1: \documentstyle[aps,prb,epsf]{revtex}
2: %\documentstyle[twocolumn,aps,prb]{revtex}
3:
4: %\documentstyle[aps]{revtex}
5:
6: %\documentstyle[aps,prb,preprint]{revtex}
7:
8: \def\batio3{BaTiO$_3$}
9:
10: \begin{document}
11:
12: %%%%%% PP Comment out next two lines for preprint
13: \twocolumn[\hsize\textwidth\columnwidth\hsize\csname
14: @twocolumnfalse\endcsname
15:
16: \title{Devonshire-Landau free energy of BaTiO$_3$ from
17: first principles}
18:
19: \author{Jorge \'I\~niguez$^{1}$\cite{byline}, S. Ivantchev$^2$,
20: J. M. Perez-Mato$^2$, and Alberto Garc\'{\i}a$^1$}
21:
22: \address{$^1$ Departamento de F\'{\i}sica Aplicada II, Universidad del
23: Pa\'{\i}s Vasco, Apdo 644, 48080 Bilbao, Spain\\ $^2$ Departamento de
24: F\'{\i}sica de la Materia Condensada, Universidad del Pa\'{\i}s Vasco,
25: Apdo 644, 48080 Bilbao, Spain}
26:
27: \maketitle
28:
29: \begin{abstract}
30: We have studied the Devonshire-Landau potential underlying the phase
31: transition sequence of BaTiO$_3$ using the first-principles effective
32: Hamiltonian of Zhong, Vanderbilt, and Rabe {[}{\it Phys. Rev. Lett.}
33: {\bf 73}, 1861 (1994){]}, which has been very successful in
34: reproducing the phase transitions and the dielectric and piezoelectric
35: properties of this compound. The configuration space (determined by
36: the polarization {\bf P} as order parameter) was explored with the
37: help of auxiliary electric fields. We show that the typically assumed
38: form of the potential, a sixth-order expansion in {\bf P} around the
39: paraelectric cubic phase, properly accounts for the behavior of the
40: system, but we find a non-trivial temperature dependence for all the
41: coefficients in the expansion, including the quadratic one, which is
42: shown to behave non-linearly. Our results also prove that the
43: sixth-order terms in the free-energy expansion (needed to account for
44: the first-order character of the transitions and the occurrence of an
45: orthorhombic phase) emerge from an interaction model that only
46: includes terms up to fourth order.
47: \end{abstract}
48:
49: %%PP%% Comment out these two lines for preprint...
50: \vskip1pc
51: ]
52:
53: \narrowtext \marginparwidth 2.7in \marginparsep 0.5in
54:
55: \section{INTRODUCTION}
56:
57: %%% Brief summary of Landau theory
58:
59: From a phenomenological point of view, the behavior of a system in the
60: vicinity of a phase transition can be described in the framework of
61: Landau theory.~\cite{landau} In this scheme, one begins by identifying
62: the so-called order parameter, a (in general multidimensional)
63: variable ${\bf Q}$ that characterizes the symmetry change in the
64: transition, and then constructs the Landau free-energy function
65: $F({\bf Q};T)$, with the property that the equilibrium value of ${\bf
66: Q}$ as a function of temperature is that which minimizes
67: $F$. Formally, the Landau free energy is an incomplete thermodynamic
68: potential, which in principle could be calculated as:
69: %
70: \begin{equation}
71: F({\bf Q};T) = -k_BT\ln \sum_{j/{\bf Q}}{e^{-U_j/k_BT}} \; ,
72: \label{eq:incomplete}
73: \end{equation}
74: %
75: in analogy with the procedure to obtain the standard free energy in
76: terms of the partition function. Here the sum is only over those
77: states with the given value of ${\bf Q}$. In a Landau phenomenological
78: treatment, one just assumes that $F({\bf Q};T)$ exists, and that it
79: can be represented as a simple power series in ${\bf Q}$ in the
80: vicinity of the transition:
81: %
82: % Adapt Eq. 136.7 in Statistical Physics
83: %
84: \begin{equation}
85: F({\bf Q};T) = F_0(T) + A(T)Q^2 +
86: B\sum_{\beta}{f_{\beta}^{(4)}}({\bf Q})+
87: \dots \; ,
88: \label{eq:landau}
89: \end{equation}
90: %
91: where $f_{\beta}^{(j)}$ are invariants of order $j$ constructed from
92: ${\bf Q}$. The quadratic coefficient $A$ is assumed to vary linearly
93: with temperature in the form $A(T)=\alpha(T-T_0)$, in such a way that
94: $T_0$ is the transition temperature (for a second-order transition) or
95: the lower metastability limit of the upper phase (for a first-order
96: one). These assumptions are known to break down in the critical region
97: close to the transition point of continuous phase transformations, but
98: they are valid when describing the approximate behavior of the system
99: in wider temperature intervals around the transition
100: temperature. Classical Landau theory ignores any temperature variation
101: of the higher-order coefficients in the expansion. Nevertheless, it
102: has been applied quite successfully to the analysis of many materials
103: in wide temperature ranges.~\cite{radescu} Nowadays Landau theory is
104: the most common approach to study the phenomenology of any
105: symmetry-breaking structural phase transition.
106: %
107:
108: One of the early successes of this phenomenological approach to phase
109: transitions was the description by Devonshire~\cite{devonshire} of the
110: sequence of transitions in barium titanate (\batio3), a sequence that
111: spans a temperature interval of more than 200 degrees. This material
112: exhibits at high temperatures a paraelectric cubic perovskite
113: structure and, as temperature decreases, it undergoes three successive
114: first-order transitions to ferroelectric phases with tetragonal,
115: orthorhombic, and rhombohedral symmetries. In these phases the
116: polarization ${\bf P}$ points along one of the $\langle 1,0,0\rangle$,
117: $\langle 1,1,0\rangle$, and $\langle 1,1,1\rangle$ cubic directions
118: respectively.~\cite{jona} The polarization ${\bf P}$ can be identified
119: as the (three-dimensional) order parameter, and the Devonshire-Landau
120: potential (per unit volume) written as:
121: %
122: \begin{equation}
123: \label{eq-cowley}
124: \begin{array}{lll}
125: F & = & F_0+\frac{1}{2}aP^2 + u P ^4 + v (P_x^4+P_y^4+P_z^4)\\
126: & & + h_1 P^6 + h_2 (P_x^6+P_y^6+P_z^6)\\
127: & & + h_3 [P_x^4(P_y ^2+P_z^2)+P_y^4(P_z^2+P_x^2)\\
128: & & + P_z^4(P_x^2+P_y^2)] \; ,
129: \end{array}
130: \end{equation}
131: %
132: where $F_0$ is the free energy of the reference cubic phase and $P^2$ stands
133: for $P_x^2+P_y^2+P_z^2$ (the notation is taken from the review paper by
134: Cowley~\cite{cowley}). This is the complete expansion of $F$ up to
135: sixth order, and its relatively simple form is due to the high symmetry of the
136: system. Sixth-order terms in ${\bf P}$ in the expansion are needed to account
137: for the first-order character of the transitions (to model non-equivalent
138: coexisting free-energy minima). Moreover, it can be seen that one needs
139: anisotropic sixth-order terms in $F$ to account for the orthorhombic phase of
140: \batio3.~\cite{vander-cohen} Devonshire showed that by assuming a linear
141: temperature dependence for $a$ and suitable {\sl constant} values for the rest
142: of the coefficients it is possible to qualitatively reproduce the transition
143: sequence as well as the dielectric properties of the system.~\cite{devon-more}
144:
145: %% Experimental facts against Devonshire's theory
146:
147: After Devonshire's work, the use of an expansion like that of
148: Eq.~\ref{eq-cowley} has become the standard way to model the
149: thermodynamic properties of ferroelectric perovskites, even though the
150: correctness of the assumption of constant high-order coefficients was
151: questioned early on. Drougard {\it et al.\/} and Huibregtse {\it et
152: al.\/}~\cite{exp} studied BaTiO$_3$ at temperatures just above the
153: cubic to tetragonal transition and in the orthorhombic phase
154: respectively; they assumed the existence of a Devonshire-Landau
155: potential but showed that there should be a significant temperature
156: dependence of the fourth-order terms.~\cite{gonzalo} It has since been
157: shown that there is a relatively large latitude to produce
158: Devonshire-like models that lead to qualitatively sensible predictions
159: for the transition sequence, divergence of the dielectric constants,
160: etc., even though these models may not be {\sl realistic} when confronted
161: with other experimental evidence.~\cite{other-forms} On the other
162: hand, it could be questioned whether such a relatively simple
163: free-energy expansion as that in Eq.~\ref{eq-cowley} exists at
164: all. For instance, Nakamura and Kinase have worked on a different
165: approach to the problem in which a non-polynomical from of the free
166: energy is used; they also discuss the connection of their work to the
167: Devonshire theory.~\cite{nakamura}
168:
169: In the past decade, first-principles methods based on
170: density-functional theory have demonstrated to be accurate enough to
171: reproduce and predict structural phase transitions in perovskite
172: oxides.~\cite{vanderbilt} In particular, the phase transition sequence
173: and other properties of BaTiO$_3$,~\cite{zhong,garcia}
174: PbTiO$_3$,~\cite{waghmare-rabe} and KNbO$_3$~\cite{krakauer} have been
175: studied using the so called {\sl effective Hamiltonian} approach. An
176: ab-initio effective Hamiltonian $H_{\rm eff}$ is a mechanical model
177: that includes the relevant microscopic degrees of freedom of the
178: system and is constructed on the basis of first-principles
179: calculations. This model can then be analyzed by statistical methods
180: (typically, Monte Carlo or molecular dynamics) to explore the
181: finite-temperature behavior of the system. Monte-Carlo simulations
182: with an ab-initio effective Hamiltonian for \batio3\ have been
183: extremely successful, replicating approximately the experimental
184: transition sequence~\cite{zhong} and succeeding in reproducing the
185: main features of the dielectric and piezoelectric properties of the
186: real system.~\cite{garcia}
187:
188: A natural question to ask then is to what extent the thermal behavior
189: resulting from these ab-initio effective Hamiltonians is compatible
190: with the Devonshire-Landau framework. In this paper we show that a
191: Devonshire-Landau expansion of the form of Eq.~\ref{eq-cowley} does
192: indeed emerge from detailed Monte Carlo simulations with the ab-initio
193: effective Hamiltonian for \batio3 of Ref.~\onlinecite{zhong}, with the
194: important qualification that the coefficients in the expansion have a
195: non-trivial temperature dependence. In particular, we show that the
196: sixth-order terms in $F({\bf P};T)$ are non-zero only in the
197: temperature range in which the transitions occur. Moreover, it becomes
198: clear that these sixth-order coefficients appear as
199: products of the statistical fluctuations of a Hamiltonian that only
200: includes terms up to fourth order in the polar degrees of freedom.
201: %
202:
203: \section{METHOD AND TECHNICAL DETAILS}
204:
205: %%% The effective Hamiltonian for barium titanate
206:
207: Zhong, Vanderbilt, and Rabe~\cite{zhong} constructed their effective
208: Hamiltonian for BaTiO$_3$ retaining as relevant degrees of freedom of the
209: system a local polar distortion ${\bf u}_{i}$ in each cell, the homogeneous
210: strain $\eta$, and an inhomogeneous strain represented by a second set of
211: local vectors ${\bf v}_{i}$. The $H_{\rm eff}$ has the form:
212: %
213: \begin{equation}
214: \label{eq-zhong}
215: \begin{array}{lll}
216: H_{\rm eff} & = & \sum_{i,j;\alpha,\beta} J(i,j;\alpha,\beta)
217: u_{i\alpha} u_{j\beta} \\ & & + \sum_{i;\alpha,\beta}
218: \Gamma(\alpha,\beta) u_{i\alpha}^2 u_{i\beta}^2 \\ & & +
219: \sum_{i;\alpha,\beta;l} B(i;\alpha,\beta;l) u_{i\alpha} u_{i\beta}
220: \eta_{l} \\ & & + \sum_{l,h} B(l,h) \eta_{l} \eta_{h},
221: \end{array}
222: \end{equation}
223: %
224: where, for clarity, we have not written the terms associated to the
225: inhomogeneous strain. Here $i$ and $j$ range over the cells in the
226: system, $\alpha$ and $\beta$ are cartesian indexes, and $l$ and $h$
227: refer to tensor components in Voigt notation. For the local polar
228: modes, the $H_{\rm eff}$ contains harmonic couplings
229: $J(i,j;\alpha,\beta)$ (on-site, $i=j$, and between modes in different
230: cells up to the third nearest neighbors) that reproduce the
231: instabilities of the cubic phase of \batio3,~\cite{uns-band} and
232: fourth-order on-site terms $\Gamma(\alpha,\beta)$ that define the
233: low-symmetry minima and stabilize the system. The effect of strain is
234: included through the standard elastic energy and through the coupling
235: coefficients $B(i;\alpha,\beta;l)$, which account for the spontaneous
236: strain in the low-symmetry phases. This model is probably the simplest
237: one that captures the essential physics of the system. As mentioned in
238: the introduction, $H_{\rm eff}$ is fourth-order in the polar
239: variables.
240:
241: %%% Simple connections between the effective Hamiltonian and
242: %%% Devonshire's theory.
243:
244: A direct link between this $H_{\rm eff}$ and the phenomenological
245: potential of Eq.~\ref{eq-cowley} can be established in the very
246: low-temperature limit $T\rightarrow$ 0. Eq.~\ref{eq:incomplete} shows
247: that $F({\bf P};0)$ is just the energy of the lowest-lying state with
248: polarization ${\bf P}$. For \batio3 this state exhibits homogeneously
249: polarized cells with the global strain adjusted so as to minimize the
250: energy (the inhomogeneous strain is zero). It can be
251: shown~\cite{king-smith} that for a given ${\bf P}$, the homogenous
252: strain is proportional to the square of the polarization, so $F({\bf
253: P};0)$ is fourth-order in the polarization. This means that
254: $h_1=h_2=h_3=0$ in the low-temperature limit. Using the $H_{\rm eff}$
255: parameters, we obtain the rest of coefficients in $F$ at 0~K:
256: $a=-4/3$, $u=0.094$, and $v=0.051$. Here, units have been chosen so
257: that the first-octant rhombohedral ground state minimum is located at
258: ${\bf P}=(1,1,1)$ and its energy per unit volume with respect to the
259: cubic phase is $-1$. (This election fixes the value of $a$ and the
260: relation $9u+3v=1$; i.e., $F({\bf P};0)$ is determined only by one
261: coefficient.) We have plotted this 0~K free-energy map in
262: Fig.~\ref{fig-maps}h), which shows the cubic phase as a local
263: free-energy maximum, tetragonal and orthorhombic saddle points (both
264: stable in the radial direction), and the rhombohedral global minima.
265: %
266: \begin{figure}
267: \epsfxsize=\hsize\epsfbox{fig1_a.eps}
268: \epsfxsize=\hsize\epsfbox{fig1_b.eps}
269: \epsfxsize=\hsize\epsfbox{fig1_c.eps}
270: \epsfxsize=\hsize\epsfbox{fig1_d.eps}
271: \epsfxsize=\hsize\epsfbox{fig1_e.eps}
272: \epsfxsize=\hsize\epsfbox{fig1_f.eps}
273: \epsfxsize=\hsize\epsfbox{fig1_g.eps}
274: \epsfxsize=\hsize\epsfbox{fig1_h.eps}
275: \caption{Free-energy maps of BaTiO$_3$ at several temperatures. Panel
276: a): 300~K, b): 290~K, c): 250~K, d): 220~K, e): 210~K, f): 190~K, g):
277: 100~K, and h): 0~K. For each temperature we show the (010) and
278: (1$\bar{1}$0) planes of the order-parameter configuration space, with
279: contour lines depicted only in the low free-energy regions. Symbols
280: characterize the relative stability of the critical points with
281: respect to displacements within a given plane: circles for free-energy
282: minima (a bigger circle corresponds to the stable-equilibrium state),
283: squares for free-energy maxima, and triangles for saddle points
284: (``up'' (resp. ``down'') triangles for points that are maxima
285: (resp. minima) along the radial direction). Note that critical points
286: with orthorhombic symmetry may need to be represented by two different
287: symbols in the $(010)$ and $(1\bar{1}0)$ planes (see panel b)).}
288: \label{fig-maps}
289: \end{figure}
290:
291: Now we tackle the calculation of $F$ at finite temperatures. A
292: straightforward approach to obtain information about the Landau
293: potential from the results of Monte Carlo~\cite{mc} simulations was
294: presented by Radescu {\sl et al.}~\cite{radescu} in the context of a
295: three dimensional $\Phi^4$ model. From Eq.~\ref{eq:incomplete} it is
296: easy to show that there exists a relation between the probability
297: distribution of the order parameter ${\cal P}({\bf Q};T)$ and the
298: corresponding Landau potential $F({\bf Q};T)$:
299: %
300: \begin{equation}
301: \label{eq-radescu}
302: F({\bf Q};T) - F_{eq}(T) = -k_B T ln [{\cal P}({\bf Q};T)],
303: \end{equation}
304: %
305: where $F_{eq}(T)$ is the equilibrium free energy of the system at the
306: temperature considered. This procedure, which can be quite efficiently
307: applied to relatively simple cases such as the $\Phi^4$ model
308: (one-dimensional order parameter, a single second-order phase
309: transition), is not well suited for our problem. We have to study a
310: three-dimensional free-energy map with coexisting non-equivalent
311: minima. A typical Monte Carlo simulation gives information only about
312: one free-energy minimum (usually, the one corresponding to the
313: stable-equilibrium state of the system), with a poor sampling of the
314: order-parameter probability-distribution function in other regions. On
315: the other hand, more sophisticated sampling strategies would be very
316: demanding from the computational point of view.
317:
318: This problem can be overcome, and $F({\bf P};T)$ computed in an
319: efficient manner, by modifying the effective Hamiltonian so as to
320: include the effect of an external electric field ${\bf
321: E}$:~\cite{garcia}
322: %
323: \begin{equation}
324: H'_{\rm eff} = H_{\rm eff} - {\bf E} \cdot Z^* \sum_i {\bf u}_i\; ,
325: \label{eq:hprime}
326: \end{equation}
327: %
328: where $Z^*$ is the Born effective charge associated to the local polar
329: distortions. As the quantity multiplying the electric field is just
330: the total dipole moment of the system, the new Landau potential is
331: simply $F'({\bf P};T,{\bf E}) = F({\bf P};T) - {\bf EP}$. An applied
332: electric field changes the location (and possibly the symmetry) of the
333: stable-equilibrium state. The new location can be determined by a MC
334: simulation of the modified effective Hamiltonian (${\bf
335: P}_{eq}=\langle {\bf P}\rangle$) and also computed
336: directly from $F'$. The idea can be mathematically expressed as:
337: %
338: \begin{equation}
339: \label{eq-def}
340: \left[\frac{\partial F'({\bf P};T,{\bf E})}{\partial {\bf P}}\right]_{{\bf
341: P}_{eq}(T,{\bf E})} = {\bf 0}\; .
342: \end{equation}
343: %
344: The left hand side of this equation depends linearly on the
345: coefficients of $F$ and on the electric field. Thus, for a given
346: temperature we can consider several electric fields, perform MC runs
347: to obtain the equilibrium values ${\bf P}_{eq}$,~\cite{tech1} and then
348: find the best solution of an overdetermined set of linear equations of
349: the form of Eq.~\ref{eq-def} to get the coefficients of the expansion
350: of $F$.
351: %
352:
353: For each temperature, we first performed a zero-field calculation to
354: obtain a snapshot of the stable-equilibrium configuration. We then
355: used this configuration as the starting point for the runs with
356: electric field applied. In order to illustrate the kind of information
357: we were pursuing, consider $T$ such that the stable-equilibrium state
358: of the system is orthorhombic with ${\bf P}_{eq} = P_{eq} (1,1,0)$. In
359: this case, we would first use fields of the form $(E,E,0)$ to obtain
360: information about the response of the system within the orthorhombic
361: phase. Fields of the forms $(E,0,0)$, $(0,0,E)$, and $(E,E,E)$ would
362: probe the relative stability of minima with different
363: symmetries. Finally, general fields leading to $P_{x,eq} \ne P_{y,eq}
364: \ne P_{z,eq}\ne P_{x,eq}$, would explore in detail the intermediate
365: regions of the configuration space. In our Monte Carlo simulations we
366: used a $14\times 14\times 14$ supercell (which corresponds to 13720
367: atoms) with periodic boundary conditions; we typically performed 15000
368: MC sweeps for the thermalization of the system and 35000 more to
369: calculate the averages of the polarization (these are very
370: well-converged calculation conditions, as can be checked in
371: Refs.~\onlinecite{zhong,garcia,waghmare-rabe}). For each temperature
372: we considered around one hundred different electric fields and
373: constructed a largely overdetermined system of equations for the
374: coefficients in $F$, which could then be reliably fitted.
375:
376: \section{RESULTS AND DISCUSSION}
377:
378: The data from our Monte Carlo simulations can indeed be represented by
379: a Landau free energy in the form of Eq.~\ref{eq-cowley} in a wide
380: temperature interval that includes all the transitions. (The
381: transition temperatures predicted by the effective Hamiltonian are
382: $T_{c1}=297$~K (cubic to tetragonal), $T_{c2}=230$~K (tetragonal to
383: orthorhombic) and $T_{c3}=200$~K (orthorhombic to rhombohedral)
384: respectively.~\cite{temps}) Our results show that the coefficients in
385: the expansion of $F$ have a significant and non-trivial temperature
386: dependence.
387:
388: Before discussing in detail the behavior of the coefficients, it is
389: helpful to present the shape of the free energy they determine at a
390: few temperatures in the range of the phase
391: transitions. Fig.~\ref{fig-maps}a) (300~K) shows how tetragonal local
392: minima appear just above the first transition, while the absolute
393: minimum of $F$ is is still located at {\bf P}=0 in the cubic well.
394: %
395: Panels b) and c) correspond to 290~K and 250~K respectively, i.e.,
396: temperatures in the range of stability of the tetragonal phase. We
397: see that the tetragonal wells are indeed the global free-energy
398: minima, and that the orthorhombic and rhombohedral wells nucleate in
399: the form of saddle points that are unstable with respect to a
400: tetragonal distortion (in the rhombohedral case, the saddle points are
401: unstable with respect to an orthorhombic distortion also).
402: %
403: Panels d) and e) refer to 220~K and 210~K respectively, i.e.,
404: temperatures in the orthorhombic range. The tetragonal wells become
405: metastable with respect to the orthorhombic ones, which are now the
406: free-energy global minima. At the same time, rhombohedral local minima
407: appear.
408: %
409: Finally, panels f) and g) correspond to 190~K and 100~K respectively,
410: both in the rhombohedral temperature range. Here we see that the
411: rhombohedral minima have finally become the deepest ones.
412: %
413: It is clear that the sequence converges to the free-energy map given
414: by the effective Hamiltonian itself (panel h)). For all the
415: transitions, the coexistence of different free-energy minima in the
416: figures evidences their first-order character.
417:
418: Our method to explore the $F$ landscape has a limitation when probing
419: the vicinity of the cubic phase. After the point {\bf P}=0 becomes a
420: local free-energy maximum a few degrees below $T_{c1}$, it is no
421: longer possible to sample the region around it using auxiliary
422: electric fields: the cubic structure is already unstable and thus no
423: longer useful as a starting point for the simulations, and attempts to
424: steer a system in a tetragonal, orthorhombic, or rhombohedral well
425: towards {\bf P}=0 only succeed in landing it in the corresponding
426: symmetry-inverted domain. As the behavior of $F$ around {\bf P}=0 is
427: basically determined by the quadratic coefficient $a$ of the
428: Devonshire-Landau expansion, the value assigned to this coefficient by
429: a fit of the simulation data should be considered suspect. (This
430: problem pertains only to the region near {\bf P}=0. Information about
431: the relative stability of tetragonal, orthorhombic, and rhombohedral
432: phases is available at any temperature, since it is always possible to
433: apply fields that lead the system to each of these symmetries.)
434: %
435: \begin{figure}[b!]
436: \epsfxsize=\hsize\epsfbox{fig2.eps}
437: \caption{Temperature behavior of the quadratic coefficient ($a$) of
438: the Devonshire-Landau potential: The dashed line connects the values
439: obtained from a direct fit to MC data, the solid line shows the
440: interpolated $a(T)$, and the solid circles are the points used to fit
441: the interpolating polynomial (see text). The transition temperatures
442: are marked with vertical lines.}
443: \label{fig-quadra}
444: \end{figure}
445: %
446: As shown in Fig.~\ref{fig-quadra}, at temperatures around and above
447: $T_{c1}$ the behavior of $a$ is fairly reasonable. It is positive at
448: high temperatures, becoming negative a few degrees below $T_{c1}$, as
449: corresponds to a first-order transition. For very low temperatures
450: $a$ also behaves well, tending smoothly to its 0~K value. This is
451: because in this region the underlying Landau potential is so simple
452: (the thermal fluctuations are relatively small and the sixth-order
453: terms almost negligible) that we can use the information obtained
454: around the low-symmetry minima to reconstruct the complete free-energy
455: map. However, in the intermediate region (from 150~K to 250~K), $a$
456: turns positive again, which would mean that the cubic phase becomes
457: metastable in a wide temperature interval. This metastability is an
458: unphysical result (explicitly ruled out within our model by zero-field
459: MC simulations starting from a cubic phase, in which we found no trace
460: of it). We thus reconsidered the fitting of our data, imposing a more
461: physical temperature evolution of $a$: we assumed that in the
462: intermediate-temperature range $a(T)$ is given by a simple
463: interpolation (Fig.~\ref{fig-quadra}) between the high-temperature
464: region, where we can reliably sample it, and the low-temperature limit
465: determined directly by the convergence to the effective Hamiltonian (a
466: third-order polynomial suffices for our purposes). With this from for
467: $a(T)$ fixed, our MC data are still well fitted to the Landau
468: potential of Eq.~\ref{eq-cowley}. For instance, at 190~K the relative
469: error obtained when $a(T)$ is freely fitted is $2.2\%$, and when
470: $a(T)$ is fixed by the interpolation the error is still very small:
471: $3.6\%$. This clearly shows that, in the intermediate-temperature
472: region, our Monte Carlo data contain little information about the
473: value of $a$.~\cite{umbrella} Higher-order terms were tried but found
474: to play no role in the fit, so they can be excluded from the
475: potential. Our final results are shown in Fig.~\ref{fig-coeffs}.
476: %
477: \begin{figure}[b!]
478: \epsfxsize=\hsize\epsfbox{fig3.eps}
479: \caption{Free-energy coefficients fitted to MC data assuming $a(T)$ is
480: given by the solid line in Fig.~\protect\ref{fig-quadra}. The
481: transition temperatures are marked with vertical lines.}
482: \label{fig-coeffs}
483: \end{figure}
484: %
485: At very low temperatures $a$, $u$, and $v$ tend to their 0~K values
486: and the sixth-order terms go to zero. On the other hand, at
487: temperatures well above the first transition ($T>350$~K) we find the
488: MC data are properly fitted to a fourth-order Landau potential, so the
489: sixth-order terms can again be excluded from the model. The sixth-order
490: terms then occur only in the temperature range in which the
491: transitions take place (as plainly shown in Fig.~\ref{fig-coeffs}). It
492: is also remarkable that the fourth-order terms exhibit a very strong
493: temperature dependence. Particularly, the anisotropic term $v$ changes
494: sign at around 90~K and takes large negative values all through the
495: intermediate-temperature range. So, apart from the influence it has,
496: together with $h_2$ and $h_3$, in determining the transition sequence,
497: we find that a negative $v$ is responsible for the first-order character of
498: the cubic to tetragonal transition.
499:
500: In Fig.~\ref{fig-thexp} we have plotted our coefficients near the
501: cubic to tetragonal transition temperature together with the available
502: experimental results.~\cite{exp,gonzalo} (We have changed the
503: temperature scale to make the experimental $T_{c1}$ coincide with the
504: theoretical one.) The agreement is only qualitative but as good as
505: could be expected given the simplifying assumptions involved in the
506: experimental work of Ref.~\onlinecite{exp} (the coefficients are
507: restricted to be constant or to depend linearly on temperature) and
508: the fact that the differences among the two experiments are comparable
509: to those between experiment and theory.
510:
511: Our work strongly suggests that the phase transitions of BaTiO$_3$ can
512: be described in terms of a single Landau potential $F$ with the form
513: of Eq.~\ref{eq-cowley}. Despite what is assumed in most of the
514: previous work on this problem, the quadratic parameter $a$ is found to
515: exhibit a strongly non-linear temperature dependence. This is clear
516: because we can reliably calculate $a$ at high (over $T_{c1}$) and very
517: low temperatures, and the two regions cannot be joined linearly. We
518: show that all the high-order terms in $F$ present a very significant
519: evolution with temperature. This conclusion is essentially opposed to
520: the temperature-independent behavior that is still assumed by some
521: authors (see, for example, Ref.~\onlinecite{other-forms}). It is also
522: very remarkable that the two features of BaTiO$_3$ that require the
523: inclusion of sixth-order terms in the Landau potential, i.e., the
524: first-order character of the transitions and the occurrence of an
525: orthorhombic phase, have been reproduced using a fourth-order
526: effective Hamiltonian.~\cite{PZT} This piece of information should be
527: taken into account when constructing mechanical models to study the
528: finite-temperature behavior of similar compounds. For instance, the
529: authors of Ref.~\onlinecite{nakamura} included sixth-order terms in
530: the underlying interaction model for \batio3\ with the explicit aim to
531: account for the first-order character of the transitions; our results
532: indicate that those terms should be unnecessary.
533: %
534: \begin{figure}[b!]
535: \epsfxsize=\hsize\epsfbox{fig4.eps}
536: \caption{Comparison of our free-energy coefficients (lines) with the
537: experimental values of Refs.~\protect\onlinecite{exp} (open symbols)
538: and~\protect\onlinecite{gonzalo} (solid symbols): solid line and
539: circles for $a$, dotted line and squares for $u+v$, dashed line and
540: ``up'' triangles for $h_1+h_2$, and dash-dotted line and ``down''
541: triangles for $h_3$. The vertical line marks the cubic-tetragonal
542: transition temperature.}
543: \label{fig-thexp}
544: \end{figure}
545: %
546:
547: \section{SUMMARY}
548:
549: We have studied the Devonshire-Landau potential underlying the phase
550: transition sequence of BaTiO$_3$ using the first-principles effective
551: Hamiltonian parametrized by Zhong, Vanderbilt, and Rabe. The
552: order-parameter configuration space was explored with the help of
553: auxiliary electric fields that change the location and relative
554: stability of the free-energy minima. Our results show that the
555: typically assumed form of the potential, an expansion up to sixth
556: order in the polarization from the paraelectric cubic phase, properly
557: accounts for the behavior of the system. But, despite what is usually
558: presumed, we find a non-trivial temperature dependence for all the
559: coefficients in the expansion, including the quadratic term $a$, which
560: is shown to behave non-linearly. Our work also shows that the
561: sixth-order terms in polarization needed to explain basic features of
562: BaTiO$_3$ in a Devonshire-Landau approach (the first-order character
563: of the transitions and the occurrence of an orthorhombic phase) are
564: properly accounted for by an interaction model that only includes
565: terms up to fourth order.
566:
567: \section{ACKNOWLEDGEMENTS}
568:
569: This work has been supported by the Spanish CICyT grant
570: No. PB98-0244. J.~I. and S.~I. acknowledge financial support from the
571: Basque regional government and the UPV, respectively. We thank
572: Javier Junquera and Jon Saenz for their help.
573:
574:
575: \begin{thebibliography}{99}
576:
577: \bibitem[*]{byline} Email address: wdbingoj@lg.ehu.es
578:
579: \bibitem{landau} L. D. Landau and E. M. Lifshitz, {\it Statistical
580: Physics} (Pergamon Press Ltd., London, 1958).
581:
582: \bibitem{radescu} S. Radescu, I. Etxebarria, and J. M. Perez-Mato,
583: {\it J. Phys.: Cond. Matt.} {\bf 7}, 585 (1995).
584:
585: \bibitem{devonshire} A. F. Devonshire, {\it Phyl. Mag.} {\bf 40}, 1040
586: (1949); {\it ibid.} {\bf 42}, 1065 (1951); {\it Advances in Physics}
587: {\bf 3}, 85 (1954).
588:
589: \bibitem{jona} See, for example: F. Jona and G. Shirane, {\it
590: Ferroelectric Crystals} (Pergamon Press, Oxford, England 1962;
591: reedition by Dover Publications, Inc., New York 1993) or M. E. Lines
592: and A. M. Glass, {\it Principles and Applications of Ferroelectric
593: Materials} (Clarendon Press., Oxford 1977).
594:
595: \bibitem{cowley} R. A. Cowley, {\it Advances in Physics} {\bf 29}, 1
596: (1980).
597:
598: \bibitem{vander-cohen} For a discussion of the conditions that a
599: Devonshire-Landau potential must fulfill to account for phases of a
600: given symmetry, see D. Vanderbilt and M. H. Cohen, cond-mat/0009337.
601:
602: \bibitem{devon-more} In the second article in
603: Ref.~\onlinecite{devonshire}, Devonshire extended the theory to
604: consider the strain and gave a proper account for the elastic and
605: piezoelectric properties of the system.
606:
607: \bibitem{exp} M. E. Drougard, R. Landauer, and D. R. Young, {\it
608: Phys. Rev.} {\bf 98}, 1010 (1955); E. J. Huibregtse and D. R. Young,
609: {\it ibib.} {\bf 98}, 1562 (1956).
610:
611: \bibitem{gonzalo} A temperature dependence of the sixth-order terms in
612: $F$ was reported by J. A. Gonzalo and J. M. Rivera in Ferroelectrics
613: {\bf 2}, 31 (1971).
614:
615: \bibitem{other-forms} See the discussion in Ref.~\onlinecite{cowley}
616: and that by K. Fujita and Y. Ishibashi, {\it Jpn. J. Appl. Phys.} {\bf
617: 36}, 254 (1997).
618:
619: \bibitem{nakamura} K. Nakamura and W. Kinase, {\it J. Phys. Soc. Jap.}
620: {\bf 61}, 4596 (1992); {\it ibid.} {\bf 61}, 2114 (1992).
621:
622: \bibitem{vanderbilt} See the short review by D. Vanderbilt in {\it
623: Current Opinions in Solid State and Materials Science} {\bf 2}, 701
624: (1997) and references therein.
625:
626: \bibitem{zhong} W. Zhong, D. Vanderbilt, and K. M. Rabe, {\it
627: Phys. Rev. Lett.} {\bf 73}, 1861 (1994); {\it Phys. Rev.} {\bf B52},
628: 6301 (1995).
629:
630: \bibitem{garcia} A. Garc\'{\i}a and D. Vanderbilt, {\it
631: Appl. Phys. Lett.} {\bf 72}, 2981 (1998); {\it Proceeding of the First
632: Principles Calculations for Ferroelectrics: Fifth Williamsburg
633: Workshop}, edited by R. E. Cohen (AIP, Woodbury, New York, 1998),
634: p. 53.
635:
636: \bibitem{waghmare-rabe} U. V. Waghmare and K. M. Rabe, {\it
637: Phys. Rev.} {\bf B52}, 13236 (1995).
638:
639: \bibitem{krakauer} H. Krakauer, R. Yu, C. Z. Wang, and C. Lasota, in
640: {\it Proceedings of the 1997 Williamsburg Workshop on Ferroelectrics},
641: edited by H. Chen, {\it Ferroelectrics} {\bf 206}, 133 (1998).
642:
643: \bibitem{uns-band} The dispersion curves associated to the harmonic
644: effective system are essentially those of the unstable band calculated
645: by Ghosez {\sl et al.} [Ph. Ghosez, E. Cockayne, U. V. Waghmare, and
646: K. M. Rabe, {\it Phys. Rev.} {\bf B60}, 836 (1999)].
647:
648: \bibitem{king-smith} R. D. King-Smith and D. Vanderbilt, {\it
649: Phys. Rev.} {\bf B49} 5828 (1994).
650:
651: \bibitem{mc} Although our emphasis is on Monte Carlo simulations,
652: the discussion applies also to molecular dynamics methods.
653:
654: \bibitem{tech1} From a MC simulation we typically obtain $\langle{\bf
655: P}\rangle$ corresponding to the stable-equilibrium state of the
656: system. Eq.~\protect\ref{eq-def} can be used also if the system gets
657: stuck in a metastable-equilibrium state, but not if the simulation
658: alternates from one free-energy minimum to another, giving a spurious
659: $\langle{\bf P}\rangle$.
660:
661: \bibitem{temps} The experimental transition temperatures for BaTiO$_3$ are
662: $T_{c1} = 403$~K, $T_{c2} = 278$~K, and $T_{c3} = 183$~K.
663:
664: \bibitem{umbrella} We tried to compute $a$ in the problematic
665: temperature interval by modifying the effective Hamiltonian to keep
666: the point ${\bf P}=0$ as a local minimum (what could be understood
667: as a kind of umbrella sampling, see M. P. Allen and D. J. Tildesley,
668: {\it Computer Simulation of Liquids} (Oxford, New York,
669: 1990)). However, this modification results in inhomogeneous
670: equilibrium states, which lie outside the Devonshire-Landau framework.
671:
672: \bibitem{PZT} For Pb(Zr$_x$Ti$_{1-x}$)O$_3$, a similar fourth-order
673: effective Hamiltonian reproduces a monoclinic phase, and thus would
674: formally lead to eigth-order terms in the Devonshire-Landau
675: potential. See Ref.~\onlinecite{vander-cohen} and references therein.
676:
677: \end{thebibliography}
678:
679: \end{document}
680: