0808.1706/d22.tex
1: \documentclass[twocolumn,prb,preprintnumbers,amsmath,amssymb,floatfix,showpacs]{revtex4}
2: 
3: \usepackage{graphicx}
4: \usepackage{subfigure}
5: \usepackage{dcolumn}
6: 
7: \setlength{\oddsidemargin}{0in}
8: \setlength{\evensidemargin}{0in}
9: \setlength{\textwidth}{6.5in}
10: \setlength{\topmargin}{0.2in}
11: \setlength{\headheight}{0.1in}
12: \setlength{\headsep}{0in}
13: \setlength{\textheight}{9.0in}
14: 
15: \begin{document}
16: 
17: \title{Anisotropy and Magnetism in the LSDA+U Method}
18: \author{Erik R. Ylvisaker}
19: \affiliation{Department of Physics, University of California, Davis, California, 95616}
20: 
21: \author{Warren E. Pickett}
22: \affiliation{Department of Physics, University of California, Davis, California, 95616}
23: 
24: \author{Klaus Koepernik}
25: \affiliation{IFW Dresden, P O Box 270116, D-01171 Dresden, Germany}
26: 
27: \date{\today}
28: %
29: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: %%  TODO: Write abstract
31: %%
32: \begin{abstract}
33: 
34: Consequences of anisotropy (variation of orbital occupation) and 
35: magnetism, and their coupling, are analyzed for LSDA+U functionals, 
36: both the commonly used ones as well as less commonly applied functionals.  
37: After reviewing and extending some earlier observations for an
38: isotropic interaction, the anisotropies are examined more fully and
39: related to use with the local density (LDA) or local spin density
40: (LSDA) approximations.  
41: The total energies of all possible integer configurations of an
42: open $f$ shell are presented for three functionals, where some
43: differences are found to be dramatic.  
44: Differences in how the commonly used ``around mean field'' (AMF) and ``fully
45: localized limit'' (FLL) functionals 
46: perform are traced to such differences.
47: The LSDA+U interaction term, applied self-consistently, usually enhances
48: spin magnetic moments and orbital polarization, and the double-counting terms of both functionals 
49: provide an opposing, moderating tendency (``suppressing the magnetic 
50: moment''). The AMF double counting term gives magnetic states a significantly 
51: larger energy penalty than does the FLL counterpart.
52: 
53: \end{abstract}
54: 
55: % End abstract.
56: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
57: 
58: \maketitle
59: 
60: \section{Introduction}
61: 
62: Density functional theory (DFT) and its associated local (spin) density 
63: approximation [L(S)DA] is used widely to describe the properties of a 
64: wide variety of materials, often with great success.
65: However there exists a class of materials which are poorly described, 
66: sometimes qualitatively, by 
67: LDA. These so-called strongly correlated materials
68: typically contain atoms with open $d$ or $f$ shells, in which the
69: corresponding orbitals are in some sense localized.  
70: % There is now evidence
71: % that oxygen $2p$ orbitals can behave in a correlated fashion.[cite RbO2, 
72: % or other interesting p orbital systems]  
73: The LSDA+U approach was introduced 
74: by Anisimov, Zaanen, and Andersen\cite{OriginalLDAU} to treat correlated 
75: materials as a modification of LDA (`on top of LDA') that adds an 
76: intra-atomic Hubbard U repulsion term in the energy functional.  Treated in
77: a self-consistent mean field (`Hartree-Fock') manner, in quite a large number
78: of cases the LDA+U result provides a greatly improved description of 
79: strongly correlated materials.  
80: 
81: At the most basic level, the LSDA+U correction tends to drive the correlated
82: orbital $m$ occupation numbers $n_{m\sigma}$ ($\sigma$ denotes spin projection)
83: to integer values 0 or 1.  This in turn produces, under appropriate conditions,
84: insulating states out of conducting LSDA states, and the Mott insulating
85: state of several systems is regarded as being well described by LSDA+U at
86: the band theory level.   Dudarev {\it et al.}\cite{Dudarev98} and Petukhov
87: {\it et al.}\cite{Petukhov03} provided some description of the
88: effect of the spin dependence of two different double counting terms 
89: within an isotropic approximation.
90: Beyond this important but simple effect, there is 
91: freedom in which of the spin-orbitals ($m\sigma$) will be occupied, which can
92: affect the result considerably and therefore makes
93: it important to understand the effects of anisotropy and spin polarization in
94: LSDA+U.  After the successes of providing realistic pictures of the Mott
95: insulating state in La$_2$CuO$_4$ and the transition metal 
96: monoxides,\cite{OriginalLDAU} the anisotropy contained in the LSDA+U method 
97: produced the correct orbitally ordered magnetic arrangement for KCuF$_3$ that provided
98: an understanding of its magnetic behavior.\cite{SashaL95}
99: 
100: The anisotropy of the interaction, and its connection to the level of spin
101: polarization, is a topic that is gaining interest and importance.  One example
102: is in the LSDA+U description of the zero temperature Mott transition under 
103: pressure in the classic Mott insulator MnO.  The first transition under pressure
104: is predicted to
105: be\cite{Kasinathan07} an insulator-insulator (not insulator-metal) transition, with a
106: $S=\frac{5}{2} \rightarrow S=\frac{1}{2}$ moment collapse and a volume collapse.
107: The insulator-to-insulator aspect is surprising, but more surprising is the form of
108: moment collapse: each orbital remains singly occupied beyond the transition, but
109: the spins of electrons in two of the orbitals have flipped direction.  This type 
110: of moment collapse is totally unanticipated (and hence disbelieved by some), but
111: it is robust against crystal structure (occurring in both rocksalt and NiAs
112: structures) and against reasonable variation of the interaction strength.  Detailed analysis indicates it is
113: a product of the anisotropy of the LSDA+U interaction and the symmetry lowering due to
114: antiferromagnetic order.
115: 
116: Another unanticipated result was obtained\cite{KWL2004} in LaNiO$_2$, which is a metal
117: experimentally. This compound is also a metal in LSDA+U over a very large range of 
118: interaction strength $U$, rather than reverting to a Mott insulating Ni$^{1+}$
119: system which would be isovalent with CaCuO$_2$.  
120: For values of $U$ in the range expected to be
121: appropriate for the Ni ion in this oxide, the magnetic system consists of an
122: atomic singlet consisting of antialigned $d_{x^2-y^2}$ and $d_{z^2}$ spins
123: on each Ni ion.  Again the anisotropy of the interaction evidently plays a
124: crucial role in the result, with its effect being coupled thoroughly with 
125: band mixing effects.
126: 
127: The addition of a Hubbard U interaction also introduces the need for 
128: ``double counting'' correction
129: terms in the energy functional, to account for the fact that the Coulomb
130: energy is already included (albeit more approximately) in the LSDA functional.
131: All double counting schemes subtract an averaged energy for the occupation
132: of a selected reference state depending only on $\{N_{\sigma}\}$,
133: which largely cancels the isotropic
134: interaction of the $E_{I}$ term Eq. (\ref{eq:LDAU-General}).
135: Several forms for these double-counting terms have been
136: proposed,\cite{OriginalLDAU,Anisimov93,CS} but primarily two are
137: commonly used.  These LDA+U functionals are most often referred to as around mean
138: field (AMF) and the fully localized limit (FLL), which is also referred to as
139: the atomic limit (AL).  The distinctions between these forms have attracted
140: some discussion, but without consideration of the full anisotropy of the
141: interaction.
142: 
143: The need for double-counting corrections is not unique to the LDA+U method; 
144: any other method that adds correlation terms to the LSDA functional, such 
145: as the dynamical LDA+DMFT (dynamical mean field theory) approach, also will 
146: require double-counting corrections.  This is an unfortunate consequence of LDA's 
147: success; LDA works too well, even in correlated systems where it usually 
148: gets interatomic charge balance reasonably, to just throw it away.\footnote{There
149: are numerous examples for strongly correlated ({\it heavy fermion}) metals where
150: the Fermi surface calculated within LDA is predicted as well as for more conventional
151: metals.  This surely requires that the charge balance between the various atoms
152: is accurate.}  The common
153: approach has been to use LSDA for correlated materials and include a 
154: double-counting correction.  There are techniques being developed which do 
155: not build on a correction to DFT-LDA, 
156: but it remains to be seen whether these approaches will be successfully 
157: applied to a broad range of solid-state materials.
158: 
159: Although there has been much study on the performance of these LDA+U 
160: functionals in the context of real materials, and an early review of the method
161: and some applications was provided by Anisimov, Aryasetiawan, and Lichtenstein,\cite{review}
162: relatively little has been done 
163: to understand, qualitatively and semi-quantitatively, how the functionals 
164: operate based solely on their energetics, distinct from DFT-LSDA effects.   
165: In this paper we analyze the functionals that are commonly used, as well as 
166: others which were introduced early on but are not so commonly used.  Some of 
167: the nomenclature in the literature is confusing, so we try to clarify these 
168: confusions where we can.
169: 
170: \section{The LSDA+U Correction $\Delta E$}
171: 
172: %
173: \begin{table*}[th]
174: \begin{centering}
175: \vskip 4mm
176: %\begin{large}
177: \begin{tabular}{c|cccc}
178: LDA+U        & $E_{dc}$ &=& $E_{dc}$    & DFT XC \\
179:   Functional &          & & (rewritten) &  Functional  \\ \hline \hline
180: Fl-nS		& $\frac12 UN^2 - \frac{U+2lJ}{2l+1}\frac{1}{4}N^2$				  	   
181:          &=&  $\frac12 UN^2 - \frac{U+2lJ}{2l+1} \frac{1}{2}\sum_\sigma (\frac{N}{2})^2$ 	& LDA	\\ \hline
182: Fl-S (AMF) 	& $\frac12 UN^2 - \frac{U +2lJ}{2l+1}\frac{1}{2}\sum_\sigma N_\sigma^2$
183:          &=&  $\frac12 UN^2 - \frac{U +2lJ}{2l+1}\frac{1}{2}\sum_\sigma N_\sigma^2$     & LSDA	\\     \hline
184: FLL			& $\frac12 UN(N-1) - \frac12 J\sum_\sigma N_\sigma(N_\sigma-1)$ 
185:          &=&  $\frac12 UN(N-1) - \frac12 J\sum_\sigma (N_\sigma^2 -N_\sigma)$      & LSDA \\   \hline
186: FLL-nS		& $\frac12 UN(N-1) - \frac14 J N(N-2)$ 						   
187:          &=&  $\frac12 UN(N-1) - \frac12 J\sum_\sigma ((\frac{N}{2})^2-N_\sigma)$& LDA	\\     \hline
188: \end{tabular}
189: %\end{large}
190: \end{centering}
191: \caption{The double-counting terms of various LDA+U functionals.  In the second expression two of
192: them are rewritten to reflect how they are (somewhat deceptively) identical in form, but in one
193: case a distinction between spin-up and spin-down (relative to half of N: 
194: $N_{\sigma}\leftrightarrow N$/2) is made.  Note that while the first two forms
195: appear to contain an isotropic self-interaction 
196: [$\frac{1}{2}UN^2$ rather than $\frac{1}{2}N(N-1)$] 
197: they are derived from a form which {\it explicitly} has {\it no} self-interaction between
198: the orbital fluctuations $\delta_{M\sigma}$. See text for more discussion.}
199: \label{tbl:Functionals}
200: \end{table*}
201: %
202: The LDA+U functional is usually coded in a form in which the choice of coordinate system is
203: irrelevant, often referred to as the rotationally-invariant form.\cite{SashaL95}  
204: This form involves Coulomb
205: matrix elements that have four orbital indices, and the orbital occupation numbers are matrices
206: in orbital space ({\it viz.} $n_{mm'}$).  One can always (after the fact) rotate into the
207: orbital Hilbert space in which the occupations are diagonal, in which case the interactions
208: have only two indices.  In our discussion we will work in the diagonal representation. 
209: 
210: The LDA+U functionals considered here can all be written in the form  
211: \begin{eqnarray}
212: \Delta E = E_I - E_{dc},
213: \label{eq:simple}
214: \end{eqnarray}
215:  where the direct interaction is
216: %
217: \begin{eqnarray}
218: E_I = \frac12 {\sum_{m\sigma \neq m'\sigma'}} W_{mm'}^{\sigma\sigma'} n_{m\sigma}n_{m'\sigma'}
219: \label{eq:LDAU-General}
220: \end{eqnarray}
221: %
222: and $E_{dc}$ is the double-counting correction.  The Coulomb matrix
223: elements are given in terms of the direct and (spin-dependent) exchange contributions as
224: \begin{eqnarray}
225: W_{mm'}^{\sigma\sigma'} = (U_{mm'} - J_{mm'} \delta_{\sigma,\sigma'}). 
226: \end{eqnarray}   
227: By the convention chosen here, $E_I$ and $E_{dc}$ are both positive quantities as long as the
228: constants $U$ and $J$ (which define the matrix elements $U_{mm'}$ and $J_{mm'}$ but are not the
229: same) as chosen conventionally, with $U$ much larger than $J$.
230: 
231: Note that the orbital+spin diagonal term has been omitted in Eq. \ref{eq:LDAU-General} -- there is
232: no self-interaction in $E_I$.  However, it is formally allowed to include the diagonal
233: `self-interaction' term, because the matrix element vanishes identically (self-interaction
234: equals self-exchange: $U_{mm} = J_{mm}$), and it can simplify
235: expressions (sometimes at a cost in clarity) if this is done.  The double counting correction
236: depends only on the orbital sum $N_{\sigma}$, which appears up to 
237: quadratic order.  A consequence is that 
238: it will contain terms in $n_{m\sigma} n_{m\sigma}$,
239: which are self-interactions.  Thus while the LSDA+U method was not intended as a self-interaction correction
240: method, it is not totally self-interaction free.  In fact, the underlying LSDA method also contains
241: self-interaction, and the double-counting term may serve to compensate somewhat this unwanted
242: effect.  We discuss self-interaction at selected points in this paper.
243: 
244: 
245: \subsection{Short formal background to the LSDA+U method.}
246: The ``LSDA+U method'' is actually a class of functionals.  Each functional
247: has the same form of interaction $E_I$, with differences specified by \\
248: (1) choice of the form of double counting term. \\
249: (2) choice of constants $U$ and $J$. For a given functional, these are
250: `universal' constants like $\hbar, m, e$, i.e. they are not functional of 
251: the density in current implementations. 
252: Possibilities for doing so, that is, determining them self-consistently
253: within the theory, have been proposed.\cite{erwin} \\
254: (3) choice of projection method to determine the occupation matrices from the
255: Kohn-Sham orbitals.  Given identical choices for
256: (1) and (2) above, there will be some (typically small) differences in
257: results from different codes due to the
258: projection method.
259: 
260: The occupation numbers (or, more generally, matrices) are functionals of the
261: density, $n_{m\sigma}[\rho]$, through their dependence on the Kohn-Sham orbitals.
262: Then, whereas in LSDA one uses the functional derivative
263: \begin{eqnarray}
264: {\rm LSDA}:~~\frac{\partial E_{LSDA}[\{\rho_{s}\}]}{\partial \rho_{\sigma}(r)}
265: \end{eqnarray}
266: in minimizing the functional, in LSDA+U the expression generalizes to
267: \begin{eqnarray}
268: {\rm LSDA+U}:~~\frac{\partial ( E_{LSDA}[\{\rho_{s}\}] + 
269:     \Delta E[\{n_{ms}[\rho_s]\}] )}{\partial \rho_{\sigma}(r)}
270: \end{eqnarray}
271: Since the resulting spin densities $\rho_{s}$ are changed by including the 
272: $\Delta E$ correction, the change in energy involves not only $\Delta E$ but also the
273: change in $E_{LSDA}$.  In practice, there is no reason to compare $E_{LSDA+U}$
274: with $E_{LSDA}$ as they are such different functionals.  However, in the 
275: following we will be assessing the importance of the choice of the double
276: counting term in the LSDA+U functional, and it is of interest
277: to compare, for fixed $U$ and $J$, the energy differences between 
278: LSDA+U functionals differing only in their double counting terms in order
279: to understand the differing results.
280: Even if the set of occupation numbers turn out to be
281: the same (a situation we consider below), 
282: the densities $\rho_{\sigma}$ will be different and the differences in
283: $E_{LSDA}$ may become important.
284: 
285: As with the non-kinetic energy terms in $E_{LSDA}$, the functional derivatives
286: of $\Delta E$ lead to potentials in the Kohn-Sham equation.  These are non-local
287: potentials, which (via the same projection used to define the occupation
288: numbers) give rise to orbital-dependent (nonlocal) potentials
289: \begin{eqnarray}
290: v_{m\sigma} \equiv \frac{\partial \Delta E}{\partial n_{m\sigma}} = 
291:   v_{m\sigma}^I - v_{m\sigma}^{dc},\nonumber \\
292:   v_{m\sigma}^I = \sum_{m'\sigma' \neq m\sigma} W_{mm'}^{\sigma\sigma'} n_{m'\sigma'}.
293: \end{eqnarray}
294: The double counting orbital potential is discussed later.
295: 
296: The corresponding contribution to the eigenvalue sum $E_{sum}$ is
297: \begin{eqnarray}
298: \Delta E_{sum} = \sum_{m\sigma} v_{m\sigma} n_{m\sigma},
299: \end{eqnarray}
300: which is subtracted from the eigenvalue sum to obtain the Kohn-Sham
301: kinetic energy.
302: However, there are indirect effects of the orbital potentials that affect
303: all of the kinetic and (LSDA) potential energies; these will be different
304: for different $\Delta E$ functionals because the orbital potentials, which depend 
305: on the derivative of $\Delta E$ and not simply on the values of $n_{m\sigma}$,
306: differ for each functional.  This makes it necessary, for understanding the
307: effects of the $\Delta E$ correction and the change in energy, to analyze the
308: orbital potentials. 
309: We provide a brief discussion in Sec. V.
310: 
311: \subsection{Fluctuation forms of LSDA+U}
312: 
313: First we consider the class of functionals that can be written in what is termed here as a 
314: {\it fluctuation form}.  The original LDA+U functional was introduced in 1991 by Anisimov, Zaanen 
315: and Andersen\cite{OriginalLDAU} and was written as
316: %
317: \begin{equation}
318: \label{eq:Fl-nS}
319: 	\Delta E^{Fl-nS} = \frac12 \sum_{m\sigma \neq m'\sigma'} W_{mm'}^{\sigma\sigma'} 
320:                             (n_{m\sigma} - \bar n)(n_{m'\sigma'} - \bar n),
321: \end{equation}
322: %
323: where $\bar n = N^\text{corr} / 2(2l+1)$ is the average occupation of the correlated orbitals.  
324: (Henceforth $N \equiv N^\text{corr}$.)  Note that the energy is changed only according
325: to {\it angular `fluctuation'} away from the (spin-independent) angular average occupation.
326: This form is properly used with LDA (the `LDA averages' $\bar n$ are the reference) and not LSDA. 
327: This form was  originally advocated with generic  $(U - J \delta_{\sigma,\sigma'})$ matrix
328: elements instead of the full Coulomb matrix, 
329: but we use the full $W_{mm'}^{\sigma\sigma'}$ here for comparison with other functionals.  
330: 
331: In 1994, Czyzyk and Sawatzky\cite{CS} introduced a change to (\ref{eq:Fl-nS}) and also
332: proposed a 
333: new functional.  The motivation for changing (\ref{eq:Fl-nS}) was to use an LSDA exchange-correlation
334: functional to treat spin splitting effects rather than LDA.  This change motivated the following equation,
335: %
336: %\begin{widetext}
337: \begin{eqnarray}
338: \label{eq:AMF}
339: 	\Delta E^{Fl-S}&=&\frac12 \sum_{m\sigma \neq m'\sigma'} W_{mm}^{\sigma\sigma'} 
340:         (n_{m\sigma} - \bar n_\sigma)(n_{m'\sigma'} - \bar n_{\sigma'})\nonumber \\
341:                    & = & \Delta E^{AMF}
342: \end{eqnarray}
343: %\end{widetext}
344: %
345: where $\bar n_\sigma = N_\sigma / (2l+1)$ is the average occupation of a single spin 
346: of the correlated orbitals.  Here the energy correction is due to {\it angular fluctuations}
347: away from the spin-dependent angular mean, and hence must be used with LSDA.   
348: We point out that the authors in 
349: Ref. [\onlinecite{CS}] refer to Eq. (\ref{eq:Fl-nS}) as 
350: $E^\text{LDA+AMF}$ and Eq. (\ref{eq:AMF}) as $E^\text{LSDA+AMF}$.  This wording may have
351: caused subsequent confusion, due to the way these terms have come to be used, and also
352: because a discussion of the ``+U'' functionals requires explicit 
353: specification of whether LDA or LSDA is being used just to understand which functional 
354: is being discussed.   Also confusing is that Solovyev, Dederichs, and 
355: Anisimov\cite{Solovyev94} rejustified Eq. \ref{eq:Fl-nS} using ``atomic limit''
356: terminology.
357: 
358: The fluctuation forms of LSDA+U are automatically particle-hole symmetric, since
359: $n_{m\sigma}\rightarrow 1-n_{m\sigma}$, $\bar n_{\sigma}\rightarrow 1-\bar n$
360: gives $ n_{m\sigma}-\bar n_{\sigma} \rightarrow -(n_{m\sigma}-\bar n_{\sigma})$
361: and the expression is quadratic in these fluctuations.  The general form of Eq.
362: \ref{eq:simple} need not be particle-hole symmetric.
363: 
364: Many authors (present authors included) have used 
365: the term LDA+U where the term 
366: LSDA+U would be more appropriate, which is especially confusing when discussing the 
367: AMF functional.  We choose to depart from this confusing nomenclature by giving 
368: (\ref{eq:Fl-nS}) and (\ref{eq:AMF}) unique names specifying their fluctuation forms, and their connection to LSDA (Fl-S) or to LDA (Fl-nS).
369: We collect the double-counting terms for the various functionals, along with their
370: connection to LDA or LSDA, in Table I.
371: 
372: \subsection{The Fully Localized Limit (FLL) Functional}
373: The second functional introduced by Czyzyk and Sawatzky\cite{CS} is the FLL functional.  
374: (A $J$=0 version of FLL was
375: introduced in 1993 by Anisimov {\em et al.}\cite{Anisimov93})  
376: The authors referred to it (confusingly, as terminology has progressed)
377: as the ``around mean field'' functional but the atomic 
378: limit double counting term; in the literature it is now referred to as 
379: the atomic limit, or fully localized limit (FLL) functional.  This functional cannot be 
380: written in the fluctuation form 
381: of the previous two 
382: functionals (the fluctuation form is exhausted by the -S and -nS cases).   
383: The FLL functional is written in the form of (\ref{eq:simple}), with 
384: the double-counting term given in Table \ref{tbl:Functionals}.  
385: 
386: There is yet another LDA+U functional that is available, which was 
387: introduced in 1993 by Anisimov {\em et al.} \cite{Anisimov93}
388: There is no clear name for it, but since it can be obtained by using 
389: $N_\sigma = N/2$ in $E_{dc}$ for FLL, one might consistently refer to 
390: it as FLL-nS, corresponding to FLL with no 
391: spin dependence.  The authors in Ref. 
392: \onlinecite{Anisimov93} indicate that this functional is to be used with LDA,
393: in accordance with the lack of spin dependence in the double counting term.
394: 
395: 
396: \subsection{Implementation of LSDA+U \\
397:             in Some Widely Used Codes}
398: 
399: The Fl-nS  functional is implemented in the {\sc Wien2k} code, as {\verb nldau } $=2$, and called HMF (Hubbard in Mean Field),\cite{Wien2k} however it is apparently not often used. \\
400: The Fl-S (AMF) functional is implemented in the {\sc Wien2k} code\cite{Wien2k} as {\verb nldau } $=0$ and the FPLO code\cite{FPLO} as AMF.  It is also available in the {\sc Abinit} code\cite{abinit} when using a PAW basis set\cite{amadon:155104} by setting {\verb usepawu } $=2$.  \\
401: The FLL functional is implemented in several general-purpose DFT codes, such as {\sc Wien2k} ({\verb nldau } $=1$)\cite{Wien2k}, FPLO (select AL in fedit)\cite{FPLO}, VASP, PW/SCF, and {\sc Abinit} ({\verb usepawu } $=1$) when using a PAW basis set. \\
402: The FLL-nS functional is available in VASP. \\
403: 
404: 
405: %Table \ref{tbl:Functionals} clarifies the similarities and differences between Fl-nS and Fl-S (AMF) on the one hand, and between FLL and FLL-nS on the other.  Here we make some observations about the two functionals that should be used with LSDA, {\it i.e.} the FLL and Fl-S double-counting terms.  Comparing their forms, the difference between the Fl-S (AMF) and the FLL double counting can be written as
406: %
407: %\begin{eqnarray}
408: %   E_{\text{dc}}^{\text{Fl-S}} &-& E_{\text{dc}}^{\text{FLL}} = \frac{(U-J)}2\sum_\sigma N_\sigma(1-\bar n_\sigma)  \nonumber \\
409: %        &=&  \frac{(U-J)}2 \left[ N - \frac{N^2 + M^2}{2(2l+1)}\right] > 0.
410: %    \label{eq:FLL-AMF}
411: %\end{eqnarray}
412: %
413: %where the strict inequality holds for any open shell. Since this double counting energy is subtracted, one has the result
414: %
415: %\begin{eqnarray}
416: %E_I^{\text{FLL}} \geq E_I^{\text{Fl-S}} \equiv E_I^{\text{AMF}}
417: %\end{eqnarray}
418: %
419: %when evaluated with the same $n_{\sigma}$ (hence the same $N$ and $M$). This quantity is largest when $N_\uparrow$ is near $N_\downarrow$ but variations in (\ref{eq:FLL-AMF}) are generally smaller (roughly by a factor of 2) than the other spin-dependent terms in the individual double counting terms.  Thus, for small $M$, these two functionals give similar results in how states are ordered energetically, especially for the lowest energy states.
420: 
421: % One should not blindly apply these relationships between the magnitudes of the double-counting terms (at given $N_{\sigma}$) to resulting total energies.  Other terms in the LSDA energy depend not only on the occupation numbers $n_{m\sigma}$, but also depend on {\it derivatives} of the double-counting terms (the orbital potentials), because these appear in the Kohn-Sham equation that is used to minimize the functional and evaluate the kinetic, potential, and LSDA exchange-correlation energies.
422: 
423: \subsection{General Remarks}
424: When the Fl-nS and Fl-S functionals are written  
425: in their fluctuation form, there is no 
426: separate double-counting term, hence one does not need the double-counting interpretation.
427: They can of course be expanded to 
428: be written in the `interaction minus double-counting' form of (\ref{eq:LDAU-General}),
429: which is useful especially for comparing with functionals that can only be written in that form.  
430: A comparison of the double-counting terms is given in Table \ref{tbl:Functionals}.
431: Reducing all to interaction minus double-counting form makes the difference between 
432: the functionals most evident; since they all have the 
433: same ``direct-interaction'' term, the {\it only difference} between the functionals 
434: is what double-counting energy is used; the uninteresting tail seems to be wagging the
435: exciting dog, which is in fact the case.  The double-counting terms can be reduced 
436: to dependence only on $N$ and $N_\sigma$  thanks to summation rules 
437: (there is at least one free index) on the 
438: $U_{mm'}$ and $J_{mm'}$ matrices,
439: %
440: \begin{eqnarray}
441: 	\sum_m U_{mm'} & = &(2l+1)U				\\
442: 	\sum_m J_{mm'} & = &U + (2l)J, 
443: \end{eqnarray}
444: %
445: that is, the sum over any column (or row) of the U and J matrices is a fixed simple value, which depends on the input parameters $U$ and $J$.  One can then simply see that a sum over a column of $W$ is $(2l+1)U$ if $\sigma \neq \sigma'$ and $2l(U - J)$ if $\sigma = \sigma'$.
446: 
447: The $U_{mm'}$ and $J_{mm'}$ matrices satisfy, by definition, $U_{mm} - J_{mm} = 0$, so that 
448: there is no self-interaction, whether or not the (vanishing)diagonal term $m\sigma = m'\sigma'$ is included 
449: in the interaction term.  As mentioned earlier, the following analysis assumes the 
450: the occupation matrix has been diagonalized.  While this can always be done, 
451: the transformed matrix elements $U_{mm'}$ and $J_{mm'}$ will not be exactly what we have used in 
452: Sec. \ref{Sec:NumericalResults}.  
453: 
454: \section{Analysis of the Functionals}
455: 
456: \subsubsection{The $J=0$ simplification.} 
457: It is not uncommon for practitioners to use `effective' values
458: ${\tilde U} = U-J, {\tilde J}=0$ and insert
459: these constants (for $U, J$) into LDSA+U.
460: For $J = 0$, of course Hund's coupling (intra-atomic exchange)
461: is lost, but $J$ also controls the
462: anisotropy of the interaction, and for $J=0$ anisotropy also is lost 
463: ($U_{mm'}\equiv U$
464: as well as $J_{mm'}\equiv 0$ for $m\neq m'$). 
465: This case is relatively simple, it seems it should provide the ``big picture''
466: of what LSDA+U does with simple Coulomb repulsion, 
467: and it has been discussed several times before.
468: With $J = 0$, the fluctuation functionals simplify to
469: \begin{eqnarray}
470: \Delta E^{Fl-\kappa}_{J=0}&=&\frac{U}{2}\sum_{m\sigma \neq m'\sigma'} \delta n_{m\sigma}
471:                                      \delta n_{m'\sigma'} \nonumber \\
472:  &=&\frac{U}2 \left[\left(\sum_{m\sigma}\delta n_{m\sigma} \right)^2
473:              - \sum_{m\sigma}(\delta n_{m\sigma})^2\right] \nonumber \\
474:  &=&-\frac{U}2 \sum_{m\sigma}(\delta n_{m\sigma})^2 \equiv -\frac{U}2 \Gamma_{\kappa}^2 \leq 0,
475: \end{eqnarray}
476: %
477: because the sum of fluctuations vanishes by definition for either form 
478: $\kappa = nS$ or $S$; note the `sign change' of this expression when the diagonal
479: terms are added, and subtracted, to simplify the expression. 
480: Here $\Gamma^2$ is the sum of the squares of the fluctuations, bounded
481: by $0 \leq \Gamma_{\kappa}^2 \leq N$.
482: For integer occupations the energy corrections for Fl-nS and Fl-S (AMF) can be written
483: \begin{eqnarray}
484: \Delta E^{Fl-S}_{J=0}&=&-\frac{U}2 \left[N(1 - \bar n) -\frac{M^2}{2(2l+1)}\right],\nonumber \\
485: \Delta E^{Fl-nS}_{J=0}&=&-\frac{U}2 N(1 - \bar n).
486: \label{eq:J=0}
487: \end{eqnarray}
488: 
489: There are two things to note here.
490: %
491: \begin{enumerate}
492: \item In Fl-nS, the energy is independent of both the spin and orbital polarization of the state,
493: which lacks the basic objective of what LSDA+U is intended to model.  
494: % The $N$ dependence is $\Delta E^{Fl-nS}$ is very simple; it is a parabola with an 
495: % energy minimum of $\frac{2l+1}{4}U$ at $N = 2l+1$ (half-filled).  i
496: Considering the form of its double counting term (see Table \ref{tbl:Functionals}) 
497: with its self-interaction
498: term (proportionality to $N^2$), Fl-nS for $J$=0 becomes simply a self-interaction correction method. 
499: \item In Fl-S (AMF), configurations with magnetic moments are energetically {\it penalized},
500: {\it proportionally to} $U$ and quadratically with $M$. 
501: In later sections we will 
502: discuss the partial cancellation with the LSDA magnetic energy.	
503: \end{enumerate}
504: %
505: 
506: Under the same conditions, the FLL functional becomes
507: \begin{eqnarray}
508: \Delta E^{FLL} = \frac{U}{2}\sum_{m\sigma} n_{m\sigma}(1 - n_{m\sigma}) \geq 0.
509: \end{eqnarray}
510: Solovyev {\it et al.}\cite{Solovyev94} noted the important and easily
511: recognizable characteristics of this expression.  Besides being non-negative,
512: for integer occupations the energy vanishes.  It is a simple inverted parabola
513: as a function of each $n_{m\sigma}$.  From the derivative, the orbital potentials
514: are linear functions of $n_{m\sigma}$, with a discontinuity of $U$ when $n_{m\sigma}$
515: crosses an integer value.  These characteristics underlie the most basic properties
516: of the LSDA+U method: integer occupations are energetically preferred, and 
517: discontinuities in the potentials model realistically the Mott insulator gap that
518: occurs in strongly interacting systems at (and only at) integer filling.
519: 
520: \subsubsection{$J \neq 0$, but Isotropic}
521: %
522: Simplification of the full expression for a functional results by separating out
523: the isotropic parts of the interaction:
524: %
525: \begin{subequations}
526: 	\label{eq:UJbar}
527: 	\begin{equation}
528: 		\label{eq:Ubar}
529: 		U_{mm'} = U + \Delta U_{mm'}
530: 	\end{equation}
531: 	\begin{equation}
532: 		\label{eq:Jbar}	
533: 		J_{mm'} =  U\delta_{mm'} + J(1-\delta_{mm'}) + \Delta J_{mm'}.	
534: 	\end{equation}
535: \end{subequations}
536: %
537: %
538: %First we examine the contribution from $U_{mm'}$, by writing $\Delta E= \epsilon^U + 
539: %\epsilon^J$ where $\epsilon^U$ and $\epsilon^J$ are the independent contributions from 
540: %$U_{mm'}$ and $J_{mm'}$, then we obtain
541: %\begin{equation}
542: %	\epsilon^U =  - \frac{U \Gamma^2}{2} + 
543: %				    \sum_{m\sigma \neq m'\sigma'} \frac{\Delta U_{mm'}}{2} \delta n_{m\sigma} \delta n_{m'\sigma'}.
544: %\end{equation}
545: %
546: %In performing a similar analysis for $\epsilon^J$, we obtain a factor of
547: %
548: %\begin{equation}
549: %	 \sum_{mm'\sigma} \delta n_{m\sigma} \delta n_{m'\sigma} 
550: %	 	 = \sum_\sigma \left(\sum_m \delta n_{m\sigma} \right)^2
551: %\end{equation}
552: %
553: %for which the quantity in parentheses is zero in Fl-S, but in Fl-nS the summation is equal to $\frac12 M^2$.
554: %
555: %Putting it all together, we have
556: %
557: %\begin{eqnarray}
558: %	\Delta E^{\text{Fl-nS}} 
559: %    & = & -\frac{\bar U - \bar J}2 N (1-\bar n) - \frac{\bar J}4 M^2  \nonumber \\
560: %		& + & \frac12 \sum_{m\sigma \neq m'\sigma'} \Delta W_{mm'}^{\sigma\sigma'} \delta n_{m\sigma} \delta n_{m'\sigma'}
561: %	\label{eq:Fl-nS-simplified}
562: %\end{eqnarray}
563: %
564: %and for Fl-S we have:
565: %
566: %\begin{eqnarray}
567: %	\Delta E^{\text{Fl-S}} 
568: %	  & = & -\frac{\bar U - \bar J}2  N(1-\bar n) + \frac{(\bar U - \bar J)}{4(2l+1)}M^2 \nonumber \\
569: %		& + &  \frac12 \sum_{m\sigma \neq m'\sigma'} \Delta W_{mm'}^{\sigma\sigma'} \delta n_{m\sigma} \delta n_{m'\sigma'}
570: %	\label{eq:AMF-simplified}	
571: %\end{eqnarray}
572: %o
573: The isotropic parts simplify, giving
574: \begin{eqnarray}
575: \Delta E^{Fl-nS}&=&-\frac{U-J}{2}\sum_{m\sigma}n_{m\sigma}^{2}-\frac{J}{4}M^{2}
576:   \nonumber \\
577:               & &+\frac{U-J}{2}N\overline{n}+\Delta E^\mathrm{aniso},\\
578: \Delta E^{Fl-S}&=&-\frac{U-J}{2}\sum_{m\sigma}n_{m\sigma}^{2}
579:            +\frac{U-J}{4}\frac{M^{2}}{2l+1} \nonumber \\
580:              & &+ \frac{U-J}{2}N\overline{n}+\Delta E^\mathrm{aniso}, \\
581: \Delta E^{FLL}&=&-\frac{U-J}{2}\sum_{m\sigma}n_{m\sigma}^{2}+\frac{U-J}{2}N
582:            \nonumber \\
583:            &  &+ \Delta E^\mathrm{aniso}
584: \label{eq:isotropic}
585: \end{eqnarray}
586: with the universal anisotropy contribution
587: \begin{eqnarray}
588: \Delta E^\mathrm{aniso}&=&\frac{1}{2}\sum_{mm^{\prime}\sigma\sigma^{\prime}}
589:   \Delta W_{mm^{\prime}}^{\sigma\sigma^{\prime}}n_{m\sigma}n_{m^{\prime}\sigma^{\prime}},\\
590: \Delta W_{mm'}^{\sigma\sigma'}&\equiv &\Delta U_{mm'} - \Delta J_{mm'} \delta_{\sigma\sigma'}.
591: \label{eqn:aniso}
592: \end{eqnarray}
593: is the anisotropic part of the interaction matrix elements.  These equations, up to the 
594: $\Delta W$ term, are the extensions of Eq. (\ref{eq:J=0}) to include isotropic exchange in
595: explicit form.
596: 
597: The first term in each of these expressions contains $-\frac{1}{2}{\tilde U}n_{m\sigma}^{2}
598: ({\tilde U}\equiv U-J)$ and hence has the appearance of a self-interaction correction.
599: Since the diagonal term of the interaction $E_I$ is specifically excluded, it does not
600: actually contain any self-interaction; in fact, the sign of the interaction $E_I$ is
601: {\it positive}.   (The double counting term does contain terms quadratic in $N$ which
602: must be interpreted as self-interaction.)  Nevertheless, the rewriting of the functional
603: leads to a self-interaction-like form, and that part of the functional will have an
604: effect related to what appears in the self-interaction-corrected LDA method, but
605: by an amount proportional to ${\tilde U}$ rather than a direct Coulomb integral, and depending 
606: on the difference of $n_{m\sigma}$ from the reference occupation, see Sec. IV.
607: % $\bar n, \bar n_{\sigma}, \frac{1}{2}$ respectively Eq. \ref{eq:isotropic}. 
608: 
609: \subsubsection{Fl-nS}
610: For Fl-nS, if we are restricted to integer occupations (so $n_{m\sigma}^2 = n_{m\sigma}$), 
611: then $\Gamma^2$ depends only on $N$, so the first term in $\Delta E^{Fl-nS}$ above 
612: depends only on $N$.  Then, up to corrections in $\Delta U$ 
613: and $\Delta J$, the state with the largest total spin moment 
614: will be favored; this is Hund's first rule.  In fact, even with the $\Delta U$ and 
615: $\Delta J$ terms, the $-{J}M^2/4$ term is still strongly dominant.  Except for $N = 7$, 
616: there are many ways to arrange electrons in orbitals which maximizes $S$.  Energy differences 
617: between these arrangements arise only from anisotropy ($\Delta U$ and $\Delta J$) 
618: and spin-orbit coupling.
619: 
620: \subsubsection{Fl-S}
621: In Fl-S, instead of having the $-{J}M^2/4$ term from Fl-nS which {\em favors} magnetism, 
622: there is a term $\frac{(U - J)}{4(2l+1)}M^2$ which {\em opposes} magnetism.  
623: This term (as in the $J=0$ case) comes from the occupation variance which wants to evenly 
624: distribute electrons across both spin channels.  Within LSDA there is something like a 
625: Stoner term of the form $-\frac14 I M^2$ which will compete with this Fl-S magnetic 
626: penalty.  We return to this aspect in later sections and the appendix.
627: 
628: 
629: \subsubsection{Spin-orbit Coupling; Particle-Hole Symmetry}
630: Without spin-orbit interaction, for a given $N$ there are many states that are 
631: degenerate for both double counting schemes.  Every value of $N$ has at least four 
632: degeneracies, those with $\pm L_z, \pm S_z$. 
633: 
634: Any state which has the same number of spin up as spin down electrons ($M=0$) gives the 
635: same energy from Fl-nS and Fl-S, since then $\bar n_\uparrow = \bar n_\downarrow 
636: = \bar n$ (the orbital potentials are distinct, however).  
637: Of course this fixed $N$, $M$=0 specification may contain many different
638: configurations.  Looking at results mentioned later, for Fl-S the ground state 
639: for an even number of electrons is $S_z = 0$ (so $\bar n_\sigma = \bar n$), thus the 
640: configuration which gives the Fl-S ground state has the same energy in Fl-S and Fl-nS. 
641: 
642: 
643: \section{Fractional Occupations}
644: 
645: Here we briefly discuss the effect of non-integer occupations in LSDA+U.  Taking a 
646: general set of occupations as $\{n_{m\sigma}\}$, we define a set of integer 
647: occupations, $\{\hat{n}_{m\sigma}\}$, and the fractional part of the occupations 
648: as $\gamma_{m\sigma} = n_{m\sigma} - \hat{n}_{m\sigma}$.  For illustration purposes 
649: we will choose the simplest possible scenario, where charge is transfered to an 
650: empty orbital $a$ from an occupied orbital $b$ both of the same spin, so that 
651: $0 < \gamma_{a\uparrow} = -\gamma_{b\uparrow}$, $\hat{n}_{a\uparrow} = 0$ and 
652: $\hat{n}_{b\uparrow} = 1$.  With this selection, $N_\sigma$ is unchanged 
653: (and therefore, $N$ and $M$ as well) so that $E_{dc}$ is unchanged.  Thus, the 
654: effect of the charge transfer is entirely contained in the $E_I$ term.  
655: Expanding $E_I$ for the general occupation set gives
656: %
657: \begin{equation}
658: 		E_I[\{n_{m\sigma}\}] - E_I[\{\hat{n}_{m\sigma}\}] =
659: 			U\gamma_{a\uparrow} (1 - \gamma_{a\uparrow})
660: \end{equation}
661: %
662: for the $J = 0$ case, and for $J \neq 0$ we find
663: %
664: \begin{eqnarray}
665: E_I[\{n_{m\sigma}\}]&-&E_I[\{\hat{n}_{m\sigma}\}]  \nonumber \\
666:   &=&
667:    \sum_{m\sigma} (W_{am}^{\uparrow\sigma} - W_{bm}^{\uparrow\sigma}) 
668:       \hat{n}_{m\sigma} \gamma_{a\uparrow} \nonumber \\
669:   & &            - W_{ab}^{\uparrow\uparrow} \gamma_{a\uparrow}^2.
670: 		\label{eq:FracOccExp}
671: \end{eqnarray}
672: %
673: 
674: The dominant term in (\ref{eq:FracOccExp}) is where $m\sigma = b\uparrow$.  
675: This term gives a contribution $W_{ab}^{\uparrow\sigma'}\gamma_{a\uparrow} 
676: \sim U\gamma_{a\uparrow}$ (since $U >> J$ for typical parameter choices, 
677: where other terms give contributions proportional to 
678: $(W_{am}^{\uparrow\sigma} - W_{bm}^{\uparrow\sigma})\gamma_{a\uparrow} \propto 
679: J\gamma_{a\uparrow}$.  The term with $m\sigma = a\uparrow$ is killed off by 
680: the factor of $\hat{n}_{a\uparrow}$, and the term in $\gamma^2$ is significantly 
681: smaller than the others for $\gamma < 0.5$.
682: 
683: This shows that there is an energy penalty for fractional occupation, 
684: proportional to $U$ and {\em linear in $\gamma$} at small $\gamma$.  Thus, in 
685: configuration space, the LSDA+U functionals have many local minima around 
686: configurations with integer occupations.  This result is fairly general.  
687: Even for charge transfer between orbitals of opposite spins, the linear energy 
688: penalty in $\gamma$ is still dominant over any additional terms coming from 
689: the double-counting or spin-orbit.
690: 
691: In practice, this gives the possibility that LSDA+U will get `stuck' in a 
692: local minimum with some configuration that may not be the true ground state.  
693: This behavior is not uncommon; LSDA+U has been reported\cite{CeLDAU} to find multiple 
694: local minima depending on the starting configuration. 
695: 
696: 
697: 
698: \section{Orbital Potential Matrix Elements}
699: Up to now only the energy functionals themselves were discussed. 
700: Now we return to the derivatives, the orbital potentials $v_{m\sigma}$.
701: It is simple to derive the exact expressions, and the interaction term $E_I$
702: common to all forms gives a potential $\Delta v_{m\sigma}$ which depends only
703: on the occupations of the {\it other} orbitals 
704:  $n_{m'\sigma'}, m'\sigma' \neq m\sigma$.
705: The potential resulting from the double counting term is functional specific, and
706: may contain a contribution from $n_{m\sigma}$ itself, {\i.e.} a self-interaction.
707:  
708: 
709: We confine our observations here to the subdivision (introduced just above) 
710: of the interaction into
711: an unitarily invariant isotropic part, and into an anisotropic part Eq.
712: (\ref{eq:UJbar}) that is much smaller and more difficult
713: to analyze.  As for the energy itself, it is convenient to add and subtract the
714: diagonal self-Coulomb and self-exchange, which makes the effect of the potential
715: much more transparent at the cost of introducing the misleading self-interaction
716: interpretation.
717: 
718: The potential matrix elements are
719: \begin{eqnarray}
720: \Delta v_{m\sigma}^{Fl-nS}&=&-\left(U-J\right)\left[n_{m\sigma} -\overline{n}\right]
721:          -\frac{J}{2}M\sigma \nonumber \\
722:                           & &+\Delta v_{m\sigma}^{aniso},\\
723: \Delta v_{m\sigma}^{Fl-S}&=&-\left(U-J\right)\left[n_{m\sigma} - \overline{n}_{\sigma}\right]
724:          +\frac{U-J}{2}\frac{M}{2l+1}\sigma \nonumber \\
725:                          & &+\Delta v_{m\sigma}^{aniso},\\
726: \Delta v_{m\sigma}^{FLL}&=&-\left(U-J\right)\left[n_{m\sigma} - \frac{1}{2}\right]
727:           +\Delta v_{m\sigma}^\mathrm{aniso},
728: \end{eqnarray}
729: with the anisotropic potential term 
730: \begin{eqnarray}
731: \Delta v_{m\sigma}^\mathrm{aniso}=\sum_{m^{\prime}\sigma^{\prime}}
732:          \Delta W_{mm^{\prime}}^{\sigma\sigma^{\prime}}
733:                 n_{m^{\prime}\sigma^{\prime}}.
734: \end{eqnarray}
735: The main occupation number dependent term, proportional to $n_{m\sigma}$,
736: has a self-interaction appearance and effect, as discussed above for the
737: functionals.  The differences in this term arise from the ``reference'' occupation
738: with which $n_{m\sigma}$ is compared to determine the potential shift.  The 
739: ``fluctuation'' $n_{m\sigma} - n_\mathrm{ref}$ is smaller for $Fl-S$ (AMF) than for $Fl-nS$
740: because the occupation for a given spin direction tends to be closer to $\bar{n}_{\sigma}$
741: than to $\bar{n}$.  The reference occupation for FLL is, like $Fl-nS$, spin-independent,
742: in fact, the reference is half-filling. In this sense, FLL seems more like a single-band
743: Hubbard model treatment than the other two functionals.
744: 
745: The other difference that is evident in this form is the spin dependence.
746: $Fl-nS$ additionally has a spin orientation dependent potential shift 
747: proportional to $J$ and to $M$ (similar to an LSDA treatment, but using $J$ instead of
748: the Stoner $I$) and enhances spin-splitting 
749: of the eigenenergies $\varepsilon$ accordingly.  
750: In Fl-S (AMF) the analogous term is $+\left(U-J\right)\frac{M}{2\left(2l+1\right)}\sigma$,
751: with a sign that impedes magnetism.
752: It can be simplified to  $\approx \frac{J}{2}M\sigma$
753: when $U\approx2\left(l+1\right)J$ . This expression illuminates the
754: reason that AFM is sometimes found to {\it decrease} the magnetic moment:
755: this term more or less cancels the spin
756: splitting of LSDA due to the opposite sign. What is left is a splitting
757: of occupied and unoccupied levels due to the $n_{m\sigma}$ term, which is almost
758: independent of $M$.  The effect is to support a spin-polarized solution,
759: but provide little discrimination between different $M$. Since the
760: spin polarization energy does not favor large $M$, we end up with
761: a tendency of a near degeneracy of different $M$ values,
762: as we already pointed out from purely energetic arguments.
763: For the case of a half-filled fully polarized shell $n_{ms}=\delta_{\sigma,1}$
764: (the case $N=M=7$ in Section \ref{Sec:NumericalResults}) the potential
765: matrix vanishes, which can be seen from $\overline{n}=\frac{1}{2}$
766: $\frac{M}{2\left(2l+1\right)}=\overline{n}=\frac{1}{2}$. However,
767: at the same time the energy contribution also vanishes $\Delta E^{Fl-S}=0$
768: (for integer occupations) and the Fl-S functional has no effect at
769: all.
770: 
771: The SIC term in FLL splits occupied
772: and unoccupied states symmetrically, while in the fluctuation functionals
773: the splitting happens with respect to the averaged occupation, which
774: is seen in the overall energy positions in Fig \ref{fig:Scatter}.
775: 
776: 
777: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
778: 
779: \section{Numerical Results}
780: \label{Sec:NumericalResults}
781: %
782: \begin{figure}[t]
783: \includegraphics[width=0.45\textwidth]{AngMom.eps}
784: \caption{(Color online) Angular momentum values of $S_z, L_z$, and $J_z$ 
785: of the lowest energy state for (a) AMF (Fl-S) and (b) FLL, with 
786: spin-orbit coupling.  Parameter values are $U = 8$, $J = 1$, $I = 0.75$.  
787: The AMF (Fl-S) curves do not follow 
788: Hund's rules, because the Stoner parameter is too small.  FLL follows 
789: Hund's rules exactly with these parameters.
790: }
791: \label{fig:AngSO}
792: \end{figure}
793: %
794: 
795: Following common terminology, for the remainder of the paper we refer to the
796: Fl-S functional simply as the AMF form.
797: We have taken values for $U_{mm'}$ and $J_{mm'}$ (used for Eu) from 
798: Ref. \onlinecite{EuN} (recalculated, to include more significant figures).  
799: These matrices are generated using 
800: $U = 8$ and $J = 1$ (values typical of rare earths) following the 
801: procedure given in the appendix of Ref. \onlinecite{CS}.
802: 
803: In our analysis of the AMF and FLL functionals, which are based on
804: an LSDA reference state, we include a Stoner term
805: %
806: \begin{eqnarray}
807: 	E(M) = -\frac14 I M^2
808: 	\label{eq:Stoner}
809: \end{eqnarray}
810: %
811: to model the magnetic effects of LSDA on the total energy.  The 
812: addition of this term helps to give a picture of the degree to which the
813: functionals reproduce Hund's first rule.  
814: Typical values of $I$ for ionized lanthanides are 0.75 eV, so we use 
815: this value for the calculations of this section.  Further discussion 
816: of the Stoner $I$ is included in the appendix.  
817: 
818: Spin-orbit interaction is included in the form
819: %
820: \begin{equation}
821: 	E_{SO} = \lambda \vec S \cdot \vec L \rightarrow
822:                 \sum_{m\sigma} S_z  L_z,
823: \end{equation}
824: %
825: where the second form applies when only $z$-components of moments
826: are treated, as is done in current implementations of the LSDA+U method.
827: Due to this restriction, LSDA+U often does not produce the correct multiplet 
828: energies in the atomic limit.  The visible result in LSDA+U band structures
829: is splittings of occupied, or unoccupied, correlated suborbitals that
830: can be as large as a few times $J$, and understanding the splittings is
831: not straightforward.  For $4f$ systems these splittings\cite{EuN,ReN}
832: may not be of much interest unless one of the correlated bands approaches
833: the Fermi level. In heavy fermion compounds, for example, LSDA+U results are
834: used to infer which parts of the Fermi surface has a larger amount of 
835: $f$ character.\cite{YbRh2Si2}  The same effects (eigenvalue splittings)
836: occur in $3d$ or $5f$ systems,
837: however, where they are expected to become more relevant but are 
838: masked by stronger banding tendencies.
839: 
840: %
841: \begin{figure}[t]
842: \includegraphics[width=0.45\textwidth]{FLL-AMF.eps}
843: \caption{(Color online) Shown here is $\Delta E^{FLL}$ plotted vs. 
844: $\Delta E^{Fl-S}$ for each of the 3432 configurations of $N=7$ electrons, using 
845: $U=8, J=1, I = 0.75$, all in eV.  The ordering of states is shown for Fl-S by 
846: counting from left to right, and for FLL by counting from bottom to top.  
847: Open squares show values for $U = 7$ and $J = 0$.     }
848: \label{fig:FLL-AMF}
849: \end{figure}
850: %
851: 
852: Here we consider values of $\lambda$ of 0 and 0.2 eV.  The magnitude 
853: of the spin-orbit interaction is not critical to the results; it mainly 
854: serves to break degeneracies.  Without the spin-orbit interaction, the 
855: ground state for any of the functionals at a given $N$ is degenerate 
856: with several other states.  For instance with $N=6$, the AMF functional
857: has states 
858: with $L_z=1, S_z=0$ and $L=11, S_z=0$ with the same lowest energy.
859: 
860: In Fig. \ref{fig:AngSO} are the ground states for both AMF and FLL with 
861: $U = 8$, $J = 1$ and $I = 0.75$.  The FLL and Fl-nS (not shown) schemes 
862: both reproduce Hund's rules exactly with these parameters.  AMF does not 
863: reproduce Hund's rules (in fact penalizes magnetism)
864: until $I$ is increased to around 1.5, which is 
865: somewhat larger than reasonable values of $I$.  If one expects LSDA+U to 
866: reproduce Hund's rules, then the AMF scheme performs rather poorly.  
867: For instance, at $N = 7$, Hund's rules ask that all electrons be 
868: spin-aligned, but the AMF ground state has only one unpaired spin  
869: due to the magnetic penalty appearing in Eq.
870: (\ref{eq:J=0}).  With these parameter choices, $U/(2l+1) > I$, so 
871: the AMF magnetic penalty wins over the Stoner energy.  This is likely to 
872: be the case for $3d$ transition metals as well, since $U_{3d}/(2l+1) 
873: \sim 1$eV, but it may not be as significant since $I$ for $3d$ elements 
874: is larger.
875: 
876: We examine the energetics in more detail in Fig. \ref{fig:FLL-AMF}, where 
877: $\Delta E$ for the AMF and FLL functional is plotted for every 
878: configuration for $N = 7$.  
879: The configurations fall into separate lines for each spin moment $M$, since
880: $E_{dc}$ depends only on $N$ and $M$ for both functionals.  
881: For the case of $J = 0$, all the states with a particular $M$ value 
882: collapse to a single energy value (the orbital index loses any impact), 
883: this is shown with the open squares.  
884: A value of $I$ was chosen so that the cancellation discussed in the 
885: previous paragraph is slightly broken.  
886: 
887: If we examine the $J = 0$ case first (the large open squares in Fig. \ref{fig:FLL-AMF}),
888: we see that the separation of states in FLL is much larger than AMF
889: (9 eV versus 3 eV), with $M = 7$ the lowest energy for FLL but highest for AMF.
890: This is a direct consequence of the magnetic penalty of AMF discussed previously.  
891: If $I$ were increased above 1 eV (keeping the other parameters fixed), then AMF would begin to 
892: favor the $M=7$ state by a small amount.  
893: 
894: Once $J$ is turned on, the degeneracy is split, and the 
895: configurations with a particular $M$ spread out around the $J = 0$ value.  The spread is especially large
896: for the highly degenerate $M=1$ value (from -5 to 8 eV), so that even if $I$ were larger than the typical LSDA value 
897: (in which case, with $J=0$ AMF would favor a high spin state) the large spread of $M=1$ values
898: would cause the low-spin states to be favored in AMF.  This spread is entirely coming from the $E_I$ term and is
899: independent of the double-counting choice.
900: Here we see for AMF a competition between $J$ and $I$:  $J$ is actually preferring 
901: a low spin configuration, in contrast to the conventional wisdom that $J$ 
902: increases the tendency for magnetism.  We see this same tendency occurs in FLL, as for $J = 0$ the separation
903: between $M=7$ and $M=1$ states is 9 eV, but with $J = 1$ this separation is reduced to 4 eV.  
904: Since in FLL the Hubbard $U$ does not penalize magnetic states the way AMF 
905: does, the presence of $J$ is not able to compete with $I$.  
906: This makes it clear why FLL is generally accepted to perform better for systems known 
907: to have high-spin states ({\it e.g.} Eu and Gd).  Conversely, FLL may be 
908: less successful at modeling low-spin states.  
909: 
910: As mentioned previously, it is fairly common for theoretical studies to replace $U$ and $J$ with effective parameters $\tilde U$ and $\tilde J$.  
911: For any double-counting term chosen, using these effective parameters will lower the energy of
912: the high-spin state relative to the low spin state as compared to using $U$ and $J$ directly.  With orbitals that are not highly localized,
913: such as $3d$ or $5f$ states it may be the case with FLL that the 
914: reduction of the energy separation between high-spin and low-spin caused by using $U$ and $J$ 
915: would allow for significant competition between magnetism and kinetic energy in LSDA+U.
916: 
917: We now have seen why and  how FLL and AMF perform differently in assigning a 
918: magnetic moment.  This may be of particular interest for studies of 
919: pressure-induced changes in magnetic moment, such as that seen in 
920: MnO\cite{Kasinathan07} without changes in orbital $M$ occupancy.  Applications of LSDA+U are more thoroughly discussed 
921: in Sec. \ref{sec:Applications}.
922: 
923: % Experimentally, MnO shows a moment collapse from $M$=5 to $M$=1 (or less) near
924: % 100 GPa.  
925: % Within LSDA, the moment decreases continuously with decreasing volume, from 
926: % high spin state to the low spin state.  {\bf Briefly describe MnO results.
927: % Maybe move this paragraph to Discussion.}
928: % One can argue that Hund's first rule (maximize $S$) is carried out well by 
929: % LSDA (for example, Mn in MnO; Gd) and that is not the job of $E_I$.  
930: 
931: 
932: %
933: \begin{figure*}[t]
934: \includegraphics[width=0.95\textwidth]{scatter-wide.eps}
935: \caption{(Color online) Scatter plot of all energies $\Delta E$ for 
936: all states in the (a) AMF(Fl-S), (b) FLL and (c) Fl-nS double-counting 
937: schemes, for $U = 8$, $J = 1$, and $I = 0.75$ (FLL and AMF only).  
938: Spin-orbit is neglected here.  For AMF, low spin states (black and 
939: red circles) appear as lowest energy configurations for all $N$, 
940: but this is not the case for FLL or Fl-nS.  The dashed lines indicate the
941: mean energy over configuration for each $N$; note that the variation with
942: $N$ is much less for FLL than for the other two functionals.
943: }
944: \label{fig:Scatter}
945: \end{figure*}
946: %
947: 
948: Shown in Fig. \ref{fig:Scatter} are scatter plots of the energies of all 
949: possible states for a given number of $f$ electrons with integer 
950: occupations.  SO is neglected, as it makes very minor changes to this 
951: picture by splitting some degeneracies.  The particle-hole symmetry of
952: each functional is apparent.
953: In Fl-nS and FLL, the ground state energy for 
954: $N = 7$ is roughly 3 eV lower than the next level, which are the 
955: (degenerate) ground states 
956: for $N = 6$ and $8$.  This is almost entirely due to the term depending on 
957: $M^2$ (either the $J$ term in (\ref{eq:Fl-nS}) or the Stoner term 
958: in FLL), because $M$ is large and at its maximum with 7 spins aligned.  
959: In AMF low spin states can be seen at the low end of the range 
960: for configurations at each $N$; the high spin states for $N$=6 and 7 are 
961: disfavored by 6-7 eV.  We see that the trend where AMF favors low-spin configurations
962: and FLL favors high-spin configurations shown for $N=7$ in 
963: Fig. \ref{fig:FLL-AMF} is present for all $N$.  The large spread of values for low spin configurations (black and red circles)
964: is seen clearly for AMF as they appear in both the lowest energy positions and the highest energy positions.  The high-spin configurations (large open symbols and triangles) 
965: are in the middle of each distribution for $N$.  For e.g. $N=5$, counting from the lowest energy, $M=1$ configurations are found first, followed by $M=3$ configurations then the $M=5$ configurations are found (with the trend reversing counting up to the highest energy states).  In FLL, the lowest energy configurations for $N \neq 7$ are still the configurations with maximum spin for a given $N$, and states with lower spins are found in succession.  Again using $N=5$ as an example, the $M=5$ configurations are lowest in energy, and then $M=3$ configurations are seen at energies lower then $M=1$ states.
966: 
967: 
968: \section{Discussion}
969: 
970: In this paper we have tried to clarify the behavior of the various functionals that are used in the 
971: LSDA+U method, we have compared the functionals formally in certain limits, 
972: we have presented the orbital potentials that arise, and we have
973: analyzed the total energy corrections that LSDA+U functionals apply 
974: to LSDA total energies, given a set of occupation numbers.  
975: The Fl-nS functional which was originally introduced 
976: strongly favors spin-polarized states as does the commonly used FLL functional. 
977: The other most commonly used functional besides FLL, Fl-S (AMF), has characteristics that
978: tend to {\it suppress} moment formation or reduce the magnitude of the moment.  
979: When analyzed, this AMF functional shows positive energy penalties to magnetism that compete 
980: with the magnetic tendencies of the LSDA functional, and when $J > 0$ non-magnetic 
981: solutions become even more
982: likely to win out.  We have provided a short analysis of the behavior when $J=0$ is used.  While this case is instructive, we advise against its use; it is just as simple to do the full $J \neq 0$ calculation.
983: 
984: When LSDA+U is applied to correlated insulators in the strong coupling regime, it provides
985: a very good picture of the system at the band structure (effective one-electron) level.
986: The initial successes include the $3d$ transition metal monoxides MnO, FeO, CoO, and NiO,
987: for which the LSDA description is very poor.  Other early successes included the insulating
988: phases of the layered cuprates that become high temperature superconductors when doped, and
989: the unusual magnetic insulator KCuF$_3$, which was the first case where crucial orbital
990: ordering was reproduced.  LSDA+U is not a satisfactory theory of
991: single particle excitations of such systems, but nevertheless provides a realistic picture
992: of the underlying electronic structure.
993: 
994: The more interesting, and more difficult, cases now lie between the strongly correlated limit of wide-gap 
995: magnetic insulators and weakly correlated regime that is well described by LSDA.  Some
996: of these are metals, some are unconventional insulators, and many lie near the metal-insulator
997: borderline.  It is for these intermediate cases that it becomes essential, if applying the
998: LSDA+U approach, to understand what the method is likely to do, and especially to understand
999: the tendencies of the various choices of functional.  This is what we have tried to clarify
1000: in this paper.  As a summary, we will provide an overview of an assortment of results that have 
1001: appeared in the literature for systems that lie somewhere in the intermediate correlation
1002: regime.
1003: 
1004: 
1005: \subsection{Examples of LSDA+U behavior from applications}
1006: \label{sec:Applications}
1007: \subsubsection{Strongly correlated insulators.}
1008: {\it Cuprates.} The insulating phase of the cuprate class of high temperature
1009: superconductors comprised the ``killer app'' that served to 
1010: popularize\cite{OriginalLDAU,VIA1992} the LSDA+U method,
1011: and in the intervening years the method has been applied to cuprates and other correlated insulators too many times to cite.  
1012: Simply put, in cuprates it produces the 
1013: Cu $d^9$ ion and accompanying insulating band structure.\cite{VIA1992,rosner}  
1014: The hole resides in the
1015: $d_{x^2-y^2}$ orbital and is strongly hybridized with the planar oxygen $p_{\sigma}$
1016: orbitals, as much experimental data was indicating.
1017: 
1018: {\it MnO.}
1019: Experimentally, MnO shows at room temperature a moment collapse 
1020: from $M=5$ to $M=1$ (or less), a volume collapse,
1021: and an insulator-to-metal transition, near 100 GPa; this is the classic Mott transition.
1022: Within LSDA, the moment decreases continuously with decreasing volume,\cite{Cohen1997} 
1023: from the high spin (HS) state 
1024: to a low spin (LS) state. The insulator-to-metal transition occurs at much too low a
1025: pressure (without any other change). A volume collapse is predicted, although 
1026: the pressure is significantly overestimated (150 GPa).
1027: 
1028: The application of LSDA+U in its FLL flavor has been applied and analyzed in
1029: detail,\cite{Kasinathan07} and provides a different picture in several ways. The ambient
1030: pressure band gap is improved compared to experiment. The volume collapse transition 
1031: occurs around 120 GPa and is accompanied by a moment collapse from $M=5$ to $M=1$. 
1032: The nature of this (zero temperature) transition is insulator to insulator, while the
1033: experimental data indicate an insulator-to-metal transition at room temperature. 
1034: The zero temperature transition might indeed be insulator-to-insulator;
1035: such a phase transition would be a type that LSDA+U should work well for.
1036: It is also possible that the static mean-field approximation underlying LSDA+U, which favors
1037: integer occupations and hence insulating solutions, has too strong a tendency and fails
1038: to describe this transition.  This question could be settled by studying experimentally
1039: the Mott transition at low temperature
1040: 
1041: Even more unexpected than the insulator to insulator aspect is the LSDA+U 
1042: prediction is that the low spin state has 
1043: an unanticipated orbital occupation pattern,\cite{Kasinathan07} being one
1044: in which every $3d$ orbital remains singly occupied (as in the high spin state).
1045: but spin in two orbitals antialign with those in the other three orbitals.
1046: This state is obtained simply from the $M=5$ high spin state by
1047: flipping the spins of two of the orbitals. The resulting density remains spherical, but
1048: the spin density exhibits an angular nodal structure leading at the same time
1049: to a high degree of polarization of the spin-density but a low total moment ($M=1$).
1050: This solution (being the high pressure ground state in LSDA+U) can be traced\cite{Kasinathan07} back
1051: to the interplay between symmetry lowering due to the antiferromagnetic order
1052: (cubic lowered to rhombohedral) 
1053: and the anisotropy part of the interaction Eq. \ref{eqn:aniso}. The symmetry lowering
1054: lifts the cubic grouping ($t_{2g}$ and $e_g$ manifolds), thus allowing a higher number
1055: of allowed occupation patterns.
1056: 
1057: The anisotropic part of the interaction is responsible\cite{Kasinathan07} 
1058: for Hund's second rule ordering
1059: of states, which has the tendency to increase the mutual distance of each pair
1060: of electrons. If the overall energetics (band broadening and kinetic effects) 
1061: reduce the gain of energy due to spin-polarization, then Hund's 
1062: first rule may become suppressed and the result is a low spin state.
1063: The anisotropic interaction is however not influenced
1064: by this suppression, since it is a local term proportional to a parameter $J$.
1065: It will enforce a Hund's second rule like separation of the electrons under the
1066: low spin condition, and thus can be shown to result exactly in the occupation pattern observed
1067: for MnO. In a sense the low spin state is an example of Hund's second rule without 
1068: Hund's first rule.
1069: 
1070: {\it FeO, CoO, NiO.} Together with MnO, these classic Mott (or `charge transfer')
1071: insulators have been prime applications of the LSDA+U 
1072: method.\cite{AnisimovNiO94,ShickLDAU,Bengone00,Dobysheva04}  The behavior of the
1073: open $3d$ shell in these compounds has not been analyzed in the detail that was
1074: done for MnO, however.
1075: 
1076: \subsubsection{Metals}
1077: Correlated metals involve carriers that can move, hence they invariably involve
1078: fluctuations, in occupation number, in magnetic moment, in orbital occupation, etc.
1079: It cannot be expected that a self-consistent mean field treatment such as LSDA+U can
1080: answer many of the questions raised by their behavior.  However, there is still 
1081: the question of whether LSDA+U can provide a more
1082: reasonable starting point than LSDA alone in understanding these metals.  
1083: In our opinion, this remains 
1084: an open question, but one for which some evidence is available.
1085: 
1086: The Fe-Al system has provided one platform for the application of LSDA+U to moderately
1087: correlated metals.  The systems treated include the Fe impurity in Al (Kondo system,
1088: experimentally), and the compounds Fe$_3$Al, FeAl, and FeAl$_3$. The calculated behavior is too 
1089: complex to summarize here.  The LSDA+U result will, generally speaking, be likely to
1090: give a good picture of a Kondo ion when it produces an integer-valent ion with a
1091: large value of U.  Both FLL and AMF
1092: functionals have been applied in this regime,\cite{Mohn01,Lechermann04} with substantially differing
1093: results, leading one to question whether either is more realistic than simple LSDA.  
1094: Results are also sensitive to volume, {\it i.e.} whether using the experimental
1095: lattice constant or the calculated equilibrium value, and the calculated equilibrium
1096: is different from LSDA and LSDA+U. One result was that, for moderate 
1097: U$_{Fe} \sim$ 3-4 eV,  AMF strongly 
1098: reduces the magnetic moment, while FLL does not.\cite{Lechermann04}
1099: Another application found that the magnetism disappeared within a certain range of
1100: intermediate values of $U_{Fe}$, that is, it was magnetic around small $U_{Fe}$ and
1101: also again at large coupling,\cite{Mohn01} but non-magnetic between.
1102: %destroys magnetic order at $U = 3.7$ eV, but magnetic order with a larger moment 
1103: %appears at $U = 5.5$ eV.  This is likely do to the complex interplay between LDA and 
1104: %the ``+U'' part of the functional, as the moment changes (with respect to $U$) 
1105: %from $e_g$ and $t_{2g}$ to being primarily $t_{2g}$.\cite{Mohn01}
1106: 
1107: % Study of Li$_{3-x}$Fe$_{x}$N.  They compare LSDA, GGA, SIC, AMF for a few different
1108: % structures.  AMF with a large U decreases the spin moment on the irons fairly
1109: % consistently.  Also it increases the orbital moment.\cite{Novak02}
1110: 
1111: \subsubsection{Moderately strongly interacting oxides.} 
1112: Trying to address seriously the electronic
1113: structure of intermediate coupling oxides, which are often near the metal-insulator
1114: transition, is a challenge that has begun to be addressed more directly.  The peculiar
1115: Na$_x$CoO$_2$ system, which becomes superconducting when hydrated (water intercalates 
1116: between CoO$_2$ layers) is one example.  One set of studies showed no appreciable 
1117: difference between FLL and AMF,\cite{KwanWoo04} with both predicting charge disproportionation on the 
1118: Co ion for $x$=$\frac{1}{3}$ and $\frac{1}{2}$ for $U\approx$ 2.5-3 eV.   It is likely
1119: that this compound presents a case where the interplay between LSDA and $U$ has effects
1120: that are not fully understood.  Also, it is unclear why there is so little difference
1121: between the FLL and AMF functionals in this system.
1122:  
1123: The compound Sr$_2$CoO$_4$ is another example.  Both functionals show a collapse of the 
1124: moment\cite{KwanWoo06} around $U$ = 2.5 eV, related to the metal-half metal transition that occurs,
1125: but the result for the moments ($M$(AMF) $<$ $M$(FLL)) bears out the tendency of AMF to
1126: penalize magnetic moments..  
1127: The fixed spin-moment calculation in Fig. 9 in Ref. \onlinecite{KwanWoo06} is instructive too, showing the competition 
1128: between LSDA magnetic energy and AMF magnetic penalty.  Also it shows the creation of 
1129: local minima around M = integer values that LDA+U introduces.
1130: 
1131: \subsubsection{f electron materials.}  
1132: {\it 4$f$ systems.}
1133: These metals often display the correlated electron physics of
1134: a magnetic insulator at the band structure level: background conduction bands provide the
1135: metallic nature, while the correlated states have integer occupation.  The LSDA+U method
1136: seems to be a  realistic method for placing the $f$ states closer to where they
1137: belong (away from the Fermi level).  Gd is a good example, which has been studied at
1138: ambient pressure and compared to photoemission data\cite{ShickLDAU} and magnetic dichroism
1139: data.\cite{HarmonGd,Alouani07}  The LSDA+U method has also been applied up to extremely high pressure
1140: to assess where the `Mott transition' in the $4f$ bands is likely to occur.
1141: The LSDA+U method has also been applied to heavy fermion metals, for example Cu and U
1142: compounds,\cite{Oppeneer97} PrOs$_2$Sb$_{12}$,[\onlinecite{Harima05}] and  
1143: YbRh$_2$Si$_2$~[\onlinecite{YinYb}].  In such systems the LSDA+U method may even 
1144: provide a good estimate of which
1145: itinerant states at the Fermi level are strongly coupled to the localized $f$ states, {\it i.e.}
1146: the Kondo coupling matrix elements.  These
1147: $4f$ systems may become heavy fermion metals (YbRh$_2$Si$_2$) or novel heavy fermion
1148: superconductors (YbAlB$_4$), or they may remain magnetic but
1149: otherwise rather uninteresting metals (Gd). 
1150: 
1151: {\it $5f$ systems.} 
1152: A variety of application of the LSDA+U method to $5f$ systems, and especially Pu, have
1153: been presented.\cite{Bouchet00,Savrasov00,Price00,Shick05,Shorikov05}  
1154: Given the complexity of the phase diagram of elemental Pu, together
1155: with claims that dynamic correlation effects must be included for any realistic
1156: description of Pu, a more critical study of Pu would be useful.
1157: 
1158: % This paper apparently uses FLL-nS as the double-counting term for 
1159: % $\delta$- and $\alpha$-Pu.  It's a very long paper and I don't know much about Pu, aside 
1160: % from its phase diagram is very complicated but they seem to get some magnetic 
1161: % results consistent with experiment. \cite{Shorikov05}
1162: 
1163: % LDA+U applied to $\delta$-Pu I don't recognize the functional, it looks more like the Dudarev version of LDA+U?\cite{Bouchet00}
1164: 
1165: % LDA+U I don't know what DC was used, maybe we can ask Savrasov.  I think it was probably FLL, since he seems to like that better than AMF.  They find large spin and orbital moments that cancel. \cite{Savrasov00}
1166: 
1167: % LDA+U (AMF) applied to $\delta$-Pu finds nearly zero orbital and spin moments, which are shown as a function of U in Fig. 1. AMF quickly kills off the moment, even for small U (although maybe it's not so small for a $5f$ system). \cite{Shick05}
1168:   
1169: 
1170: \section{Acknowledgments}
1171: We have benefited from discussion on various aspects of this work with
1172: M. Johannes, J. Kune\v{s}, A. K. McMahan, I. I. Mazin, and G. Sawatzky.
1173: This project was supported by DOE through the Scientific Discovery through 
1174: Advanced Computing (grant DE-FC02-06ER25794), by DOE grant DE-FG02-04ER46111, and by the Computational Materials Science Network.
1175: 
1176: \appendix
1177: \section{Calculation of the Stoner $I$ for $3d$ and $4f$ Shells}
1178: \label{Appendix}
1179: 
1180: %
1181: \begin{figure}
1182: \begin{centering}
1183: \includegraphics[width=0.30\textwidth]{I}
1184: \par\end{centering}
1185: 
1186: \caption{\label{fig:Inl}Shell-Stoner integrals for the $3d$ and $4f$ atoms.
1187: For explanations see text.}
1188: 
1189: \end{figure}
1190: %
1191: 
1192: 
1193: The Stoner parameter $I$ is a well established quantity. For 
1194: metals its value is obtained by a second order expansion
1195: of the LSDA xc-energy around the non-magnetic solution, resulting
1196: in a Fermi surface averaged integral of the radial wave functions
1197: with the xc-kernel.\cite{Janak77} LSDA+U is usually
1198: applied to describe insulating states, where the Fermi surface vanishes.
1199: In the context of discussing the LSDA contribution to the
1200: energy of a correlated $d$- or $f$-shell it is more natural to consider
1201: the energy contribution from the localized shell. This leads to a
1202: derivation of the Stoner-$I$ similar to the formulation of Janak
1203: but adapted to atom-like situations.
1204: 
1205: Seo presented\cite{Seo06} the second order perturbation
1206: theory of the spin polarization in DFT, which results
1207: in explicit expressions for the shell exchange parameter $I_{nl}$ that are
1208: applicable to atom like situations. In this work a numerical estimate
1209: for $I_{nl}$ was derived indirectly from exchange splittings and
1210: spin polarization energies taken from DFT calculations. The idea behind
1211: this perturbation theory, the expansion the xc-energy around the
1212: spherically averaged non-magnetic density of the shell under consideration,
1213: was also discussed in the appendix of Kasinathan {\it et al.}\cite{Kasinathan07}
1214: and leads to 
1215: $\Delta E_{xc}\approx-\frac{1}{4}I_{nl}M^{2}$
1216: with the shell-Stoner integral 
1217: \begin{eqnarray}
1218: I_{nl}=-\frac{1}{2\pi}\int K_{0}\left(r\right)\left[R_{nl}\left(r\right)\right]^{4}r^{2}dr
1219: \label{eq:shellStoner}
1220: \end{eqnarray}
1221: \begin{eqnarray}
1222: K_{0}\left(\vec r,\vec r'\right)&=&\left.\frac{\delta^{2}E_{xc}}{\delta m\left(\vec r\right) 
1223:   \delta m\left(\vec r'\right)}\right|_{n^{\mathrm{spher}},m=0} \nonumber\\
1224:  &\rightarrow& K_{0}(\vec r) \delta(\vec r - \vec r').
1225: \end{eqnarray}
1226: The last expression applies for a local approximation ({\it viz.} LSDA) to
1227: $E_{xc}$. $K_{0}(\vec r,\vec r')$ is a magnetization-magnetization
1228: interaction, directly analogous to the second functional derivative of the
1229: DFT potential energy with respect to $n(\vec r)$, which is the Coulomb
1230: interaction $e^2/|\vec r - \vec r'|$ plus an `xc interaction' arising from
1231: $E_{xc}$.
1232: 
1233: For a more detailed discussion of the parameter $I_{nl}$ we performed
1234: LSDA calculations for free atoms and ions and explicitly calculated
1235: $I_{nl}$ from Eq. (\ref{eq:shellStoner}). It turns out that
1236: $\Delta E_{xc}(M)$ given above is by far the largest $M$-dependent term of
1237: the energy expansion. The spin polarization energy of isolated atoms/ions
1238: with spherical $M$ is well described by this estimate with an error
1239: smaller than $5-10\%$. The resulting shell-Stoner integrals $I_{nl}$
1240: have very similar values compared to the ones obtained from the theory
1241: for the metallic situation. (Note, however, that there is a factor
1242: of 2 difference in the definition of the Stoner $I$ in some
1243: of the publications.) 
1244: 
1245: 
1246: For the $3d$ transition element series we get values $I_{nl}$ ranging from
1247: 0.62 eV for Sc to 0.95 eV for Zn
1248: (see Fig. 4). These values increase across the series by $\approx$ 0.15 -- 0.20 eV,
1249: when the exchange only LSDA is used, pointing to a reduction due to
1250: (LDA-type) correlation effects when the full xc-kernel is used. For
1251: the $4f$ series the shell-Stoner integrals vary from 0.58 eV for Ce to
1252: 0.75 eV for Yb. The LDA correlation effects amount to $10\%$ of these
1253: values. The values obtained depend on the choice of the reference
1254: system, which serves as zeroth order in the functional expansion.
1255: For instance for the $3^{+}$-ions of the $4f$-series $I_{4f}$ is
1256: increased by $6-20\%$ with respect to the neutral atoms.
1257: 
1258: \bibliographystyle{apsrev}
1259: \bibliography{ldau}
1260: 
1261: \end{document}
1262: 
1263: