1: \documentclass[prl,twocolumn,showpacs]{revtex4}
2: \usepackage[dvips]{graphicx}
3: \usepackage{latexsym}
4: \usepackage{bm}
5: %\renewcommand{\topmargin}{.0in}
6: \begin{document}
7: \newcommand{\fig}[2]{\includegraphics[width=#1]{#2}}
8: %\newcommand{\fig}[2]{}
9: \newcommand{\pprl}{Phys. Rev. Lett. \ }
10: \newcommand{\pprb}{Phys. Rev. {B}}
11: \newcommand{\be}{\begin{equation}}
12: \newcommand{\ee}{\end{equation}}
13: \newcommand{\bea}{\begin{eqnarray}}
14: \newcommand{\eea}{\end{eqnarray}}
15: \newcommand{\nn}{\nonumber}
16: %
17: \newcommand{\la}{\langle}
18: \newcommand{\ra}{\rangle}
19: \newcommand{\dg}{\dagger}
20: \newcommand{\upa}{\uparrow}
21: \newcommand{\dna}{\downarrow}
22:
23:
24: \title{Electron correlation and Fermi surface topology of Na$_x$CoO$_2$}
25:
26: \author{Sen Zhou$^1$, Meng Gao$^1$, Hong Ding$^1$, Patrick A. Lee$^2$,
27: and Ziqiang Wang$^1$}
28: \affiliation{$^1$ Department of Physics, Boston College, Chestnut Hill,
29: MA 02467}
30: \affiliation{$^2$ Department of Physics, Massachusetts Institute of Technology,
31: Cambridge, MA 02139}
32:
33:
34: \date{\today}
35:
36: \begin{abstract}
37:
38: The electronic structure of Na$_x$CoO$_2$ revealed by recent photoemission
39: experiments shows important deviations from band theory predictions.
40: The six small Fermi surface pockets predicted by LDA calculations have not
41: been observed as the associated $e_g^\prime$ band fails to cross the
42: Fermi level for a wide range of sodium doping concentration $x$. In addition,
43: significant bandwidth renormalizations of the $t_{2g}$ complex
44: have been observed. We show that these discrepancies are due to strong
45: electronic correlations by studying the multi-orbital Hubbard model
46: in the Hartree-Fock and strong-coupling Gutzwiller approximation.
47: The quasiparticle dispersion and the Fermi surface topology obtained
48: in the presence of strong local Coulomb repulsion are in good agreement
49: with experiments.
50:
51: \typeout{polish abstract}
52: \end{abstract}
53:
54: \pacs{71.27.+a, 71.18.+y, 74.25.Jb, 74.70.-b}
55:
56: \maketitle
57: The cobaltates (Na$_x$CoO$_2$) are doped 3d transition metal oxides in which
58: the Co atoms form a layered hexagonal lattice structure. In contrast to
59: the high-$T_c$ cuprates, where the Cu$^{2+}$ has a 3d$^9$ configuration and
60: occupies the highest {\it single} $e_g$ ($d_{x^2-y^2}$) orbital near the
61: Fermi level, the cobaltates are {\it multi-orbital} systems where
62: the Co$^{4+}$ is in the 3d$^5$ configuration, occupying the three
63: lower $t_{2g}$ orbitals, similar to the ruthenates (Sr$_2$RuO$_4$).
64: The unexpected discovery \cite{takada03} of a superconducting phase of yet
65: unknown origin in hydrated Na$_x$CoO$_2$ around $x\sim0.3$ has generated
66: renewed interests in this material. However, such basic issues as the
67: low energy electronic structure and Fermi surface topology in the
68: cobaltates have not been well understood.
69: Band structure (LDA) calculations \cite{singh00}
70: find that the trigonal symmetry of the Co site in the triangular
71: lattice splits the three $t_{2g}$ complex into an $a_{1g}$
72: and two degenerate $e_g^\prime$ states at the zone center ($\Gamma$ point).
73: The LDA predicts a large Fermi surface (FS) associated with the $a_{1g}$
74: band enclosing the $\Gamma$ point and
75: six small FS pockets of mostly $e_g^\prime$ character near the K points
76: \cite{singh00,pickett}.
77:
78: However, recent angle-resolved photoelectron spectroscopy (ARPES)
79: measurements on the cobaltates revealed
80: only a {\it single} hole-like FS centered around the $\Gamma$ point for
81: a wide range of Na concentration $x$ \cite{hbyang2,hasan2,hasan,hbyang1}.
82: The area enclosed
83: by the FS exhausts the Luttinger volume, which is consistent with the
84: observation that the dispersion of the $e_g^\prime$ band associated
85: with the LDA FS pockets lies below and
86: never crosses the Fermi level \cite{hbyang2}.
87: The absence of the FS pockets is unexpected and puts serious
88: constraints on several proposed theories of non-phonon mediated
89: superconductivity as well as magnetic properties based on the
90: nesting conditions of the FS pockets \cite{kuroki04,johannes04,yanase}.
91: Furthermore, the measured quasiparticle bandwidths are significantly
92: narrower than the LDA predictions \cite{hbyang2,hasan2}.
93: These fundamental discrepancies between ARPES and LDA suggest that the
94: effects of strong electronic correlations are important in the cobaltates.
95: The effects of local Coulomb repulsion $U$ has been considered
96: in the LSDA+U approach, which indeed finds the absence of the
97: small FS pockets \cite{PHZhang}. However, the latter is tied to
98: the fully polarized ferromagnetic state in the LSDA+U theory which
99: gives spin-split bands and a spin polarized FS with an area twice as large.
100: This is inconsistent with ARPES
101: and likely an artifact of the LSDA+U approximation.
102: A recent calculation based on the multi-orbital Hubbard model
103: and the dynamical mean-field theory finds that the FS pockets
104: become even larger in size than the LDA predications \cite{liebsch}.
105:
106: The focus of the present work is to explain how strong correlations
107: drive orbital polarization and the band narrowing observed in ARPES.
108: We adopt a multi-orbital
109: Hubbard model description where the noninteracting part is determined by
110: fitting the LDA band structure. The interacting part contains both
111: the intra ($U$) and the inter-orbital ($U^\prime$) local Coulomb repulsion
112: as well as the Hund's rule coupling $J_H$.
113: First, a basis independent Hartree-Fock (HF) calculation is performed which is
114: in essence a LDA+U calculation in the paramagnetic phase. We find that
115: for $U^\prime$ much less than U, multi-orbital occupation is favored in order
116: to reduce the cost of double occupation. As a result, the HF self-energy
117: renormalizes the atomic level spacing in such a way that the size of the
118: FS pockets associated with the $e_g^\prime$ band grows. This trend is however
119: reversed when $U^\prime$ grows and becomes
120: comparable to U. In the HF theory, the size of the $e_g^\prime$ FS
121: pockets begins to shrink for $U^\prime/U > 3/5$. To correctly capture
122: the physics of strong correlation for large $U$ and $U^\prime$, we generalize
123: the Gutzwiller approximation to the case of multi-orbitals.
124: We find that in the strong-coupling regime, orbital polarization is
125: tied to Pauli-blocking, i.e. the orbital occupation dependence of the
126: Gutzwiller band renormalization factors. We obtain both band narrowing
127: and the disappearance of the FS pockets in good agreement with the
128: ARPES experiments.
129:
130: We start with the multi-orbital tight-binding model on a two-dimensional
131: triangular lattice,
132: %
133: \begin{equation}
134: H_0=-\sum_{ij,\sigma}\sum_{\alpha \beta}t_{ij,\alpha \beta}
135: d^{\dagger}_{i\alpha \sigma}d_{j\beta \sigma}
136: -{\Delta \over 3}\sum_{i,\sigma}\sum_{\alpha \neq \beta}
137: d^{\dagger}_{i\alpha \sigma}d_{i\beta \sigma},
138: %\nonumber \\
139: %&-&\mu_f \sum_{i,\alpha,\sigma}d^{\dagger}_{i\alpha \sigma}d_{j\alpha \sigma},
140: \label{h-tb}
141: \end{equation}
142: %
143: where the operator $d^{\dagger}_{i\alpha\sigma}$ creates an electron in the
144: $\alpha$ orbital with spin $\sigma$ on the Co site and
145: $t_{ij,\alpha\beta}$ is the hopping integral between the $\alpha$
146: orbital on site $i$ and the $\beta$ orbital on site $j$.
147: The relevant valence bands near the FS consist of
148: the Co $t_{2g}=\{d_{xy},d_{yz},d_{zx}\}$ orbitals and have an electron
149: occupancy of $5+x$. The $\Delta$ in Eq.~(\ref{h-tb}) describes
150: the crystal field due to trigonal distortion that splits the $t_{2g}$
151: complex into a lower $a_{1g}$ singlet and a higher $e'_g$ doublet,
152: where $a_{1g}=(d_{xy}+d_{yz}+d_{zx})/\sqrt{3}$,
153: and $e'_g=\{(d_{zx}-d_{yz})/\sqrt{2}$, $(2d_{xy}-d_{yz}-d_{zx})/\sqrt{6}\}$.
154: For convenience, we will work in the hole-picture via a particle-hole
155: transformation $d \rightarrow \tilde d^\dagger$, in which the band
156: filling of holes is $1-x$. The structure of
157: the tight-binding Hamiltonian in k-space on the triangular lattice is
158: \begin{equation}
159: H_0=\sum_{k,\sigma,\alpha\beta}
160: {\bf K}^d_{\alpha\beta}(k)
161: \tilde d^{\dagger}_{k\alpha\sigma}\tilde d_{k\beta\sigma}
162: +{\Delta\over 3}\sum_{k,\sigma,\alpha\neq\beta}
163: \tilde d^{\dagger}_{k\alpha\sigma}\tilde d_{k\beta\sigma}.
164: %\nonumber \\
165: %&+&\mu_f\sum_{k,\alpha,\sigma}d^{\dagger}_{k\alpha\sigma}d_{k\alpha\sigma}.
166: \label{hk-tb}
167: \end{equation}
168: %
169: The hopping matrix $\bf K$ in the $t_{2g}$ basis is given by
170: %
171: \begin{equation}
172: {\bf K}^d(k)=\left ( \begin{array}{clcr}
173: \varepsilon(t,1,2,3) &\varepsilon(t',3,1,2) &\varepsilon(t',2,3,1)\\
174: \varepsilon(t',3,1,2) &\varepsilon(t,2,3,1) &\varepsilon(t',1,2,3)\\
175: \varepsilon(t',2,3,1) &\varepsilon(t',1,2,3) &\varepsilon(t,3,1,2)
176: \end{array} \right),
177: \label{htk}
178: \end{equation}
179: with $(1,2,3)=(k_1,k_2,k_3)$, $k_1=\sqrt{3}k_x/2-k_y/2$,
180: $k_2=k_y$, $k_3=-k_1-k_2$, and
181: $\varepsilon(t,1,2,3)=2t_1\cos{k_1}+2t_2(\cos{k_2}+\cos{k_3})
182: +2t_3\cos{(k_2-k_3)}+2t_4[\cos{(k_3-k_1)}+\cos{(k_1-k_2)}]
183: +2t_5\cos{(2k_1)}+2t_6[\cos{(2k_2)}+\cos{(2k_3)}]+\cdots$.
184: The $(t,t')$ denote the (intra,inter)-orbital hopping.
185:
186:
187: Fig.{\ref{fig1}} shows the fitting of the tight-binding dispersions
188: obtained by diagonalizing Eq.~(\ref{hk-tb}) to the LDA band structure
189: at $x$=1/3 \cite{pickett}. We note that the fit with up to
190: third-nearest-neighbor (NN) hopping or more describes the LDA bands
191: quite well. On the other hand, the
192: tight-binding model cannot reproduce completely the LDA dispersions
193: even with up to eighth-NN hopping.
194: The discrepancy is most pronounced along the M-K
195: direction where the two $e'_g$ bands cross in the tight-binding fit
196: (Fig.~1a). Similar disagreement can be traced back to the previous
197: tight-binding fits \cite{yanase,liebsch,johannes}.
198: We believe the difficulty arises from the hopping path via the O 2$s$ and 2$p$
199: orbitals. Nevertheless, the tight-binding model works
200: very well at low energies near the Fermi level.
201: The FS consists of a cylindrical sheet around the
202: $\Gamma$-point and six hole pockets near the K-points as shown in Fig.~1b-d.
203: The central FS has a dominant $a_{1g}$ character while the six FS
204: pockets are mainly of the $e_g^\prime$ character.
205: The hopping integrals obtained from the fit with up to third-NN are
206: $t$=(-44.6, -9.0, 36.2, 5.9, 57.9, 36.7)meV and
207: $t'$=(-157.8, -30.2, 37.1, 9.2, -11.9, -21.0)meV.
208: The crystal-field splitting $\Delta$ is chosen to be 0.01eV.
209: In the rest of the paper, we use these parameters for $H_0$.
210: Our results are insensitive to these values provided that
211: they provide a good fit of the LDA band structure near the FS.
212: Note that although holes are evenly distributed among the three
213: $t_{2g}$ orbitals, in the $a_{1g}$ and $e_{g}^\prime$ basis
214: (hereafter referred to as the $\{a\}$ basis), the hole occupations
215: are 0.123 ($e_g^\prime$) and 0.421 ($a_{1g}$) respectively.
216: Despite its higher orbital energy, the $a_{1g}$ hole
217: orbital has a higher occupation due to its larger bandwidth.
218: %
219: \begin{figure}
220: \begin{center}
221: %\vskip-2.8cm
222: %\hspace*{-3.8cm}
223: \fig{3.3in}{fig1.eps}
224: \vskip-2.4mm
225: \caption{Tight-binding fits to the LDA band structure at $x=1/3$.
226: (a) The fitted band dispersions with up to 3rd, 5th and 8th NN hopping.
227: The corresponding Fermi surfaces are plotted in (b), (c), and (d)
228: respectively.}
229: \label{fig1}
230: \vskip-9mm
231: \end{center}
232: \end{figure}
233: %
234:
235: The correlation effects are described by
236: the multi-orbital Hubbard model $H=H_0+H_I$, where
237: $H_0$ is the tight-binding Hamiltonian in Eq.(\ref{hk-tb}),
238: $H_I$ represents the local Coulomb repulsion $U$ (intra-orbital) and
239: $U'$ (inter-orbital) and Hund's rule coupling $J_H$.
240: For $t_{2g}$ orbitals, $H_I$ has been shown to take the
241: form \cite{castellani}
242: \begin{eqnarray}
243: H_I&=&U\sum_{i,\alpha}{\hat n}_{i\alpha\uparrow}{\hat n}_{i\alpha\downarrow}
244: +(U'-{1\over 2}J_H)\sum_{i,\alpha>\beta}{\hat n}_{i\alpha}{\hat n}_{i\beta}
245: \label{h-int} \\
246: &-&J_H\sum_{i,\alpha\neq\beta}{\bf S}_{i\alpha}\cdot {\bf S}_{i\beta}
247: +J_H\sum_{i,\alpha\neq\beta}a^{\dagger}_{i\alpha\uparrow}
248: a^{\dagger}_{i\alpha\downarrow}a_{i\beta\downarrow}a_{i\beta\uparrow}.
249: \nonumber
250: \end{eqnarray}
251: with $U^\prime = U - 2J_H$.
252: Here ${\hat n}_{i\alpha}$ and ${\bf S}_{i\alpha}$ are the density and
253: the spin operators in the $\{a\}$ basis where
254: the tight-binding part $H_0$ is
255: %
256: \begin{equation}
257: H_0=\sum_{k,\sigma}\sum_{\alpha\beta}
258: {\bf K}^a_{\alpha\beta}(k)
259: a^{\dagger}_{k\alpha\sigma}a_{k\beta\sigma}
260: +\sum_{k,\alpha,\sigma}\Delta_{\alpha}
261: a^{\dagger}_{k\alpha\sigma}a_{k\alpha\sigma}.
262: \label{hk-tba}
263: \end{equation}
264: %
265: Here $\Delta_{\alpha}$ = $-\Delta/3$, $2\Delta/3$ for the $e'_g$ and $a_{1g}$
266: orbitals respectively. The hopping matrix
267: ${\bf K}^a(k)={\bf O}^{T}{\bf K}^d(k){\bf O}$,
268: with ${\bf O}$ the orthogonal rotation from the $t_{2g}$ to
269: the $\{a\}$ basis. $H_I$ is identical in
270: these two bases. The hierarchy of the interaction
271: strength is $U>U'>J_H\geq 0$.
272: \begin{figure}
273: \begin{center}
274: %\vskip-2.8cm
275: %\hspace*{-3.8cm}
276: \fig{2.8in}{fig2.eps}
277: \vskip-2.4mm
278: \caption{HF results for $U$ = 3.0\textit{eV} at $x=1/3$.
279: The band dispersions are shown for $\eta= 1/3, 1/2, 2/3$.
280: }
281: \label{fig2}
282: \vskip-9mm
283: \end{center}
284: \end{figure}
285:
286: We first study the effects of interactions in the HF theory in the
287: orbital sector. In the paramagnetic phase, the interacting Hamiltonian
288: is given by,
289: \begin{eqnarray}
290: H_I^{HF} &=& \sum_{k,\sigma,\alpha}
291: ({1\over 2}Un_{\alpha}+U_{\rm eff}^\prime
292: \sum_{\beta\neq\alpha}n_{\beta})
293: a^{\dagger}_{k\alpha\sigma}a_{k\alpha\sigma}\nonumber \\
294: &-&{U\over 4}\sum_{k,\alpha}n^2_{\alpha}
295: -{U_{\rm eff}^\prime\over 2}\sum_{k,\alpha\neq\beta}
296: n_{\alpha}n_{\beta}
297: \label{hartree} \\
298: &+&(U-2U_{\rm eff}^\prime)\sum_{k,\alpha\neq\beta}
299: \bigl[n_{\alpha\beta}
300: a^{\dagger}_{k\alpha\sigma}a_{k\beta\sigma}
301: -{n^2_{\alpha\beta}\over2}\bigr],
302: \nonumber
303: \end{eqnarray}
304: where $n_{\alpha\beta}
305: =(1/N_s)\sum_{k,\sigma}n_{\alpha\beta}^\sigma(k)$,
306: $n_{\alpha\beta}^\sigma(k)=\langle a^{\dagger}_{k\alpha\sigma}a_{k\beta\sigma}
307: \rangle$, $n_{\alpha}=n_{\alpha\alpha}$, and $U_{\rm
308: eff}^\prime=U^\prime-J_H/2$.
309: In essence, this HF analysis, also discussed in Ref.~\cite{liebsch},
310: is equivalent to the LDA+U theory \cite{LDA+U}.
311: Since we are interested in the orbital dependent corrections, we have not
312: displayed in Eq.~(\ref{hartree}) the double-counting term which corrects
313: for the energy already included in the LDA, because it depends only on the
314: total density.
315:
316: Note that the HF theory is basis independent. It is thus convenient to
317: stay in the $\{a\}$ basis where the local density matrix, and thus the
318: HF self-energies are diagonal in orbital space, i.e.
319: $n_{\alpha\ne\beta}=0$.
320: Then the average energy has contributions only from the first two
321: terms in Eq.~(\ref{hartree}). Recall that most of the holes reside in
322: the $a_{1g}$ orbitals. If $U^\prime_{\rm eff} < U/2$, it is
323: favorable to increase the population of the $e_g^\prime$ orbitals to
324: take advantage of the smaller inter-orbital repulsion $U^\prime_{\rm
325: eff}$. On the other hand for $U^\prime_{\rm eff} > U/2$, the
326: tendency is to empty the $e_g^\prime$ orbitals in favor of $a_{1g}$
327: occupation. The crucial factor of $1/2$ in front of $U$ comes
328: from the fact that due to exchange, intra-orbital repulsion operates
329: only between holes with opposite spin, whereas both spins contribute
330: to $U_{\rm eff}^\prime$.
331:
332: We proceed to calculate the HF self-energy
333: in terms of the average ${\bar n}=(n_{a_{1g}}+2n_{e_g^\prime})/3=(1-x)/3$
334: and the difference $\delta=(n_{a_{1g}}-n_{e'_g})/3$ between the
335: hole occupation of the $a_{1g}$ and $e_g^\prime$ orbitals,
336: \begin{eqnarray}
337: \Sigma^{\text{HF}}_{e'_g} &=&{1\over2}{\bar n}U(1+4\eta)+\delta U
338: (\eta-1/2)
339: \\
340: \Sigma^{\text{HF}}_{a_{1g}}&=&{1\over2}{\bar n}U(1+4\eta)-2\delta U
341: (\eta-1/2)
342: \end{eqnarray}
343: where $\eta=U_{\rm eff}^\prime/U$ is the relative strength of the
344: inter-orbital interaction.
345: The interaction effect in the paramagnetic HF theory
346: is to simply shift the atomic levels
347: by $\Sigma^{\text{HF}}_{e'_g}$ and $\Sigma^{\text{HF}}_{a_{1g}}$
348: respectively, resulting in a renormalization of the atomic level spacing
349: $\Delta^\prime=-3\delta U(\eta-1/2)$.
350: As expected, the direction
351: of the charge transfer depends on the ratio $\eta$.
352: Since the majority of the holes resides in
353: the $a_{1g}$ orbital in the noninteracting
354: limit, $\delta>0$.
355: Thus, for $\eta<1/2$, the level splitting renormalizes upward,
356: $\Delta^\prime >0$, and interactions induce a transfer of carriers
357: from the $a_{1g}$ to the $e_{g}^\prime$ orbital. The self-consistent
358: HF results are shown in Fig.~2 at $x=1/3$ for $U=3$eV which is close
359: to the value ($3.7$ eV) estimated from the LDA
360: \cite{johannes}. For $\eta=1/3$, the size of the hole pockets
361: indeed becomes larger than that of the noninteracting/LDA ones.
362: At $\eta=1/2$, the
363: self-energy corrections are equal among the orbitals and
364: the noninteracting (LDA) band dispersions are unchanged as shown in Fig.~2.
365: When $\eta>1/2$, i.e. for $U^\prime/U>3/5$ or $J_H/U<1/5$ which is
366: reasonable for the cobaltates \cite{PHZhang}, the level splitting
367: renormalizes downward, $\Delta^\prime <0$,
368: triggering a transfer of holes from the $e'_g$ to the $a_{1g}$ orbital.
369: The six FS pockets continues to shrink as the $e_g^\prime$ band
370: sinks with increasing $\eta$ and disappears beyond a critical ratio $\eta_c$,
371: as shown in Fig.~2 for $\eta=2/3$.
372: We find that $\eta_c(U=3.0{\rm eV})\simeq 5/8$.
373: %
374: \begin{figure}
375: \begin{center}
376: %\vskip-0.8cm
377: %\hspace*{-3.8cm}
378: \fig{2.8in}{fig3.eps}
379: \vskip-2.4mm
380: \caption{The doping $x$ dependence of the level spacing and bandwidth
381: renormalization for the $a_{1g}$ and $e_g^\prime$ orbitals.}
382: \label{fig3}
383: \vskip-9mm
384: \end{center}
385: \end{figure}
386: %
387:
388: The HF analysis shows that the disappearance of the six
389: FS pockets near the K-points is the physics of large $U$ and $U'$
390: compared with $J_H$.
391: In this case, the HF theory itself becomes unreliable. Moreover, the
392: localization tendency leading to the bandwidth reduction cannot be
393: captured by the HF theory. It is therefore instructive to study the
394: problem in the strong-coupling limit by projecting out the states of
395: double-occupation prohibited by the large on-site Coulomb repulsions.
396: This can be achieved by the Gutzwiller projection
397: $\vert \Psi(\{n_\alpha\})\rangle
398: =P_G \vert \Psi_0(\{n_\alpha\})\rangle$, where $P_G$ is the
399: projection operator that operates on
400: the noninteracting state $\vert\Psi_0(\{n_\alpha\})\rangle$
401: in a given orbital occupation scheme $\{n_\alpha\}$. It removes the
402: double-occupancy by electrons from both the same and different orbitals.
403: This variational procedure is most conveniently implemented in
404: the Gutzwiller approximation where the effect of projection is
405: taken into account by the statistical weighting factor multiplying
406: the quantum coherent state \cite{fczhang}. Specifically,
407: we approximate the hopping term by
408: \begin{equation}
409: \langle\Psi\vert a_{i\alpha\sigma}^\dagger
410: a_{j\beta\sigma}\vert \Psi\rangle
411: =g_t^{\alpha\beta} \langle \Psi_0\vert
412: a_{i\alpha\sigma}^\dagger a_{j\beta\sigma}
413: \vert \Psi_0\rangle,
414: \label{gwt}
415: \end{equation}
416: where the Gutzwiller renormalization factor $g_t$ is given by the
417: ratio of the probabilities in the hopping process in the
418: projected $\vert\Psi\rangle$ and
419: the unprojected $\vert\Psi_0\rangle$. We find,
420: \begin{equation}
421: g_t^{\alpha\beta}={x\over\sqrt{(1-n_{i\alpha\sigma})(1-n_{j\beta\sigma})}}.
422: \label{gt}
423: \end{equation}
424: It is important to note that in a multi-orbital system $g_t^{\alpha\beta}$
425: depends on the occupation of the orbitals connected by the hopping
426: integral as seen in the denominator in Eq.~(\ref{gt}). The latter
427: originates from the Pauli principle.
428: It compensates for the effects of
429: ``Pauli-blocking'' of double-occupation by electrons in the same
430: quantum states which already operate in the free fermion term on the RHS of
431: Eq.~(\ref{gwt}), while the numerator $x$ in Eq.~(\ref{gt})
432: describes the ``Coulomb-blocking''
433: due to the large on-site $U$ and $U^\prime$. It turns out that
434: the denominator is crucial for carrier transfer and orbital-polarization in the
435: strong-coupling limit. In the uniform paramagnetic phase, $g_t^{\alpha\beta}
436: =2x/\sqrt{(2-n_\alpha)(2-n_\beta)}$. The orbital occupations are variational
437: parameters determined by minimizing the ground state energy
438: of the Hamiltonian,
439: \begin{eqnarray}
440: H_{GW}&=&\sum_{k,\sigma,\alpha\beta}
441: g_t^{\alpha\beta}
442: {\bf K}^a_{\alpha\beta}(k)
443: a^{\dagger}_{k\alpha\sigma}a_{k\beta\sigma}
444: +\sum_{i,\alpha,\sigma}\Delta_{\alpha}
445: a^{\dagger}_{i\alpha\sigma}a_{i\alpha\sigma}
446: \nonumber \\
447: &+& \sum_{i,\alpha}\varepsilon_{i\alpha}
448: (\sum_\sigma a_{i\alpha\sigma}^\dagger a_{i\alpha\sigma}
449: -n_{i\alpha}),
450: \label{hgw}
451: \end{eqnarray}
452: where $\varepsilon_{i\alpha}$ are the Lagrange multipliers enforcing
453: the occupation $n_{i\alpha}=\sum_\sigma\langle
454: a_{i\sigma\alpha}^\dagger a_{i\sigma\alpha}
455: \rangle$. They are determined by the self-consistency equation
456: \begin{equation}
457: \varepsilon_{i\alpha}=\varepsilon_\alpha
458: ={1\over 2-n_\alpha}{1\over N_s}\sum_{k,\beta,\sigma}
459: g_t^{\alpha\beta} {\bf K}^a_{\alpha\beta}
460: \langle a^{\dagger}_{k\alpha\sigma}a_{k\beta\sigma}\rangle.
461: \label{epsilon}
462: \end{equation}
463: The RHS of this equation is the derivative of the kinetic energy of
464: band $\alpha$ with respect to $n_\alpha$, and can be understood by
465: the following argument. The transfer of a hole from band $\alpha$ to
466: $\beta$ causes the respective bandwidths to decrease and increase by
467: ${\cal O} (1/N_s)$. However, the kinetic energies of the occupied
468: states in each band are changed by order unity, and this energy
469: difference must be reflected in the equilibrium condition by an
470: energy shift $\varepsilon_\alpha$. Thus in contrast to HF theory,
471: there is no energy cost proportional to $U$. Instead,
472: both the band-narrowing and the renormalization of the level spacing
473: $\Delta^\prime=\varepsilon_{a_{1g}}-\varepsilon_{e_g^\prime}$
474: contribute to the redistribution of holes among the orbitals.
475: In Fig.~3, we show the self-consistently determined band-narrowing factor
476: $z_\alpha=g_t^{\alpha\alpha}$ and the renormalized level spacing.
477: For orbitals with a larger hole occupation, the bandwidth reduction
478: is smaller and the renormalized band energy is lower, resulting
479: in the transfer of more holes into these bands. The combined
480: effects cause the holes to move out of the $e_g^\prime$ band into
481: the $a_{1g}$ band.
482: The calculated band dispersions and the FS topology at
483: $x=0.3, 0.5, 0.7$ are shown in Fig.{\ref{fig4}} in the strong
484: coupling theory. The corresponding non-interacting case is also shown
485: for comparison. The six hole pockets are completely absent at all
486: levels of sodium doping due to strong correlation, leaving a single hexagonal
487: Fermi surface centered around the $\Gamma$ point satisfying the
488: Luttinger theorem. This, as well as the band-narrowing due to strong
489: Coulomb repulsion, is in very good agreement with the photoemission
490: experiments \cite{hbyang2,hasan2}.
491: \begin{figure}
492: \begin{center}
493: %\vskip-0.8cm
494: %\hspace*{-3.8cm}
495: \fig{3.3in}{fig4.eps}
496: \vskip-2.4mm
497: \caption{The band dispersions and the Fermi surfaces in the strong
498: coupling limit (red solid lines) for doping $x$ = 0.3, 0.5 and 0.7.
499: The noninteracting dispersions (blue dashed lines) are also plotted
500: for comparison.}
501: \label{fig4}
502: \vskip-9mm
503: \end{center}
504: \end{figure}
505:
506: In conclusion, we have shown that strong correlation plays
507: an important role in understanding the electronic structure of Na$_x$CoO$_2$.
508: It pushes the $e_g^\prime$ band below
509: the Fermi level, leading to an orbital polarized state with a single
510: hole-like FS. The absence of the small FS pockets, which
511: would have contributed significantly to the density of states in band
512: structure calculations, further suggests that the large mass enhancement
513: observed in the specific heat measurement \cite{chou}
514: is due to strong correlation.
515:
516: This work is supported by DOE grants DE-FG02-99ER45747 and DE-FG02-03ER46076,
517: ACS grant 39498-AC5M, and NSF grants DMR-0072205 and DMR-0201069.
518: \begin{thebibliography}{99}
519: \bibitem{takada03}
520: K. Takada, et. al., Nature (London) {\bf422}, 53 (2003).
521: \bibitem{singh00}
522: D. J. Singh, Phys. Rev. B{\bf61}, 13397 (2000).
523: \bibitem{pickett}
524: K.-W. Lee, J. Kune\v{s}, and W. E. Pickett,
525: Phys. Rev. B{\bf70}, 045104 (2004).
526: \bibitem{hbyang2}
527: H.~B. Yang, et. al., cond-mat/0501403 (2005).
528: \bibitem{hasan2}
529: M.~Z. Hasan, et. al., cond-mat/0501530 (2005).
530: \bibitem{hasan}
531: M.~Z. Hasan, et. al., Phys. Rev. Lett. {\bf92}, 246402 (2004).
532: \bibitem{hbyang1}
533: H.~B. Yang, et. al., Phys. Rev. Lett. {\bf92}, 246403 (2004).
534: \bibitem{kuroki04}
535: K. Kuroki, Y. Tanaka, and R. Arita,
536: \pprl{\bf93}, 077001 (2004).
537: \bibitem{johannes04}
538: M.~D. Johannes,
539: I.~I. Mazin, D.~J. Singh, and D.~A. Papaconstantopoulos,
540: Phys. Rev. Lett. {\bf93}, 097005 (2004).
541: \bibitem{yanase}
542: Y. Yanase, M. Mochizuki, and M. Ogata, J. Phys. Soc. Jpn. {\bf74}, 430 (2005);
543: M. Mochizuki, Y. Yanase, and M. Ogata, cond-mat/0407094.
544: \bibitem{PHZhang}
545: P. Zhang \textit{et al}., Phys. Rev. Lett. \textbf{93}, 236402 (2004);
546: P. Zhang \textit{et al}., Phys. Rev. B{\bf70}, 085108 (2004).
547: \bibitem{liebsch}
548: H. Ishida, M. D. Johannes, and A. Liebsch, cond-mat/0412654.
549: \bibitem{johannes}
550: M.~D. Johannes, I. I. Mazin, and D. J. Singh, cond-mat/0408696.
551: \bibitem{castellani}
552: C. Castellani, C.~R. Natoli, and J. Ranninger,
553: Phys. Rev. B{\bf18}, 4945 (1978)
554: \bibitem{LDA+U}
555: A.~I. Liechtenstein, V.~I. Anisimov, and J. Zaanen,
556: Phys. Rev. B{\bf52}, R5467 (1995).
557: \bibitem{fczhang}
558: F.~C. Zhang, \textit{et al}.,
559: %C. Gros, T.~M. Rice, and H. Shiba,
560: Supercond. Sci. Technol. {\bf1}, 36 (1988).
561: \bibitem{chou}
562: F.C. Chou, et. al., Phys. Rev. Lett. \textbf{92}, 157004
563: (2004).
564: \end{thebibliography}
565: \end{document}
566:
567: