0807.2498/tuj.tex
1: %\documentclass[aps,english,preprint]{revtex4}
2: 
3: 
4: \documentclass[aps,english,twocolumn]{revtex4}
5: 
6: \usepackage{graphicx}
7: \usepackage{amsmath, amssymb}
8: \usepackage{babel}
9: 
10: \setcounter{MaxMatrixCols}{10}
11: 
12: 
13: \begin{document}
14: 
15: \title{
16: Doping dependent charge transfer gap and realistic electronic model
17: of n-type cuprate superconductors }
18: 
19: \author{T. Xiang$^{1,2}$, H. G. Luo$^2$, D. H. Lu$^3$, K. M. Shen$^4$, Z. X. Shen$^3$}
20: 
21: \address{$^1$Institute of Physics, Chinese Academy of
22: Sciences, P.O. Box 603, Beijing 100080, China}
23: 
24: \address{$^2$Institute of Theoretical Physics, Chinese Academy of
25: Sciences, P.O. Box 2735, Beijing 100080, China}
26: 
27: \address{$^3$Department of Physics, Applied Physics,
28: and Stanford Synchrotron Radiation Laboratory, Stanford University,
29: Stanford, California 94305, USA}
30: 
31: \address{$^4$Department of Physics and Astronomy, University of British
32: Columbia, Vancouver, British Columbia, V6T 1Z4, Canada }
33: 
34: 
35: \date{\today}
36: 
37: \begin{abstract}
38: Based on the analysis of the measurement data of angle-resolved
39: photoemission spectroscopy (ARPES) and optics, we show that the
40: charge transfer gap is significantly smaller than the optical one
41: and is reduced by doping in electron doped cuprate superconductors.
42: This leads to a strong charge fluctuation between the Zhang-Rice
43: singlet and the upper Hubbard bands. The basic model for describing
44: this system is a hybridized two-band $t$-$J$ model. In the symmetric
45: limit where the corresponding intra- and inter-band hopping
46: integrals are equal to each other, this two-band model is equivalent
47: to the Hubbard model with an antiferromagnetic exchange interaction
48: (i.e. the  $t$-$U$-$J$ model). The mean-field result of the
49: $t$-$U$-$J$ model gives a good account for the doping evolution of
50: the Fermi surface and the staggered magnetization.
51: \end{abstract}
52: 
53: \maketitle
54: 
55: 
56: 
57: 
58: The evolution of the Fermi surface and the Mott insulating gap with
59: hole or electron doping is a central issue in elucidating the
60: mechanism of high-T$_c$ superconductivity. In the hole doped case, a
61: small Fermi surface arc appears first near $(\pi/2 ,\pi /2)$ and
62: then extends towards $(\pi , 0)$ and $(0, \pi )$ with increasing
63: doping. In contrast, in the electron dope case, electrons are first
64: doped into the upper Hubbard band (Cu $ 3d^{10} $ band) near ($\pi $
65: ,0) and equivalent points. With further doping but still in the
66: antiferromagnetic phase, in-gap spectral weight develops below the
67: Fermi level. These in-gap states move upwards and eventually form a
68: hole-like Fermi surface pocket around ($\pi /2$ , $\pi
69: /2$)\cite{Armitage02}. In the heavily overdoped sample, these two
70: Fermi pockets merge together and form a large Fermi surface with a
71: volume satisfying the Luttinger theorem.
72: 
73: 
74: The peculiar doping dependence of the Fermi surface topology in
75: electron-doped cuprates is a manifestation of correlation effects.
76: To understand the physics behind, much of the theoretical study has
77: been carried out with the one-band Hubbard
78: model\cite{Kusko02,Kusunose03,Senechal04}. In this model, a metallic
79: band is split into two effective bands, namely the upper and lower
80: Hubbard bands, by a correlation energy $U$ that represents the
81: energy cost for a site to be doubly occupied. Under the mean-field
82: approximation, this model gives a good account for the experimental
83: data if $U$ is assumed to fall strongly with doping. However, this
84: strong reduction of $U$ by doping is not usually
85: expected\cite{Yuan04}. The nominal Hubbard $U$-term could arise
86: either from the on-site Coulomb repulsion between two electrons in a
87: Cu $3d_{x^2-y^2}$ orbital or from the charge transfer (CT) gap
88: between O $2p$ and Cu $3d^{10}$ bands. The on-site Coulomb repulsion
89: in a Cu $3d_{x^2-y^2}$ states in high-T$_c$ cuprates is generally
90: larger than $5$ eV. The CT gap in electron doped cuprates quoted
91: from the optical data is also quite large ($\sim 1.5$
92: eV)\cite{Onose01,Wang06}. It seems that in neither case $U$ can be
93: dramatically suppressed by only 15\% doping.
94: 
95: 
96: An alternative interpretation to the two Fermi pockets is based on
97: the notion of band folding induced by the antiferromagnetic
98: interaction\cite{Yuan04,Matsui05}. This interpretation is consistent
99: with the measurement data in the overdoped regime ($x > 0.15$).
100: However, in the low doping antiferromagnetic phase, it breaks down.
101: The band folding assumes implicitly a band with large Fermi surface
102: exists and it is the antiferromagnetic interactions between the hot
103: spots split this band into a conduction electron and a shadow hole
104: band. However, in the antiferromagnetic phase at low doping, these
105: bands with the folding gap at the hot spots were not observed and
106: the band near $(\pi /2, \pi /2)$ is well below $(\pi
107: ,0)$.\cite{Armitage02} Furthermore, the antiferromagnetic
108: interaction is too small to account for the energy splitting between
109: the lower and upper CT bands at least in the low doping limit.
110: 
111: To resolve the above problems, it is important to understand
112: correctly how the hole-like Fermi pockets develop with doping. In a
113: nominally undoped Nd$_2$CuO$_4$, a dispersive band structure is
114: observed by ARPES at roughly $1.2$ eV below the chemical potential.
115: As shown in Ref. \cite{Armitage02}, the energy-momentum dispersion
116: of this spectral peak behaves almost the same as the lower CT band
117: observed in Ca$_2$CuO$_2$Cl$_2$, except in the latter case the band
118: lies at only $\sim 0.7$ eV below the chemical potential. This
119: suggests that these two bands have the same physical origin. The
120: difference is probably due to the intrinsic doping and the Fermi
121: energy is pinned near the bottom of conduction band (i.e. Cu
122: $3d^{10}$ band) in Nd$_2$CuO$_4$ and near the top of the valence
123: band (\emph{i.e.} Zhang-Rice singlet band\cite{Zhang88}) in
124: Ca$_2$CuO$_2$Cl$_2$.
125: 
126: \begin{figure}[ht]
127: %\begin{center}
128: \includegraphics[width=\columnwidth]{fig1.eps}
129: %\end{center}
130: \caption{ARPES spectra near (a) the nodal and (b) antinodal regions,
131: reproduced from the data published in Ref.~\cite{Armitage02}. (c)
132: The infrared conductivity reproduced from the data published in Ref.
133: \cite{Wang06}. The insets of (b) and (c) illustrate the indirect and
134: direct CT gaps. } \label{fig:infra}
135: \end{figure}
136: 
137: 
138: 
139: Doping electrons into Nd$_2$CuO$_4$ results in a spectral weight
140: transfer from the main spectral peak at $\sim 1.2$eV to a ``in-gap"
141: state. This in-gap state first appears as a week low energy ``foot"
142: at $\sim 0.5$ eV below the Fermi level $\varepsilon_F$ along the
143: zone diagonal in the undoped Nd$_2$CuO$_4$ (Fig.~\ref{fig:infra}).
144: It moves towards the Fermi level with doping and becomes a broad
145: hump just below the Fermi level at optimal doping. The hole Fermi
146: pockets observed at high doping originate from these in-gap states.
147: In contrast, the states near $(\pi,0)$ resides at $\varepsilon_F$ as
148: they are derived from the bottom of the upper Hubbard band. The fact
149: that the broad maximum is slightly below $\varepsilon_F$ is caused
150: by the Frank-Condon broadening as discussed below.
151: 
152: It should be pointed out that, same as for the dispersive high
153: energy band, the in-gap states behave similarly as the low energy
154: coherent states observed in hole-doped
155: Ca$_2$CuO$_2$Cl$_2$\cite{Shen04}. Near half filling, the in-gap
156: state in Nd$_2$CuO$_4$ lies also at $\sim 0.7$ eV above the high
157: energy spectral peak. This suggests that, similar as in hole doped
158: materials, the high energy hump structure in the spectra results
159: from the Franck-Condon broadening and the in-gap states are the true
160: quasiparticle excitations located at the top of the lower CT
161: band\cite{Shen04}. At half filling, the in-gap state is not observed
162: because its quasiparticle weight is vanishingly small\cite{Rosch05}.
163: 
164: 
165: The spectral weight transfer induced by doping has also been
166: observed in the optical measurements (Fig.~\ref{fig:infra}). At zero
167: doping, the optical CT gap appears at $\sim 1.5$ eV. Upon doping, a
168: mid-infrared conductivity peak develops. This mid-infrared peak
169: appears at $\sim 0.5$ eV at low doping\cite{Onose01,Wang06}, and
170: then moves towards zero energy with increasing doping. The doping
171: dependence of the mid-infrared peak is consistent with the doping
172: evolution of the ``in-gap'' states observed by ARPES. It suggests
173: that the mid-infrared peak results mainly from the optical
174: transition between the ``in-gap'' states and the upper Hubbard band.
175: The polaron effect may also have some contribution to these
176: mid-infrared peaks.\cite{Mishchenko08}
177: 
178: 
179: 
180: The above discussion indicates that the true CT gap, measured as the
181: minimum excitation energy between the lower and upper Hubbard bands,
182: is only $0.5$ eV at half filling, much lower than the optically
183: measured CT gap, which is usually believed to be about 1.5 eV. This
184: difference between the true quasiparticle gap that determines the
185: transport and thermodynamics and optically measured CT gaps has also
186: been found in hole doped materials. For La$_2$CuO$_4$, Ono \emph{et
187: al.}\cite{Ono07} found recently that the CT gap obtained from the
188: high temperature behavior of the Hall coefficient is only 0.89 eV,
189: while the corresponding optical CT gap is about 2 eV. This means
190: that the optical CT gap, which is generally determined from the peak
191: energy of the optical absorption, does not correspond to the true
192: gap between the two bands in high-T$_c$ oxides. The indirect nature
193: of the gap (insets of Fig.~\ref{fig:infra}) and the Frank-Condon
194: effect lead to the overestimate of the gap by optics. It also means
195: that the charge fluctuation in high-T$_c$ materials is much stronger
196: than usually believed and should be fully considered in the
197: construction of the basic model of high-T$_c$ superconductivity
198: \cite{Zaanen85, Varma06, Luo05}.
199: 
200: The doping dependence of low energy peaks observed by both APRES and
201: optics indicates that there is a gap closing with doping. This gap
202: closing may result from the Coulomb repulsion between O $2p$ and Cu
203: $3d$ electrons. Doping electrons will increase the occupation number
204: of Cu $3d$ states, which in turn will add an effective potential to
205: the O $2p$ states and raise their energy level. If $U_{pd}$ is the
206: energy of the Coulomb interaction between neighboring O and Cu ions,
207: then the change in the O $2p$ energy level will be $\delta
208: \varepsilon_{p} \approx + 2 x U_{pd} $, where $x$ is the doping
209: concentration and the factor 2 appears since each O has two Cu
210: neighbors. $U_{pd}$ is generally estimated to be of order 1-2 eV.
211: Thus a 15\% doping of electrons would reduce the CT gap by 0.3-0.6
212: eV, within the range of experimentally observed gap reduction.
213: Furthermore, the electrostatic screening induced by doping can
214: reduce the on-site Coulomb interaction of Cu $3d_{x^2-y^2}$
215: electrons. This can also reduce the gap between the O $2p$ and the
216: upper Hubbard bands.
217: 
218: 
219: \begin{figure}[ht]
220: %\begin{center}
221: \includegraphics[width=0.8\columnwidth]{fig2.eps}
222: %\end{center}
223: \caption{Fermi surface density map at different dopings $x$,
224: obtained by integrating the spectral function from -40 to 20 meV
225: around the Fermi level, for the $t$-$U$-$J$ model. }
226: \label{fig:band}
227: \end{figure}
228: 
229: 
230: Now let us consider how to characterize the low energy charge and
231: spin dynamics of the system. For simplicity, we focus on the
232: electronic structure and leave the additional electron-phonon
233: interaction effect for future study. The lower CT band behaves
234: similarly as the Zhang-Rice singlet band. Thus if its charge
235: fluctuation with the upper Hubbard band is ignored, this band should
236: be described by an effective one-band $t$-$J$ model\cite{Zhang88}.
237: Similarly, the upper Hubbard band should also be described by an
238: effective one-band $t$-$J$ model if there is no charge fluctuation.
239: However, in the case the hybridization or charge transfer between
240: these bands is important, it can be shown from a three-band model
241: that these two $t$-$J$ models should be combined together and
242: replaced by the following hybridized two-band $t$-$J$
243: model\cite{Baskara}
244: \begin{eqnarray}
245: H &=&\sum_{ij\sigma }t_{ij}^{e}e_{i}^{\dagger }d_{i\sigma }d_{j\sigma
246: }^{\dagger }e_{j}+\sum_{ij\sigma }t_{ij}^{h}h_{i}^{\dagger }d_{i\sigma
247: }d_{j\sigma }^{\dagger }h_{j}  \notag \\
248: &&+\sum_{ij\sigma }t_{ij}\left( \sigma d_{i\sigma }^{\dagger }d_{j\overline{
249: \sigma }}^{\dagger }e_{i}h_{j}+h.c.\right) +J\sum_{\langle ij\rangle
250: }\mathbf{S}_{i}\cdot \mathbf{S}_{j}  \notag \\
251: &&+\sum_{i}\left( \varepsilon _{e}e_{i}^{\dagger }e_{i}+\varepsilon
252: _{h}h_{i}^{\dagger }h_{i}\right) -V_{pd}\sum_{\langle ij\rangle
253: }e_{i}^{\dagger }e_{i}h_{j}^{\dagger }h_{j},  \label{eq:tj}
254: \end{eqnarray}
255: where $h_{i}$, $e_{i}$ and $d_{i\sigma }$ are the annihilation
256: operators of a Zhang-Rice singlet hole, a doubly occupied
257: $d_{x^{2}-y^{2}}$ state (doublon), and a pure Cu$^{2+}$ spin,
258: respectively. At each site, these three states cannot coexist and
259: the corresponding number operators should satisfy the constraint
260: \begin{equation}
261: e_{i}^{\dagger }e_{i}+h_{i}^{\dagger }h_{i}+\sum_{\sigma }d_{i\sigma
262: }^{\dagger }d_{i\sigma }=1.
263: \end{equation}
264: The difference between the number of doubly occupied
265: $d_{x^{2}-y^{2}}$ states and Zhang-Rice singlet holes is the doping
266: concentration of electrons, $\langle e_{i}^{\dagger
267: }e_{i}-h_{i}^{\dagger }h_{i} \rangle =x$.
268: 
269: 
270: In Eq. (\ref{eq:tj}), $\mathbf{S}_{i}= d^\dagger_i \mathbf{\sigma}
271: d_i/2$ is the spin operator and $\mathbf{\sigma}$ is the Pauli
272: matrix. $\varepsilon _{e}$ and $\varepsilon_{h}$ are the excitation
273: energies of a doublon and a Zhang-Rice singlet, respectively.
274: $t_{ij}^{e}$ and $t_{ij}^{h}$ are the hopping integrals of the upper
275: Hubbard and Zhang-Rice singlet bands. In Eq. (\ref{eq:tj}), if
276: $\varepsilon_{h}\gg \varepsilon_{e}
277: > 0$, then $\langle h_{i}^{\dagger }h_{i} \rangle
278: \approx 0$ and $H$ simply reduces to the one-band $t$-$J$ model of
279: doubly occupied electrons in the doublon-spinon representation. On
280: the other hand, if $\varepsilon _{e}\gg \varepsilon_{h}>0$ , then
281: $\langle e_{i}^{\dagger }e_{i} \rangle \approx 0$ and $H$ becomes
282: simply the one-band $t$-$J$ model of Zhang-Rice singlets in the
283: holon-spinon representation. The $t_{ij}$ term describes the
284: hybridization between the upper Hubbard and Zhang-Rice singlets. The
285: last term results from the Coulomb repulsion between a Cu
286: $3d_{x^{2}-y^{2}}$ and its neighboring O $2p_{x,y}$ electrons.
287: $V_{pd}$ is proportional to the Coulomb repulsion between Cu and O
288: ions $U_{pd}$.
289: 
290: 
291: 
292: 
293: The above Hamiltonian can be simplified if $t_{ij}^{e} = t_{ij}^{h}
294: = t_{ij}$. In this case, by using the holon-doublon representation
295: of an electron operator $c_{i\sigma }= \sigma h_{i}^{\dagger }
296: d_{i\sigma } + e_{i} d_{i \overline{\sigma }}^{\dagger }$, and
297: taking a mean-field approximation for the $V_{pd}$-term,
298: $e_{i}^{\dagger }e_{i}h_{j}^{\dagger }h_{j} \approx \langle
299: e_{i}^{\dagger }e_{i} \rangle h_{j}^{\dagger }h_{j} + e_{i}^{\dagger
300: }e_{i} \langle h_{j}^{\dagger } h_{j} \rangle - \langle
301: e_{i}^{\dagger }e_{i} \rangle \langle h_{j}^{\dagger }h_{j} \rangle
302: $, one can then express $H$ as
303: \begin{equation}\label{eq:tUJ}
304: H=\sum_{ij\sigma }t_{ij}c_{i\sigma }^{\dagger }c_{j\sigma
305: }+U\sum_{i}n_{i\uparrow }n_{i\downarrow }+J\sum_{\langle ij\rangle
306: }S_{i}\cdot S_{j},
307: \end{equation}
308: where $n_{i\sigma }=c_{i\sigma }^{\dagger }c_{i\sigma }$ and
309: $U=\varepsilon_{e} + \varepsilon _{h}-4V_{pd}( \langle
310: e_{i}^{\dagger } e_{i}\rangle +\langle h_{i}^{\dagger } h_{i}
311: \rangle )$. In electron doped materials, as the induced hole
312: concentration is very small, $\langle h_{i}^{\dagger } h_{i}\rangle
313: \ll \langle e_{i}^{\dagger } e_{i}\rangle \approx x $, we have
314: $U\approx \varepsilon_{e} + \varepsilon _{h}-4 x V_{pd}$. It should
315: be emphasized that the spin exchange term in Eq. (\ref{eq:tUJ}) is
316: not a derivative of the one-band Hubbard model in the strong
317: coupling limit. It actually arises from the antiferromagnetic
318: superexchange interaction between two undoped Cu$^{2+}$ spins via an
319: O $2p$ orbital. This term, as shown in Ref. \cite{Daul00}, can
320: enhance strongly the superconducting pairing potential.
321: 
322: 
323: The $t$-$U$-$J$ model defined by Eq. (\ref{eq:tUJ}) is obtained by
324: assuming $t_{ij}^{e}=t_{ij}^{h}=t_{ij}$. This is a strong
325: approximation which may not be fully satisfied in real materials.
326: Nevertheless, we believe that this simplified model still catches
327: qualitatively the low energy physics of high-T$_c$ cuprates. It has
328: already been used, as an extension of either the Hubbard or the
329: $t$-$J$ model, to explore physical properties of strongly correlated
330: systems, such as the gossamer superconductivity.\cite{Zhang03}
331: 
332: 
333: 
334: The above Hamiltonian reveals two features about the effective
335: Hubbard interaction. First, $U$ is determined by the CT
336: gap\cite{Zaanen85}, not the Coulomb interaction between two
337: electrons in a Cu $3d_{x^2-y^2}$ orbital. It is in the intermediate
338: or even weak coupling regime, rather than the strong coupling limit
339: as usually believed. Second, $U$ is doping dependent. It drops with
340: doping. These are in fact the two key features that are needed in
341: order to explain the experimental results with the Hubbard model
342: \cite{Kusko02, Kusunose03, Senechal04}.
343: 
344: 
345: We have calculated the single-particle spectral function and the
346: staggered magnetization for the $t$-$U$-$J$ model using the
347: mean-field approximation. In the calculation, $t_{ij}$ are
348: parameterized by the first, second and third nearest neighbor
349: hopping integrals $(t,t^{\prime },t^{\prime \prime })$. The
350: parameters used are $t=0.326$ eV, $t^\prime = -0.25 t$, $t^{\prime
351: \prime} =0.15 t$, $J = 0.3 t$ eV, $\varepsilon_{e} + \varepsilon
352: _{h} = 4 t$ and $V_{pd} = 2.7 t$.
353: 
354: 
355: Fig. \ref{fig:band} shows the intensity plot of the spectral
356: function at the Fermi level. The doping evolution of the Fermi
357: surface agrees with the ARPES measurements.\cite{Armitage02} It is
358: also consistent with the mean-field calculation of the Hubbard model
359: by Kusko et al.,  while our calculation has the same shortcoming of
360: the mean field calculation in providing too large band
361: width.\cite{Kusko02} The difference is that, in our calculation, the
362: Hubbard interaction $U$ is not an adjustable parameter of doping. It
363: decreases almost linearly with doping. For the parameters given
364: above, $U \approx 4t - 10.8xt$.  Whereas in the calculation of Kusko
365: et al.\cite{Kusko02}, $U$ is determined by assuming the mean-field
366: energy gap to be equal to the experimentally observed value of the
367: ``pseudogap".
368: 
369: \begin{figure}[ht]
370: \begin{center}
371: \includegraphics[width=0.8\columnwidth]{fig3.eps}
372: \end{center}
373: \caption{Comparison of the mean field result (open circles) of the
374: staggered magnetization $m$ as a function of doping with the
375: experimental data (solid circles)\cite{stagger}. The theoretical
376: data obtained by Yan \emph{et al.}\cite{Yan06} and by Yuan \emph{et
377: al.}\cite{Yuan04} are also shown for comparison. }
378: \label{fig:stagger}
379: \end{figure}
380: 
381: 
382: Fig. \ref{fig:stagger} shows the theoretical result of the
383: staggered magnetization $m=(\langle n_{i\uparrow} -
384: n_{i\downarrow}\rangle )/2$. The simple mean-field result agrees
385: well with the experimental
386: data\cite{Mang04,Ross91,Matsuda90,stagger}, especially in the low
387: doping range. It is also consistent qualitatively with other
388: theoretical calculations\cite{Yan06,Yuan04}. $m$ decreases almost
389: linearly at low doping. However, it shows a fast drop above $\sim
390: 0.14$, when the lower Zhang-Rice singlet holes begin to emerge
391: above the Fermi surface. This abrupt change of $m$ is an
392: indication of a significant renormalization of the Fermi surface.
393: It may result from the quantum critical fluctuation as suggested
394: in Ref. \cite{Dagan04}. $m$ does not vanish above the optimal
395: doping, this is probably due to the mean-field approximation.
396: 
397: 
398: 
399: In conclusion, we have shown that the minimal CT gap is much smaller
400: than the optical gap and the charge fluctuation between the
401: Zhang-Rice singlet and the upper Hubbard bands is strong in electron
402: doped copper oxides. The low-lying excitations of the system are
403: governed by the hybridized two-band $t$-$J$ model defined by Eq.
404: (\ref{eq:tj}) or approximately by the $t$-$U$-$J$ model defined by
405: Eq. (\ref{eq:tUJ}). This conclusion is drawn based on the analysis
406: of electron doped materials. However, we believe it can be also
407: applied to hole doped cuprate superconductors. Our mean-field
408: calculation for the $t$-$U$-$J$ model gives a good account for the
409: doping evolution of the Fermi surface as well as the staggered
410: magnetization. It sheds light on the further understanding of
411: high-T$_c$ superconductivity.
412: 
413: 
414: We wish to thank N. P. Armitage and  N. L. Wang for providing the
415: ARPES and infrared conductivity data shown in Fig.~\ref{fig:infra}.
416: Support from the NSFC and the national program for basic research of
417: China is acknowledged. The Stanford work was supported by DOE Office
418: of Science, Division of Materials Science, with contract
419: DE-AC02-76SF00515.
420: 
421: \begin{thebibliography}{9}
422: 
423: \bibitem{Armitage02} N. P. Armitage, F. Ronning, D. H. Lu, C. Kim, A.
424: Damascelli, K. M. Shen, D. L. Feng, H. Eisaki, Z. X. Shen, P. K.
425: Mang, N. Kaneko, M. Greven, Y. Onose, Y. Taguchi, and Y. Tokura,
426: Phys. Rev. Lett. \textbf{88}, 257001 (2002).
427: 
428: \bibitem{Kusko02} C. Kusko, R. S. Markiewicz, M. Lindroos, and A. Bansil,
429: Phys. Rev. B \textbf{66}, 146513(R)(2002)
430: 
431: \bibitem{Kusunose03} H. Kusunose and T. M. Rice, Phys. Rev. Lett. \textbf{91}
432: , 186407 (2003).
433: 
434: \bibitem{Senechal04} D. S\'{e}n\'{e}chal and A.-M. S. Tremblay, Phys. Rev.
435: Lett. \textbf{92}, 126401 (2004).
436: 
437: \bibitem{Yuan04} Q. Yuan, Y. Chen, T. K. Lee, and C. S. Ting, Phys. Rev. B
438: \textbf{69}, 214523 (2004).
439: 
440: 
441: \bibitem{Onose01} Y. Onose, Y. Taguchi, K. Ishizaka, and Y. Tokura, Phys.
442: Rev. Lett. \textbf{87}, 217001 (2001)
443: 
444: 
445: \bibitem{Wang06} N. L. Wang, G. Li, D. Wu, X. H. Chen, C. H. Wang,
446: and H. Ding, Phys. Rev. B \textbf{73}, 184502 (2006).
447: 
448: 
449: \bibitem{Matsui05} H. Matsui, K. Terashima, T. Sato, T. Takahashi, S. C.
450: Wang, H. B. Yang, H. Ding, T. Uefuji, and K. Yamada, Phys. Rev. Lett.
451: \textbf{94}, 047005 (2005).
452: 
453: \bibitem{Shen04} K. M. Shen, F. Ronning, D. H. Lu, W. S. Lee, N.
454: J. C. Ingle, W. Meevasana, F. Baumberger, A. Damascelli, N. P.
455: Armitage, L. L. Miller, Y. Kohsaka, M. Azuma, M. Takano, H. Takagi,
456: and  Z. X. Shen, Phys. Rev. Lett. \textbf{93}, 267002 (2004).
457: 
458: \bibitem{Rosch05} O. R\"{o}sch, O. Gunnarsson, X. J. Zhou,
459: T. Yoshida, T. Sasagawa, A. Fujimori, Z. Hussain, Z.-X. Shen, and S.
460: Uchida Phys. Rev. Lett. \textbf{95}, 227002 (2005).
461: 
462: 
463: 
464: \bibitem{Zhang88} F. C. Zhang and T. M. Rice, Phys. Rev. B
465: \textbf{37}, 3759 (1988).
466: 
467: \bibitem{Mishchenko08}
468: A.S. Mishchenko, N. Nagaosa, Z.-X. Shen, G. De Filippis, V.
469: Cataudella,  T. P. Devereaux, C. Bernhard, K.W. Kim and J. Zaanen,
470: Phys. Rev. Lett. 100, 166401 (2008)
471: 
472: \bibitem{Ono07} S. Ono, S. Komiya, and Y. Ando, Phys. Rev. B
473: \textbf{75}, 024515 (2007).
474: 
475: \bibitem{Zaanen85} J. Zaanen, G. A. Sawatzky, and J. W. Allen,
476: Phys. Rev. Lett. \textbf{55}, 418  (1985).
477: 
478: \bibitem{Varma06} C. M. Varma, Phys. Rev. B \textbf{73}, 155113
479: (2006).
480: 
481: \bibitem{Luo05} H. G. Luo and. T. Xiang, Phys. Rev. Lett.
482: \textbf{94}, 027001 (2005).
483: 
484: \bibitem{Baskara} A two-species $t$-$J$ model with equal $+e$ and $-e$
485: charge carriers but without hybridization was proposed by G.
486: Baskaran, cond-mat/0505509 (unpublished).
487: 
488: \bibitem{Daul00} S. Daul, D. J. Scalapino, and S. R. White, Phys.
489: Rev. Lett. \textbf{84}, 4188 (2000).
490: 
491: \bibitem{Zhang03} F. C. Zhang, Phys. Rev. Lett. \textbf{90}, 207002
492: (2003); R. Laughlin, cond-mat/0209269.
493: 
494: \bibitem{Mang04} P. K. Mang, O. P. Vajk, A. Arvanitaki, J. W. Lynn,
495: and M. Greven, Phys. Rev. Lett. 93, 027002(2004).
496: 
497: \bibitem{Ross91} M. J. Rosseinsky, K. Prassides,
498: and P. Day, Inorg. Chem. \textbf{30}, 2680 (1991)).
499: 
500: \bibitem{Matsuda90} M. Matsuda \emph{et al.}, Phys. Rev. B
501: \textbf{42}, 10098(1990).
502: 
503: \bibitem{stagger} The staggered magnetization $m$ is determined by
504: multiplying the value of $m(x)/m(0)$ given in Ref. \cite{Mang04}
505: with the staggered magnetization at zero doping, $m(0)=0.4$, given
506: in Ref. \cite{Matsuda90}.
507: 
508: \bibitem{Yan06} X. Z. Yan and C. S. Ting, Phys. Rev. Lett.
509: \textbf{97}, 067001 (2006).
510: 
511: \bibitem{Dagan04} Y. Dagan, M. M. Qazilbash, C. P. Hill, V. N. Kulkarni,
512: and R. L. Greene, Phys. Rev. Lett. \textbf{92}, 167001 (2004).
513: 
514: \end{thebibliography}
515: 
516: \end{document}
517: 
518: 
519: 
520: \begin{figure}[ht]
521: \begin{center}
522: \includegraphics[width=0.18\columnwidth]{nccodispersion.eps}
523: \includegraphics[width=0.18\columnwidth]{nccodispersion.eps}
524: \includegraphics[width=0.4\columnwidth]{lowerband.eps}
525: \end{center}
526: \caption{Left and Middle: EDC for Nd$_2$CuO$_4$ and
527: Ca$_2$CuO$_2$Cl$_2$, respectively. Right: Energy dispersion of the
528: lower CT band along the $(\pi, 0)-(0, \pi )$ direction for
529: Nd$_2$CuO$_4$ and Ca$_2$CuO$_2$Cl$_2$. The minimum binding energy is
530: shifted to zero.\cite {Armitage02} Due to slight oxygen
531: nonstoichiometry, the Fermi level is pined at the top of the lower
532: Oxygen $2p$ band in Ca$_2$CuO$_2$Cl$_2$ and the bottom of the upper
533: Hubbard band (Cu $ed^{10}$), respectively. This leads to the
534: difference between the Fermi level related to the lower CT band. }
535: \label{fig:arpes}
536: \end{figure}
537: 
538: 
539: 
540: \begin{figure}[ht]
541: \begin{center}
542: %\includegraphics[width=0.8\columnwidth]{franck.eps}
543: \end{center}
544: %\caption{Cartoon Schematic of Spectral Weight in (Polaronic) Mott insulator.}
545: \end{figure}
546: