cond-mat0409513/pap.tex
1: \documentstyle[prb,aps,eqsecnum,manuscript]{revtex}
2: %\documentclass[prb,aps,manuscript,showpacs]{revtex}
3: %\usepackage{epsfig}
4: %\usepackage{latexsym}
5: %\usepackage{amssymb}     
6: %\usepackage{graphicx}
7: 
8: \begin{document}
9: 
10: \title{Temperature effects on dislocation core energies \\ in silicon and 
11: germanium}
12:  
13: \author{Caetano R. Miranda}
14: 
15: \address{Instituto de F\'{\i}sica ``Gleb Wataghin'', 
16: Universidade Estadual de Campinas,\\
17: CP 6165, CEP 13083-970, Campinas, SP, Brazil}       
18: 
19: \author{R. W. Nunes }
20: 
21: \address{Departamento de F\'{\i}sica, Universidade Federal de Minas Gerais, \\
22: Belo Horizonte, Minas Gerais , CEP 30123-970, Brazil}       
23: 
24: \author{A. Antonelli}
25: 
26: \address{Instituto de F\'{\i}sica ``Gleb Wataghin'', 
27: Universidade Estadual de Campinas,\\
28: CP 6165, CEP 13083-970, Campinas, SP, Brazil} 
29: 
30: \date{\today}
31: \maketitle
32:    
33: \begin{abstract}
34: 
35: Temperature effects on the energetics of the 90$^\circ$ partial
36: dislocation in silicon and germanium are investigated, using
37: non-equilibrium methods to estimate free energies, coupled with Monte
38: Carlo simulations. Atomic interactions are described by Tersoff and
39: EDIP interatomic potentials. Our results indicate that the vibrational
40: entropy has the effect of increasing the difference in free energy
41: between the two possible reconstructions of the 90$^\circ$ partial,
42: namely, the single-period and the double-period geometries. This effect
43: further increases the energetic stability of the double-period
44: reconstruction at high temperatures. The results also indicate that
45: anharmonic effects may play an important role in determining the
46: structural properties of these defects in the high-temperature regime.
47: 
48: \end{abstract}
49: 
50: %\maketitle
51: 
52: \pacs{61.72.Lk, 61.72.Bb, 62.20.Fe}
53: 
54: 
55: 
56: %\begin{multicols}{2} 
57: 
58: %
59: \section{Introduction}              
60: %
61: 
62: The study of the atomic structure of dislocation cores in
63: semiconductors is not free from outstanding
64: issues.~\cite{hirsch,duesbery,alexan,bulatov} At low deformations,
65: dislocations in these systems are known to occur along $<$110$>$
66: directions on $\{111\}$ slip planes of the diamond (or zinc-blend)
67: lattice. Due to the presence of a two-atom basis, this lattice
68: supports two possible $\{111\}$ dislocation slip planes, the so-called
69: glide and shuffle sets.  In the glide set, the slip plane lies between
70: atomic layers separated by a third of the nearest-neighbor bond
71: length, and geometrically it would appear that dislocation glide
72: should require breaking three covalent bonds.  In the case of the
73: shuffle set, the slip plane lies between atomic layers separated by a
74: full bond length, and glide appears to involve breaking only a single
75: covalent bond. This would suggest that lattice friction should be
76: smaller in the shuffle set, and hence that dislocations would glide in
77: that configuration. However, the prevalent opinion is in favor of the
78: glide set, based primarily on the experimental evidence showing that
79: dislocations in Si are dissociated into Shockley partials bounding
80: stacking-fault ribbons, and on the fact that stacking faults are not
81: stable in the shuffle plane.~\cite{hirsch,duesbery,alexan,bulatov}
82: Nevertheless, this conclusion has been disputed by Louchet and
83: Thibault-Desseaux,~\cite{louchet} who argued that partials in the
84: shuffle position could occur in conjunction with glide-plane stacking
85: faults.  The mechanism controlling dislocation glide is also a matter
86: of debate. Many theoretical and experimental works have either
87: supported or assumed the Hirth and Lothe mechanism,~\cite{hl} based on kink
88: nucleation and migration, but an impurity-pinning
89: mechanism proposed by Celli et al.~\cite{celli} has been shown to be
90: consistent with recent experimental results.~\cite{kolar} This issue
91: is far from being resolved.
92: 
93: Another unresolved issue has recently arisen, regarding the core
94: structure of the 90$^\circ$ partial dislocation. In $\{111\}$ planes,
95: dislocations lining along $<$110$>$ directions have their Burgers
96: vectors either parallel (the screw orientation) or making a 60$^\circ$
97: angle with the dislocation line. Both dissociate into Shockley
98: partials, the former into two 30$^\circ$ partials, and the latter into
99: a 30$^\circ$ partial and a 90$^\circ$ partial.  Both partials are
100: believed to undergo core reconstruction, as indicated by the low
101: density of unpaired spins found in EPR
102: experiments.~\cite{hirsch,duesbery,alexan,bulatov}
103: 
104: In the glide set, the nature of the reconstruction of the 90$^\circ$
105: partial appears straightforward. The symmetrically-reconstructed
106: geometry shown in Fig.~\ref{cores}(a) displays mirror symmetries along the
107: $[110]$ direction. A symmetry-breaking reconstruction shown in
108: Fig.~\ref{cores}(b) lowers the core energy by $\sim$200 meV/\AA, according to
109: {\it ab initio} and tight-binding (TB)
110: calculations.~\cite{Bigger,nbv,hansen} This structure retains the
111: periodicity of the lattice along the dislocation direction, thus being
112: called the single-period (SP) structure. Until recently, it had been a
113: matter of consensus that the SP core was the ``ground-state''
114: structure of the 90$^\circ$ partial, since the earliest independent
115: works of Hirsch and Jones.~\cite{hirsch2,jones} However, the structure
116: shown in Fig.~\ref{cores}(c) has been recently proposed by Bennetto and
117: collaborators,~\cite{Nunes_2} and shown to have lower internal energy
118: than the SP core, by means of total-energy tight-binding (TETB), and
119: {\it ab initio} local-density-approximation (LDA) calculations. In
120: this new geometry, along with the mirror symmetry, the translational
121: symmetry of the lattice by a single period is also broken, and the
122: periodicity along the core is doubled. For that reason, this
123: reconstruction has been called the double-period (DP) structure.
124: 
125: Both the SP and DP geometries are consistent with all available
126: experimental information about the 90$^\circ$ partial. EPR
127: measurements in Si indicate a rather small density of dangling bonds
128: in the core of the dislocation.~\cite{hirsch,duesbery,alexan} The two
129: structures are fully reconstructed, meaning that neither would give
130: rise to deep-gap states which would show an EPR signal. Moreover, both
131: cores consist entirely of fivefold, sixfold, and sevenfold rings, both
132: being consistent with images produced by transmission electron
133: microscopy, at the current level of resolution of this
134: technique.~\cite{kolar} So far, the experiments appear unable to
135: decide clearly on the issue. Recent experimental work by
136: Batson~\cite{batson} indicates a DP-derived structure (called the
137: ``Extended DP structure'' by this author) to be more consistent with
138: STEM and EELS experiments.
139: 
140: At the theoretical level, several calculations have addressed the
141: energetics of the SP and DP cores in diamond (C), silicon (Si), and
142: germanium (Ge), using different
143: methods.~\cite{Nunes_2,Vall,nv,Blase,Nunes_1,Nunes_3,Lehto} Virtually
144: all of the more accurate {\it ab initio} and TETB
145: studies~\cite{Nunes_2,nv,Blase,Nunes_1,Nunes_3} agree that the DP core
146: is more stable at 0 K, with the exception of the {\it ab initio}
147: cluster calculations for Si in Ref.~\onlinecite{Lehto}. For example,
148: LDA and TETB calculations in Refs.~\onlinecite{Nunes_2} and
149: \onlinecite{Nunes_3} indicate that the internal energy of the DP
150: structure is lower than that of the SP core in all three materials (C,
151: Si, and Ge). A possible influence of supercell boundary conditions on
152: these results was raised in Ref.~\onlinecite{Lehto} (using
153: Keating-potential calculations) and later refuted by supercell-size
154: converged TETB calculations in Ref.~\onlinecite{nv}. The issue was
155: further revisited in Ref.~\onlinecite{Blase}, with {\it ab initio}
156: calculations showing that the DP core remains more stable in C even
157: under relatively severe stress regimes.  Ideally, supercell-size
158: converged calculations like the TETB ones in Ref.~\onlinecite{nv}
159: would properly address the issue for isolated dislocations in the
160: bulk. However, in that study very large supercells (containing up to
161: 1920 atoms) were employed, which makes such calculations prohibitive
162: for the more accurate {\it ab initio} techniques.
163: 
164: Closely related to the present study is the work of Valladares and
165: co-workers,~\cite{Vall} in which temperature effects on the energetics
166: of the SP and DP cores in Si were investigated using a Tersoff
167: potential,~\cite{Ters} within a harmonic approximation treatment of
168: vibrational entropic effects in the dislocation free energy, at finite
169: temperatures. They find that the free-energy difference between the
170: two cores, at a temperature of 800 K, becomes smaller than the
171: zero-temperature internal energy split, suggesting the possibility of
172: the two cores playing a role in the plasticity of Si, in the
173: temperature range of most plastic deformation experiments ($\sim$700-1000
174: K). However, the calculations in Ref.~\onlinecite{Vall} did not
175: include anharmonic effects, which are likely to be relevant in that
176: range of temperatures.
177: 
178: In this paper, we present the results of a study of anharmonic
179: vibrational effects in the free-energy difference between the SP and
180: DP cores in Si and Ge. As discussed in the following, we arrive at
181: different conclusions with respect to the work of Valladares {\it et
182: al.}, with our results indicating that the effect of increasing
183: temperature is to enhance the thermodynamical stability of the DP core
184: with respect to the SP geometry.  In our calculations, anharmonic
185: effects are fully taken into account by using a methodology that
186: allows to estimate the free energy through non-equilibrium
187: simulations.  The details of this methodology are discussed in
188: Sec. II. In Sec. III our results are presented and discussed. We end
189: with a summary and conclusions in Sec. IV.
190: 
191: 
192: %
193: \section{Methodology}              
194: 
195: \subsection{Interatomic Potentials}
196: \label{secA}
197: 
198: There are several empirical potentials available in the literature to
199: model the interactions between Si atoms.~\cite{Potential1} In the
200: present work, our aim is to account for vibrational entropic effects
201: on the energetics of the SP and DP reconstructions of the 90$^{\circ}$
202: partial dislocation in Si. One key ingredient for this purpose is that
203: both the SP and DP structures be a local minimum of the potential. It
204: is essential then that the potential be able to describe the
205: reconstruction-driven symmetry breaking that leads to the SP
206: geometry.~\cite{Bigger,nbv,hansen} This is an important aspect, since
207: both core structures consist of fully reconstructed bonds, and a
208: realistic description of vibrational properties requires a proper
209: treatment of bonding for the minima in question. For example,
210: potentials such as the Stillinger-Weber(SW) and the Kaxiras-Pandey(KP)
211: fail in that regard.~\cite{Potential2} Both the Tersoff~\cite{Ters}
212: and the Environment Dependent Interatomic Potential
213: (EDIP)~\cite{edip1,edip2} models describe correctly the energetics of
214: the SP reconstruction with respect to a symmetric
215: (so-called quasi-fivefold) structure.~\cite{Bigger,edip2}
216: 
217: The focus of our paper is on the effects of the dislocation
218: vibrational modes on the free energies of the two core models at
219: finite temperatures.  In a previous work we have used these two
220: potentials to estimate the free energy of Si in several different
221: phases (amorphous, liquid, and crystalline, as well as clathrate
222: structures), obtaining results which are in good agreement with those
223: of experimental and other theoretical works.~\cite{Caetano} For a more
224: specific analysis of the description of anharmonic effects by the EDIP
225: and Tersoff potentials, we show in Fig.~\ref{bulk} the results we obtained for
226: the free energy as a function of temperature, in a 216-atom bulk
227: supercell.  The figure includes also the Tersoff-potential harmonic
228: approximation values from Ref.~\onlinecite{harm} and the experimental
229: results from Ref.~\onlinecite{fexp}. As expected, the experimental
230: numbers start deviating appreciably from the harmonic approximation
231: curve around the Debye temperature of Si ($\sim$640 K), where
232: anharmonic effects begin manifesting themselves. From the figure, we
233: see that both the Tersoff and the EDIP potentials can account for
234: these anharmonic effects in the free-energy curve at a quantitative
235: level.
236: 
237: Moreover, it is clear that the EDIP results are in better agreement
238: with the experimental ones. Of particular relevance to the present
239: work is the scale of the difference in free energy between
240: harmonic-approximation and experimental results, which is
241: $\sim$10~meV/atom at 640 K. As a first approximation, we may consider
242: ten atoms per period as constituting the core in the SP and DP
243: structures (those taking part in the fivefold and sevenfold rings,
244: seen on [110] projections, surrounding the geometric center of the
245: core), which gives an initial estimate of $\sim$30~meV/\AA\ for the
246: scale of the anharmonic effects, from the bulk results. From this, one
247: might expect that anharmonic effects should be important in the case
248: of the 90$^\circ$ partial, since the difference in enthalpy at 0 K
249: between the SP and DP cores is $\sim$60~meV/\AA.
250: 
251: On the basis of these results, we argue that both potentials give a
252: realistic description of vibrational effects, in particular of the
253: anharmonic contributions which become important in the range of
254: temperatures of our study. For the Ge case, only a Tersoff model was
255: employed, since there are no available parameters for Ge in the EDIP
256: form. The functional form and the parameters we use for the two models
257: are given in Refs.~\onlinecite{Ters} (Tersoff) and \onlinecite{edip2}
258: (EDIP).
259: 
260: \subsection{Simulations and Free Energy Calculations}
261: \label{secB}
262: 
263: In our statistical computation of dislocation free energies, we use a
264: Metropolis-algorithm Monte-Carlo method. Two types of simulations were
265: performed, one employed a canonical-ensemble (constant-NVT) treatment,
266: and the other simulated an isobaric-isothermal ensemble
267: (constant-NPT).~\cite{Frenkel} A simulation cell with 192 atoms was
268: used, containing a 90$^\circ$ partial-dislocation dipole which was
269: introduced by displacing the atoms according to the
270: continuous-elasticity solution for the displacement field of a
271: dislocation dipole.~\cite{hl} The introduction of a dipole enables the
272: use of periodic boundary conditions to avoid surface effects.  It
273: should be pointed out that the free energy calculations presented in
274: this work are far more demanding in terms of computer time than
275: harmonic-approximation calculations, which only require one
276: diagonalization of the dynamical matrix, in order to determine the
277: phonon frequencies. Therefore, the complexity of our calculations
278: imposes a limitation on the size of the cell used in the
279: simulations. The cell size chosen reflects the best compromise between
280: a good description of the essential physics and computational
281: feasibility.
282: 
283: The free energy calculations were performed using a non-equilibrium
284: method, namely the Reversible-Scaling method (RS).~\cite{RS_1} This
285: method consists of simulating a quasi-static process where the
286: potential-energy function of the system is dynamically scaled. It
287: allows for an accurate and efficient evaluation of the free energy
288: over a wide range of temperatures, using a single Monte Carlo
289: simulation, starting from a reference temperature for which the free
290: energy of the system is known. In order to obtain the free energy at
291: this reference temperature, we used the Adiabatic Switching method
292: (AS).~\cite{AS_1} Using the AS approach, the absolute free energy can
293: be computed by determining the work done to switch the potential
294: energy of the system from the configuration of interest to that of a
295: reference system for which the free energy is known analytically. The
296: systematic error in the free energy, caused by dissipation, can be
297: estimated by reversing the switching procedure and computing the
298: overall work performed in this round-trip procedure. The statistical
299: error can be estimated by performing several trajectories in the
300: configuration space of the system.
301: 
302: In this work, the reference system was an Einstein crystal, i.e., a
303: collection of independent harmonic oscillators. The angular frequency
304: of the oscillators ($\omega$) was taken to be approximately the
305: vibration modes of Si ($\omega = 28$~Trad/s) and Ge ($\omega = 15$~
306: Trad/s), which correspond to the acoustic phonon modes from the
307: edge of the Brillouin zone that give rise to a pronounced peak in the 
308: phonon density of states of these materials.~\cite{Kane,Nilsson}
309: It should be pointed out that the final value for the Gibbs free energy
310: is not sensitive to the choice of the oscillator frequency,
311: provided that the chosen frequency is not very far from the
312: relevant frequencies of the physical system.
313: Monte Carlo constant-NVT simulations were used to generate
314: switching trajectories, with a typical "switching time" of 3.0
315: $\times$ 10$^{5}$ MC/steps. The volume in these simulations was chosen
316: to be that obtained from a constant-NPT simulation for the given
317: reference temperature at zero pressure. In general, 25 switching
318: trajectories were used to estimate the statistical error. The
319: estimated error was $\pm$ 1.0 $\times$ 10$^{-3}$ eV/atom. RS
320: calculations were carried out at zero pressure, covering a range from
321: T=100 K to T=2000 K, using a scaling time of 5.0 $\times$ 10$^{5}$
322: MC/steps. Five different switching trajectories were sampled, in order
323: to estimate the statistical errors. In this way, we were able to
324: estimate the absolute Gibbs free energy (G$_{abs}$) at zero pressure
325: of the SP and the DP reconstructions in the interval from 300K to 1200
326: K using EDIP and Tersoff models for Si, and for Ge in the same range
327: of temperatures using a Tersoff potential. The absolute vibrational
328: entropy was calculated by taking the numerical derivative of the
329: absolute Gibbs free energy with respect to temperature: $S_{abs}=
330: -(\frac{\partial G_{abs}}{\partial T})$. 
331: 
332: 
333: \section{Results and Discussion}              
334: \label{enthal}
335: \subsection{Enthalpies at zero Kelvin}
336: 
337: Imposing periodic boundary conditions in a dislocation calculation
338: using supercells requires the net Burgers vector in the cell to
339: vanish. This is done by using cells which contain two dislocations
340: with opposite Burgers vectors. Hence, besides the core enthalpies that
341: interest us, supercell calculations also include elastic interactions
342: between the two dislocations, and between the dislocations and their
343: periodic images. In order to extract meaningful core energy values,
344: these elastic interactions must be properly accounted for. Previous
345: calculations have shown that dislocation cores in covalent
346: semiconductors are very narrow, with the elastic fields reaching the
347: linear behavior predicted by the continuum linear elasticity solution
348: already at a distance of about two lattice periods from the center of
349: the core.~\cite{hansen,Blase} This indicates that, to the extent that
350: one is interested in differences in energies between different core
351: models, long-range elastic effects should mostly cancel out for
352: sufficiently large cells. This is because these long-range fields are
353: defined by the dislocation slip system which is obviously the same in
354: our case, since we are comparing different core structures of the same
355: dislocation. We must be careful, however, in providing numerical
356: evidence that the supercells we use in our calculations are
357: sufficiently large, in the above sense.
358: 
359: In the particular case of the SP geometry, it is known that a
360: relatively strong dipolar elastic interaction is present, which
361: depends on the relative senses of reconstruction of the two
362: dislocations in the cell, with respect to the broken mirror
363: symmetry. Depending on the size of the cell, this interaction may be
364: of the same order of the core-energy difference between the SP and DP
365: geometries. (A similar effect is present in the DP core, but this is a
366: weaker quadrupolar interaction which can be neglected.~\cite{Nunes_2}) As
367: discussed in the literature,
368: ~\cite{Nunes_2,Vall,nv,Blase,Nunes_1,Nunes_3} this dipolar interaction
369: can be properly handled by considering the SP-core energy as the
370: average energy between two supercell calculations, one in which the
371: two SP dislocations are reconstructed in the same sense, and the other
372: where they have opposite reconstruction
373: senses.~\cite{nv,Nunes_1,Nunes_3} In order to further investigate the
374: accuracy of this procedure and also to address the numerical
375: convergence level of our calculations for the cell sizes we use, we
376: show in Table I the 0 K enthalpies computed with the Tersoff and EDIP
377: potentials for several different cell sizes. For comparison, we also
378: include the TETB results from Ref.~\onlinecite{nv}, to which we add
379: numbers for a 240-atom supercell. In the notation introduced in
380: Ref.~\onlinecite{Lehto}, the relevant supercell parameters are $D$,
381: the height of the cell in the $<111>$ direction (perpendicular to the
382: slip plane), and $L$ the width of the cell in the $<11\bar{2}>$
383: direction (perpendicular to the dislocation line) on the slip
384: plane. We consider here only supercells with ``dipolar'' boundary
385: conditions,~\cite{bc} which were shown by Lehto and \"Oberg~\cite{Lehto} to
386: minimize the strains associated with the stacking of the infinite
387: array of dislocations. We will comment below on the comparison of our
388: results with those from Ref.~\onlinecite{Vall}, which were obtained
389: with ``quadrupolar'' boundary conditions.~\cite{bc}
390: 
391: Table I shows our results for $E_{SP-DP} = \overline{E}_{SP} -
392: E_{DP}$, the energy difference between the average energy of the two
393: SP cells ($\overline{E}_{SP}$) and the energy of the DP cell, and for
394: $\Delta E_{SP} = E_{SP}^{same} - E_{SP}^{opposite}$, the energy
395: splitting between the two SP cells (as described in the previous
396: paragraph), for four different cell sizes. Note that $E_{SP-DP}$
397: converges for relatively small cell sizes, and faster than $\Delta
398: E_{SP}$, due to the cancellation of elastic effects that happens when
399: we take the average of the two SP calculations. For the free-energy
400: calculations we discuss in the next section, we used the smaller cell,
401: with $L = 13.3$~\AA\ and $D = 18.8$~\AA, for a total of 192 atoms. In
402: the 240 atom cell, $D$ is the same as in the 192-atom cell, and $L$ is
403: only 25\% larger.  We can see from Table I that, already for this
404: 240-atom cell, the TETB result for $E_{SP-DP}$ is converged to within
405: the numerical accuracy of the method. This is also what can be
406: observed for the EDIP potential. The EDIP result for $E_{SP-DP}$ in
407: the 192-atom cell is within 13\% of the EDIP converged value, and for
408: the 240-atom cell $E_{SP-DP}$ is converged to within 3\%.  The Tersoff
409: numbers converge more slowly, with the 192-atom cell numbers being
410: more than twice as large as the converged value.  In the case of Ge,
411: we only calculated $E_{SP-DP}$ using the 192-atom cell. We obtained
412: $E_{SP-DP}$~= -8~meV/\AA\ in disagreement with the LDA and
413: Keating-potential results from Ref.~\onlinecite{Nunes_1}. However, as
414: discussed in Sec.~\ref{free-ener}, the behavior of the free energy as
415: a function of temperature for Ge is the same as that for Si, i.e.,
416: $E_{SP-DP}$ increases as the temperature is raised.  If we were to
417: extrapolate to the free-energy calculations these convergence trends
418: observed for the 0 K enthalpy results in Table I, we would expect our
419: EDIP free-energy values to be converged to within $\sim$10\%, and the
420: Tersoff values to be about twice as large as what we would obtain with
421: larger cells.
422: 
423: The TETB calculations in Table I were done at the fixed cell volume
424: (corresponding to the experimental lattice constant). As a test, we
425: also performed calculations for volumes 1\% larger and smaller,
426: obtaining the same convergence trends, i.e., energies were essentially
427: converged for the 240-atom cell. More importantly, the convergence
428: calculations with the Tersoff and EDIP potentials in Table I were done
429: at constant pressure, resulting in very small changes from the initial
430: volumes.
431: 
432: Before proceeding to the discussion of the free-energy results, let us
433: put the results in Table I in perspective. Previous studies have shown
434: that the Tersoff potential does give the right ordering of 0 K
435: enthalpies, despite underestimating the difference in enthalpies
436: between the two cores. Our converged Tersoff values of 6~meV/\AA\ are
437: in very good agreement with the converged results obtained in
438: Ref.~\onlinecite{chrzan} using supercell and cylindrical cluster
439: calculations. However, they are smaller by one order of magnitude,
440: when compared with the converged TETB values. The EDIP potential has
441: the shortcoming of predicting the wrong ordering of 0 K enthalpies for
442: the SP and DP cores. When comparing our results with those obtained by
443: Valladares {\it et al.}, we must observe that they used cells with $D
444: = 9.4$~\AA, which is half the height of our cells, and quadrupole
445: boundary conditions. As a check, we did run Tersoff calculations with
446: the same cell parameters and obtained the same results as in their
447: work. The value for $E_{SP-DP}$ in this case is 32 meV/\AA, which is
448: more than five times as large as the converged Tersoff value in Table
449: I.  This shows that, at least for the 0 K enthalpy, their results are
450: not converged at the same level as our calculations.
451: 
452: \subsection{Free energies}
453: \label{free-ener}
454: 
455: In order to test whether the empirical potentials support the SP and
456: DP reconstructions at high temperatures, we performed constant-NPT
457: simulations of all structures in a broad range of temperatures, from 0
458: K up to 2000 K (1500 K) for Si (Ge). The structures were further
459: characterized for several temperatures (at every 500 K between 0 K and
460: the highest temperature) by evaluating structural properties such as
461: the pair-correlation function, the bond-angle distribution, and the
462: atomic-coordination number. Only the natural thermal broadening of
463: these functions was observed, with no structural changes and all atoms
464: remaining fourfold coordinated. From these results, we conclude that
465: for both the Tersoff and the EDIP potentials, the SP and DP
466: reconstructions remain stable and do not undergo restructuring over
467: the entire range of temperatures we tested for each of the two
468: semiconductors. 
469: 
470: The metastability of the SP core, even at temperatures as high as 2000
471: K, is not surprising. Unlike the spontaneous transformation of the
472: symmetric quasi-fivefold core into the SP core, which happens at 0 K
473: due to the absence of an energy barrier,~\cite{Bigger,nbv,hansen} the
474: SP $\rightarrow$ DP transformation involves breaking the strong
475: reconstructed covalent bonds in the SP core. The same bond breaking
476: mechanism regulates the kink migration barriers.~\cite{nbv} We have
477: computed an energy barrier of $\sim$1.5 eV for the process involving
478: the conversion of a segment of two SP periods into a segment of one DP
479: double period.
480:     
481: Let us now turn to the focus of this work, the Gibbs free-energy
482: difference, $\Delta G$, between the SP and the DP cores as a function
483: of temperature, obtained by the Reversible Scaling Method within the
484: Monte Carlo method (RS-MC). In the previous section we discussed the
485: numerical convergence of our calculations with respect to the two
486: dimensions of the supercell which are perpendicular to the dislocation
487: line. When we introduce thermal effects, periodicity along the line is
488: disrupted by atomic vibrations, and one must be careful about the
489: sampling of the phonon modes along the dislocation direction. In our
490: free-energy calculations, we used 192-atom supercells with twice the
491: lattice period along this direction. To test the phonon sampling in
492: this cell, we also calculated free energies for a supercell with four
493: times the lattice period along the line, using the Tersoff
494: potential. In Fig.~\ref{tamanho} we show the results for the
495: free-energy difference $\Delta G$ between the SP and DP cores, for the
496: two types of SP cells and also for the average between the two
497: cells. Note that, while the results for $\Delta G$ for the individual
498: SP calculations differ quite appreciably for a given cell size and
499: vary substantially when we go from the smaller (192 atoms and twice
500: the lattice period along the dislocation line) to the larger cell (384
501: atoms and four times the period along the line), the results for
502: $\Delta G$ between the average of the SP cells and the DP cell are
503: very similar for the two cell sizes. This shows that our simulations
504: with 192 atoms capture most of the relevant differences between the
505: vibrational modes of the two cores and also lends further confirmation
506: of the cancellation of elastic interaction effects that occurs when
507: the average between the two SP cells is considered, as discussed in
508: the previous section.
509: 
510: In Fig.~\ref{G-si} we show the results for $\Delta G_{SP-DP}$ for the
511: Tersoff and EDIP models in a 192-atom supercell, showing now only the
512: average of the SP calculations. Fig.~\ref{G-ge} shows the results for
513: Ge using the Tersoff potential. In order to analyze the entropic
514: effects, Figs.~\ref{G-si} and \ref{G-ge} also display the entropic
515: contribution ($-T \Delta S$) to $\Delta G$. The Tersoff results for Si
516: in Fig.~\ref{G-si} show $\Delta G$ increasing with temperature from
517: 19~meV/\AA\ at 300 K to 37~meV/\AA\ at 1200 K.  The entropic term also
518: increases with temperature, being the dominant contribution for
519: $\Delta G$ in the high temperature regime.  For Si using the EDIP
520: model, we observe the same behavior, with $\Delta G$ increasing with
521: temperature from -32~meV/\AA\ at 300 K to -23~meV/\AA\ at 1200 K,
522: notwithstanding the fact that for this potential the SP core remains
523: more stable over the entire temperature range. This is mostly due to
524: the fact that the EDIP potential does not describe properly the
525: energetics of the two reconstructions at 0 K, as discussed above. But
526: even for this potential, the behavior of the entropic term is
527: determinant of the temperature trend observed for $\Delta G$, which is
528: qualitatively the same trend observed with the Tersoff model. As
529: indicated by the convergence trends we discussed in the previous
530: section for the 0 K enthalpies, a quantitative comparison between the
531: changes in $\Delta G$ produced by the Tersoff and EDIP potentials
532: would suffer form the fact that, for the 192-atom cell, the Tersoff
533: results could be overestimated by as much as a factor of two. Note
534: that, for silicon, the overall change in $\Delta G$ over the 300 K -
535: 1200 K interval is 18~meV/\AA\ for Tersoff and 9~meV/\AA\ for
536: EDIP. While speculative, it seems reasonable to expect that the overall
537: change in $\Delta G$ for the two potentials, over the temperature
538: range of our calculations, would be in good agreement if we used cells
539: with larger values of $D$ and $L$.
540: 
541: Regarding the contribution of the entropic term depicted in
542: Figs.~\ref{G-si} and \ref{G-ge}, two points have to be
543: emphasized. First of all, the entropy is calculated by taking the
544: numerical derivative of the free energy, and the numerical data for
545: this quantity presents some statistical fluctuations. Most certainly,
546: these fluctuations are bound to be enhanced when one takes numerical
547: derivatives. This explains some of the oscillations observed in the
548: entropic term.  Thus the qualitative behavior of the total free-energy
549: is more representative than the behavior of the isolated entropic term
550: shown in the figures.  Another point has to do with the behavior of
551: the entropic term as a function of temperature, in particular for the
552: computations using the Tersoff potential. The free energy difference
553: $\Delta G$ is quite small, when compared with the total free energy of
554: each system, and the same is true for the entropic contribution to
555: this difference.  These energy differences will be affected by
556: anharmonic effects, and the discrepancy between the entropic term
557: computed using the two models (EDIP and Tersoff) above 600 K is
558: related to the way anharmonic effects are accounted for in each
559: model. It is clear from Fig.~\ref{bulk} that the results obtained
560: using the Tersoff potential tend to deviate more from the experimental
561: data than those obtained using the EDIP, and that the difference
562: between the two potentials increase with increasing temperature. The
563: more accentuated increase in the entropic term computed using the
564: Tersoff potential has to do with its tendency to overestimate the
565: anharmonic effects, and represents a small amount of energy that is
566: not related to any changes in the structure of the defect (which would
567: imply in much bigger changes in the free energy).
568: 
569: In the case of Ge, $\Delta G$ increases with temperature faster than
570: in Si (compared to either of the Si potentials used), from 2~meV/\AA\
571: at 300 K to 33~meV/\AA\ at 1200 K. While the enthalpy difference at 0
572: K marginally favors the SP reconstruction (at variance with the LDA
573: result in Ref.~\onlinecite{Nunes_3}), the behavior of the free energy
574: suggests that at room temperature the DP reconstruction would be more
575: stable than the SP structure. It is important to emphasize the
576: similarity between our results for Si and Ge, i.e., the free energy
577: difference $\Delta G$ increases with temperature in both cases (the
578: increase rate is larger in Ge than in Si), meaning that the entropic
579: term leads to an enhancement of the thermodynamical stability of the
580: DP structure with respect to the SP, as the temperature increases.
581: 
582: At this point it is interesting to compare the RS-MC results for Si
583: with the harmonic-approximation calculations of Valladares {\it et
584: al.}.~\cite{Vall} In Fig.~\ref{vallad}, we show our RS-MC calculations
585: using the Tersoff (full line) and EDIP (dashed line) models, and the
586: Tersoff harmonic-approximation results from Ref.~\onlinecite{Vall}
587: (dotted line). For a better comparison of the temperature trends in
588: each calculation, our values for $\Delta G$ for both potentials have
589: been shifted such that they extrapolate to the same zero-temperature
590: $\Delta G$ value obtained by Valladares {\it et al.}. The same
591: qualitative behavior is observed in Fig.~\ref{vallad} for the RS-MC
592: calculations using both the Tersoff and the EDIP models. In both
593: cases, $\Delta G$ {\it increases} with temperature, while the opposite
594: is observed for the harmonic-approximation calculations, which show
595: $\Delta G$ {\it decreasing} with increasing temperature.  This points
596: to an important role of anharmonic effects in describing thermal
597: effects on dislocation cores in semiconductors. 
598: 
599: It should also be pointed out that at low temperatures our
600: results cannot be directly compared to those in
601: Ref.~\onlinecite{Vall}, because their calculations take into account
602: quantum effects which are absent in our approach. While quantum
603: effects may be relevant at low temperatures, we are mostly concerned
604: with the anharmonic effects, which become more relevant in the
605: high-temperature regime. Valladares {\it et al.} point out that the
606: order of magnitude of the difference in free energies between the SP
607: and the DP reconstructions is smaller than the thermal energy
608: ($k_{B}T$) over the entire range of temperature in their study. This
609: would suggest that the two structures would be nearly equally
610: stable. Our results, however, indicate that at high temperatures the
611: free energy difference between the SP and the DP cores is of the same
612: order of the thermal energy. Therefore, considering that the thermal
613: fluctuations are smaller than the thermal energy itself, our results
614: suggest that the DP reconstruction should be dominant in the high
615: temperature regime. In other words, while thermal fluctuations may
616: account for the creation of local defects such as kinks, it is
617: unlikely that these fluctuations may cause substantial portions of the
618: dislocation to switch from the DP to the SP reconstruction (barring a
619: possible dependence of the core energies on the stress acting on the
620: dislocation,~\cite{Blase} an issue that has not yet been investigated
621: in Si and Ge).
622: 
623: %
624: \section{Conclusions}              
625: % 
626: In this work, we studied the differences in free energies and
627: vibrational entropies between the SP and DP core reconstructions of
628: the 90$^{\circ}$ partial dislocation as a function of temperature in
629: silicon and germanium, using a Reversible Scaling Method and Monte
630: Carlo simulations, with Tersoff and EDIP models for the
631: energetics. This methodology is a fully-classical simulation which
632: includes all anharmonic vibrational effects. Our results indicate that
633: the difference in the free energies (and in the vibrational-entropy
634: contribution to this quantity) between the two core structures
635: increases with temperature in both materials. In the case of Si, this
636: behavior occurs for both potentials used in the calculations. This is
637: in contrast with the free-energy calculations by Valladares and
638: co-workers~\cite{Vall} which incorporates vibrational-entropy effects
639: only at the harmonic-approximation level. Our calculations indicate
640: that the DP reconstruction, which is lower in energy at 0 K, becomes
641: even more stable with respect to the SP structure in the high
642: temperature regime. Moreover, our results also suggest that the
643: anharmonic effects may play an important role in the description of
644: the thermal behavior of extended defects in semiconductors.
645:   
646: \begin{acknowledgments} 
647: 
648: C. R. Miranda and A. Antonelli acknowledge the support from the 
649: Brazilian funding agencies: FAPESP, CNPq, and FAEP.
650: R. W. Nunes acknowledges support from the Brazilian agencies: CNPq and
651: FAPEMIG.
652: 
653: \end{acknowledgments}  
654: 
655: %\begin{thebibliography}{99}
656: \begin{references}
657: 
658: \bibitem{hirsch} P.~B.~Hirsch, Mater. Sci. Tech. {\bf 1}, 666 (1985).
659: 
660: \bibitem{duesbery} M.~S.~Duesbery and G.~Y.~Richardson,
661: Crit. Rev. Solid State Mater. Sci. {\bf 17}, 1 (1991).
662: 
663: \bibitem{alexan} H.~Alexander and H.~Teichler, in {\it Materials
664: Science and Technology}, edited by R.~W.~Cahn, P.~Hassen, and
665: E.~J.~Kramer (VCH Weiheim, Cambridge, 1993), Vol.~4, p.~249.
666: 
667: \bibitem{bulatov} V.~V.~Bulatov, J.~F.~Justo, W.~Cai, S.~Yip,
668: A.~S.~Argon, T.~Lenosky, M.~de~Koning, and T.~Diaz de la Rubia,
669: Philos. Mag. A {\bf 81}, 1257 (2001).
670: 
671: \bibitem{louchet} F.~Louchet and J.~Thibault-Desseaux,
672: Rev. Phys. Appl. {\bf 22}, 207 (1987).
673: 
674: \bibitem{hl} J. P. Hirth and J. Lothe, {\it Theory of Dislocations}
675: (Wiley, New York, 1982).
676: 
677: \bibitem{celli} V.~Celli, M.~Kabler, T.~Ninomiya, and R.~Thomson,
678: Phys. Rev. {\bf 131}, 58 (1963).
679:  
680: \bibitem{kolar} H.~R.~Kolar, J.~C.~H.~Spence, and H.~Alexander,
681: Phys. Rev. Lett. {\bf 77}, 4031 (1996).
682: 
683: \bibitem{Bigger} J. R. K. Bigger, D. A. McInnes, A. P. Sutton, M. C. Payne, I. 
684: Stich,
685: R. D. King-Smith, D. M. Bird, and L. J. Clarke,
686: Phys. Rev. Lett {\bf 69}, 2224 (1992).
687: 
688: \bibitem{nbv} R. W. Nunes, J. Bennetto, and David Vanderbilt, 
689: Phys. Rev. Lett {\bf 77}, 1516 (1996).
690: 
691: \bibitem{hansen} L.~B.~Hansen, K.~Stokbro, B.~I.~Lundqvist,
692: K.~W.~Jacobsen, and D.~M.~Deaven, Phys. Rev. Lett {\bf 75}, 4444
693: (1995).
694: 
695: \bibitem{hirsch2} P.~B.~Hirsch, J. Phys. (Paris), Colloq. {\bf 40},
696: C6-27 (1979).
697: 
698: \bibitem{jones} R.~Jones, J. Phys. (Paris), Colloq. {\bf 40},
699: C6-33 (1979).
700: 
701: \bibitem{Nunes_2} J. Bennetto, R. W. Nunes, and David Vanderbilt, 
702: Phys. Rev. Lett {\bf 79}, 245 (1997).
703: 
704: \bibitem{batson} P.~E.~Batson, Phys. Rev. Lett {\bf 83}, 4409 (1999).
705: 
706: \bibitem{Vall} A. Valladares, A. K. Petford-Long, and A. P. Sutton,
707: Phil. Mag. Lett. {\bf 79}, 9 (1999).
708: 
709: \bibitem{nv} R. W. Nunes and David Vanderbilt, Phys. Rev. Lett {\bf 85}, 3540 
710: (2000).
711: 
712: \bibitem{Blase} X. Blase, Karin Lin, A. Canning, S. G. Louie, and D. C. Chrzan,
713: Phys. Rev. Lett {\bf 84}, 5780 (2000).
714: 
715: \bibitem{Nunes_1} R. W. Nunes and David Vanderbilt, J. Phys.:
716: Condens. Matter {\bf 12}, 10021 (2000).
717: 
718: \bibitem{Nunes_3} R. W. Nunes, J. Bennetto, and David Vanderbilt, 
719: Phys. Rev. B {\bf 58}, 12563 (1998).
720: 
721: \bibitem{Lehto} Niklas Lehto and Sven \"{O}berg,
722: Phys. Rev. Lett {\bf 80}, 5568 (1998).
723: 
724: \bibitem{Ters} J. Tersoff, Phys. Rev. B {\bf 39}, 5566 (1989).
725: 
726: \bibitem{Potential1} H. Balamane, T. Halicioglu, and W. A. Tiller,
727: Phys. Rev. B {\bf 46}, 2250 (1992).
728: 
729: \bibitem{Potential2} M. S. Duesbery, B. Joos, and D. J. Michel,
730: Phys. Rev. B {\bf 43}, 5143 (1991).
731: 
732: \bibitem{edip1} M. Z. Bazant, E. Kaxiras, and J. F. Justo, 
733: Phys. Rev. B {\bf 56}, 8542 (1997).
734: 
735: \bibitem{edip2} J. F. Justo, M. Z. Bazant, E. Kaxiras, V. V. Bulatov, 
736: and S. Yip, Phys. Rev. B {\bf 58}, 2539 (1998).
737: 
738: \bibitem{Caetano} Caetano R. Miranda and A. Antonelli, unpublished.
739: 
740: \bibitem{harm} K. Moriguchi, S. Munetoh, A. Shintani, T. Motooka, 
741: Phys. Rev. B {\bf 64}, 195409 (2001).
742: 
743: \bibitem{fexp}  D. R. Lide and H. V. Kehiaian, {\it CRC Handbook of Thermophysical and
744: Thermochemical Data}, (CRC Press, Boca Raton, 1994).
745: 
746: \bibitem{Frenkel} D. Frenkel and B. Smit, {\it Understanding Molecular
747: Simulations}, (Academic Press, San Diego, 2001).
748: 
749: \bibitem{RS_1} Maurice de Koning, A. Antonelli, and Sidney Yip,
750: Phys. Rev. Lett {\bf 83}, 3973 (1999).
751: 
752: \bibitem{AS_1} Maurice de Koning and A. Antonelli
753: Phys. Rev. E {\bf 53}, 465(1996); Phys. Rev. B {\bf 55}, 735 (1997).
754: 
755: \bibitem{Kane} E. O. Kane, Phys. Rev. B {\bf 31}, 7865 (1985).
756: 
757: \bibitem{Nilsson} G. Nilsson and G. Nelin, Phys. Rev. B {\bf 6}, 3777 (1972).
758: 
759: \bibitem{bc} The issue of supercell boundary conditions in dislocation
760: calculations has been properly addressed in
761: Ref.~\onlinecite{Lehto}. When the choice of supercell vectors is such
762: that when the cells are stacked in the direction perpendicular to the slip
763: plane, the dislocation with Burgers vector {\bf b} ({\bf -b}) in
764: one cell lies on top of the dislocation with the same Burgers vector
765: {\bf b} ({\bf -b}), in the next cell, we have a dipolar
766: configuration. On the other hand, when the cell vectors are chosen
767: such that the dislocation with Burgers vector {\bf b} ({\bf -b}) lies
768: on top of the dislocation with the opposite Burgers vector {\bf -b}
769: ({\bf b}), in the next cell, we have a quadrupolar configuration. As
770: shown in Ref.~\onlinecite{Lehto}, the former choice minimizes the spurious
771: strains associated with the periodic array of dislocations, in the
772: case of the 90$^{\circ}$ partial.
773: 
774: \bibitem{chrzan} K.~Lin and D.~C.~Chrzan, Mater. Sci. and Eng. {\bf
775: A319-321}, 115 (2001).
776: \end{references}
777: 
778:            
779: \begin{table}
780: \centering{
781: \caption{Calculated enthalpy differences at 0 K between core
782: reconstructions of the 90$^{\circ}$ partial dislocation, for four
783: different cell sizes, in meV/\AA. Results are shown for TETB (from
784: Ref.~17), Tersoff, and EDIP. Cells are defined by their height $D$ and
785: width $L$ (in \AA), as defined in the text. The number of atoms
786: ($N_{at}$) in each cell is also included. $\Delta{E}_{SP}$ is the
787: difference in enthalpy between the cells with the same senses and
788: opposite senses of the SP reconstruction (see text).  $\Delta E_{SP -
789: DP}$ is the difference in enthalpy between the average of the SP cells
790: and the DP structure.}
791: \label{tab1}
792: \begin{tabular}{lcccccccc}  
793: & & &\multicolumn{2}{c}{Tersoff} &\multicolumn{2}{c}{EDIP} &\multicolumn{2}{c}{TBTE} \\
794: $~N_{at}$ &$L$ &$D$ &$E_{SP-DP}$ &$\Delta E_{SP}$ &$E_{SP-DP}$ &$\Delta E_{SP}$ &$E_{SP-DP}$ &$\Delta E_{SP}$ \\
795: \hline
796: ~192 &13.3 &18.8 &~13 &24 &-35 &~6 &~62 &39 \\
797: ~240 &16.6 &18.8 &~~9 &16 &-39 &~4 &~55 &26 \\
798: ~576 &26.6 &28.2 &~~7 &~6 &-39 &~2 &~55 &~9 \\
799: 1920 &53.2 &47.0 &~~6 &~2 &-40 &~0 &~55 &~4 
800: \end{tabular}
801: }
802: \end{table}         
803:            
804:  
805: \begin{figure}
806: \caption{Atomic structure of the 90$^{\circ}$ partial dislocation
807: viewed from above the $\{111\}$ slip plane. (a) symmetrically
808: reconstructed core; (b) SP reconstruction; and (c) DP
809: reconstruction.
810: \label{cores}}
811: \end{figure} 
812:           
813: \begin{figure}
814: \caption{Gibbs free-energy for the bulk of silicon. Full and dashed
815: lines show our RS-MC calculations, including anharmonic effects, for
816: the EDIP and Tersoff potentials, respectively. The dotted line shows the
817: Tersoff-potential harmonic-approximation results from Ref.~28,
818: and the squares are the experimental results from Ref.~29.
819: \label{bulk}}
820: \end{figure} 
821:           
822: \begin{figure}
823: \caption{Gibbs free-energy difference per unit length $\Delta G$,
824: between the SP and the DP geometries of the 90$^\circ$ partial
825: dislocation in silicon, with different samplings of the phonon modes
826: along the dislocation direction. For the 192-atom calculations, the
827: results for the SP cores with same reconstruction senses, with
828: opposite reconstruction senses, and the average between the two (as
829: explained in the text) are shown by the squares, circles, and
830: triangles, respectively. For the 384-atom cell, same reconstruction
831: senses, opposite reconstruction senses, and the average are
832: represented by the solid line, the dashed line, and the dotted line,
833: respectively.
834: \label{tamanho}}
835: \end{figure} 
836: 
837: \begin{figure}
838: \caption{Gibbs free-energy difference per unit length ($\Delta G$) and
839: the entropic contribution $-T\Delta S$, between the SP and the DP
840: geometries in silicon, in $\rm{meV/\AA}$, using the Tersoff
841: potential (full line for $\Delta G$ and squares for $-T\Delta S$), and
842: the EDIP potential (dashed line for $\Delta G$ and circles for
843: $-T\Delta S$).
844: \label{G-si}}
845: \end{figure} 
846: 
847: \begin{figure}
848: \caption{Gibbs free-energy difference per unit length $\Delta G$, and
849: the entropic contribution $-T\Delta S$, in $\rm{meV/\AA}$, between the
850: SP and the DP geometries in germanium, using the Tersoff
851: potential (full line for $\Delta G$ and squares for $-T\Delta S$).
852: \label{G-ge}}
853: \end{figure} 
854: 
855: \begin{figure}
856: \caption{Difference in free energy between SP and DP cores in silicon for
857: RS-MC calculations using the Tersoff (full line) and EDIP (dashed
858: line) potentials, compared with the harmonic-approximation results
859: from Ref.~10. RS-MC energies at 0 K were shifted, for better
860: comparison (see text).
861: \label{vallad}}
862: \end{figure} 
863: 
864: \end{document}
865: