1: \documentstyle[11pt,psfig]{article}
2:
3: \addtolength{\oddsidemargin}{-1in}
4: \addtolength{\textwidth}{2.in}
5: \addtolength{\topmargin}{-0.5in}
6: \addtolength{\textheight}{1.5in}
7:
8: \title{The Ab Initio Melting Curve of Aluminium}
9:
10: \author{ Lidunka Vo\v{c}adlo$^1$ and Dario Alf\`e$^{1,2}$\\
11: \small{$^1$Dept. of Geological Sciences, University College London,
12: Gower Street, London, WC1E 6BT}\\ \small{$^2$Dept. of Physics and
13: Astronomy,University College London, Gower Street, London, WC1E 6BT} }
14:
15: \begin{document}
16: \maketitle
17:
18: \begin{abstract}
19: The melting curve of aluminium has been determined from 0 to $\sim
20: $150 GPa using first principles calculations of the free energies of
21: both the solid and liquid. The calculations are based on density
22: functional theory within the generalised gradient approximation using
23: ultrasoft Vanderbilt pseudopotentials. The free energy of the harmonic
24: solid has been calculated within the quasiharmonic approximation using
25: the small-displacement method; the free energy of the liquid and the
26: anharmonic correction to the free energy of the solid have been
27: calculated {\em via} thermodynamic integration from suitable reference
28: systems, with thermal averages calculated using {\em ab-initio}
29: molecular dynamics. The resulting melting curve is in good agreement
30: with both static compression measurements and shock data.
31: % but we calculate a zero pressure melting temperature
32: % of 785 K which is 15 \% lower than the experimental one (933 K) and
33: % not in very good agreement with the previously calculated DFT local
34: % density approximation (LDA) result (890 K)~\cite{dewijs98}. We show
35: % how this disagreement is probably due to the incorrect pressure
36: % predicted by GGA and discuss a possible way of correcting the
37: % calculations, which result in a corrected a zero pressure melting
38: % temperature of 912 K. In this respect aluminium is one of the few
39: % cases for which the LDA yelds better results than the GGA.
40:
41: \end{abstract}
42:
43: \section{Introduction}
44:
45: The determination of the melting curves of materials to very high pressures is of fundamental importance to our understanding
46: of the properties of planetary interiors; however, obtaining such melting curves remains
47: a major challenge to experimentalists and theorists alike. In
48: particular, the melting behaviour of iron is of great interest to the
49: Earth science community, since knowledge of this melt transition would
50: help constrain the temperature at the inner core boundary (about 1200
51: km from the centre of the Earth) which is currently uncertain to
52: within a few thousand degrees. Although several attempts have been
53: made to obtain the melting curve of iron, experimentally and
54: theoretically determined melting curves vary widely with significant
55: disagreement between static compression
56: measurements~\cite{williams87,boehler93,shen98}, shock
57: data~\cite{brown86,yoo93} and first principles
58: calculations~\cite{alfe99,alfe01a,alfe01b,laio00,belonoshko00}. Consequently,
59: the true nature of the melting curve of iron remains in some dispute.
60:
61: In order to test the reliability of the theoretical techniques used in
62: our previous work on iron and to validate further the reported melting
63: curve~\cite{alfe99,alfe01b}, we have calculated the melting curve of
64: aluminium, for which there is a plethora of ambient experimental data
65: (e.g. ~\cite{crc97}), and for which the experimental melting curve has
66: recently been measured~\cite{boehler97,hanstrom00,shaner84}.
67:
68: In the past, a number of theoretical approaches have been used to
69: investigate the melting behaviour of aluminium. Moriarty {\em et
70: al.}~\cite{moriarty84} used the generalised pseudopotential theory
71: (GPT) to calculate the free energy of both the solid and liquid. They
72: treated the solid harmonically within the quasiharmonic approximation
73: and for the liquid they used fluid variational theory, where an upper
74: bound for the free energy is calculated from a reference system
75: constructed within GPT. They obtained a melting curve to 200 GPa in
76: fair agreement with more recently determined experiment
77: data~\cite{boehler97,hanstrom00,shaner84}, predicting a zero pressure
78: melting temperature of 1050K compared to the experimental value of
79: 933K~\cite{crc97}.
80: %P\'elisier~\cite{pelisier84} also used a
81: %pertubational approach based on a local model pseudopotential, whereby
82: %the free energy of solid (including an approximate anharmonic
83: %contribution) and liquid were calculated using a variational approach,
84: %resulting in a melting curve to 200 GPa in close agreement with
85: %Moriarty et al.~\cite{moriarty84}.
86: Mei and Davenport~\cite{mei92} used the embedded atom model (EAM)
87: based on an analytical potential fitted to the structural properties
88: of aluminium. They calculated the free energies of the solid and
89: liquid and obtained a melting temperature at zero pressure of
90: 800K. Morris {\em et al.}~\cite{morris94} employed the same EAM model
91: but they used phase coexistence to determine the melting temperature
92: as a function of pressure, with results considerably lower than
93: previous theoretical and experimental estimates; they obtained a zero
94: pressure melting temperature of $\sim $720 K. Straub
95: {\em et al.}~\cite{straub94} used first principles calculations to construct
96: an optimal classical potential, and used this potential to calculate
97: the free energies of the solid and the liquid using molecular
98: dynamics; they obtained a zero pressure melting temperature of 955 K.
99:
100: The first fully {\em ab-initio} determination of aluminium melting
101: behaviour is that of de Wijs {\em et al.}~\cite{dewijs98}, who
102: obtained the zero pressure melting point by calculating the free
103: energy of the solid and the liquid entirely from first
104: principles. Their calculations were based on density functional theory
105: (DFT)~\cite{generaldft} using the local-density approximation (LDA)
106: for the exchange-correlation energy. The free energy of the solid was
107: obtained as the sum of the free energy of the harmonic solid, within
108: the quasiharmonic approximation, and the full anharmonic contribution,
109: calculated using thermodynamic integration~\cite{frenkel96} using the
110: harmonic solid as the reference system. For the liquid they used
111: thermodynamic integration with a Lennard-Jones fluid as the reference
112: system. They obtained a melting temperature of 890 K. More recently,
113: Jesson and Madden~\cite{jesson2000} used the orbital-free (OF) variant
114: of {\em ab-initio} molecular dynamics and thermodynamic integration to
115: calculate the free energy of liquid and solid aluminium. They found a
116: melting temperature of 615 K, attributing the discrepancy with the
117: DFT-LDA value of de Wijs {\em et al.}~\cite{dewijs98} to either the OF
118: approximation or the pseudopotential used.
119:
120: In this paper we present the first fully {\em ab-initio} calculations
121: of the entire melting curve of aluminium from 0 to 150 GPa. Our
122: calculations are similar in the general principles to those of de Wijs
123: {\em et al.}~\cite{dewijs98} in the sense that we calculate the {\em
124: ab-initio} free energies of both liquid and solid using thermodynamic
125: integration, although we use the generalised gradient approximation
126: (GGA)~\cite{wang91,perdew92} for the exchange-correlation energy. In
127: addition to extending the calculations to a wide range of pressures,
128: we also present a more efficient approach to the thermodynamic
129: integration scheme, in which additional intermediate steps are
130: introduced in order to minimise the computational effort. Finally, we
131: discuss some possible limitations of the GGA.
132:
133: The paper is organised as follows: in section~\ref{sec:techniques} we
134: describe the {\em ab-initio} simulation techniques and the strategy to
135: calculate the melting curve; in section~\ref{sec:liquid}
136: and~\ref{sec:solid} we describe the calculations of the free energy of
137: the liquid and the solid respectively, and in
138: section~\ref{sec:discussion} we present the melting properties of
139: aluminium.
140:
141: \section{{\em Ab-initio} simulation techniques and strategy for melting}\label{sec:techniques}
142:
143: In the present work, the aluminium system was represented by a
144: collection of Al$^{3+}$ ions and $3N$ electrons, where $N$ is the
145: number of atoms. The ions were treated as classical particles, and
146: their motion was adiabatically decoupled from that of the electrons
147: {\em via} the Born-Oppenheimer approximation. For each position of the
148: ions, the electronic problem was solved within the framework of
149: DFT~\cite{generaldft} using the GGA of Perdew and Wang
150: ~\cite{wang91,perdew92}. Thermal electronic excitations were included
151: using the standard methods of finite-temperature DFT developed by
152: Mermin~\cite{mermin65,gillan89,wentzcovitch92}.
153: %In exact DFT the
154: %exchange-correlation energy has an explicit dependence on
155: %temperature; however, little is known about this dependence and we
156: %therefore use its zero temperature form.
157: The present calculations were performed with the code
158: VASP~\cite{kresse96a} which is exceptionally efficient for metals.
159: The interaction between electrons and nuclei was described with the
160: ultrasoft pseudopotential (USPP) method~\cite{vanderbilt90}. We used
161: plane-waves with a cut-off of 130~eV. The Brillouin-zone was sampled
162: using Monkhorst-Pack special points~\cite{monkhorst76} (the detailed
163: form of sampling will be noted where appropriate). The extrapolation
164: of the charge density from one step to the next in the {\em ab-initio}
165: molecular dynamics (AIMD) simulations was performed using the
166: technique described by Alf\`e~\cite{alfe99b}, which improves the
167: efficiency of the calculations by almost a factor of two. The time
168: step used in our simulations was 1 femto-second.
169:
170: To calculate the melting temperature we calculated the Gibbs free
171: energy of both the solid and the liquid as a function of pressure and
172: temperature, $G_s(P,T)$ and $G_l(P,T)$ and, at each chosen $P$,
173: obtained the melting temperature, $T_m$, from $G_s(P,T_m) = G_l(P,T_m)$.
174: In fact, we calculated the Helmholtz free energy $F(V,T)$ as a
175: function of volume and temperature, and the Gibbs free energy was
176: obtained from the usual expression $ G = F + PV $, where $P =
177: -(\partial F / \partial V)_T$ is the pressure. The main problem in
178: determining melting curves with this technique is the high precision
179: with which the free energies need to be calculated. This is because
180: the Gibbs free energy of the liquid crosses the Gibbs free energy of the
181: solid at a shallow angle, the difference in the slopes being the
182: entropy change on melting. For aluminium this is about 1.4 $k_{\rm
183: B}$/atom at zero pressure, which means that an error of 0.01 eV/atom
184: in either $G_s$ or $G_l$ results in an error of $\approx 80$ K in the
185: melting temperature. Therefore, it is important to reduce
186: non-cancelling errors between the liquid and the solid to an absolute
187: minimum. In the next sections we give a detailed discussion of the
188: techniques that we have used to calculate the free energies of the
189: liquid and the solid, and report what the {\em controllable} errors are:
190: those due to {\bf k}-point sampling, finite size, and statistical
191: sampling. We also try to give an estimate of what the {\em
192: uncontrollable} errors due to DFT-GGA may be.
193:
194: %In these calculations {\bf k}-point sampling and plane-wave cut-offs
195: %were carefully chosen for thorough convergence and/or cancellation of
196: %errors between the solid and liquid free energies such that the
197: %precision of the calculations were generally better than only a few
198: %meV.
199:
200: \section{Free energy of the liquid}\label{sec:liquid}
201:
202: The Helmholtz free energy $F$ of a classical system containing $N$
203: particles is:
204: \begin{equation}
205: \label{eqn:helmholtz}
206: F = - k_{\rm B} T \ln \left\{ \frac{1}{N ! \Lambda^{3 N}}
207: \int_V d {\bf R}_1 \ldots d {\bf R}_N \,
208: e^{ - \beta U ( {\bf R}_1 , \ldots {\bf R}_N ; T )}
209: \right\},
210: \end{equation}
211: where $\Lambda = h / ( 2
212: \pi M k_{\rm B} T )^{1/2}$ is the thermal wavelength, with $M$ the
213: nuclear mass, $h$ the Plank's constant, $k_{\rm B}$ the Boltzmann constant
214: and $\beta = 1 / k_{\rm B} T$. The multidimensional integral extends
215: over the total volume of the system $V$.
216:
217: A direct calculation of $F$ using the equation above is impossible,
218: since it would involve knowledge of the potential energy $ U (
219: {\bf R}_1 , \ldots {\bf R}_N ; T )$ for all possible positions of the
220: $N$ atoms in the system. We have used instead the technique known as
221: thermodynamic integration~\cite{frenkel96}, as developed in earlier
222: papers~\cite{sugino95,smargiassi95a,dewijs98,alfe00}. This is a
223: general scheme to compute the free energy difference $F - F_0$ between
224: two systems whose potential energies are $U$ and $U_0$ respectively.
225: In what follows we will assume that $F$ is the unknown free energy of
226: the {\em ab-initio} system and $F_0$ is the known free energy of a
227: reference system. The free energy difference $F - F_0$ is the
228: reversible work done when the potential energy function $U_0$ is
229: continuously and reversibly switched to $U$. To do this switching, a
230: continuously variable energy function $U_\lambda$ is defined such that
231: for $\lambda=0, U_\lambda=U_0$ and for $\lambda=1, U_\lambda=U$. We
232: also require $U_\lambda$ to be differentiable with respect to
233: $\lambda$ for $0 \le \lambda \le 1$. A convenient form is:
234: \begin{equation}
235: U_\lambda = ( 1 - f(\lambda) ) U_0 + f(\lambda) U,
236: \end{equation}
237: where $f(\lambda)$ is an arbitrary continuous and differentiable
238: function of $\lambda$ with the property $f(0)=0$ and $f(1)=1$. The
239: Helmholtz free energy of this {\em hybrid} system is:
240: \begin{equation}
241: F_\lambda = - k_{\rm B} T \ln \left\{ \frac{1}{N ! \Lambda^{3 N}}
242: \int_V d {\bf R}_1 \ldots d {\bf R}_N \,
243: e^{- \beta U_\lambda ( {\bf R}_1 , \ldots {\bf R}_N ; T )}
244: \right\},
245: \end{equation}
246: Differentiating this with respect to $\lambda $ gives:
247: \begin{equation}
248: \frac{dF_\lambda }{d\lambda }=-k_BT\frac{\frac {1}{N!\Lambda^{3N}}\int_V
249: d{\bf R}_1 \ldots d{\bf R}_N e^{-\beta U_\lambda ({\bf R}_1, \ldots
250: {\bf R}_N;T)}(-\beta\frac{\partial U_\lambda }{\partial \lambda })}{{\frac
251: {1}{N!\Lambda^{3N}}\int_V d{\bf R}_1 \ldots d{\bf R}_N e^{-\beta U_\lambda
252: ({\bf R}_1, \ldots {\bf R}_N;T)}}}=\left\langle \frac{\partial U_\lambda
253: }{\partial \lambda } \right\rangle_\lambda,
254: \end{equation}
255: so
256: \begin{equation}
257: \label{eqn:ti}
258: \Delta F=F-F_0=\int\limits_0^1d\lambda \left\langle \frac{\partial
259: U_\lambda }{\partial \lambda }\right\rangle _\lambda .
260: \end{equation}
261: For our calculations we defined $U_\lambda $ thus:
262: \begin{equation}
263: U_\lambda =\left( 1-\lambda \right) U_0 + \lambda U.
264: \end{equation}
265: Differentiating U$_\lambda $ with respect to $\lambda $ and substituting
266: into Equation~\ref{eqn:ti} yields:
267: \begin{equation}\label{eqn:ti_with_linear_switching}
268: \Delta F=\int\limits_0^1d\lambda \left\langle U-U_0\right\rangle
269: _\lambda.
270: \end{equation}
271: Under the ergodicity hypothesis, thermal averages are equivalent to time
272: averages, so we calculated $ \left\langle \cdot \right\rangle _\lambda
273: $ using AIMD, taking averages over time, with the evolution of the system
274: determined by the potential energy function $U_\lambda$. The
275: temperature was controlled using a Nos\'e
276: thermostat~\cite{nose84,ditolla93}. It is important to stress that
277: the choice of the reference system does not affect the final answer
278: for $F$, although it does affect the efficiency of the calculations.
279: The latter can be understood by analysing the quantity $\langle U -
280: U_0 \rangle_\lambda$. If this difference has large fluctuations then
281: one would need very long simulations to calculate the average value to
282: a sufficient statistical accuracy. Moreover, for an unwise choice of
283: $U_0$ the quantity $\langle U - U_0 \rangle_\lambda$ may strongly
284: depend on $\lambda$ so that one would need a large number of
285: calculations at different $\lambda$'s in order to compute the integral
286: in Eq.~\ref{eqn:ti_with_linear_switching} with sufficient accuracy. It
287: is crucial, therefore, to find a good reference system, where "good"
288: means a system for which the fluctuations of $U - U_0$ are as small as
289: possible. In fact, if the fluctuations are small enough, we can
290: simply write $F - F_0 \simeq \langle U - U_0 \rangle_0$, with the
291: average taken in the reference ensemble. If this is not good enough,
292: the next approximation is readily shown to be~\cite{alfe01a}:
293: \begin{equation}
294: F - F_0 \simeq \langle U -
295: U_0 \rangle_0 - \frac{1}{2 k_{\rm B} T}
296: \left\langle \left[ U - U_0 -
297: \langle U - U_0 \rangle_0
298: \right]^2 \right\rangle_0.
299: \label{eqn:secondorder}
300: \end{equation}
301: This form is particularly convenient since one only needs to sample
302: the phase space with the reference system, and perform a number of
303: {\em ab-initio} calculations on statistically independent configurations
304: extracted from a long classical simulation.
305:
306: To evaluate the integral in
307: Equation~\ref{eqn:ti_with_linear_switching} one can calculate the
308: integrand $\langle U - U_0 \rangle_\lambda$ at a sufficient number
309: of $\lambda$ and calculate the integral numerically.
310:
311: Alternatively, one can adopt the dynamical method described by
312: Watanabe and Reinhardt~\cite{watanabe90}. In this approach the
313: parameter $\lambda$ depends on time, and is slowly (adiabatically)
314: switched from 0 to 1 during a single simulation. The switching rate
315: has to be slow enough so that the system remains in thermodynamic
316: equilibrium, and adiabatically transforms from the reference to the
317: {\em ab-initio} system. The change in free energy is then given by:
318: \begin{equation}\label{eqn:adiabatic_switching}
319: \Delta F=\int\limits_0^{T_{\rm sim}} dt\frac{d\lambda }{dt}\left(
320: U-U_0\right),
321: \end{equation}
322: where $T_{\rm sim}$ is the total simulation time, $\lambda(t)$ is an
323: arbitrary function of $t$ with the property of being continuous and
324: differentiable for $0\le t \le 1$, $\lambda(0)=0$ and
325: $\lambda(T_{\rm sim}) = 1$.
326:
327: When using this second method, it is important to ensure that the switching is
328: adiabatic, i.e. that $T_{\rm sim}$ is sufficiently large. This can be achieved by changing
329: $\lambda$ from 0 to 1 in the first half of the simulation, and then
330: from 1 back to 0 in the second half of the simulation, evaluating
331: $\Delta F$ in each case; the average of the two values is then taken as
332: the best estimate for $ \Delta F$, and the difference is a measure of the non-adiabaticty. If this difference is less than the desired statistical uncertainty, one can be confident that the simulation time is sufficiently long.
333:
334: In our calculations we chose a total simulation time of
335: sufficient length such that the difference in $\Delta F$
336: between the two calculations was less than a few meV/atom. We return later to estimate the
337: errors in our calculations in Section~\ref{sec:liquid_errors}.
338:
339: As pointed out by Jesson and Madden~\cite{jesson2000}, a possible
340: problem in the calculation of the thermodynamic integral is that the
341: system $U_\lambda$ may be in the solid region of the phase diagram,
342: even though the two end members $U_0$ and $U$ are in the liquid
343: region. If this happens, the system can freeze during the
344: switching, and the integration path is not reversible, leading to an
345: incorrect result. For small systems the situation is even more
346: problematic, since the phase diagram is not defined by sharp
347: boundaries, and the system can freeze even if it is above the melting
348: temperature of the corresponding system in the thermodynamic limit.
349: We have ourselves experienced freezing of the system for some
350: simulations at temperatures very close to the melting point; in order to avoid
351: including the results from these simulations, we carefully
352: monitored the mean square displacement and the structure factor of the
353: system, and included only those simulations in which these two quantities
354: clearly indicated liquid behaviour throughout the whole simulation.
355:
356:
357: \subsection{The reference system}
358:
359: We mentioned earlier that the efficiency of the calculations is
360: entirely determined by the quality of the reference system, i.e. by
361: the strength of the fluctuations of $\Delta U = U - U_0$. The key to
362: the success of these simulations, therefore, is being able to find a
363: reference system such that the fluctuations in $\Delta U$ are as small
364: as possible. Based on the experience of previous work on liquid
365: Al~\cite{dewijs98} and liquid Fe~\cite{alfe99,alfe01b} we experimented
366: with the Lennard-Jones (LJ) system and an inverse power potential
367: (IP). Analysis of the fluctuations in $\Delta U$ indicated that the
368: system which best represented the liquid was the IP:
369: \begin{equation}
370: U_{\rm IP} = \frac{1}{2} \sum_{I \ne J} \phi ( \mid {\bf R}_I -
371: {\bf R}_J \mid ),
372: \label{eqn:uip}
373: \end{equation}
374: where
375: \begin{equation}
376: \phi(r)= \frac {B} {r^\alpha}.
377: \end{equation}
378: The potential parameters $B$ and $\alpha$ were chosen by minimising
379: the quantity:
380: \begin{equation} \label{eqn:fluctuations}
381: \left\langle \left [ U_{\rm IP}\left( B ,\alpha\right) - U - \langle
382: U_{\rm IP}\left( B ,\alpha\right) - U \rangle \right ]^2\right\rangle
383: \end{equation}
384: with respect to $B$ and $\alpha$, where $\langle \rangle$ means
385: the thermal average in the ensemble generated by the {\em ab-initio}
386: potential. To investigate whether the optimum values for the
387: potential parameters depended strongly on thermodynamic state, we
388: performed the optimisation at the three thermodynamic states of the
389: extremes of high $P/T$ and low $P/T$, and also a point in between; we
390: found that the single choice of $B =246.67$ and $\alpha=6.7$ (units of
391: eV and \AA) was equally good for all states and we therefore used
392: these two parameters for all our calculations.
393:
394: It may be surprising that such a simple inverse power potential can
395: reproduce the energetics of the liquid with sufficient accuracy, since
396: simple repulsive potentials cannot describe metallic bonding. One may
397: think that a more realistic potential such as those based on the
398: EAM~\cite{mei92,daw93,baskes92,belonoshko97} would be more
399: appropriate, since these potentials explicitly contain a repulsive and
400: a bonding term. However, in our recent work on iron~\cite{alfe01a} we
401: tested the use of an EAM potential as a reference system and found
402: that the bonding term is almost independent of the positions of the
403: atoms, depending only on the volume and temperature of the system, and
404: the fluctuations of the energy are almost entirely due to the
405: repulsive term. Since the only relevance in this work is the strength
406: of the fluctuations (Eq.~\ref{eqn:fluctuations}), little is gained by
407: using an EAM rather then a much simpler inverse power potential.
408:
409:
410: \subsection{Free energy of the reference system}
411:
412: Consider the excess free energy of the IP, $F^{\rm ex}_{\rm IP} =
413: F_{\rm IP} - F_{\rm PG}$, where $F_{\rm PG}$ is the Helmholtz free
414: energy of the perfect gas and $F_{\rm IP}$ the total Helmholtz free
415: energy of the IP system. The very simple functional form
416: of $U_{\rm IP}$ makes it easy to show that the adimensional quantity
417: $F^{\rm ex}_{\rm IP}/k_{\rm B}T$ can only depend non-trivially on a
418: single thermodynamic variable, rather then separately on $V$ and $T$:
419: \begin{equation}
420: F^{\rm ex}_{\rm IP}/k_{\rm B}T = f(\zeta)
421: \end{equation}
422: with
423: \begin{equation}\label{eqn:zeta}
424: \zeta = B /V^{\alpha / 3} k_{\rm B} T.
425: \end{equation}
426: The free energy of the IP has been studied extensively in the
427: past~\cite{laird92}, but only for special values of the exponent
428: $\alpha$, which did not include our own $\alpha=6.7$. We have
429: therefore explicitly calculated the free energy of our inverse power
430: potential using thermodynamic integration as before, but this time we
431: started from a system of known free energy, the Lennard-Jones liquid, whose
432: potential function is given by
433: \begin{equation}
434: U_{\rm LJ}=4\varepsilon \left[ \left( \frac \sigma r\right) ^{12}-\left( \frac
435: \sigma r\right) ^6\right].
436: \end{equation}
437: The free energy of the Lennard-Jones liquid, $F_{\rm LJ}$, has been
438: accurately tabulated by Johnson {\em et al.}~\cite{johnson93}.
439: % These classical calculations
440: %can be highly accurate since the simulations can be very long and can
441: %be performed on very large systems.
442: % as a function of reduced temperature ($
443: %T^{*}=k_{\rm B}T/\varepsilon $) and density $(\rho ^{*}=(N/V)\sigma
444: %^3)$
445: %In order to get the free energy difference between the Lennard-Jones
446: %system and our inverse-power reference system, we used thermodynamic
447: %integration:
448: %\begin{equation}\label{eqn:lj_ip}
449: %\Delta F_{\rm LJ\rightarrow IP}=F_{\rm IP}-F_{\rm LJ}=\int\limits_0^{T_{\rm sim}}
450: %dt\frac{d\lambda }{ dt}\left( U_{\rm IP} -U_{\rm LJ} \right).
451: %\end{equation}
452: To calculate $F_{\rm IP} - F_{\rm LJ} = \Delta F_{\rm LJ\rightarrow
453: IP}$ we used simulation cells containing 512 atoms with periodic
454: boundary conditions and a simulation time $T_{\rm sim}= 200$ ps. We
455: performed the calculations for $\zeta$ ranging from 2.5 to 6.25, with
456: steps of 0.25. The calculations were done at a fixed volume of 14
457: \AA$^3$/atom and varing temperatures according to
458: Equation~\ref{eqn:zeta}. We carefully checked that the results were
459: converged to better than 1 meV/atom with respect to the size of the
460: simulation cell and the length of the simulations. To avoid
461: truncating the inverse power potential at a finite distance we used
462: the Ewald technique. Our results were fitted to a third order
463: polynomial in $\zeta$:
464: \begin{equation}
465: f(\zeta) = \sum_{i=0}^3 c_i \zeta^i.
466: \end{equation}
467: The coefficients are: $c_0 = 2.4333; c_1 = 27.805; c_2 = -5.0704; c_3
468: = 1.5177$, and the fitting function reproduced the calculated data
469: such that the errors in $F_{\rm IP}$ were generally less than 1 meV per
470: atom.
471:
472: % ~\cite{ewald21,ewald37}.
473:
474: As an additional check on the calculated free energy, we repeated most
475: of the simulations using the perfect gas as the reference system,
476: thereby avoiding the inclusion of any possible errors that may exist
477: in the free energy of the LJ system reported in the
478: literature~\cite{johnson93}. For these calculations we used a
479: different form for $U_\lambda$, namely:
480: \begin{equation}
481: %U_\lambda = ( 1 - \lambda^2 ) U_{\rm PG} + \lambda^2 U_{\rm IP} \; ,
482: U_\lambda = \lambda^2 U_{\rm IP} \;
483: \end{equation}
484: (the potential energy of the perfect gas is zero, so does
485: not appear in the formula). So Eq.~\ref{eqn:ti} becomes
486: \begin{equation}
487: F_{\rm IP} - F_{\rm PG} = \int_0^1 d \lambda \, 2\lambda \langle
488: U_{IP} \rangle_\lambda.
489: \end{equation}
490: The advantage of using this different functional form for $U_\lambda$
491: is that the value of the integrand does not need to be computed for
492: $\lambda=0$, where the dynamics of the system is determined by the
493: perfect gas potential. In this case, since there are no forces in the
494: system there is nothing stopping the atoms from overlapping, and the
495: potential energy $U_{\rm IP}$ diverges. Not computing the integrand at
496: $\lambda=0$ partially solves this problem, but for small values of
497: $\lambda$ where the forces on the atoms are small, the atoms can come
498: close together and the potential energy, $U_{\rm IP}$, fluctuates
499: violently. However, we found that by performing long enough
500: simulations, typically 1~ns, we could calculate the integral with an
501: accuracy of $\approx 1$ meV/atom, and, within the statistical
502: accuracy, we found the same results as those obtained using the LJ
503: reference system.
504:
505: \subsection{Free energy of the {\em ab-initio} system}
506:
507: To calculate the full {\em ab-initio} free energy of the liquid,
508: $F_{\rm liq}$, we used thermodynamic integration, starting from the IP
509: system. The calculations were performed at 18 different thermodynamic
510: states over a range of volumes (9.5-19.5 \AA$^3$/atom) and
511: temperatures (800-6000 K). To address the issue of {\bf k}-point
512: sampling and cell size errors in the free energy difference $F_{\rm
513: liq} - F_{\rm IP}$, tests were carried out on cells containing up to
514: 512 atoms and a $5 \times 5 \times 5$ {\bf k}-point grid (calculations
515: on the largest 512 atoms cell were only performed with
516: $\Gamma$-point sampling), at $V=16.5$ and $T=1000K$. The free energy
517: difference $F_{\rm liq} - F_{\rm IP}$ was calculated using the
518: perturbational approach (Eq.~\ref{eqn:secondorder}), with sets of
519: configurations generated using the IP potential. We found that a
520: 64-atom cell with a $3\times 3 \times 3$ {\bf k}-point grid was
521: sufficient to get convergence to within 2 meV/atom. However, we were
522: reluctant to perform simulations using the desired $3\times 3 \times
523: 3$ {\bf k}-point grid (14 points in the Brillouin Zone (BZ)) since
524: these calculations are extremely expensive. We found it more
525: efficient to add one further step to our thermodynamic integration
526: scheme:
527: \begin{equation}\label{eqn:gammakappa}
528: \Delta F_{\Gamma \rightarrow 333}=F_{333}-F_\Gamma
529: =\int\limits_1^0d\lambda \left\langle U_{333}-U_\Gamma
530: \right\rangle_\lambda,
531: \end{equation}
532: where $U_{333}$ and $U_\Gamma$ are the {\em ab-initio} total energies
533: calculated using the $3\times 3 \times 3$ {\bf k}-point grid and
534: $\Gamma$-point sampling respectively, and $F_{333}$ and $F_{\Gamma}$
535: are the corresponding free energies. To evaluate the free energy
536: difference $\Delta F_{\Gamma \rightarrow 333}$ we noticed that the
537: difference $U_{333}-U_\Gamma$ did not depend significantly on the
538: position of the atoms, so the integral in Eq.~\ref{eqn:gammakappa}
539: could be evaluated using the second order formula
540: (Eq.~\ref{eqn:secondorder}). Using a long
541: $\Gamma$-point {\em ab-initio} simulation, we extracted up to 25 statistically
542: independent configurations and calculated the {\em ab-initio} energies
543: using the $3\times 3 \times 3$ {\bf k}-point grid. To test this, we performed spot checks at two thermodynamic
544: states, where we calculated the full thermodynamic integral $F_{333} -
545: F_{\rm IP} $ using adiabatic switching with a switching time of
546: $\approx 2$ ps, and found the same results to within a few meV/atom.
547: %In fact, the
548: %fluctuations in the energy differences $U_{333}-U_\Gamma$ were so
549: %small that the second term in Eq.~\ref{eqn:secondorder} was completely
550: %negligible.
551:
552: The free energy difference $\Delta F_{\rm IP \rightarrow \Gamma} =
553: F_\Gamma - F_{\rm IP}$ was obtained by full thermodynamic integration
554: between the {\em ab initio} and reference system using
555: adiabatic switching (Eq.~\ref{eqn:adiabatic_switching}) with a
556: switching time of 5 ps, which resulted in errors of 1(4) meV/atom in
557: the low(high) $P/T$ region. To test this, we also calculated this free energy difference at several state points by numerical evaluation of the
558: thermodynamic integral (Eq.~\ref{eqn:ti}), with ~$\lambda$ = 0, 0.5 and 1; we found that this gave the same numerical answer to within our statistical errors.
559:
560:
561: In summary, the free energy of the liquid was obtained from a
562: series of thermodynamic integration calculations:
563: \begin{equation}
564: F_{\rm liq} = F_{333} = F_{\rm LJ}+\Delta F_{\rm LJ\rightarrow
565: IP}+\Delta F_{\rm IP\rightarrow \Gamma }+\Delta F_{\Gamma \rightarrow
566: 333}.
567: \end{equation}
568:
569:
570: \subsection{Representation of the free energy of the liquid}
571:
572: The results of the calculations described in the previous section were
573: fitted to a suitable function of $T$ and $V$. In order to do that
574: efficiently we expressed the free energy in the following way:
575: \begin{equation}
576: F_{\rm liq} = F_{\rm IP} + \Delta F = F_{\rm IP} + \Delta U^s +
577: (\Delta F - \Delta U^s)
578: \end{equation}
579: where $\Delta U^s = U^s - U^s_{\rm IP}$, with $U^s $ the zero
580: temperature ab-initio (free) energy of the face-centred-crystal (fcc)
581: and $U^s_{\rm IP}$ the inverse power energy. $U^s$ can be calculated
582: very accurately, details of which will be given below in
583: Section~\ref{sec:perfect_freeenergy}; $U^s_{\rm IP}$ has no
584: errors. The remaining quantity $\Delta F - \Delta U^s$ is a
585: weak function of $V$ and $T$, and was fitted to a polynomial in $V$
586: and $T$:
587: \begin{equation}
588: \Delta F - \Delta U^s = \sum_{j=0}^1 \left ( \sum_{i=0}^{3} a_{ij} V^i
589: \right ) T^j
590: \end{equation}
591: %The values of the parameters are (units of eV, \AA$^3$ and K), $a_{00}
592: %= -4.842 10^{-1}; a_{10} = 1.606 10^{-1}; a_{20} = -1.229 10^{-2};
593: %a_{30} = 2.696 10^{-4}; a_{01} = -1.931 10^{-4}; a_{11} = 4.020
594: %10^{-5}; a_{21} = -2.417 10^{-6}; a_{31} = 4.118 10^{-8}$.
595: The fitting reproduced the calculated data to within $\approx 2$
596: meV/atom.
597:
598: \subsection{Error estimates for $F_{\rm liq}$}\label{sec:liquid_errors}
599: The errors on $F_{\rm IP}$ and $\Delta U^s$ are each less than 1
600: meV/atom (see Section~\ref{sec:perfect_freeenergy} below). The part of
601: the free energy that carries the largest errors is $\Delta F - \Delta
602: U^s$, which we estimate to be 2(5) meV/atom at
603: low(high) $P/T$.
604:
605: \section{Free energy of the solid}\label{sec:solid}
606:
607: The free energy of the solid can be represented as the sum of two
608: contributions: the free energy of the perfect
609: non-vibrating fcc crystal and that arising from atomic
610: vibrations above zero Kelvin:
611: \begin{equation}
612: F_{\rm sol} = F_{\rm perf} + F_{\rm vib}.
613: \end{equation}
614: The contribution to the free energy due to the vibrations of the atoms
615: may be written:
616: \begin{equation}
617: F_{\rm vib} = F_{\rm harm} + F_{\rm anharm}
618: \end{equation}
619: where $F_{\rm harm}$ is the free energy of the high temperature
620: crystal in the harmonic approximation and $F_{\rm anharm}$ is the
621: anharmonic contribution.
622:
623: \subsection{Free energy of the perfect crystal}\label{sec:perfect_freeenergy}
624:
625: The free energy of the perfect crystal, $F_{\rm perf}$, was calculated
626: as a function of volume and temperature. Calculations were performed on a
627: fcc cell at a series of volumes (9.5-19.5~\AA$^3$/atom representing
628: compression up to $\sim 150$ GPa) and temperatures (up to 6000K) with a
629: 24x24x24 {\bf k}-point grid (equivalent to 1300 points in the
630: irreducible wedge of the Brilloiun zone (IBZ)), which ensures
631: convergence of the (free) energies to better than 1 meV/atom. At each
632: different temperature we calculated the {\em ab-initio} (free)
633: energy as a function of volume, and then performed a least-square fit of the
634: results to a third-order Birch-Murnaghan equation of state:
635: \begin{eqnarray}\label{murna}
636: E(V) = E_0 + \frac{3}{2}V_0K \left [ \frac{3}{4}(1+2\xi)\left
637: (\frac{V_0}{V}\right )^{4/3} - \frac{\xi}{2}
638: \left ( \frac{V_0}{V} \right )^{2}
639: -\frac{3}{2}(1+\xi) \left ( \frac{V_0}{V}
640: \right )^{2/3} +
641: \frac{1}{2} \left ( \xi + \frac{3}{2}\right ) \right ] \\
642: \xi = \frac{3}{4}(4 - K'). \hspace{12cm} \nonumber
643: \end{eqnarray}
644: The parameters $E_0, V_0, K_0,$ and $K'$ were fitted to a
645: fourth order polynomial as function of temperature:
646: \begin{equation}
647: E_0(T) = \sum_{i=0}^{4} e_{0,i} T^i; \hskip 20pt
648: V_0(T) = \sum_{i=0}^{4} v_{0,i} T^i; \hskip 20pt
649: K_0(T) = \sum_{i=0}^{4} k_{0,i} T^i; \hskip 20pt
650: K'(T) = \sum_{i=0}^{4} k_{0,i}' T^i.
651: \end{equation}
652: The fitting reproduced the calculated energies to better than 1
653: meV/atom in the whole $P/T$ range.
654:
655: \subsection{Free energy of the harmonic crystal}
656:
657: The free energy of the harmonic crystal is given by:
658: \begin{equation}
659: F_{\rm harm}(V,T)=-\left( \frac{3k_{\rm B}T}{\Omega _{\rm BZ}N_i}\right)
660: \sum\limits_i\int\limits_{\rm BZ}\left( \ln \left[ \frac{k_BT}{\hbar
661: \omega _{{\bf q},i}(V,T)}\right] -\frac {1}{24}\left[ \frac{\hbar \omega_{{\bf
662: q},i}(V,T)}{k_{\rm B}T}\right] ^2+\dots \right) d{\bf q}
663: \end{equation}
664: where $\omega _{{\bf q},i}(V,T)$ are the phonon frequencies of branch
665: {\it i} and wavevector {\bf q}, $\Omega _{\rm BZ}$ is the volume of
666: the Brillouin zone, $N_i$ is the total number of phonon branches and
667: the dependence on temperature of $\omega _{{\bf q},i}$ is due to
668: electronic excitations. We truncate the summation after the first term, which
669: is the classical limit of the free energy:
670: \begin{equation}
671: F_{\rm harm}=-\left( \frac{3k_{\rm B}T}{\Omega _{\rm BZ}N_i}\right)
672: \sum\limits_i\int\limits_{\rm BZ}\left( \ln \frac{k_{\rm B}T}{\hbar \omega
673: _{{\bf q},i}} \right) d{\bf q}.
674: \end{equation}
675: This is a justifiable approximation to make for two reasons: (i) the
676: error in making such a truncation is very small ($<$1 meV/atom), and
677: (ii) neglecting the higher order terms, i.e., the quantum corrections,
678: is consistent with the liquid calculations where the motions of the
679: atoms were treated classically.
680:
681: It is useful to express the harmonic free energy in terms of the
682: geometric average $\bar{\omega}$ of the phonon frequencies, defined
683: as:
684: \begin{equation}
685: \ln \bar{\omega} = \frac{1}{N_{{\bf q}} N_i} \sum_{{\bf q}, i}
686: \ln ( \omega_{{\bf q} i} ) ,
687: \end{equation}
688: where we have replaced the integral $\frac{1}{\Omega_{\rm
689: BZ}}\int\limits_{\rm BZ} d{\bf q}$ with the summation $\frac{1}{N_{\bf
690: q}}\sum\limits_{\bf q}$.
691: This allows us to write:
692: \begin{equation}
693: F_{\rm harm} = 3 k_{\rm B} T \ln ( \beta \hbar \bar{\omega} ).
694: \end{equation}
695:
696: To calculate the vibrational frequencies $\omega_{{\bf q},i}$, we used
697: our own implementation~\cite{darioweb} of the small displacement
698: method~\cite{kresse95,alfe01a}.
699:
700: The central quantity in the calculation of the phonon frequencies
701: is the force-constant matrix $\Phi_{i s \alpha , j t \beta}$,
702: since the frequencies at wavevector ${\bf q}$ are
703: the eigenvalues of the dynamical matrix $D_{s \alpha , t \beta}$,
704: defined as:
705: \begin{equation}
706: D_{s \alpha , t \beta} ( {\bf q} ) = \frac{1}{\sqrt{M_s M_t}} \sum_i
707: \Phi_{i s \alpha , j t \beta} \exp \left[ i {\bf q} \cdot ( {\bf
708: R}_j^0 + {\bf \tau}_t - {\bf R}_i^0 - {\bf \tau}_s) \right] \; .
709: \end{equation}
710: where ${\bf R}_i^0$ is a vector of the lattice connecting different
711: primitive cells, ${\bf \tau}_s$ is the position of the atom $s$ in the
712: primitive cell and $M_s$ its mass. If we have the complete
713: force-constant matrix, then $D_{s \alpha , t \beta}$, and hence the
714: frequencies $\omega_{{\bf \rm q} l}$, can be obtained at any ${\bf q}$. In
715: principle, the elements of $\Phi_{i s \alpha , j t \beta}$ are
716: non-zero for arbitrarily large separations $\mid {\bf R}_j^0 + {\bf
717: \tau}_t - {\bf R}_i^0 - {\bf \tau}_s \mid$, but in practice they decay
718: rapidly with separation, so a key issue in achieving our target
719: precision is the cut-off distance beyond which the elements can be
720: neglected.
721:
722: In the harmonic approximation the $\alpha$ Cartesian component of
723: the force exerted on the atom at position ${\bf R}_i^0 + {\bf \tau}_s$
724: is given by:
725: \begin{equation}
726: F_{i s \alpha} = - \sum_{j t \beta} \Phi_{i s \alpha , j t \beta}~u_{j t \beta}
727: \end{equation}
728: where $u_{j s \beta}$ is the displacement of the atom in ${\bf R}_j^0
729: + {\bf \tau}_t$ along the direction $\beta$. The force constant matrix can
730: be calculated {\em via}:
731: \begin{equation}\label{displ}
732: \Phi_{i s \alpha, j t \beta}=-\frac{F_{i s \alpha, j t \beta}}{ u_{j t \beta} }
733: \end{equation}
734: where all the atoms of the lattice are displaced one at a time along the
735: three Cartesian components by $u_{j t \beta}$, and the
736: forces $F_{i s \alpha, j t \beta}$ induced on the atoms in ${\bf
737: R}_i^0 + {\bf \tau}_s$ are calculated. Since the crystal is invariant under
738: translations of any lattice vector, it is only necessary to displace
739: the atoms in one primitive cell and calculate the forces induced on
740: all the other atoms of the crystal, so that we can simply put $j = 0$.
741: The fcc crystal has only one atom in the
742: primitive cell, so only three displacements are needed. However, a
743: displacement along the $x$ direction is equivalent by symmetry to a
744: displacement along the $y$ or the $z$ direction, and therefore only
745: one displacement along an arbitrary direction is needed. It is
746: convenient to displace the atom along a direction of high symmetry, so
747: that the supercell has the maximum possible number of symmetry
748: operations. These can be used to reduce the number of {\bf k}-points
749: in the IBZ, minimising the computational effort. For an
750: fcc crystal this is achieved by displacing the atom along the
751: diagonal of the cube.
752:
753: %It is important to appreciate that the $\Phi_{l s \alpha , l^\prime t
754: %\beta}$ in the formula for $D_{s \alpha , t \beta} ( {\bf q} )$ is the
755: %force-constant matrix in the infinite lattice, with no restriction on
756: %the wavevector ${\bf q}$, whereas the calculations of
757: %$\Phi_{l s \alpha , l^\prime t \beta}$ can only be done in supercell
758: %geometry. Without a further assumption, it is strictly impossible to
759: %extract the infinite-lattice $\Phi_{l s \alpha , l^\prime t \beta}$
760: %from supercell calculations, since the latter deliver information only
761: %at wavevectors that are reciprocal lattice vectors of the superlattice.
762:
763: %The further assumption needed is that the infinite-lattice $\Phi_{l s
764: %\alpha , l^\prime t \beta}$ vanishes when the separation ${\bf
765: %R}_{l^\prime t} - {\bf R}_{l s}$ is such that the positions ${\bf
766: %R}_{l s}$ and ${\bf R}_{l^\prime t}$ lie in different Wigner-Seitz
767: %(WS) cells of the chosen superlattice. More precisely, if we take the
768: %WS cell centred on ${\bf R}_{l^\prime t}$, then the infinite-lattice
769: %value of $\Phi_{l s \alpha , l^\prime t \beta}$ vanishes if ${\bf
770: %R}_{l s}$ is in a different WS cell; it is equal to the supercell
771: %value if ${\bf R}_{l s}$ is wholly within the same WS cell, and it is
772: %equal to the supercell value divided by an integer $P$ if ${\bf R}_{l
773: %s}$ lies on the boundary of the same WS cell, where $P$ is the number
774: %of WS cells having ${\bf R}_{l s}$ on their boundary. With this
775: %assumption, the $\Phi_{l s \alpha , l^\prime t \beta}$ elements will
776: %converge to the correct infinite-lattice values as the dimensions of
777: %the supercell are systematically increased.
778:
779: Tests for cell-size (64-512 atoms), displacement length (0.01-0.0005
780: fraction of nearest neighbours distance) and {\bf k}-point grid (up to
781: $9 \times 9\times 9$) were performed at the two extremes of high $P/T$
782: and low $P/T$ state points. Convergence of the free energy to within less
783: than 3 meV/atom was achieved using a 64-atom cell with a 0.001
784: fractional displacement and a $9 \times 9\times 9$ k-point grid
785: (equivalent to 85 points in the IBZ of the supercell). Calculations
786: were performed for $V= 9.5-18.5~{\rm \AA}^3$ and $T=500-6000K$, and
787: $\ln (\bar{\omega})$ has been fitted to the following polynomial in
788: $V$ and $T$:
789: \begin{equation}
790: \ln (\bar{\omega}) = \sum_{j=0}^3 \left ( \sum_{i=0}^{3} a_{ij} V^i \right ) T^j.
791: \end{equation}
792: %The values of the parameters are (units of eV, \AA$^3$ and K), $a_{00}
793: %= 5.046; a_{10} = 2.156 10^{-1}; a_{20} = 7.054 10^{-4}; a_{30} =
794: %-9.401 10^{-5}; a_{01} = 1.908 10^{-4}; a_{11} = -4.432 10^{-5};
795: %a_{21} = 3.360 10^{-6}; a_{31} = -8.321 10^{-8}; a_{02} = -2.665
796: %10^{-8}; a_{12} = 6.341 10^{-9}; a_{22} = -5.062 10^{-10}; a_{32} =
797: %1.320 10^{-11}; a_{03} = 1.097 10^{-12}; a_{13} = -2.767 10^{-13};
798: %a_{23} = 2.423 10^{-14}; a_{33} = -6.902 10^{-16}$.
799: The fitting reproduced the calculated data within $\approx 1$
800: meV/atom.
801:
802:
803: \subsection{Anharmonicity}
804:
805: To obtain the anharmonic contribution to the free energy of the solid
806: we have again used thermodynamic integration. In this case a natural
807: choice for the reference system could be the harmonic
808: solid~\cite{dewijs98}, but unfortunately this does not reproduce the
809: {\it ab initio} anharmonic system with sufficient accuracy. A much
810: better reference system is a linear combination of the harmonic {\em
811: ab-initio} and the same IP used for the liquid calculations~\cite{alfe01a}:
812: \begin{equation}
813: U_{\rm ref} = a U_{\rm IP} + b U_{\rm harm},
814: \end{equation}
815: where the harmonic potential energy is:
816: \begin{equation}
817: U_{\rm harm} = \frac 1 2 \sum_{i s \alpha, j t \beta}~u_{i s \alpha}
818: \Phi_{i s \alpha , j t \beta}~u_{j t \beta},
819: \end{equation}
820: and where $u_{j s \beta}$ is the displacement of the atom in ${\bf R}_j^0 +
821: {\bf \tau}_t$ along the direction $\beta$, and $\Phi_{i s \alpha , j t
822: \beta}$ is the force constant matrix. The parameters $a$ and $b$ are
823: determined by minimising the fluctuations in the energy differences
824: $U_{\rm ref} - U$ on a set of statistically independent configurations
825: generated with $U_{\rm ref}$. However, when we start our optimisation
826: procedure we do not know $U_{\rm ref}$, so we cannot use it to
827: generate the configurations. We could use the {\em ab-initio}
828: potential, but this would involve very expensive calculations. We
829: used instead an iterative procedure, like in our previous work on
830: iron~\cite{alfe01a}. We generated a set of configurations using the
831: harmonic potential $U_{\rm harm}$ and calculated the {\em ab-initio}
832: energies. By minimising the fluctuations of $U_{\rm ref} - U$ we found
833: a first estimate for $a$ and $b$, and we constructed a first estimate
834: of $U_{\rm ref}$. We generated a second set of configurations using
835: this $U_{\rm ref}$, calculated the {\em ab-initio} energies and
836: minimised again the fluctuations of $U_{\rm ref} - U$ with respect to
837: $a$ and $b$. This procedure could be continued until the values of
838: $a$ and $b$ no longer changed, but in practice we stopped after the
839: second step, and found $a=0.95$ and $b=0.12$. We did not use the
840: extra freedom in the choice of the inverse power parameters since we
841: found that this reference system already described the energetics of
842: the solid very accurately.
843:
844: The calculation of the anharmonic part of the free energy required,
845: once more, two thermodynamic integration steps. In the first step we
846: calculated the free energy difference $F_{\rm ref} - F_{\rm
847: harm}$. These are cheap calculations since they involve only the classical
848: potentials $U_{\rm IP}$ and $U_{\rm harm}$; the simulations were
849: performed with cells containing 512 atoms for 10 ps, which ensured
850: convergence of the free energy difference $F_{\rm ref} - F_{\rm harm}$
851: to within 1 meV/atom. In the second step we calculated $F_{\rm vib} -
852: F_{\rm ref}$ where, since the fluctuations in the energy differences $U - U_{\rm ref}$
853: were very
854: small, we were able to use the second order formula
855: (Eq.~\ref{eqn:secondorder}).
856:
857: %; i.e. we generated a number of
858: %statistically independent configurations using the reference potential
859: %$U_{\rm ref}$ and calculated the {\rm ab-initio} total energies of
860: %these configurations.
861: % We then used the average and the fluctuations of
862: %the energy differences to estimate the free energy difference between
863: %the two systems.
864:
865: The problem in the calculation of thermal averages for a nearly
866: harmonic system is that of ergodicity. For an harmonic system
867: different degrees of freedom do not exchange energy, so in a system
868: which is close to being harmonic the exploration of phase space using
869: molecular dynamics can be a very slow process. We solved this problem
870: following Ref.~\cite{dewijs98} whereby the statistical sampling was
871: performed using Andersen molecular dynamics~\cite{andersen80}, in
872: which the atomic velocities are periodically randomised by drawing
873: them from a Maxwellian distribution. This type of simulation generates
874: the canonical ensemble and overcomes the ergodicity problem.
875:
876: All the calculations were performed on a 64-atom cell with kpoints in
877: a $7\times 7 \times 7$ grid for the high P/T state points and a
878: $9\times 9 \times 9$ grid for the low P/T state points equivalent to
879: 172 or 365 points in the IBZ respectively.
880:
881: The anharmonic contribution to the free energy of the solid turns out
882: to be very small, being positive and equal to only a few meV/atom at
883: low pressure and approximatively -20 meV/atom at high pressure.
884:
885: \subsection{Error estimates for $F_{\rm sol}$}
886: The errors in $F_{\rm perf}$ are less than 1 meV/atom, the errors in
887: $F_{\rm harm}$ are $\approx 3(4)$ meV/atom at low(high) $P/T$ and the
888: errors in $F_{\rm ahnarm}$ are $\approx 1(4)$ meV/atom at low(high)
889: $P/T$; the total errors in $F_{\rm sol}$ are $\approx 3(6)$ meV/atom
890: at low(high) $P/T$.
891:
892: \section{Results and discussion}\label{sec:discussion}
893:
894: We display in Figure 1 our calculated melting curve compared with the
895: experimental zero pressure value~\cite{crc97}, the DAC high pressure
896: results~\cite{boehler97, hanstrom00} and the high pressure shock
897: datum~\cite{shaner84}. We also report in Figures 2a, 2b and 2c the
898: volume change on melting, $V_m$, the entropy change on melting, $S_m$,
899: and the melting gradient, $dT_m/dP$, respectively. The errors in the
900: melting curve arise from the errors in the calculated free energies
901: and are $\approx 50(100) $ K in the low(high) pressure part of the
902: diagram respectively.
903:
904: The overall agreement with the experiments is extremely good; however,
905: the low pressure results differ by more than 15 \% (at zero pressure,
906: 786 K compared with the experimental value of 933 K). Indeed, at zero
907: pressure the agreement between the calculated and experimental volume
908: change on melting and $dT_m/dP$ is rather poor (see
909: Table~\ref{tab:meltprops}). In addition, our calculations are not in
910: very good agreement with the previous calculations of de Wijs {\em et
911: al.}~\cite{dewijs98}, although this is not necessarily surprising,
912: since these latter calculations were based on LDA, while ours are
913: based on GGA. Nevertheless, one might expect the results from LDA and
914: GGA to be similar, since Al is a nearly free-electron like metal and
915: therefore one would expect a very good DFT description with both LDA
916: and GGA.
917: %This inaccuracy of the predicted zero
918: %pressure volume affects quite significantly dynamical properties like
919: %the viscosity for example: in our previous work on Al~\cite{alfe98} we
920: %showed that the viscosity of Al calculated at $T=1000$ K and the
921: %corresponding zero pressure volume is $\eta = 2.2 \pm 0.1 $ mPa~s,
922: %which should be compared with the experimental value of $1.25$ mPa~s
923: %\cite{shimoi}. In that work~\cite{alfe98} we also calculated the
924: %viscosity at the experimental zero pressure volume (which gives a
925: %negative LDA pressure of $\approx 2$ GPa) and we found $\eta = 1.4 \pm
926: %0.15 $ mPa~s, which is in excellent agreement with the experimental
927: %data.
928: To explore a possible reason why GGA does not predict the melting
929: properties of aluminium very accurately we consider the zero pressure
930: crystal equilibrium volume. This is predicted by GGA to be $\approx 2
931: \%$ larger than the experimental value; this means that the calculated
932: pressure for the experimental zero pressure volume is $\approx +1.6$
933: GPa.
934: To see how this error propagates in melting properties we may
935: devise a correction to the Helmholtz free energy such that the
936: pressure is rectified:
937: \begin{equation}
938: F_{\rm corr} = F + \delta P V,
939: \end{equation}
940: with $\delta P = 1.6$ GPa. Using $F_{\rm corr}$ in our calculations we
941: found the {\em corrected} melting curve, represented by the dotted
942: line in Fig. 1, where we assumed $\delta P$ to be the same in the
943: whole $P/T$ range. The zero pressure corrected melting temperature is
944: 912 K, which is in very good agreement with the experimental value 933
945: K. The corrected volume change on melting, entropy change on melting
946: and $dT_m/dP$ are also in much better agreement with the experimental
947: numbers. The correction is less important at high pressure, where
948: $dT_m/dP$ is smaller.
949:
950: This point may be further illustrated by looking at the zero pressure
951: phonon dispersion curves for Al. Since phonon frequencies depend on
952: the interatomic forces, their correctness is surely important in the
953: context of melting. In Figure 3 we display the GGA calculated phonon
954: dispersion curves compared with experimental
955: data~\cite{stedman66}. Our calculations were performed both at the GGA
956: zero pressure equilibrium volume and the experimental volume (both
957: at 80 K). We notice that the agreement is good (though not perfect)
958: if the calculations are performed at the experimental volume, and
959: rather poor if the calculated zero pressure GGA volume is used
960: instead. This indicates that GGA will probably yield better results
961: if the GGA pressure is corrected in order to match the experimental
962: data.
963:
964: In their work, de Wijs {\em et al.}~\cite{dewijs98} found good
965: agreement between LDA and experiments. In their case a {\em corrected}
966: LDA would lower the zero pressure melting point below 800 K. In order
967: to understand this apparent different behaviour between LDA and GGA we
968: have also calculated the phonons using LDA at the calculated
969: equilibrium volume and also at the experimental volume (both at
970: 80K). These are also reported in Fig. 3. In accord with previous LDA
971: calculations~\cite{degironcoli95} we found very good agreement with
972: the experiments when the phonons are calculated at the LDA zero
973: pressure volume, but the agreement becomes poor at the experimental
974: volume, which is consistent with the result for the melting
975: temperature~\cite{dewijs98}.
976:
977: In conclusion, both GGA and LDA predict an incorrect equilibrium volume
978: at a fixed pressure, although LDA yields very good results for both
979: the phonon dispersion curves and the zero pressure melting properties
980: (which is probably accidental). For GGA the incorrect equilibrium volume
981: propagates to an incorrect description of the phonon frequencies and
982: the melting properties. If the GGA pressures are corrected so as to
983: match the experimental data, the phonon dispersion and the melting
984: properties come out in very good agreement with the experiments. These
985: two behaviours are internally consistent, but point to an intrinsic
986: error due to the use of GGA. Quantum Monte-Carlo (QMC)
987: techniques~\cite{hammond94} have been shown to predict the energetics
988: with much higher accuracy than DFT~\cite{rajagopal95}, and
989: calculations for systems containing more than 100 atoms have already
990: been reported~\cite{kent99}. We believe that in the near future it
991: will be possible to use QMC for more accurate calculations of free
992: energies.
993:
994: To summarise, we have calculated the melting curve of aluminium
995: entirely from first-principles within the DFT-GGA framework. Our work
996: is based on the calculation of the Gibbs free energy of liquid and
997: solid Al, and for each fixed pressure the melting temperature is
998: determined by the point at which the two free energies cross. Our
999: results are in good agreement with the available experimental data,
1000: although they reveal an intrinsic DFT-GGA error which is responsible
1001: for an error of $\approx 150 $K in the low pressure melting
1002: curve. This error is probably due to the incorrectly predicted
1003: pressure by GGA, and it becomes less important in the high
1004: pressure region, as $dT_m/dP$ becomes smaller.
1005:
1006: %Ab initio calculations can now provide quantitative static,
1007: %thermodynamic and dynamic properties of high $P/T$ phases with a
1008: %precision comparable with experiments. We hope to use these techniques
1009: %to constrain the true composition and thermal structure of the Earth's
1010: %core by investigating iron alloy phases, predicting Tm curves and
1011: %thermodynamic behaviour~\cite{vocadlo00,alfe99c,alfe00b}.
1012:
1013: %V0=16.562 K=73 cf 76 expt
1014:
1015: %Figure 1 shows the Aluminium melting curve compared with DAC experiments and
1016: %shock data. The agreement is excellent.
1017:
1018:
1019: \section*{Acknowledgements}
1020:
1021: We both acknowledge the support of Royal Society University Research
1022: Fellowships; we also thank Mike Gillan and John Brodholt for useful discussions. LV thanks Humphrey Vocadlo for his assistance during the course of this research.
1023:
1024: \begin{thebibliography}{99}
1025:
1026: \bibitem{williams87} Q. Williams, R. Jeanloz, J. D. Bass,
1027: B. Svendesen, T. J. Ahrens, Science {\bf 286}, 181 (1987).
1028:
1029: \bibitem{boehler93} R. Boehler, Nature {\bf 363}, 534 (1993).
1030:
1031: \bibitem{shen98} G. Shen, H. Mao, R. J. Hemley, T. S. Duffy and
1032: M. L. Rivers, Geophys. Res. Lett. {\bf 25}, 373 (1998).
1033:
1034:
1035:
1036: \bibitem{brown86} J. M. Brown and R. G. McQueen, J. Geophys. Res. {\bf
1037: 91}, 7485 (1986).
1038:
1039: \bibitem{yoo93} C. S. Yoo, N. C. Holmes, M. Ross, D. J. Webb and
1040: C. Pike, Phys. Rev. Lett. {\bf 70}, 3931 (1993).
1041:
1042: \bibitem{alfe99} D. Alf\`e, M. J. Gillan and G. D. Price, Nature {\bf
1043: 401}, 462 (1999).
1044:
1045: \bibitem{alfe01a} D. Alf\`e, G. D. Price, M. J. Gillan, Phys. Rev. B,
1046: in press.
1047:
1048: \bibitem{alfe01b} D. Alf\`e, G. D. Price, M. J. Gillan, unpublished.
1049:
1050: \bibitem{laio00} A. Laio, S. Bernard, G. L. Chiarotti, S. Scandolo and
1051: E. Tosatti, Science {\bf 287}, 1027 (2000).
1052:
1053: \bibitem{belonoshko00} A. B. Belonoshko, R. Ahuja, and B. Johansson,
1054: Phys. Rev. Lett. {\bf 84}, 3638 (2000).
1055:
1056: \bibitem{crc97} {\it CRC Handbook of Chemistry and Physics}, Editor
1057: D. R. Lide, New York, 77th edition (1996-1997).
1058:
1059: \bibitem{boehler97} R. Boehler, M. Ross, Earth Planet. Sci. Lett. {\bf
1060: 153}, 223 (1997)
1061:
1062: \bibitem{hanstrom00} A. H\"{a}nstr\"{o}m, P. Lazor, J. of Alloys and
1063: Compounds {\bf 305}, 209 (2000).
1064:
1065: \bibitem{shaner84} J. W. Shaner, J. M. Brown, R. G. McQueen, in {\rm
1066: High pressure in Science and Technology}. Editors C. Homan, R. K. Mac
1067: Crone, E. Whalley, North Holland, Amsterdam, 137 (1984).
1068:
1069: \bibitem{moriarty84} J. A. Moriarty, D. A. Young, and M. Ross,
1070: Phys. Rev. B {\bf 30}, 578 (1984).
1071:
1072: %\bibitem{pelisier84} J. L. P\'elisier, Physica A {\bf 128}, 363 (1984).
1073:
1074: \bibitem{mei92} J. Mei, J. W. Davenport, Phys. Rev. B {\bf 46}, 21
1075: (1992).
1076:
1077: \bibitem{morris94} J. R. Morris, C. Z. Wang, K. M. Ho, and C. T. Chan,
1078: Phys. Rev. B {\bf 49}, 3109 (1994).
1079:
1080: \bibitem{straub94} G. K. Straub, J. B. Aidun, J. M. Willis,
1081: C. R. Sanchez-Castro, and D. C. Wallace, Phys. Rev. B {\bf 50}, 5055
1082: (1994).
1083:
1084: \bibitem{dewijs98} G. A. de Wijs, G. Kresse and M. J. Gillan,
1085: Phys. Rev. B {\bf 57}, 8223 (1998).
1086:
1087: \bibitem{generaldft} P. Hohenberg and W. Kohn, Phys. Rev. {\bf 136},
1088: B864 (1964); W. Kohn and L. Sham, Phys. Rev. {\bf 140}, A1133 (1965);
1089: R. O. Jones and O. Gunnarsson, Rev. Mod. Phys. {\bf 61}, 689 (1989);
1090: M. J. Gillan, Contemp. Phys. {\bf 38}, 115 (1997).
1091:
1092: \bibitem{frenkel96} For a discussion of thermodynamic integration see
1093: e.g. D. Frenkel and B. Smit, {\em Understanding Molecular
1094: Simulation}, Academic Press, San Diego (1996).
1095:
1096: \bibitem{jesson2000} B. J. Jesson and P. A. Madden, J. of
1097: Chem. Phy. {\bf 113} 5924 (2000).
1098:
1099: \bibitem{wang91} Y. Wang and J. Perdew, Phys. Rev. B {\bf 44}, 13298
1100: (1991).
1101:
1102: \bibitem{perdew92} J. P. Perdew, J. A. Chevary, S. H. Vosko,
1103: K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais,
1104: Phys. Rev. B {\bf 46}, 6671 (1992).
1105:
1106: \bibitem{mermin65} N. D. Mermin, Phys. Rev. {\bf 137}, A1441 (1965).
1107:
1108: \bibitem{gillan89} M. J. Gillan, J. Phys. Condens. Matter {\bf 1}, 689
1109: (1989).
1110:
1111: \bibitem{wentzcovitch92} R. M. Wentzcovitch, J. L. Martins and
1112: P. B. Allen, Phys. Rev. B {\bf 45}, 11372 (1992).
1113:
1114: \bibitem{kresse96a} G. Kresse and J. Furthm\"{u}ller, Phys. Rev. B
1115: {\bf 54}, 11169 (1996); a discussion of the ultra-soft
1116: pseudopotentials used in the VASP code is given in G. Kresse and
1117: J. Hafner, J. Phys. Condens. Matter {\bf 6}, 8245 (1994).
1118:
1119: \bibitem{vanderbilt90} D. Vanderbilt, Phys. Rev. B {\bf 41}, 7892
1120: (1990).
1121:
1122: \bibitem{monkhorst76} H. J. Monkhorst and J. D. Pack, Phys. Rev. B
1123: {\bf 13}, 5188 (1976).
1124:
1125: \bibitem{alfe99b} D. Alf\`{e}, Comp. Phys. Commun. {\bf 118}, 31
1126: (1999).
1127:
1128: \bibitem{sugino95} O. Sugino and R. Car, Phys. Rev. Lett. {\bf 74},
1129: 1823 (1995).
1130:
1131: \bibitem{smargiassi95a} E. Smargiassi, P. A. Madden, Phys Rev B {\bf
1132: 51}, 117 (1995).
1133:
1134: \bibitem{alfe00} D. Alf\`{e}, G. A. de Wijs, G. Kresse and
1135: M. J. Gillan, Int. J. Quant. Chem. {\bf 77}, 871 (2000).
1136:
1137: \bibitem{nose84} S. Nos\'e, Molec. Phys., {\bf 52} 255 (1984);
1138: J. Chem. Phys. {\bf 81}, 511 (1984).
1139:
1140: \bibitem{ditolla93} F. D. Di Tolla and M. Ronchetti, Phys. Rev. B {\bf
1141: 48}, 1726 (1993).
1142:
1143: \bibitem{watanabe90} M. Watanabe and W. P. Reinhardt,
1144: Phys. Rev. Lett. {\bf 65}, 3301 (1990).
1145:
1146: \bibitem{daw93} M. S. Daw, S. M. Foiles, and M. I. Baskes,
1147: Mat. Sci. Rep. {\bf 9}, 251 (1993).
1148:
1149: \bibitem{baskes92} M. I. Baskes, Phys. Rev. B {\bf 46}, 2727 (1992).
1150:
1151: \bibitem{belonoshko97} A. B. Belonoshko, and R. Ahuja, Phys. Earth
1152: Planet. Inter. {\bf 102}, 171 (1997).
1153:
1154: \bibitem{laird92} B. B. Laird and A. D. J. Haymet, Mol. Phys. {\bf
1155: 75}, 71 (1992).
1156:
1157: \bibitem{johnson93} K. Johnson, J. A. Zollweg, and E. Gubbins,
1158: Mol. Phy. {\bf 78}, 591 (1993).
1159:
1160: \bibitem{darioweb} Program available at {\tt
1161: http://chianti.geol.ucl.ac.uk/\~dario}
1162:
1163: \bibitem{kresse95} G. Kresse, J. Furthm\"{u}ller and J. Hafner,
1164: Europhys. Lett. {\bf 32} 729 (1995).
1165:
1166: \bibitem{andersen80} H. C. Andersen, J. Chem. Phys. {\bf 72}, 2384
1167: (1980).
1168:
1169: \bibitem{stedman66} R. Stedman, and G. Nilsson, Phys. Rev. {\bf 145},
1170: 492 (1966).
1171:
1172: %\bibitem{shimoi} M. Shimoji and T. Itami, {\it Atomic Transport in
1173: %Liquid Metals}, Trans Tech Publications, Aedermannsdorf (1986).
1174:
1175: %\bibitem{alfe98}
1176: %D. Alf\`{e} and M. J. Gillan, Phys. Rev. Lett., {\bf 81}, 5161 (1998).
1177:
1178: \bibitem{degironcoli95} S. de Gironcoli, Phys. Rev. B. {\bf 51}, 6773
1179: (1995)
1180:
1181: \bibitem{hammond94} B. L. Hammond, W. A. Lester, Jr., \&
1182: P. J. Reynolds, {\em Monte Carlo Methods in Ab Initio Quantum
1183: Chemistry}, World Scientific, Singapore, (1994).
1184:
1185: \bibitem{rajagopal95} G. Rajagopal, R. J. Needs, A. James,
1186: S. D. Kenny, \& W. M. C. Foulkes, Phys. Rev. B {\bf 51}, 10591 (1995).
1187:
1188: \bibitem{kent99} P. R. C. Kent, R. Q. Hood, A. J. Williamson,
1189: R. J. Needs, W. M. C Foulkes, G. \& Rajagopal, Phys. Rev. B {\bf 59},
1190: 1917 (1999).
1191:
1192: \bibitem{cannon74} J. F. Cannon, J. Phys. Chem. Ref. Data 3, 781 (1974).
1193: \bibitem{chase85} M. W. Chase, Jr., C. A. Davies, J. R. Downey, Jr., D. J. Frurip, R. A. McDonald and A. N. Syverud, J. Phys. Chem. Ref. Data Suppl. 14, 1 (1985).
1194: \end{thebibliography}
1195:
1196: \pagebreak
1197: \newpage
1198:
1199:
1200: \begin{table}
1201: \begin{tabular}{l|ccccc}
1202: % & \multicolumn{4}{c}{$K_S$ (GPa) }\\
1203: % & \multicolumn{2}{c}{This work} & \multicolumn{2}{c}{Experiments} \\
1204: %\hline
1205: & Experiment & LDA & GGA & GGA - corrected \\
1206: \hline
1207: $T_m$ (K)& 933 & 890 (20) & 786 (50) & 912 (50) \\
1208: $S_m$ ($k_{\rm B}$)& 1.38 & 1.36 (4) & 1.35 (6) & 1.37 (6) \\
1209: $V_m$ (\AA$^3$)& 1.24 & 1.26 (20) & 1.51 (10) & 1.35 (10) \\
1210: $dT_m/dP$ (K~GPa$^{-1}$) & 65 & 67 (12) & 81 & 71 \\
1211: \end{tabular}
1212: \caption{Comparison of {\em ab initio} and experimental melting
1213: properties of Al at zero pressure. Values are given for the melting
1214: temperature, $T_m$, entropy change on melting, $S_m$, volume change on
1215: melting, $V_m$, and melting gradient $dT_m/dP$. The LDA results are
1216: from Ref.~\protect\cite{dewijs98}; the experimental values for $T_m$, $S_m$ and $dT_m/dP$ are from Refs. ~\cite{crc97}, ~\cite{chase85} and ~\cite{cannon74} respectively, and the experimental melting volume, $V_m$, is calculated using the Clapeyron relation, $V_m = S_mdT_m/dP$.}
1217: \label{tab:meltprops}
1218: \end{table}
1219:
1220: \pagebreak
1221: \newpage
1222:
1223:
1224: \begin{figure}
1225: \psfig{figure=Tm.ps,height=6in}
1226: \caption{Comparison of melting curve of Al from present calculations
1227: with previous experimental results. Solid curve: present work; dotted
1228: curve: present work with pressure correction (see text); diamonds and
1229: triangles: DAC measurements of Refs.~\protect\cite{boehler97} and
1230: ~\protect\cite{hanstrom00} respectively; square: shock experiments of
1231: Ref.~\cite{shaner84}.}
1232: \label{fig:melting_curve}
1233: \end{figure}
1234:
1235: \begin{figure}
1236: \psfig{figure=dVdSdTm.ps,height=6in}
1237: \caption{Calculated pressure dependence of the melting properties of
1238: Al: a) volume change on melting, b) entropy change on melting and c)
1239: melting gradient. Solid curve: present work; dotted curve: present
1240: work with pressure correction (see text).}
1241: \label{fig:melting_props}
1242: \end{figure}
1243:
1244: \begin{figure}
1245: \psfig{figure=phon2.ps,height=6in}
1246: \caption{Comparison of the phonon dispersion curve for Al from present
1247: calculations with previous experimental results. Solid curves: present
1248: work with GGA; dotted curves: present work with GGA and with pressure
1249: correction (see text); dashed curves: present work with LDA;
1250: dot-dashed curves: present work with LDA and with pressure correction
1251: (see text); diamonds: experiments from Ref.~\cite{stedman66}.}
1252: \label{fig:phonon_curve}
1253: \end{figure}
1254:
1255: \end{document}
1256: