation time is then possible existence of a metastable and transient hexatic defined through tR = tmax + (tmin − tmax)/2. phase when simulations are performed at a high external Within the accuracy of the simulation, this time grows pressure p = 20, as presented in26. without limit as the melting temperature is approached, To test the behavior of the system at a vanishing exter- consistent with an approach to criticality and a second nal pressure, we performed a second set of simulations in order phase transition. The range of values that was which a system with N = 36, 864 particles starting from explored, however, was not large enough to detect a pos- initial conditions IC1 and IC2 was simulated at several sible power law behavior. These results did not depend different temperatures and pressure p = 0 (more pre- on the initial conditions used (IC1 or IC2) and on the cisely, with vanishing normal and tangential stresses). critical temperature used (Tc = 0.40725 or Tc = 0.4095, As with the first set of simulations, the enthalpy h of as there is no exact melting temperature but rather an the system as a function of time (Figure 1) was sta- interval [T−,T+]). 4

6 T=0.40725 IC1 Tc=0.40725 T=0.41000 5000 IC2 Tc=0.40725 5 IC1 Tc=0.40950 IC2 Tc=0.40950

4

3 g(r)

2

500 1

0 0 5 10 15 20 25 30 35 4045 Relaxation Time r[σ] 10 T=0.40725 T=0.41000 1 0.01 (T-Tc)/Tc 0.1 FIG. 4: Relaxation time as a function of temperature as the

6 0.01

g (r) transition temperature Tc is approached from above. There are four set of points, corresponding to the two different ini- 0.001 tial conditions (IC1 and IC2) indicated in Figure 1, and two possible values for Tc, determined by the finite interval in 0.0001 which Tc is found. The time increases without limit as Tc is approached, consistent with criticality. Error bars are given -5 10 by tmin − tmax. 1 r[σ] 10

FIG. 2: Upper panel: Pair correlation function at T = C. Behavior of elastic constants is consistent with 0.40725 indicates long-range translational order, and at T = KTHNY theory 0.41000, absence of it. Lower panel, orientational correlation function at T = 0.40725 indicates long range order, and at T = 0.41000, lack of it. Another way to determine the nature of the phase tran- sition at p = 0 and whether it is consistent with theoreti- cal predictions consist on monitoring the behavior of the -72000 elastic constants of the system. The elastic response of -74000 an isotropic, homogeneous, continuum solid is character-

-76000 ized by two Lam´ecoefficients λ and µ. They appear in the compliance tensor as -78000

-80000

-82000 1  λ  Sijkl = δikδjl + δilδjk − δijδkl (2) -84000 4µ λ + µ Enthalpy -86000

-88000 where -90000

-92000 0.37 0.38 0.39 0.4 0.41 0.42 Temperature ij = Sijklσkl,

FIG. 3: Final enthalpy as a function of temperature. It presents a significant increase at Tc = 0.40815 ± 0.00090. The ij is the strain and σkl the stress. There are several low temperature phase is a solid, as evidenced by the behav- possible ways of extracting λ and µ from this tensor. In ior of pair and orientational correlation functions. The high the continuum theory they are of course equivalent. But temperature phase is a liquid (Figure 2). in a numerical calculation involving a finite number of atoms, this will no longer necessarily be the case. They should coincide, however, within numerical accuracy and error bars (see below). The following relations hold for 5 the combination of Lam´ecoefficients: 4µ(λ + µ) K ≡ (3) 2µ + λ 1 = ≡ K1 (4) S0000 1 = ≡ K2 (5) S1111 1 FIG. 5: Elastic constant K1/Kc (left-hand-side panel), = ≡ K3 (6) K2/Kc (middle panel) and K3/Kc (right-hand -side panel) as S0011 + 2S0101 a function of temperature. Within numerical accuracy they coincide, as they should. Near the transition their value is It is a significant prediction of KTHNY theory that K consistent with 1, as predicted by KTHNY theory. approaches a universal value as the critical temperature Tc is approached from below:

kBTc lim K = 16π ≡ Kc (7) − 2 T →Tc b lattice, P6 = 1 and Pk = 0 for n 6= 6. A dislocation is characterized by two neighboring sites, one with five, where b is the Burgers vector of the dislocations (given and the other with seven, neighbors. The KTNHY the- by the lattice constant at zero temperature). K vanishes ory predicts that the loss of long range translational order above Tc. is due to the proliferation, and subsequent unbinding, of We have computed the strain of our system following thermally generated dislocation pairs. Such pairs will be Ray and Rahman39,40, characterized then by clusters of four sites, two of them with five, and two of them with seven, neighbors. Figure 1  −1 t t −1  6 shows that across the solid-to-liquid transition there is  = (hR ) h hhR − I (8) 2 a significant decrease in P6, and a corresponding increase where h is a two-by-two matrix whose column vectors in P5 and P7, consistent with the KTHNY theory. define the simulation box, hR is a reference box (here taken as the time average of the simulation box) and I is the identity matrix. The compliance tensor is given in 1 P4 terms of strain fluctuations through P5 0.9 P6 P7 P 0.8 8 Sijkl = βVR(hijkli − hijihkli) (9)

k 0.7 where VR is the volume of the reference box. 0.6 The compliances were calculated from simulations of 2.5 × 106 time steps, after 5 × 105 equilibration steps. 0.5

The error bars were estimated by performing blocking 0.4 averages33,34 and then propagating the error in equation 0.3

9. The simulation data was divided into 5 data blocks. Defect Fraction, P The value of the Burgers vector was estimated as the lat- 0.2 tice constant that minimizes the energy of a triangular 0.1 crystal, b0 = 1.11145σ. Figure 5 shows the ratios K1/Kc, 0 K2/Kc and K3/Kc as a function of temperature as the 0.4 0.405 0.41 0.415 0.42 0.425 transition is approached from below. The computed val- Temperature ues are consistent with the KTHNY theory. FIG. 6: Pk, the fractional number of sites with k neighbors, as a function of temperature for k = 4, 5, 6, 7, 8. There is a sig- nificant decrease in P6 across the solid-liquid transition with D. Proliferation of dislocations is consistent with a corresponding increase in P5 and P7, consistent with a tran- KTHNY theory sition driven by the unbinding of dislocation pairs.

Finally, and to fully characterize the transition ob- served at p = 0, we monitored the number of disloca- tions as a function of temperature, counting the number Figure 7 also provides a visual illustration of the num- of neighbors for each point using a Voronoi construction ber of dislocations, monitored with the number of sites after equilibrium had been reached. Figure 6 shows Pk, having 5 or 7 nearest neighbors, within the simulation the fractional number of sites with k neighbors, as a func- box for different temperatures. Their proliferation is ap- tion of temperature. Of course, for a perfect triangular parent, consistent with the KTNHY theory. 6

1 1 erably when an external pressure p is included. Indeed, the energy U of a dislocation dipole with Burgers vector ~b, whose components are separated by R~ in this case is38

y 0.5 y 0.5 b2K  R 1  U(R) = log + C − cos 2θ + P bR sin θ 4π τ 2 (10) 0 0 0 0.5 1 0 0.5 1 where τ is the dislocation core size; C, roughly, deter- x x mines the core energy (i.e., the minimum energy needed 1 1 to generate a dipole); R = |R~|; b = |~b| and θ is the an- gle between R~ and ~b. Clearly, when sin θ < 0 the last term on the right hand side turns the dislocation-dipole y y 0.5 0.5 unbinding process into a thermally activated one, with an activation energy UA (taking θ = −π/2, the most favorable case, for illustration purposes) 0 0 0 0.5 1 0 0.5 1 b2K   1 K b  1 x x U ≡ log + C − . A 4π 4π P τ 2 FIG. 7: Dislocation dipole population as a function of tem- perature, monitored with the sites having 5 (red) and 7 (blue) Consequently, and given the logarithmic dependence on neighbors. Upper left panel, T = 0.400; upper rigth panel, the ratio of external pressure to elastic constant, even a T = 0.405; lower left panel, T = 0.40725, all three temper- modest value of external pressure P , compared to K, will ature values below Tc. Lower rigth panel, T = 0.4095 just give values for the activation energy in the same ball park above Tc. The number of dislocation dipoles steadily increases as the chemical potential for the dislocation dipole, and a as the transition is approached. Sites with 8 neighbors (green) proliferation of isolated dislocations will ensue. Thus, it are also indicated. will not be surprising in those circumstances to observe an hexatic phase. However, Eqn. (10) shows that, at any given non vanishing pressure there will be a finite IV. DISCUSSION: THE ROLE OF EXTERNAL rate of dislocation generation, driving the system away PRESSURE from equilibrium. A quantitative study of this interesting phenomenon is outside the scope of the present paper. An hexatic phase has been observed in simulations pre- In the absence of external pressure, disclination pairs sented by Chen et al.26 and also reproduced here. How- above the dislocation unbinding transition interact via ever, this phase is transient, not in equilibrium, and oc- an energy that depends on the logarithm of their mu- curs when the Lennard-Jones system is subjected to a tual distance, with a coupling (called KA by Nelson and significant external pressure of 20. The use of a constant Halperin2) that is finite due to the screening effect of the pressure ensemble molecular dynamics ensure the exis- free dislocations. A second transition towards the liquid tence of a homogeneous phase. However, we have been state thus occurs because of two distinct screenings: free unable to observe the hexatic phase at a constant van- dislocations screen the interaction between disclination ishing external pressure. Is there a reason, within the pairs to an effective logarithmic interaction, and then KTHNY theory, not to observe an hexatic phase at zero this interaction is renormalized because of the interac- external pressure with a finite number of particles? tion between disclination pairs at different length scales. The hexatic phase arises2 after (i.e., at a higher tem- So, three length scales should be needed for this scale de- perature) a triangular lattice undergoes a dislocation un- pendent interaction among disclination pairs to become binding transition but before (i.e., at a lower temper- operative. In addition, a further length scale would seem ature) a disclination unbinding transition occurs. The to be necessary in order to have enough free dislocations latter is possible because the plasma of free dislocations in between a disclination pair for their interaction to be screens the disclination-disclination interaction, allowing effectively screened. According to this reasoning, at least a transition much like the one originally considered by ∼ 106 particles would be needed to observe an hexatic Kosterlitz and Thouless1. As mentioned in the Intro- phase as an equilibrium phase at zero external pressure duction, it is critical, in a numerical simulation, to have in a numerical simulation. several length scales available, since the KTHNY mech- anism involves the interaction among defects of different sizes, at many scales. This is in addition to the fact that V. CONCLUSIONS the theoretical analysis is carried out in the thermody- namic limit. We now argue that at least 106 particles are The KTNHY theory of melting in two dimensions1–3 needed in simulations, in order to have three decades in involves the interaction among dislocation dipoles whose length scales. sizes span different length scales. Thus, a numerical sim- The physics of dislocation unbinding changes consid- ulation that aims at verifying the theory should involve 7 enough different length scales for this interaction to be simulations. There could be an abrupt discontinuity, as possible. In two dimensions, 104 particles thus appears as in a first order transition, or there could be a smooth an a absolute minimum. Increasing this number should change, including a temperature range with an hexatic 6 improve the statistics. By the same token, 10 particles phase. Above Tc, the relaxation time increases as Tc is should be a minimum number to capture this type of approached, consistent with criticality. effect in three dimensions. Our simulation has been car- ried at constant (vanishing) external pressure in order to prevent phase coexistence. We find a solid-to-liquid transition, and the behavior of the solid phase as the Acknowledgments transition temperature is approached is consistent with the predictions of KTNHY. The behavior of enthalpy as a We thank Rodrigo Arias and Patricio Cordero for use- function of temperature is less conclusive: it changes sig- ful discussions. This work was supported by Fondecyt nificantly across a narrow temperature range ∆T , from a Grants 1100100 and 1100198, Fondap Grant 11980002, solid low temperature phase to a liquid high temperature Anillo ACT 127 and a Conicyt 2002 fellowship (M.S.). phase. The behavior within ∆T could not be resolved be- M. S. is a Howard Hughes Medical Institute Fellow of cause of the limited time-scale that can be reached with the Helen Hay Whitney Foundation.

1 J. M. Kosterlitz and D. J. Thouless, J. Phys. C 6, 1181 22 A. F. Bakker, C. Bruin and H. J. Hilhorst, Phys. Rev. Lett. (1973). 52, 449 (1984). 2 B. I. Halperin and D. R. Nelson, Phys. Rev. Lett. 41, 121 23 J. Lee and K. J. Strandburg, Phys. Rev. B 46, 11190 (1978); D. R. Nelson and B. I. Halperin, Phys. Rev. B 19, (1992). 2457 (1979). 24 M. Li, Phys. Rev. B 62, 13979 (2000). 3 A. P. Young, Phys. Rev. B 19, 1855 (1979). 25 J. F. Fern´andez,J. J. Alonso and J. Stankiewicz, Phys. 4 D. R. Nelson, Defects and Geometry in Condensed Matter Rev. E 55, 750 (1997). Physics (Cambridge University Press, 2002). 26 K. Chen, T. Kaplan and M. Mostoller, Phys. Rev. Lett. 5 C. A. Murray and D. H. Van Winkle, Phys. Rev. Lett. 58, 74, 4019 (1995). 1200 (1987). 27 F. L. Somer, G. S. Canright, T. Kaplan, K. Chen and M. 6 R. E. Kusner, J. A. Mann, J. Kerins and A. J. Dahm, Phys. Mostoller, Phys. Rev. Lett. 79, 3431 (1997); F. L. Somer, Rev. Lett. 73, 3113 (1994). G. S. Canright and T. Kaplan, Phys. Rev. E 58, 5748 7 K. Zahn, R. Lenke and G. Maret, Phys. Rev. Lett. 82, (1998). 2721 (1999). 28 H. Shiba, A. Onuki and T. Araki, EPL 86, 66004 (2009). 8 B.-J. Lin and L.-J. Chen, J. Chem. Phys. 126, 034706 29 F. Lund, Phys. Rev. Lett. 69, 3084 (1992) (2007). 30 R. Arias and F. Lund, Defect and Diffusion Forum 150– 9 H. H. von Gr¨unberg, P. Keim, K. Zahn and G. Maret, 151, 121 (1997). Phys. Rev. Lett. 93, 255703 (2004). 31 Z. H. Hin, P. Gumbsch, K. Lu and E. Ma, Phys. Rev. Lett. 10 A. H. Marcus and S. A. Rice, Phys. Rev. Lett. 77, 2577 87, 055703 (2001). (1996). 32 L. G´omez,A. Dobry, Ch. Geuting, H. T. Diep, and L. 11 D. E. Angelescu, C. K. Harrison, M. L. Trawick, R. A. Burakovsky, Phys. Rev. Lett. 90, 095701 (2003). Register and P. M. Chaikin, Phys. Rev. Lett. 95, 025702 33 D. Frenkel and B. Smit, Understanding Molecular Simu- (2005). lation: From Algorithms to Applications, Academic Press 12 Y. Peng, Z. Wang, A. M. Alsayed, A. G. Yodth and Y. (2002). Han, Phys. Rev. Lett. 104, 205703 (2010). 34 W. Janke, Statistical Analysis of Simulations: Data Corre- 13 A. Hashimoto, K. Suenaga, A. Gloter, K. Urita and S. lations and Error Estimation, published in Quantum Sim- Iijima, Nature 430, 870 (2004). ulations of Complex Many-Body Systems: From Theory to 14 A. Carpio, L L. Bonilla, F. de Juan, and M. A. H. Vozme- Algorithms, Edited by J. Grotendorst , D. Marx and A. diano, New Journal of Physics 10, 053021 (2008). Muramatsu. John von Neumann Institute for Computing, 15 B. W. Jeong, J. Ihm, and G.-D. Lee, Phys. Rev. B 78, Juelich, NIC Series, Vol. 10 (2002). 165403 (2008). 35 M. P. Allen and D. J. Tildesley, Computer Simulations of 16 A. Mesaros, D. Sadri, and J. Zaanen, Phys. Rev. B 79, Liquids, Oxford Science Publications (1987). 155111 (2009). 36 G. J. Martyna, D. J. Tobias and M. L. Klein, J. Chem 17 E. Ertekin, D. C. Chrzan and M. S. Daw, Phys. Rev. B Phys. 101, 4177 (1994). 79, 155421 (2009). 37 P. C. Hohenberg and B. I. Halperin, Rev. Mod. Phys. 49, 18 O. V. Yazyev and S. G. Louie, Phys. Rev. B 81, 195420 436 (1977). (2010). 38 J. P. Hirth and J. Lothe, Theory of Dislocations (Kriger, 19 S. Toxvaerd, Phys. Rev. A 24, 2735 (1981). 1992). 20 M. P. Allen, D. Frenkel, W. Gignac and J. P. McTague, J. 39 J. R. Ray, J. Appl. Phys. 53, 6441 (1982). Chem. Phys. 78, 4206 (1983). 40 J. R. Ray and A. Rahman, J. Chem. Phys., 80, 4423 21 K. J. Strandburg, J. A. Zollweg and G. V. Chester, Phys. (1984). Rev. B 30, 2755 (1984).