cond-mat0505393/ed.tex
1: \documentclass[twocolumn,aps,prl,amsmath,floatfix]{revtex4}
2: \usepackage{graphicx}
3: \begin{document}
4: 
5: \title{Novel Mott Transitions in the Nonisotropic Two-Band Hubbard Model}
6: \author{A. Liebsch} 
7: \affiliation{Institut f\"ur Festk\"orperforschung, 
8:              Forschungszentrum J\"ulich, 
9:              52425 J\"ulich, Germany}
10: \begin{abstract}
11: The Mott transition in a two-band Hubbard model involving subbands of 
12: different widths is studied as a function of temperature using dynamical 
13: mean field theory combined with exact diagonalization. The phase diagram 
14: is shown to exhibit two successive first-order transitions if the full
15: Hund's rule coupling is included. In the absence of spin-flip and 
16: pair-exchange terms the lower transition remains first-order while the
17: upper becomes continuous.
18: \\  \\
19: PACS numbers: 71.20.Be, 71.27.+a.
20: \end{abstract}
21: \maketitle
22: 
23: The nature of the metal insulator transition in materials involving
24: subbands of different widths has been intensively debated during the
25: recent years \cite{anisimov,epl,prl,koga1,prb,koga2,koga3,ferrero,%
26: medici,arita,song,knecht}. 
27: This issue is relevant for the understanding of the effect of strong
28: local Coulomb interactions in systems such as Ca$_{2-x}$Sr$_x$RuO$_4$. 
29: In this layer perovskite the partially filled Ru $t_{2g}$ bands for $x=2$ 
30: consist of wide $d_{xy}$ and narrow $d_{xz,yz}$ bands \cite{oguchi}. 
31: As pointed out in \cite{prl2000}, the onsite Coulomb energy lies between 
32: the subband widths: $W_{xz,yz}<U<W_{xy}$. The usual criterion with the 
33: parameter $U/W$ as a measure of the importance of correlations must then 
34: be generalized. The pure Sr compound is superconducting below 1.5 K 
35: \cite{maeno}. 
36: Iso-electronic replacement of Sr by Ca leads to an effective band 
37: narrowing due to octahedral distortions \cite{fang} 
38: and a metal insulator transition \cite{nakatsuji}. 
39: As a consequence of non-cubic crystal fields many other transition metal 
40: oxides also involve non-equivalent partially occupied subbands.   
41: 
42: A key question in these materials is whether the wide and narrow subbands
43: exhibit separate Mott transitions or whether single-particle hybridization
44:  and inter-orbital Coulomb interactions ensure a single transition for all 
45: bands simultaneously. This issue was studied initially by Anisimov et 
46: al.~\cite{anisimov} and \mbox{Liebsch} \cite{epl} for Ca$_{2-x}$Sr$_x$RuO$_4$ 
47: using simplified band structure models which did not yet include the full 
48: complexity due to Ca-induced octahedral distortions. Correlations were
49: treated in the dynamical mean field theory \cite{dmft} (DMFT) combined 
50: with the non-crossing approximation (NCA) and Quantum Monte Carlo (QMC) 
51: method, respectively. Since the NCA calculations (at $T=0$) neglected 
52: interorbital Coulomb interactions the results showed separate, 
53: `orbital-selective' Mott transitions for the narrow and wide subbands 
54: \cite{anisimov}. In contrast, the QMC calculations (at $T=0.125$~eV) 
55: included interorbital Coulomb interactions and suggested a common transition 
56: for all $t_{2g}$ bands \cite{epl}. 
57: 
58: Recent theoretical studies of the Mott transition in a paramagnetic 
59: two-band model system have led to apparent contradictions
60: \cite{prl,koga1,prb,koga2,koga3,ferrero,medici,arita}.  
61: Including the full Hund's rule coupling Koga {\it et al.}~\cite{koga1} 
62: found orbital-selective metal insulator transitions at $T=0$. 
63: Neglecting spin-flip and pair-exchange terms Liebsch \cite{prb} obtained 
64: at $T>0$ a single first-order transition, followed by a 
65: bad-metallic/non-Fermi-liquid (NFL) phase. 
66: As we argue below, the various results and conclusions are 
67: consistent provided that the choice of crucial parameters such as 
68: temperature and Hund's rule coupling is properly taken into account.
69: 
70: In the present work we combine finite temperature DMFT with exact 
71: diagonalization \cite{ed} (ED) to determine the $T/U$ phase diagram 
72: of a two-band system consisting of non-hybridizing, half-filled subbands 
73: with semi-circular density of states of width $W_1=2$~eV and $W_2=4$~eV. 
74: The subbands interact 
75: via intra- and interorbital Coulomb matrix elements $U$ and $U'=U-2J$, 
76: where $J$ is the Hund's rule exchange integral. In contrast to the 
77: multi-band QMC approach, which includes only Ising-like exchange terms to 
78: avoid sign problems at low temperatures \cite{held}, ED permits also the
79: consideration of spin-flip and pair-exchange interactions. To be specific 
80: we take $J=U/4$ which is approximately satisfied in several transition 
81: metal oxides. 
82: 
83: The main result of this work is that the $T/U$ phase diagram in the 
84: presence of the full Hund's rule exchange exhibits 
85: {\it two successive first-order phase transitions}, with separate  
86: hysteresis loops and coexistence regions. The intermediate region 
87: corresponds to the $T>0$ analog of the orbital-selective Mott 
88: (OSM) phase obtained at \mbox{$T=0$} in \cite{koga1}. On the other hand, 
89: if spin-flip and pair-exchange terms are omitted, we find a 
90: {\it single first-order transition succeeded by a non-Fermi-liquid 
91: phase}, in agreement with previous QMC results \cite{prb}. 
92: Both trends are consistent with those obtained by several groups 
93: \cite{koga1,koga2,ferrero,medici,arita} for the same two-band model 
94: at $T=0$.
95: 
96: The ED/DMFT results are derived from a two-band generalization of the 
97: approach employed for single-bands \cite{dmft,ed,gracias}. Since at $T>0$ 
98: all states of the impurity Hamiltonian are used in the construction 
99: of the subband Green's functions $G_i(i\omega_n)$, two bath 
100: levels per impurity level are taken into account ($n_s=6$ per spin). 
101: To check the accuracy of this approximation we have evaluated the $T/U$ 
102: phase diagram of a single band for $n_s=3,\dots,6$. As shown in Fig.~1,
103: the stability boundaries $U_{c1}(T)$ and $U_{c2}(T)$ for $n_s=3$ are 
104: slightly too low. Nevertheless, the overall shape of the phase diagram 
105: agrees qualitatively with the converged results for $n_s=6$ \cite{tong}. 
106: Thus, we are confident that in the two-band case $n_s=6$ also yields a 
107: reasonable picture of the phase diagram. Preliminary results for $n_s=8$ 
108: will also be presented. 
109: 
110: \begin{figure}[t!]%1
111:   \begin{center}
112:   \includegraphics[width=4.5cm,height=8cm,angle=-90]{Fig1.ps}
113:   \end{center}
114:   \vskip-2mm
115: \caption{
116: Phase diagram for one-band Hubbard model, calculated within ED/DMFT
117: for $n_s=3,4,6$. Symbols (x): $T=0$ transitions obtained in \cite{bulla}.
118: }\end{figure}
119: 
120: To analyze the metal/insulator transition we study the quasiparticle weights 
121: $Z_i = 1/[1-d{\rm Re}\,\Sigma_i(\omega)/d\omega\vert_{\omega=0}]$, 
122: which in the metallic range can be represented as
123: $Z_i \approx 1/[1- {\rm Im}\,\Sigma_i(i\omega_0)/ \omega_0]$, where 
124: $\Sigma_i(i\omega_0)$ is the subband self-energy at the first Matsubara 
125: frequency. Fig.~2(a) shows $Z_i(U)$ for $J'=J=U/4$, where $J'$ 
126: denotes spin-flip and pair-exchange terms \cite{fit}. Two critical regions 
127: can be identified, each with hysteresis loops characteristic of first-order 
128: phase transitions. The coexistence areas are $U^<_{c1}(T) < U < U^<_{c2}(T)$ 
129: near 2.0~eV and $U^>_{c1}(T) < U < U^>_{c2}(T)$ near 3.0~eV, where 
130: $U^<_{cn}(T)$ and $U^>_{cn}(T)$ are the stability boundaries obtained for 
131: increasing ($n=2$) and decreasing ($n=1$) $U$.
132: Let us denote the true critical energies of these transitions
133: as $U^<_c(T)$ and $U^>_c(T)$. Below $U^<_c(T)$ both bands are metallic 
134: while above $U^>_c(T)$ both are insulating. At the lower transition 
135: both bands undergo first-order transitions -- but in fundamentally 
136: different ways: $Z_1(U)$ becomes very small while $Z_2(U)$ drops to a 
137: finite value. The narrow band therefore undergoes a `complete' 
138: metal/insulator transition, whereas the wide band exhibits an `incomplete' 
139: transition to a new, considerably more correlated phase. This band becomes 
140: fully insulating near 3.0~eV, where it exhibits a weak second hysteresis loop. 
141: To our knowledge this is the first time that {\it sequential first-order 
142: transitions} are identified in the $T/U$ phase diagram of a Hubbard model 
143: involving non-equivalent bands \cite{kawakami}.
144: If such a material could be encountered experimentally, the conductivity 
145: as a function of pressure would show two consecutive jumps. 
146: 
147: Fig.~2(b) shows $Z_i(U)$ for $J'=0$, $J=U/4$, i.e., in the absence of 
148: spin-flip and pair-exchange terms.  The results are similar to those in 
149: Fig.~2(a), with the important exception that the wide band above the lower
150: transition is even more correlated and the upper transition is now continuous 
151: at smaller $U$ \cite{continuous}. 
152: Remarkably, the lower transition remains first-order for both subbands. 
153: In this case the conductivity shows a jump at $U^<_c(T)$ but a change of slope 
154: at $U^>_c(T)$. The results in Fig.~2(b) confirm the picture obtained previously 
155: within the QMC for $T>0$, $J'=0$ which showed the existence of a single 
156: first-order transition followed by a mixed insulating/bad-metallic phase 
157: \cite{prb,ipt,bluemer1}. They also demonstrate that the two-band ED/DMFT 
158: for $n_s=6$ is qualitatively accurate \cite{hyst}. 
159: 
160: \begin{figure}[t!]%2
161:   \begin{center}
162:   \includegraphics[width=8cm,height=8cm,angle=-90]{Fig2a.ps}
163:   \includegraphics[width=8cm,height=8cm,angle=-90]{Fig2b.ps}
164:   \end{center}
165:   \vskip-2mm
166: \caption{
167: $Z_i(U)$ of nonisotropic two-band Hubbard model, calculated within ED/DMFT. 
168: (a) $J'=J=U/4$, (b) $J'=0$, $J=U/4$. Solid (dashed) curves: narrow (wide)
169: band. Results for different temperatures are displaced vertically by $0.3$.
170: }\end{figure}
171: 
172: The phase diagrams deduced from the ED results for $T\ge2.5$~meV are shown 
173: in Fig.~3. For $J'=J$ as well as $J'=0$ the transition at $U^<_c(T)$ is 
174: first-order for both subbands. The subsequent transition of the wide band at 
175: $U^>_c(T)$ is first-order for $J'=J$ but continuous for $J'=0$.  
176: At $U^<_c(T)$ the metal/insulator transition is complete only for the narrow 
177: band. The wide band first undergoes a transition to a more strongly correlated 
178: phase and becomes truly insulating at the second transition at $U^>_c(T)$. 
179: The overall shape of the phase diagram for $J'=J$ agrees with the one 
180: recently obtained by Inaba {\it et al.} \cite{kawakami}. 
181: 
182: \begin{figure}[t!]%3
183:   \begin{center}
184:   \includegraphics[width=5cm,height=8cm,angle=-90]{Fig3a.ps}
185:   \includegraphics[width=5cm,height=8cm,angle=-90]{Fig3b.ps}
186:   \end{center}
187:   \vskip-2mm
188: \caption{
189: Phase diagram for nonisotropic two-band Hubbard model, calculated within 
190: ED/DMFT. (a) $J'=J=U/4$, (b) $J'=0$, $J=U/4$. Solid dots in (a), (b): 
191: stability boundaries of both subbands near lower first-order transition.
192: Open dots in (a): stability boundaries of wide band near upper first-order 
193: transition. Dashed line in (b): approximate location of continuous transition 
194: of wide band. Symbols (x): $T=0$ transitions obtained in \cite{koga1,koga2,arita};
195: (+): transitions at $T=31$~meV  obtained in \cite{prb,knecht}.    
196: Lines are guides to the eye.
197: }\end{figure}
198: 
199: Fig.~3 also shows the critical Coulomb energies obtained at $T=0$ 
200: \cite{koga1,koga2,arita} and $T=31$~meV \cite{prb,knecht}. As illustrated 
201: in Fig.~1 for the one-band case, the ED scheme with small $n_s$ underestimates 
202: the critical Coulomb energies by about 0.1 to 0.2~eV. Preliminary results 
203: for $n_s=8$ indicate similar shifts in the two-band case. Taking these 
204: displacements into account the ED results shown in Fig.~3 are in excellent 
205: correspondence with those obtained at $T=0$ and $T=31$~meV.
206: Evidently the conflicting conclusions reached in 
207: \cite{anisimov,epl,prl,koga1,prb,koga2,ferrero,medici,arita} concerning 
208: the nature of the Mott transition in multi-band systems were caused by 
209: different behaviors obtained for $T=0$ vs. $T>0$ and $J'=J$ vs. $J'=0$. 
210: Accounting for these different parameter choices, the DMFT treatments 
211: are consistent. 
212: 
213: Quasiparticle spectra derived within the QMC/DMFT for $J'=0$ \cite{prb} 
214: showed that in the intermediate phase the self-energy of the wide band  
215: at small $\omega_n$ deviates significantly from metallic $\sim\omega_n$ 
216: behavior \cite{biermann}. Accordingly, the quasi-particle spectra show 
217: a pseudogap which for larger $U$ gets more pronounced \cite{bad}, indicating 
218: progressive non-Fermi-liquid properties. A true gap opens at $U^>_c(T)$. 
219: Thus, $Z_2(U)>0$ 
220: in the region $U^<_c(T)<U<U^>_c(T)$ does not imply existence of quasiparticles. 
221: As indicated in Fig.~3(b) the intermediate `orbital-selective Mott' phase in 
222: the absence of spin-flip and pair-exchange is in fact a mixed insulating/NFL 
223: phase. The same trend is found using the $T>0$ ED/DMFT \cite{future}. 
224: 
225: \begin{figure}[t!]%4
226:   \begin{center}
227:   \includegraphics[width=5cm,height=4cm,angle=-90]{Fig4a.ps}\hskip-3mm
228:   \includegraphics[width=5cm,height=4cm,angle=-90]{Fig4b.ps}
229:   \end{center}
230:   \vskip-2mm
231: \caption{
232: Self-energy of wide band for $J'=J$ in intermediate phase at $U=2.4$~eV. 
233: (a) $n_s=6$; (b) $n_s=8$. Solid dots: $T=10$~meV; empty dots: $T=5$~meV; 
234: (+): $T=2.5$~meV. 
235: }\end{figure}
236: 
237: For $J'=J$, the low-frequency analysis of $\Sigma_2(i\omega_n)$ is more
238: intricate. As shown in Fig.~4(a) for $n_s=6$, $\Sigma_2(i\omega_n)$ 
239: reveals deviations from metallic $\sim\omega_n$ variation, giving rise to 
240: small pseudogaps in the quasiparticle spectra. This behavior is incompatible 
241: with Im\,$\Sigma_2(\omega)\sim\omega^2$ and suggests that, as for $J'=0$, in 
242: the intermediate phase the wide band at $T>0$ violates Fermi-liquid theory. 
243: % \cite{0505393}.
244: The extension of the present ED approach to $n_s=8$ indicates, however, that 
245: the additional bath levels are important for the low-frequency variation of 
246: $\Sigma_2(i\omega_n)$. As can be seen in Fig.~4(b), the deviations are absent 
247: and the $\sim\omega_n$ variation is consistent with Fermi-liquid behavior.
248: In fact, the shoulder near $\omega_0=0.06$ suggests that Fermi-liquid 
249: properties persist up to about $T=\omega_0/\pi\approx 20$~meV. Thus, 
250: the OSM phase in Fig.~3(a) is the $T>0$ analog of the orbital-selective 
251: Mott phase identified first by Koga {\it el al.}~\cite{koga1} at $T=0$.
252: A more complete discussion of the results for $n_s=8$ will be given elsewhere 
253: \cite{future}. Because of finite size limitations of the present ED/DMFT 
254: scheme associated with the small number of bath levels, a precise 
255: determination of low-temperature properties is not possible. Nevertheless,      
256: an approximate extrapolation of $\Sigma_2(i\omega_n)$ indicates that the 
257: $T\rightarrow0$ limit for $J'=J$ satisfies Fermi-liquid criteria, in 
258: agreement with previous $T=0$ studies \cite{koga1,koga2,ferrero,medici,arita}.
259: 
260: In view of the importance of spin-flip and pair-exchange terms for 
261: the Mott transition in multi-band materials \cite{pruschke} it is 
262: desirable to investigate the $T/U$ phase diagram by methods 
263: which permit adequate treatment of the complete Hund's rule matrix, 
264: for instance, a two-band extension of $T>0$ numerical renormalization
265: group studies \cite{costi}; see also \cite{kawakami}.   
266: Because of sign problems, recent QMC extensions including spin-flip 
267: and pair-exchange terms are limited to $T\ge 1/6$~eV$\ \gg T_c$ 
268: \cite{koga3} or $T=0$ \cite{arita}. Also, other recent works 
269: \cite{ferrero,medici} employing a variety of quantum impurity 
270: methods deal so far mainly with $T=0$ and do not yet allow the 
271: identification of the multi-band Mott transition at general $T,U$ 
272: values. 
273: 
274: In summary, the $T/U$ phase diagram of the Hubbard model involving 
275: half-filled, non-equivalent subbands is shown to be remarkably rich.
276: The competing kinetic energy scales, coupled via Coulomb and exchange 
277: energies, give rise to {\it sequential first-order phase transitions}. 
278: The lower transition separates a purely metallic phase from a mixed
279: phase with insulating narrow and strongly-correlated wide subbands. 
280: The wide band becomes insulating at the second first-order transition. 
281: Omission of spin-flip and pair-exchange terms enhances the correlations 
282: in the wide band in the intermediate phase so that it no 
283: longer satisfies Fermi-liquid criteria, and modifies the upper phase 
284: transition from first-order to continuous.  
285: 
286: For the analysis of experimental data of materials such as 
287: Ca$_{2-x}$Sr$_x$RuO$_4$ it is necessary to account also for hybridization 
288: between orbitals. Preliminary studies of this effect within two-band 
289: models for $T\gg T_c$ \cite{koga3} and $T=0$ \cite{medici,song} suggest 
290: significant changes. Moreover, spatial fluctuations \cite{kotliar} and 
291: deviations from half-filling might play a decisive role close to the 
292: Mott transition. More work is needed to investigate whether both 
293: first-order transitions persist in the presence of these effects, or 
294: whether the weak first-order behavior of the upper transition disappears 
295: and only the dominant lower transition survives as the common first-order 
296: Mott transition for all bands. 
297: 
298: I like to thank Th. Costi, K. Held, N. Kawakami, and E. Koch for valuable 
299: discussions.
300: 
301: Email: a.liebsch@fz-juelich.de
302: \begin{thebibliography}{99}
303: \bibitem{anisimov}
304:    V. I. Anisimov, I. A. Nekrasov, D. E. Kondakov, T. M. Rice 
305:       and M. Sigrist,
306:    Euro. Phys. J. B {\bf 25}, 191 (2002). 
307: 
308: \bibitem{epl} 
309:    A. Liebsch, 
310:       Europhys. Lett. {\bf 63}, 97 (2003). 
311: 
312: \bibitem{prl}  
313:    A. Liebsch, 
314:       Phys. Rev. Lett. {\bf 91}, 226401 (2003).
315: 
316: \bibitem{koga1}
317:    A. Koga, N. Kawakami, T. M. Rice, and  M. Sigrist,
318:       Phys. Rev. Lett. {\bf 92}, 216402 (2004).
319: 
320: \bibitem{prb}  
321:    A. Liebsch, 
322:       Phys. Rev. B  {\bf 70}, 165103 (2004).
323: 
324: \bibitem{koga2}
325:    A. Koga, N. Kawakami, T. M. Rice, and  M. Sigrist,
326:       cond-mat/0406457.
327: 
328: \bibitem{koga3}
329:    A. Koga, N. Kawakami, T. M. Rice, and  M. Sigrist,
330:       cond-mat/0503651.
331: 
332: \bibitem{ferrero}
333:    M. Ferrero, F. Becca, M. Fabrizio, and M. Capone,
334:       cond-mat/0503759
335: 
336: \bibitem{medici}
337:    L. de' Medici, A. Georges and S. Biermann,
338:       cond-mat/0503764.
339: 
340: \bibitem{arita}
341:    R. Arita and K. Held,
342:       cond-mat/0504040. 
343: 
344: \bibitem{song}
345:    Yun Song and L.-J. Zou, 
346:       cond-mat/0504420. 
347: 
348: \bibitem{knecht}
349:    C. Knecht, N. Bl\"umer, and P.G.J. van Dongen,
350:       cond-mat/0505106.
351: 
352: \bibitem{oguchi}
353:    T. Oguchi, 
354:       Phys. Rev. B {\bf 51}, 1385 (1995);
355:    I. I. Mazin and D. Singh,
356:       Phys. Rev. Lett. {\bf 79}, 733  (1997).
357: 
358: \bibitem{prl2000}  
359:    A. Liebsch and A. I. Lichtenstein,
360:       Phys. Rev. Lett. {\bf 84}, 1591 (2000).  
361: 
362: \bibitem{maeno}
363:    Y. Maeno, T. M.  Rice and M. Sigrist,
364:       Physics Today {\bf 54}(1), 42 (2001).
365: 
366: \bibitem{fang}  
367:    A. Fang and K. Terakura, 
368:       Phys. Rev. B {\bf 64}, R020509 (2001);
369:    A. Fang, N. Nagaosa, and K. Terakura, 
370:       Phys. Rev. B {\bf 69}, 045116  (2004).
371: 
372: \bibitem{nakatsuji}
373:    S. Nakatsuji and Y. Maeno,
374:       Phys. Rev. Lett. {\bf 84}, 2666 (2000). 
375: 
376: \bibitem{dmft} 
377:    For a review, see:
378:    A. Georges, G. Kotliar, W. Krauth and M. J. Rozenberg, 
379:    Rev. Mod. Phys. {\bf 68}, 13 (1996).
380: 
381: \bibitem{ed}  
382:    M. Caffarel and W. Krauth, 
383:       Phys. Rev. Lett. {\bf 72}, 1545 (1994).
384: 
385: \bibitem{held}  
386:    For a discussion of this problem, see: 
387:    K. Held and D. Vollhardt, 
388:       Eur. Phys. J. B {\bf 5}, 473 (1998). 
389: 
390: \bibitem{gracias}
391:    I like to thank the authors of \cite{dmft} for making the single-band ED 
392:    code available, and A.I. Lichtenstein for parts of the two-band ED code.
393: 
394: \bibitem{bulla}
395:    See: R. Bulla, T. A. Costi, and D. Vollhardt, 
396:       Phys. Rev. B {\bf 64}, 045103 (2001) and references herein.
397: 
398: \bibitem{tong}
399:    N.H. Tong, S.Q. Shen, and F.C. Pu, 
400:       Phys. Rev. B {\bf 64}, 235109 (2001).
401: 
402: \bibitem{fit}
403:    More accurate results at low $\omega_n$ and $T$ are obtained by minimizing 
404:    differences between impurity and Anderson model Green's functions via 
405:    $\sum_n \vert\Delta G_i(i\omega_n)\vert^2$ rather than 
406:    $\sum_n \vert\Delta G_i(i\omega_n)\vert$. I thank E. Koch for this suggestion.
407:  
408: \bibitem{kawakami}
409:    After completion of this work we learned of similar results 
410:    by K. Inaba {\it et al.}, cond-mat/0506151.  
411: 
412: \bibitem{continuous}
413:    The term `continuous' indicates here that there is no jump; it may
414:    also signify crossover behavior.
415: 
416: \bibitem{ipt}
417:    Surprisingly, two-band DMFT using iterated perturbation theory for 
418:    $J'=J$ reveals the lower transition as first-order but the upper as
419:    continuous \cite{prb}. 
420: 
421: \bibitem{bluemer1}
422:    Recent QMC/DMFT results \cite{knecht} for $J=U/4,\,J'=0$ are in 
423:    agreement with those of \cite{prb}. Compare $Z_i(U)$ and $N_i(\omega)$ 
424:    in Figs.~9(b) and 11 \cite{prb} with Figs.~1(b)(inset) and 3,4 
425:    \cite{knecht}. 
426:    Thus, the claims in \cite{knecht} that ``the second transition was 
427:    not seen in earlier QMC and IPT studies'' and ``our high-precision 
428:    data correct earlier QMC results'' are unfounded. 
429:    See also: A. Liebsch, cond-mat/0506138.
430: 
431: \bibitem{hyst}
432:    The QMC hysteresis region is slightly wider than in ED.
433:    This could be due to slow convergence of QMC at $T=0.02$~eV 
434:    and too small $n_s$ in ED. 
435: 
436: \bibitem{biermann}
437:    Similar QMC results were obtained for $J'=0$, $T>0$, $W_1=0.1W_2$ by 
438:    Biermann {\it et al.}, cond-mat/0505737.
439: 
440: \bibitem{bad}
441:    See also Fig.~3 of \cite{knecht}. 
442: 
443: %\bibitem{0505393}
444: %   A.Liebsch, cond-mat/0505393.
445: 
446: \bibitem{future}
447:    A. Liebsch, to be published.
448: 
449: \bibitem{pruschke}
450:    See also: Th. Pruschke and R. Bulla, cond-mat/0411186.
451: 
452: \bibitem{costi}
453:    Th. Costi, to be published.   
454: 
455: \bibitem{kotliar}
456:    O. Parcollet, G. Biroli, and G. Kotliar,
457:       Phys. Rev. Lett. {\bf 92}, 226402 (2004).
458: \end{thebibliography}
459: \end{document}
460: 
461: 
462: