cond-mat0603548/ed.tex
1: \documentclass[twocolumn,aps,prl,amsmath,floatfix]{revtex4}
2: \usepackage{graphicx}
3: \begin{document}
4: \title{
5: Non-Fermi-liquid phases in the two-band Hubbard model:\\
6: Finite-temperature exact diagonalization study of Hund's rule coupling}
7: \author{A.~Liebsch and T.~A.~Costi} 
8: \affiliation{Institut f\"ur Festk\"orperforschung, 
9:              Forschungszentrum J\"ulich, 
10:              52425 J\"ulich, Germany}
11: \begin{abstract}
12: The two-band Hubbard model involving subbands of different widths is 
13: investigated via finite-temperature exact diagonalization (ED) and 
14: dynamical mean field theory (DMFT). In contrast to the quantum Monte Carlo 
15: (QMC) method which at low temperatures includes only Ising-like exchange 
16: interactions to avoid sign problems, ED permits a treatment of Hund's 
17: exchange and other onsite Coulomb interactions on the same footing.
18: The role of finite-size effects caused by the limited number of bath 
19: levels in this scheme is studied by analyzing the low-frequency behavior 
20: of the subband self-energies as a function of temperature, and by 
21: comparing with numerical renormalization group (NRG) results for an
22: effective one-band model. 
23: For half-filled, non-hybridizing bands, the metallic and insulating 
24: phases are separated by an intermediate mixed phase with an insulating  
25: narrow and a bad-metallic wide subband. The wide band in this phase 
26: exhibits different degrees of non-Fermi-liquid behavior, depending on 
27: the treatment of exchange interactions. Whereas for complete Hund's 
28: coupling, infinite lifetime is found at the Fermi level, in the 
29: absence of spin-flip and pair-exchange, this lifetime becomes finite. 
30: Excellent agreement is obtained both with new NRG and previous 
31: QMC/DMFT calculations. These results suggest that-finite temperature 
32: ED/DMFT might be a useful scheme for realistic multi-band materials.\\
33: \\
34: PACS. 71.20.Be  Transition metals and alloys - 71.27+a Strongly correlated
35: electron systems 
36: \end{abstract}
37: \maketitle
38: 
39: \section{1 \ Introduction}
40: 
41: Strongly correlated materials exhibit a wealth of fascinating 
42: physical phenomena associated with complex single-electron and 
43: many-electron interactions. Transition metal oxides, for example,
44: tend to have partially filled shells of highly correlated $d$ 
45: electrons, surrounded by complicated lattice geometries, with 
46: many atoms and electrons per unit cell. The theoretical description 
47: of the electronic properties of these materials is a challenging 
48: topic in condensed matter physics. Significant advances were 
49: achieved during recent years via the dynamical mean field theory 
50: (DMFT) \cite{rmp,reviews} which provides a treatment of 
51: single-electron and many-electron interactions on the same footing. 
52: 
53: For realistic materials, DMFT has been used extensively in 
54: combination with the quantum Monte Carlo (QMC) method \cite{qmc}. 
55: This approach has the advantage that it can be readily applied to 
56: systems consisting of two or more subbands. It has the drawback,
57: however, that, to avoid sign problems at low temperatures, it 
58: includes only Ising-like exchange interactions \cite{held}.
59: Improvements of the QMC method to include the full Hund's coupling 
60: are so far limited to $T=0$ or rather high temperatures 
61: \cite{arita,koga3,rubtsov}.  
62: In view of this limitation there is clearly a need to explore 
63: alternative methods that are applicable to multi-band materials 
64: and complete onsite exchange interactions.
65: 
66: The aim of the present work is two-fold: First, exact diagonalization 
67: (ED) \cite{ed,rmp} is proposed as a potentially highly useful impurity 
68: solver for finite-$T$ DMFT studies of realistic systems. The attractive 
69: feature of this approach is that, in contrast to finite-$T$ QMC, onsite 
70: Coulomb and exchange interactions are treated on the same basis. 
71: Second, we apply finite-$T$ ED/DMFT to a highly nontrivial system which 
72: has recently received considerable attention, namely, the two-band 
73: Hubbard model consisting of subbands of different widths
74: \cite{anisimov,epl,prl,koga1,prb70,koga2,koga3,prl05,inaba,knecht,
75: biermann05,ferrero,medici,arita,song}. The phase diagram of this 
76: system was recently evaluated in Refs.~\cite{prl05,inaba}. Here we focus 
77: on the electronic properties of the so-called orbital-selective phase
78: in which the narrow band is insulating while the wide band is still 
79: metallic. To examine the influence of finite-size effects in the ED
80: approach, and to investigate the electronic properties in the limit 
81: of low temperatures, we have also performed numerical renormalization 
82: group (NRG) DMFT calculations for a simplified two-band model which 
83: is particularly suited for the intermediate phase.
84: 
85: The main result of this paper is that the two-band ED/DMFT calculations 
86: provide a correct picture of the electronic properties of the Hubbard 
87: model involving nonequivalent subbands, including the unusual 
88: non-Fermi-liquid properties of the orbital-selective phase.  
89: This outcome is remarkable since, for computational reasons, the number 
90: of bath levels per impurity orbital is necessarily smaller than in 
91: analogous one-band models. Nevertheless, despite this limitation, 
92: the ED results are in good qualitative or, in some cases, quantitative 
93: agreement with the NRG results. In the Ising case they also agree very 
94: well with previous QMC/DMFT results \cite{prb70,biermann05}. 
95: 
96: The overall consistency with the NRG and QMC calculations suggests 
97: that finite-$T$ ED/DMFT might indeed be very useful in the future 
98: to analyze realistic materials, especially, if full diagonalization
99: is replaced by finite-$T$ Lanczos methods \cite{jaklic,capone}
100: in order to be able to deal with larger cluster sizes.
101: 
102: Hubbard models involving nonequivalent subbands are relevant for 
103: compounds such as Ca$_{2-x}$Sr$_x$RuO$_4$ and Na$_x$CoO$_2$, where, 
104: as a result of the nearly two-dimensional lattice structure, coexisting 
105: narrow and wide bands arise naturally. Thus, onsite Coulomb energies 
106: can be simultaneously large and small relative to the widths of 
107: important subbands. As a function of doping concentration, both 
108: materials give rise to a remarkably rich sequence of phases, including 
109: superconductivity and Mott insulating behavior \cite{maeno,nacoo}. 
110: Evidently, the competition between multiple kinetic energy 
111: scales and local Coulomb and exchange energies is an important 
112: feature of these strongly correlated systems.
113: 
114: A consistent treatment of Coulomb and exchange interactions in materials
115: of this kind is important since it has recently become clear that the 
116: Hund's coupling has a decisive influence on the nature of the Mott 
117: transition \cite{koga2,prl05,ferrero,medici,arita,pruschke}. 
118: In fact, the different treatments of exchange terms in earlier finite-$T$ 
119: QMC \cite{epl,prl,prb70} and zero-$T$ ED \cite{koga1} calculations have 
120: given rise to some confusion, with apparently contradictory results. 
121: As was clarified in Ref.~\cite{koga2} for $T=0$ and in Ref.~\cite{prl05} 
122: for $T>0$, however, the QMC and ED results are in agreement provided 
123: that exchange interactions are treated in the same manner. Thus, for 
124: full Hund's coupling the two-band Hubbard model exhibits successive 
125: first-order transitions. In striking contrast, in the absence of 
126: spin-flip and pair-exchange only the lower transition remains 
127: first-order \cite{prb70,prl05,knecht}. 
128: 
129: 
130: Moreover, as will be discussed in detail below, the nature of the 
131: intermediate phase depends in a subtle manner on the treatment of 
132: exchange interactions. In this regard the ED and NRG results yield 
133: the following picture: For full Hund's coupling, the wide band has 
134: infinite lifetime at $E_F$ but does not satisfy Fermi-liquid criteria 
135: at finite frequencies. This finding is consistent with recent results 
136: obtained by Biermann {\it et al.} \cite{biermann05} in $T=0$ two-band 
137: ED/DMFT calculations. For Ising exchange, instead, correlations 
138: are significantly enhanced and the lifetime becomes finite even at 
139: $E_F$, in agreement with QMC calculations \cite{prb70,biermann05}.   
140: 
141: {\it
142: Thus, the Mott transition in the Hubbard model involving different
143: subbands does not consist of equivalent sequential transitions. Instead,
144: when the narrow band becomes insulating, the wide band is forced into a
145: bad-metallic state whose deviations from Fermi-liquid behavior depend
146: in a qualitative manner on the treatment of exchange interactions.} 
147:          
148: In the past, finite-$T$ ED/DMFT methods have been applied mainly 
149: to single-band cases, where the incorporation of an appropriate 
150: cluster of bath levels is feasible. Throughout this paper we consider 
151: two impurity levels, each surrounded by either 2 or 3 bath levels, 
152: giving total cluster sizes $n_s=6$ or $n_s=8$ per spin. Below we 
153: demonstrate that even a cluster size of $n_s=6$ provides qualitatively 
154: correct subband self-energies. For instance, at $T=10$~meV the critical 
155: Coulomb energies of the orbital-selective Mott transitions for 
156: $n_s=6$ differ by only about {0.1\ldots0.2}~eV from those derived 
157: for $n_s=8$. This finding is interesting since it suggests that 
158: finite-$T$ ED/DMFT calculations for more realistic three-band models 
159: using a cluster size of $n_s=9$ (three impurity levels, each coupled
160: to two bath levels) should be useful. This would allow one to 
161: re-examine the Mott transition in systems that have been studied 
162: previously using QMC/DMFT for Ising exchange interactions.
163: 
164: A more accurate representation of low-frequency properties at low
165: temperatures requires three bath levels per impurity orbital: one 
166: near the Fermi level to provide adequate metallicity, and two for 
167: the upper and lower Hubbard bands, giving $n_s=8$ per spin. This 
168: extension leads to a significant reduction of finite-size effects.
169: Several comparisons of results for $n_s=6$ and $n_s=8$ are provided 
170: below to illustrate the range of applicability of the two-band 
171: ED/DMFT approach.    
172: 
173: The outline of the paper is as follows: In Section 2 the multiband
174: Hubbard model is specified and its numerical solution via the 
175: finite-$T$ exact diagonalization method is discussed. Section 3
176: describes the effective one-band model used in the NRG approach
177: to evaluate the electronic properties of the wide band in the 
178: intermediate phase when the narrow band is insulating. 
179: In Section 4 the ED/DMFT is applied to the purely metallic phase
180: just below the first Mott transition. 
181: Section 5 deals with the intermediate phase in the presence of full
182: Hund's coupling, where wide band exhibits infinite lifetime at
183: $E_F$, combined with non-Fermi-liquid behavior at finite frequencies.
184: In Section 6 this intermediate phase is considered in the absence of
185: spin-flip and pair-exchange terms, giving rise to finite lifetime 
186: even at $E_F$. In Section 7 the two-band ED approach is applied 
187: to the case studied previously within QMC/DMFT \cite{prb70}.
188: Section 8 presents analogous results for the case considered 
189: recently within QMC/DMFT by Biermann {\it et al.} \cite{biermann05}. 
190: Section 9 contains a brief discussion of iterated perturbation 
191: theory (IPT) DMFT results for the two-band model and Section 10 
192: provides a summary and outlook.             
193: 
194: \section{2 \ Multiband exact diagonalization method} %II
195: 
196: The two-band Hubbard model for non-equivalent subbands is represented
197: by the Hamiltonian:
198: \begin{eqnarray}
199:    H &=&  H_0 \ + \ H_1 \ + \ H_2 \             \nonumber\\ 
200:  H_0 &=& \sum_{lmi\sigma}  t_{lmi} 
201:              c_{li\sigma}^+ c_{mi\sigma}  \nonumber\\
202:  H_1 &=& \sum_{li} U n_{li\uparrow} n_{li\downarrow} \ + \
203:          \sum_{l\sigma\sigma'} (U'-J\delta_{\sigma\sigma'}) 
204:                  n_{l1\sigma} n_{l2\sigma'}                \nonumber
205: \end{eqnarray}
206: \begin{eqnarray}
207:  H_2 &=& -J'\sum_l [c_{l1\uparrow}^+ c_{l1\downarrow}
208:             c_{l2\downarrow}^+ c_{l2\uparrow}    +  {\rm H.c.} ] 
209:                                                             \nonumber\\
210:      &&  -J'  \sum_l [c_{l1\uparrow}^+ c_{l1\downarrow}^+                   
211:                   c_{l2\uparrow} c_{l2\downarrow}    + {\rm H.c.} ]\, ,
212: \end{eqnarray}
213: where $c_{li\sigma}^+$ and $c_{li\sigma}$ are creation and
214: annihilation operators for electrons at site $l$ in orbital $i=1,2$
215: with spin $\sigma$ and  $n_{li\sigma}=c_{li\sigma}^+ c_{li\sigma}$. 
216: H.c.~denotes Hermitian conjugate terms.
217: $H_0$ is the single-particle Hamiltonian which we represent by 
218: half-filled, non-hybridizing bands with density of states 
219: $N_i(\omega)= 2/(\pi D_i)\sqrt{1- (\omega/D_i)^2}$ and widths 
220: $W_1=2$~eV, $W_2=4$~eV, where  $W_i=2D_i$. $H_1$ represents the 
221: anisotropic, Ising-like onsite Coulomb and exchange interactions 
222: and $H_2$ additional spin-flip and pair-exchange contributions. For 
223: isotropic Hund's coupling one has $J'=J=(U-U')/2$. Below, we discuss 
224: results for $J'=J$ as well as $J'=0$. The latter case corresponds to 
225: the Ising-like exchange treatment in previous QMC/DMFT calculations 
226: \cite{prl,prb70}.
227: 
228: The ED/DMFT results are derived from a two-band generalization of the 
229: approach employed for single bands \cite{ed,rmp}. 
230: The effective Anderson impurity Hamiltonian includes impurity levels 
231: $\epsilon_{1,2}$ and bath levels $\epsilon_{k=3\ldots n_s}$ 
232: (see Fig.~1). For particle-hole symmetry $\epsilon_{1,2}=0$. 
233: Each impurity level interacts with its own bath, so that for $n_s=6$  
234: impurity level 1 couples to bath levels 3,4 with $\epsilon_4=-\epsilon_3$, 
235: while level 2 couples to bath levels 5,6 with $\epsilon_6=-\epsilon_5$. 
236: For $n_s=8$ impurity levels 1,2 couple in addition to the bath levels 
237: $\epsilon_7=0$ and $\epsilon_8=0$, respectively.
238: 
239: \begin{figure}[b!]%1  see Gnuz in  ED8/Jump  and   ED6/T=0.010
240:   \begin{center}
241:   \includegraphics[width=6.0cm,height=3cm,angle=-0]{fig1.ps}
242:   \end{center}
243:   \vskip-2mm
244: \caption{\label{ED}
245: Level diagram for ED scheme. Blue lines: impurity levels;
246: red solid (plus dashed) lines: bath levels for $n_s=6$ ($n_s=8$). 
247: Green lines: Coulomb and exchange interactions between impurity
248: levels; red dotted lines: hopping interactions between impurity
249: and bath levels.
250: }\end{figure}
251: 
252: Since at $T>0$ all states of the impurity Hamiltonian are used in the 
253: construction of the subband Green's functions, for $n_s=6$ the largest 
254: matrix to be diagonalized corresponds to the sector $n_\uparrow=
255: n_\downarrow=3$, with dimension $[n_s!/((n_s/2)!)^2]^2=400$. 
256: For $n_s=8$ this dimension increases to $4900$. 
257: We neglect hybridization between bands, so that self-energies and Green's 
258: functions are diagonal in orbital space. Denoting eigenvalues and 
259: eigenvectors of the impurity Hamiltonian by $E_\nu$ and $|\nu \rangle$, 
260: the 
261: subband Green's functions are evaluated from the expression 
262: \begin{equation}
263:  G_i(i\omega_n) = \frac{1}{Z} \sum_{\nu\mu} 
264:             \frac{\vert\langle \nu|c_{0i\sigma}^+ |\mu \rangle \vert^2}
265:                     {E_\nu - E_\mu - i\omega_n}
266:    [e^{-\beta E_\nu} + e^{-\beta E_\mu}].    
267: \end{equation}
268: %\begin{eqnarray}
269: % G_i(i\omega_n) &=& \frac{1}{Z} \sum_{\nu\mu} 
270: %            \frac{\vert\langle \nu|c_{0i\sigma}^+ |\mu \rangle \vert^2}
271: %                    {E_\nu - E_\mu - i\omega_n}
272: %   [e^{-\beta E_\nu} + e^{-\beta E_\mu}]\, .    
273: %                    {E_\nu - E_\mu - i\omega_n} \nonumber \\ 
274: %   &&\  \times\ [e^{-\beta E_\nu} + e^{-\beta E_\mu}]\, .    
275: %\end{eqnarray}
276: Here, $\beta=1/k_B T$, $\omega_n=(2n+1)\pi/\beta$ are Matsubara 
277: frequencies, and $Z=\sum_\nu {\rm exp}(-\beta E_\nu)$ is the partition 
278: function. $l=0$ denotes the impurity site. Since we consider only 
279: paramagnetic phases the spin index of Green's functions and 
280: self-energies is omitted for convenience. Because of the particle-hole
281: symmetry of the present system, Green's functions and self-energies
282: are purely imaginary quantities. 
283: 
284: \begin{figure}[t!]%1  see Gnuz in  ED8/Jump  and   ED6/T=0.010
285:   \begin{center}
286: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{ZU8.ps}
287: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{ZU6.ps}
288:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig2a.ps}
289:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig2b.ps}
290:   \end{center}
291:   \vskip-2mm
292: \caption{\label{ZU}
293: $Z_i(U)$ as a function of $U$ for $J=J'=U/4$ at $T=10$~meV, 
294: calculated within ED/DMFT for different cluster sizes:
295: (a) $n_s=8$; (b) $n_s=6$. 
296: Solid (red) curves: narrow band, dashed (blue) curves: wide band.  
297: }\end{figure}
298: 
299: To provide an overview of the Mott transitions in this model and 
300: to illustrate the sensitivity of the critical Coulomb energies 
301: to the cluster size used in the ED/DMFT, we show in Fig.~\ref{ZU} 
302: the quantities $Z_i(U)=1/[1- {\rm Im}\,\Sigma_i(i\omega_0)/\omega_0]$
303: for $n_s=6$ and $n_s=8$ at $T=10$~meV, assuming $J'=J=U/4$.
304: Here $\Sigma_i(i\omega_0)$ are the subband self-energies at the 
305: first Matsubara frequency. Thus, for isotropic Hund's 
306: exchange and $n_s=8$, $U_{c1}\approx2.1$~eV and $U_{c2}\approx3.1$~eV.
307: (In our notation the $U_{ci=1,2}$ refer here to the subband critical
308: Coulomb energies for increasing $U$ and should not be confused with 
309: the stability boundaries for increasing vs decreasing $U$ at the 
310: individual Mott transitions.) Near both critical Coulomb
311: energies $Z_i(U)$ exhibit hysteresis behavior characteristic of 
312: first-order transitions. For $n_s=6$, the common metallic phase 
313: is stable only up to about $U_{c1}\approx2$~eV.
314: Thus, inclusion of zero-energy bath levels supports the metallic 
315: character of the DMFT solution. A shift of about 0.2~eV is found 
316: for $U_{c2}$ where the wide band becomes insulating. These shifts 
317: are consistent with single-band ED/DMFT results\cite{prl05} upon 
318: increasing $n_s$ from 3 to 4; only minor additional shifts occur 
319: in this case between $n_s=4$ and the fully converged results for 
320: $n_s=6$.
321: 
322: According to the QMC/DMFT results discussed in Ref.~\cite{prb70}, 
323: the Mott transitions for $J=U/4$ and $J'=0$, i.e., for Ising-like 
324: exchange, are located at $U_{c1}\approx2.1$~eV and 
325: $U_{c2}\approx2.7$~eV. Moreover, apart from the shift of $U_{c2}$, 
326: this transition is no longer first-order. The two-band ED/DMFT 
327: results in Ref.~\cite{prl05} for $n_s=6$ confirmed this fundamental 
328: difference between the $J'=J$ and $J'=0$ treatments.  
329: 
330: As will be discussed in detail below, the wide band in the intermediate
331: phase, i.e., for $U_{c1}<U<U_{c2}$, does not satisfy Fermi-liquid 
332: criteria. Thus, the fact that $Z_2(U)$ as defined above is finite 
333: in this region does not imply existence of ordinary quasiparticles.
334: In the case of full Hund's coupling, the imaginary part of the 
335: self-energy of the wide band vanishes in the low-frequency limit, 
336: but does not increase linearly at small $i\omega_n$, implying 
337: non-quadratic variation at real frequencies. For Ising-like exchange,
338: Im\,$\Sigma_2(i\omega_n)$ remains finite for $\omega_n\rightarrow0$,
339: implying finite lifetime even for states at $E_F$. Plots like those 
340: in Fig.~1 are nevertheless useful since they permit a convenient
341: identification of phase transitions. Thus, $Z_i(U)\rightarrow0$ 
342: at $U_{ci}$ indicates that the narrow or wide subbands become 
343: insulating for $U>U_{ci}$, respectively. (Of course, due to the 
344: discrete representation along the imaginary frequency axis at finite 
345: $T$, $Z_i(U)$ does not fully vanish.) 
346: 
347: \begin{figure}[t!]%2
348:   \begin{center}
349: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{W1W2J.ps}
350: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{W1W2J0.ps}
351:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig3a.ps}
352:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig3b.ps}
353:   \end{center}
354:   \vskip-2mm
355: \caption{\label{W1W2}
356: Imaginary part of self-energy of wide band in intermediate phase
357: for decreasing widths $W_1$ of narrow band, with fixed $W_2=4$~eV, 
358: $U=2.4$~eV, $n_s=8$, $T=10$~meV. Alternating solid (red) and dashed
359: (blue) curves: $W_1=0.4,\,0.8,\,\,1.2,\,1.6,\,2.0$~eV (from bottom).
360: (a) $J=J'=U/4$; (b)  $J=U/4$, $J'=0$.
361: }\end{figure}
362: 
363: To check the convergence of the ED results with cluster size $n_s$
364: we compare them with NRG calculations for an effective one-band model
365: which is most suitable for the intermediate phase and which is
366: specified in the following section. Since the narrow band is insulating 
367: in this phase one of the key features of the effective model is the 
368: omission of one-electron hopping in the narrow band. Accordingly, the
369: upper and lower Hubbard peaks will be centered at about $\pm U/2$, but
370: the influence of their width is neglected. (In a local description, 
371: the width of the Hubbard bands is approximately given by $W$ \cite{rmp}.)
372: In Fig.~\ref{W1W2} we show that this approximation indeed has only a 
373: minor effect on the electronic properties of the metallic wide band. 
374: Keeping $W_2=4$~eV fixed and changing $W_1$ between $0.1\,W_2$ and 
375: $0.5\,W_2$, we notice that, both for isotropic and anisotropic exchange 
376: coupling, the self-energy of the wide band at $U=0.6\,W_2=2.4$~eV
377: is almost unaffected by the original single-particle width of the 
378: insulating narrow band. The detailed electronic properties of these 
379: phases will be discussed in Sections 5 and 6.    
380: 
381: \section{3 \ Numerical renormalization group approach for 
382:           effective one-band model}
383: 
384: In the limit of an insulating narrow band interacting with a metallic 
385: wide band, the two-band model can be simplified to an effective one-band 
386: model by eliminating the high energy states associated with the upper 
387: and lower Hubbard bands of the narrow band \cite{biermann05}.
388: Neglecting the one-electron hopping in the narrow band, i.e., assuming 
389: $W_1=0$, and fixing the occupations of these orbitals, the terms in 
390: Eq.~(1) involving interorbital Coulomb and pair-exchange interactions
391: disappear. The effective Hamiltonian then reduces to 
392: \begin{eqnarray}
393:   H  & = & \sum_{lm\sigma}t_{lm}c^{+}_{l\sigma}c_{m\sigma}
394:      + U\sum_l n_{l\uparrow} n_{l\downarrow}\nonumber\\ 
395:      &&\  - \sum_{l}[2 J S^{z}_{l}s^{z}_{l} +
396:       2J' (S^{+}_{l}s^{-}_{l}+S^{-}_{l}s^{+}_{l})]\, . \label{eq:fkl}
397: \end{eqnarray}
398: In this model, the low energy degrees of freedom of the narrow band 
399: are represented by local moments $\vec{S}_{l}$. These interact 
400: ferromagnetically with the local spin density $\vec{s}_{l}$ of the 
401: wide-band conduction electrons which are also subject to a local 
402: Coulomb repulsion $U$. The model corresponds to a double-exchange 
403: model with Coulomb interactions amongst the itinerant electrons or, 
404: equivalently, to the ferromagnetic Kondo lattice model with interactions 
405: in the band. The antiferromagnetic case, $J<0$, for $U=0$, has already 
406: been investigated in the context of heavy fermions  \cite{costi.02}.
407: We adapt these calculations to the case of interest here, namely,
408: ferromagnetic exchange, $J>0$, and $U>0$. The equivalence between the 
409: full two-band model and the effective model holds provided $U$ is large
410: enough so that a description of the low energy degrees of freedom of 
411: the narrow band in terms of local moments is possible and provided
412: that $J' \ll U/2$. The latter condition corresponds to eliminated
413: excited states of the full model being far from the first excited 
414: state of the effective model. Unless $J$ is chosen to be very small, 
415: we see that for realistic values of $J\approx U/4$,
416: the equivalence will hold well for Ising exchange coupling ($J'=0$) 
417: but less well for isotropic Hund's coupling ($J'=J$). In either case,
418: the effective model should be increasingly accurate in the limit 
419: $\omega, T \rightarrow 0$.
420: 
421: %\begin{figure}[t]
422: %\centering 
423: %\includegraphics[height=3.5cm,width=7.5cm]{low-energy-states.eps}
424: %\caption{\label{low-energy-states}
425: %Low lying two-electron states of the two-band and effective one-band
426: %models in the atomic limit. The $n=2,S_{z}=0$ states at $J+J'$ above the
427: %groundstate correspond to $\frac{1}{\sqrt{2}}(|\uparrow\rangle|\downarrow\rangle-
428: %|\downarrow\rangle|\uparrow\rangle)$ in both models, whereas the higher $n=2,S_{z}=0$
429: %state in the two-band model (at $J+U/2$ above the groundstate) corresponds to 
430: %$|0\rangle|\uparrow\downarrow\rangle$ (or $|\uparrow\downarrow\rangle|0\rangle$).
431: %For finite $W_{1}$, the latter would correspond to excitations to the upper and lower 
432: %``Hubbard bands'' of the narrow band, and are eliminated in the 
433: %effective model. The equivalence between the models holds when the 
434: %neglect of these states is a good approximation,
435: %i.e. for $J' \ll \frac{U}{2}$. 
436: %}
437: %\end{figure}
438: 
439: Within DMFT, we need to solve an effective quantum impurity model 
440: corresponding to a $S=1/2$ Kondo impurity coupled ferromagnetically 
441: with conduction electrons subject to a local Coulomb repulsion:
442: \begin{eqnarray}
443:   H  & = & \sum_{k,\sigma}\varepsilon_{k}c^+_{k\sigma}c_{k\sigma}
444:   + U n_{0\uparrow}n_{0\downarrow}\nonumber\\ 
445:   &&\   - 2J S_0^{z}s^{z}_{0} -
446:       2J' (S_0^{+}s^{-}_{0}+S_0^{-}s^{+}_{0})\, . \label{eq:quantum-imp}
447: \end{eqnarray}
448: We solve this model using the numerical renormalization 
449: group method \cite{wilson.74}, which allows
450: calculation of dynamical quantities on the real energy axis at both zero
451: and finite temperature \cite{costi.94,hofstetter.00}.
452: The impurity self-energy
453: $\Sigma(\omega)$ is calculated using the method
454: described in Ref.~\cite{bulla.98}. For comparison with ED, it is then 
455: evaluated on the imaginary axis $z=i\omega$ by analytic continuation
456: \begin{eqnarray}
457:  \Sigma(i\omega) = -\frac{1}{\pi}\int_{-\infty}^{+\infty} d\omega'\ 
458:     \frac{{\rm Im}\,\Sigma(\omega')}{i\omega-\omega'}\, .
459: \label{eq:matsubara-se}
460: \end{eqnarray}
461: The calculations use a logarithmic discretization of the conduction band 
462: $\varepsilon_{k_n}\rightarrow \pm D_2 \Lambda^{-\frac{n-1}{2}}$ with
463: $\Lambda=1.5$ and we retain of order 600 states per NRG iteration. 
464: Details of the calculation of spectra and other dynamical quantities 
465: can be found in Refs.~\cite{costi.94,bulla.01}.
466: 
467: The effective model specified in Eq.~(\ref{eq:fkl}) 
468: allows the non-Fermi-liquid physics of the intermediate phase of the 
469: two-band model to be understood qualitatively \cite{biermann05}. 
470: The case of Ising exchange ($J'=0$) describes interacting 
471: conduction electrons scattering from a disordered static configuration 
472: of local spins with $S_{l}^{z}=\pm 1/2$. The disorder potential
473: is proportional to the Ising exchange $J$. The conduction electrons
474: therefore have a finite lifetime at the Fermi level even at $T=0$, 
475: as is characteristic of a disordered metal. Switching on the
476: spin-flip part of the Hund's exchange gives the disordered spin 
477: configuration some dynamics at finite temperature, but as long as 
478: $J' < J$, we expect from the ferromagnetic Kondo model that the 
479: spin-flip part of the Hund's exchange will renormalize to zero 
480: at low temperature with the Ising part of the Hund's exchange remaining
481: finite. At $T=0$, the system will again be disordered with a disorder 
482: potential given by the finite Ising part of the Hund's exchange. 
483: This results in a finite lifetime for the conduction electrons at $T=0$. 
484: 
485: The situation changes for isotropic Hund's exchange ($J'=J$). 
486: In this case, the ferromagnetic exchange in the effective impurity 
487: model (\ref{eq:quantum-imp}) is known to be marginally irrelevant 
488: \cite{cragg.79,koller.05}. 
489: Thus, both Ising and spin-flip parts of the Hund's exchange renormalize
490: to zero and there will be no disorder scattering at $T=0$, giving rise
491: to an infinite lifetime for the conduction electrons at $T=0$. The vanishing
492: of the self-energy at $\omega=T=0$ also implies that the narrow-band
493: spectral function satisfies a pinning condition at $\omega = T =0$.
494: In Sections 5 and 6, these qualitative considerations will be seen 
495: to be in accord with the numerical results from ED for the full 
496: two-band model and NRG results for the effective one-band model.
497: 
498: \section{4 \ Metallic phase}
499: 
500: We begin our discussion of the multi-band ED results with the region 
501: in which both subbands are metallic, just below the Mott transition 
502: associated with the narrow band. To illustrate the applicability of 
503: the ED approach this region is very useful since it involves 
504: important redistribution of spectral weight between low and high 
505: frequencies that must be captured properly in order to describe
506: the Mott transition. 
507: 
508: 
509: \begin{figure}[t!]%3
510:   \begin{center}
511: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{M8.ps}
512: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{M6.ps}
513:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig4a.ps}
514:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig4b.ps}
515:   \end{center}
516:   \vskip-2mm
517: \caption{\label{M8M6}
518: Imaginary part of subband self-energies in metallic region 
519: for $J=J'=U/4$ at $T=2.5$~meV. 
520: Solid (red) curves: narrow band, dashed (blue) curves: wide band. 
521: From top: $U=1.8,\ 1.9,\,2.0,\,2.1$~eV.   (a) $n_s=8$; (b) $n_s=6$.
522: }\end{figure}
523: 
524: Fig.~\ref{M8M6} shows a comparison of the subband self-energies for 
525: $n_s=8$ and $n_s=6$ in the region close to the lower critical Coulomb 
526: energy for $T=2.5$~meV. For $n_s=8$, the $\Sigma_i(i\omega_n)$ vary 
527: roughly linearly at low frequencies, indicating that both bands are 
528: metallic. Of course, the narrow band is more strongly correlated than 
529: the wide band. The spacing between the lowest excited state and the 
530: ground state energy in these calculations is typically $1-2$~meV, 
531: i.e., evaluation of the self-energies in this low temperature range 
532: is indeed meaningful.
533: 
534: The results for $n_s=6$ are similar, except that the self-energy of the 
535: narrow band at $U=2.1$~eV is inversely proportional to $i\omega_n$,
536: indicating that the Mott transition in this case is located between
537: $U=2.0$~eV and $U=2.1$~eV. In addition, at the lowest Matsubara 
538: frequencies the self-energy of the wide band at 2.1~eV exhibits  
539: deviations from linear $i\omega_n$ variation. These deviations 
540: are related to finite-size effects and will be analyzed in more 
541: detail in the next section when we discuss the intermediate 
542: orbital-selective phase. Just below the transition for $n_s=6$, at  
543: $U=2$~eV, very small deviations from approximately linear $i\omega_n$ 
544: variation can also be seen in both $\Sigma_i(i\omega_n)$. 
545: 
546: It is remarkable that the ED results for $n_s=8$ capture the correlation 
547: effects in the common metallic phase of both subbands very well until 
548: close to $U_{c1}$ and down to rather low temperatures. This result 
549: is not at all obvious since so close to the Mott transition a large 
550: fraction of the spectral weight of the narrow band is transfered from 
551: the Fermi level to the Hubbard bands. Even the results for $n_s=6$ are  
552: qualitatively correct. The main effect of the cluster size $n_s=6$ is 
553: the underestimate of $U_{c1}$ by about $0.1$~eV. According to the 
554: single-band results \cite{prl05} inclusion of the zero energy bath levels 
555: in the $n_s=8$ cluster should provide the most important part of the 
556: shift towards the correct $U_{c1}$. In addition these extra bath levels 
557: yield an excellent representation of the metallic properties of both 
558: subbands right up to the critical Coulomb energy. 
559: 
560: Of course, slight finite-size effects should be manifest also in the 
561: low-temperature ED results for $n_s=8$. The precise form of  
562: $\Sigma_i(i\omega_n)$ at very low frequencies and temperatures,
563: in particular, the range of true Fermi-liquid behavior in the immediate 
564: vicinity of $U_{c1}$, can only be investigated by more accurate methods, 
565: such as a full two-band extension of the NRG approach which is applicable 
566: at zero and finite $T$ \cite{costi}. 
567: 
568: We emphasize that this region of the $T/U$ phase diagram for isotropic 
569: Hund's exchange is not yet accessible using QMC calculations. 
570: It would therefore be of great interest to extend the present results 
571: to $n_s=9$ and $n_s=12$ in order to explore the strongly correlated 
572: metallic phase of realistic three-band materials which have so far been 
573: investigated only for Ising-like exchange interactions. A more detailed 
574: comparison of the role of Hund vs. Ising exchange treatments in the 
575: metallic phase of the present two-band model will be given elsewhere
576: \cite{metallic}.     
577: 
578: \section{5 \ Intermediate phase: isotropic Hund's exchange}
579: 
580: As shown first by Koga {\it et al.}~\cite{koga1} within ED/DMFT 
581: calculations at $T=0$, the two-band Hubbard model with full Hund's exchange 
582: interaction exhibits successive, orbital-selective Mott transitions. 
583: In Ref.~\cite{prl05} we proved that analogous ED calculations at 
584: finite temperature are consistent with these findings, revealing 
585: sequential first-order transitions associated with the two subbands. 
586: In this section we discuss in more detail the low temperature properties 
587: of the intermediate phase between the fully metallic and insulating phases.
588: 
589: \begin{figure}[t!]%4
590:   \begin{center}
591: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{OSMTa.ps}
592: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{OSMTb.ps}
593:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig5a.ps}
594:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig5b.ps}
595:   \end{center}
596:   \vskip-2mm
597: \caption{\label{OSMT}
598: Imaginary part of subband 
599: self-energies in intermediate phase, for $U=2.4$~eV and $J=J'=U/4$.  
600: $T=2.5,\ 5,\ 10,\ 20,\ 30$~meV (from bottom). 
601: (a) $n_s=8$; (b) $n_s=6$; 
602: }\end{figure}
603: 
604: Fig.~\ref{OSMT} shows the subband self-energies for a Coulomb energy 
605: $U=2.4$~eV, i.e., between the lower and upper Mott transitions. 
606: The results 
607: for $n_s=8$ demonstrate that the narrow band is insulating, i.e., its
608: self-energy is inversely proportional to $i\omega_n$ at low frequencies. 
609: $\Sigma_1(i\omega_n)$ is also nearly independent of temperature in the 
610: range $T=2.5\ldots 30$~meV, suggesting that the Mott gap is considerably 
611: larger. The self-energy of the wide band, however, extrapolates to zero 
612: at low frequencies, indicating metallic behavior. These results 
613: suggest that the pinning condition $N_2(0)=A_2(0)$ is satisfied, i.e., 
614: the spectral density at $E_F$ of the interacting system coincides 
615: with that of the non-interacting system. Because of the rapid increase 
616: of Im\,$\Sigma_2(i\omega_n)$ at small $\omega_n$, the spectral 
617: distribution of the wide band in the intermediate phase consists of 
618: a very narrow peak at $E_F$ and large upper and lower Hubbard bands.
619: Nevertheless, as will be discussed further below, the comparison 
620: with the NRG results indicates that this band does not satisfy 
621: true Fermi-liquid behavior.
622: 
623: The results for $n_s=6$ shown in Fig.~\ref{OSMT}(b) agree qualitatively 
624: with those for $n_s=8$. The self-energies of the narrow band are nearly 
625: identical, with only a slightly larger dependence on temperature in 
626: the case $n_s=6$. The main difference is that $\Sigma_2(i\omega_n)$ 
627: at low $T$ and small $i\omega_n$ exhibits larger deviations as a result 
628: of finite-size effects. This is plausible since, as discussed above, 
629: the wide band is metallic, but highly correlated. The pronounced 
630: three-peak structure of its spectral function is therefore not 
631: represented accurately by only two non-zero bath levels per 
632: impurity orbital in the ED calculation for $n_s=6$. The extra bath
633: level with zero energy included for $n_s=8$ therefore plays a crucial 
634: role for a proper description of the metallic behavior of the wide
635: band for full Hund's coupling.   
636: 
637: \begin{figure}[t!]%5
638:   \begin{center}
639: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{GnuGJ.ps}
640:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig6.ps}
641:   \end{center}
642:   \vskip-2mm
643: \caption{\label{GreenJ}
644: Imaginary part of subband 
645: Green's functions in intermediate phase, for $U=2.4$~eV, $J=J'=U/4$,
646: $n_s=8$. $T=2.5,\ 5,\ 10,\ 20,\ 30$~meV. 
647: The dot indicates the pinning condition for the wide band.
648: }\end{figure}
649: 
650: Fig.~\ref{GreenJ} shows the subband Greens functions $G_i(i\omega_n)$ 
651: in the intermediate phase at various temperatures. Im\,$G_1$ is linear in 
652: $\omega_n$ at low frequencies, as expected for the insulating subband. 
653: Although the wide band exhibits clear signs of finite-size effects at 
654: small $\omega_n$ it is nevertheless seen to approximately satisfy the 
655: pinning condition Im\,$G_2(i\omega_n)\rightarrow -\pi N_2(0) = -1$ in 
656: the limit $\omega_n\rightarrow0$.   
657: 
658: For computational reasons the role of finite-size effects in the 
659: present finite-temperature ED approach based on full matrix 
660: diagonalization can at present not be checked by going beyond the 
661: cluster size $n_s=8$. To analyze the low-frequency properties of 
662: the wide band in the intermediate phase more closely, we have 
663: carried out NRG calculations for the effective one-band model  
664: discussed in Section 3. Although this model is more appropriate
665: for the anisotropic exchange treatment discussed in the next section,
666: we use it here (with some caution) since it is so far the only scheme 
667: capable of providing a guideline for the two-band Hubbard model 
668: at low finite temperatures.
669: 
670: \begin{figure}[t!]%6
671:   \begin{center}
672: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NRG3.ps}
673: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NRG4.ps}
674:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig7a.ps}
675:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig7b.ps}
676:   \end{center}
677:   \vskip-2mm
678: \caption{\label{NRG34}
679: Imaginary part of self-energy $\Sigma_2$ of wide band for $U=2.4$~eV 
680: and $J=J'=U/4$. (a) Solid (red) and dashed (blue) curves: ED results 
681: for $n_s=8$ at  $T=2.5,\,5,\,10,\,20,\,30$~meV (from bottom); 
682: green curves: NRG results for $T=3,\ 10,\ 15,\ 34$~meV (from top). 
683: (b) Same quantities on expanded low-frequency scale. Black curve (+):
684: self-energy $\Sigma_1$  of narrow band at $2.0$~eV. 
685: }\end{figure}
686: 
687: In Fig.~\ref{NRG34}(a) the ED results for Im\,$\Sigma_2(i\omega_n)$ are 
688: compared with the corresponding self-energy derived within the NRG. 
689: The overall frequency variation of $\Sigma_2$ is seen to be remarkably 
690: similar for both methods, implying similar spectral distributions.
691: The most noticeable difference is the larger dependence on temperature 
692: in the case of the ED results.
693: According to Fig.~\ref{OSMT}, the increase in cluster size from $n_s=6$ 
694: to $n_s=8$ diminishes the variation of $\Sigma_2$ with temperature and 
695: makes the minimum near $\omega_n\approx0.9$ less deep. This trend 
696: indicates that larger cluster sizes would bring the ED results
697: into better agreement with the NRG data. We point out, however, 
698: that part of the difference should be caused by the approximate 
699: nature of the NRG model in the case $J'=J$. 
700: As shown in the following section, for $J'=0$ the ED and NRG results 
701: for $\Sigma_2$ at $T=20\ldots30$~meV are in perfect agreement.  
702: 
703: The low-frequency region of $\Sigma_2(i\omega_n)$ is plotted in 
704: greater detail in Fig.~\ref{NRG34}(b). In the ED calculation, the 
705: spacing between the lowest excited level and the ground state energy 
706: is typically $10$~meV. Thus, temperatures below this range will give
707: rise to increasing finite-size effects, precluding any meaningful 
708: analysis at low frequencies. Despite this limitation, the ED data 
709: follow the frequency dependence of the NRG results rather well. 
710: 
711: In order to provide some insight into the variation of $\Sigma_2$ 
712: at finite $\omega_n$, we compare it in Fig.~\ref{NRG34}(b) also 
713: with the self-energy of the narrow band  $\Sigma_1$
714: at $U=2.0$~eV, i.e., 
715: just below the Mott transition at $U_{c1}\approx2.1$~eV. 
716: According to Fig.~1(a), we find: 
717: $Z_1(U=2.0~{\rm eV})\approx Z_2(U=2.4~{\rm eV})\approx 0.11$. 
718: Nevertheless, whereas $\Sigma_1$ at $U=2.0$~eV exhibits the typical 
719: behavior expected for a strongly correlated Fermi-liquid phase, 
720: $\Sigma_2$ evidently has much larger nonlinear corrections. A similar 
721: systematic difference is found between $\Sigma_2$ in the 
722: intermediate phase and the behavior the self-energy in a 
723: single-band model for $U$ just below the Mott transition.
724: 
725: The NRG results shown in Fig.~\ref{NRG34} indicate that   
726: $\Sigma_2(i\omega_n)\rightarrow0$ in the limit $\omega_n\rightarrow0$ 
727: and for low $T$. Thus, particles at $E_F$ have infinite lifetime. 
728: In addition, analogous NRG calculations at $T=0$ yield \cite{costi}  
729: %Im\,$\Sigma_2(\omega)\sim -1/({\rm ln}\omega)^2$, in contrast to 
730: non-Fermi-liquid behavior, in agreement with Ref.~\cite{biermann05}.
731: Thus, at small finite $i\omega_n$ Im\,$\Sigma_2$ does not vary 
732: linearly. The qualitative agreement between the finite-$T$ NRG and 
733: ED results seen in Fig.~\ref{NRG34} shows that the 
734: ED data are consistent with non-Fermi-liquid behavior.  
735: 
736: \begin{figure}[t!]%7
737:   \begin{center}
738: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NRG1.ps}
739:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig8.ps}
740:   \end{center}
741:   \vskip-2mm
742: \caption{\label{NRG12}
743: Spectral distribution of wide band in intermediate phase,
744: calculated within NRG approach for $U=2.4$~eV, $J=J'=U/4$. 
745: Green curve: $T=3$~meV; red curve: $T=30$~meV;  
746: dashed blue curve: bare density of states.
747: }\end{figure}
748:     
749: Fig.~\ref{NRG12} shows the spectral function of the wide band in the
750: intermediate phase, calculated within the NRG. Spectra of this kind 
751: were also obtained by Arita and Held \cite{arita} for $U$ near $U_{c1}$ 
752: within QMC/DMFT calculations for the same two-band model for $T=0$ and
753: $J'=J=U/4$.
754: As mentioned above, the distribution exhibits a very narrow peak which 
755: satisfies the pinning condition at $E_F$ for temperatures up to about 
756: 10~meV. 
757: % This finding is consistent with the ED/DMFT results reported in 
758: % Ref.~\cite{prl05}, which revealed 
759: % $T_{c1}\approx 20$~meV and $T_{c2}\approx 15$~meV.    
760: Because of the sharp central peak of $A_2(\omega)$, the real and 
761: imaginary parts of the self-energy of the wide band $\Sigma_2(\omega)$ 
762: exhibit structure on a similar scale. Nevertheless, the variation 
763: of Im\,$\Sigma_2(i\omega_n)$ obtained in the NRG is very smooth. 
764: Thus, the weak shoulder seen in the ED results for $\Sigma_2$ near 
765: $\omega_n\approx0.06$ in Figs.~\ref{W1W2}(a), \ref{OSMT}(a) and 
766: \ref{NRG34}  must be attributed to finite-size effects.   
767: 
768: The ED and NRG results discussed above demonstrate that, in the 
769: presence of full Hund's exchange, the wide band above the main 
770: Mott transition at $U_{c1}$ remains metallic, without satisfying
771: Fermi-liquid criteria. In the following section it will 
772: be shown that the absence of spin-flip and pair-exchange terms
773: enhances this trend towards non-Fermi-liquid behavior, so that 
774: even particles at $E_F$ acquire a finite lifetime. Thus, in both
775: cases, the intermediate phase is bad-metallic.  
776: 
777: We point out that the differences between the self-energies for 
778: $n_s=8$ and $n_s=6$ appear mainly at rather low temperatures, below
779: about $T=20$~meV. Since QMC calculations also become difficult to 
780: converge in this range, three-band ED calculations with 
781: two bath levels per impurity orbital, i.e., cluster size $n_s=9$, 
782: should be competitive with analogous QMC calculations -- with the 
783: important additional benefit, that the ED approach can handle 
784: full Hund's coupling.          
785: 
786: \section{6 \ Intermediate phase: Ising exchange}
787: 
788: As shown in Ref.~\cite{prb70}, in the absence of spin-flip and 
789: pair-exchange contributions, the first-order Mott transition at 
790: $U_{c1}\approx 2.1$~eV affects the different subbands of the 
791: Hubbard model in fundamentally different ways: The narrow band 
792: undergoes a complete metal insulator transition, but the wide
793: band changes from a normal metal to a bad metal in the sense
794: that its self-energy exhibits progressive deviations from 
795: Fermi-liquid linear $\omega_n$ variation at low frequencies.
796: Thus, Im\,$\Sigma_2(i\omega_n)\rightarrow c(U)$ for     
797: $\omega_n\rightarrow 0$, where  $c(U)\approx 0\rightarrow-\infty$ 
798: for $U = U_{c1}\rightarrow U_{c2}$. Spectral functions obtained via 
799: the maximum entropy method showed that this breakdown of Fermi-liquid
800: behavior in the wide band in the intermediate phase leads to a narrow 
801: pseudogap near $E_F$ and to a violation of the pinning condition, 
802: i.e., $A_2(0)<N_2(0)$. The spectral function of the wide band then 
803: acquires a characteristic four-peak structure, with two maxima flanking 
804: the pseudogap, in addition to the Hubbard bands at higher energies.  
805: With increasing $U$, the pseudogap becomes deeper and wider, until 
806: it turns into a true insulating gap at $U_{c2}\approx 2.7$~eV.
807: As also shown in Ref.~\cite{prb70}, the upper bad-metal to insulator
808: transition is not a first-order transition, in contrast to the main
809: transition at $U_{c1}$.   
810: 
811: 
812: In this section we discuss the temperature dependence of the subband
813: self-energies in the intermediate phase and the finite-size effects 
814: associated with the limited number of bath levels included in our 
815: ED/DMFT approach. 
816: 
817: 
818: \begin{figure}[t!]%8
819:   \begin{center}
820: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NRG3J0.ps}
821: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NRG4J0.ps}
822: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NRG5J0.ps}
823:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig9a.ps}
824:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig9b.ps}
825:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig9c.ps}
826:   \end{center}
827:   \vskip-2mm
828: \caption{\label{NRG345J0}
829: Imaginary part of subband self-energies $\Sigma_i(i\omega_n)$ 
830: in intermediate phase for $n_s=8$, $U=2.4$~eV, $J=U/4$, $J'=0$;
831: Solid (red) and dashed (blue) curves:  ED results for    
832: $T=2.5,\ 5,\ 10,\ 20,\ 30$~meV (from bottom). 
833: (a)  $\omega_n\Sigma_i(i\omega_n)$;
834: (b)  $\Sigma_2(i\omega_n)$; green curves: NRG results as in (c).
835: (c) Low-frequency behavior of self-energy of wide band: 
836: red and blue curves: ED results as in (b); green curves: NRG results 
837: for $T=2,\,4.4,\,10,\,20,\,30$~meV (from bottom). 
838: For clarity, successive curves in (c) are displaced vertically by 1.
839: }\end{figure}
840: 
841: Fig.~\ref{NRG345J0}
842: shows the subband self-energies for $n_s=8$ and $U=2.4$~eV,
843: assuming again $J=U/4$, but $J'=0$. The narrow band is fully insulating, 
844: so that $\omega_n\,{\rm Im}\,\Sigma_1(i\omega_n)\rightarrow {\rm const.}$ 
845: nearly independently of temperature, similarly to the case $J'=J$ plotted 
846: in Fig.~\ref{OSMT}. The self-energy of the wide band, however, differs 
847: qualitatively from the behavior found for isotropic Hund's coupling. 
848: Rather than vanishing in the limit $\omega_n\rightarrow0$, 
849: Im\,$\Sigma_2(i\omega_n)$ now approaches a sharp minimum, which gets 
850: more pronounced towards low temperatures. Clearly, since    
851: $\omega_n\,{\rm Im}\,\Sigma_2(i\omega_n)\rightarrow0$ at small 
852: $\omega_n$, this band is not yet insulating.  
853: Thus, instead of satisfying the pinning condition at $E_F$, the 
854: spectral function of the wide band exhibits a dip or pseudogap at
855: the Fermi level (see below).
856: 
857: Fig.~\ref{NRG345J0}(b) also shows the NRG results for 
858: $\Sigma_2(i\omega_n)$. 
859: They are seen to be in very good agreement with the ED data, 
860: except at low frequencies for $T\le10$~meV. 
861: The fact that for $T=20\ldots30$~meV there is now much better
862: coincidence between the ED and NRG results than in Fig.~\ref{NRG34} 
863: for $J'=J$ indicates that, as discussed in Section 3, the 
864: effective one-band model employed 
865: for the NRG is more appropriate in the case $J'=0$. The low-frequency 
866: behavior of Im\,$\Sigma_2(i\omega_n)$ is shown in more detail 
867: in Fig.~\ref{NRG345J0}(c). Both the ED and NRG results reveal that the 
868: sharp minimum of $\Sigma_2$ at small $\omega_n$ gets progressively 
869: deeper at low temperatures, suggesting that the pseudogap in the spectral
870: function becomes accordingly deeper. 
871: 
872: 
873: \begin{figure}[t!]%9
874:   \begin{center}
875: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NFL6a.ps}
876: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NFL6b.ps}
877:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig10a.ps}
878:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig10b.ps}
879:   \end{center}
880:   \vskip-2mm
881: \caption{\label{NFL6}
882: Imaginary part of subband self-energies $\Sigma_i(i\omega_n)$ 
883: in intermediate phase for $n_s=6$, $U=2.2$~eV and $J=U/4$, $J'=0$.   
884: $T=2.5,\ 5,\ 10,\ 20,\ 30$~meV (from bottom).  
885: }\end{figure}
886: 
887: Analogous results for $n_s=6$ are given in Fig.~\ref{NFL6}. Since 
888: according to Ref.~\cite{prl05} in this case the wide band becomes 
889: insulating near $U\approx2.4$~eV, we choose $U=2.2$~eV to illustrate 
890: the non-Fermi-liquid behavior of this band in the intermediate phase. 
891: As for $n_s=8$, the narrow band is insulating, i.e., 
892: $\omega_n\,{\rm Im}\,\Sigma_1(i\omega_n)\rightarrow{\rm const.}$ 
893: at low frequencies. The frequency variation of $\Sigma_2(i\omega_n)$ 
894: is seen to be more strongly affected by finite-size effects. 
895: Nevertheless, as in the case $n_s=8$, a sharp minimum is found 
896: for $\omega_n\rightarrow0$.
897: 
898: \begin{figure}[t!]%10
899:   \begin{center}
900: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{GnuGJ0.ps}
901:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig11.ps}
902:   \end{center}
903:   \vskip-2mm
904: \caption{\label{GreenJ0}
905: Imaginary part of subband 
906: Green's functions in intermediate phase, for $U=2.4$~eV, $J=U/4$, 
907: $J'=0$, $n_s=8$. $T=2.5,\ 5,\ 10,\ 20,\ 30$~meV (from top). 
908: }\end{figure}
909: 
910: Fig.~\ref{GreenJ0} shows the subband Green's functions $G_i(i\omega_n)$ 
911: for anisotropic Hund's exchange at various temperatures. 
912: As for the case $J'=J$ shown in Fig.~\ref{GreenJ}, Im\,$G_1$ is linear 
913: in $\omega_n$ at low frequencies, since the narrow band is insulating.
914: On the other hand, because of the more severe breakdown of Fermi-liquid 
915: behavior in the wide band for $J'=0$, with Im\,$\Sigma_2(i\omega_n)\ne 0$ 
916: in the low-frequency limit, the Green's function no longer satisfies the 
917: pinning condition. Thus, Im\,$G_2(i\omega_n)\rightarrow -c(T)$ with 
918: $c(T)<\pi N_2(0) = 1$, and $c(T)\rightarrow0$ for decreasing temperature.
919: This behavior implies that the spectral function of the wide band exhibits
920: a pseudogap at $E_F$ which becomes progressively deeper towards low $T$.  
921: 
922: This picture is fully confirmed by the NRG results shown in 
923: Fig.~\ref{NRG12J0}. In contrast to the narrow peak at $E_F$ in the
924: three-peak structure seen in Fig.~\ref{NRG12} for isotropic Hund's 
925: coupling, the spectra for $J'=0$ show a pseudogap which becomes deeper 
926: as $T$ decreases. As a result, the spectral distribution now exhibits a 
927: characteristic four-peak structure, with two maxima limiting the pseudogap 
928: and two shoulders associated with the Hubbard peaks. Spectra of this kind 
929: were first observed in the QMC/DMFT results at $T=31$~meV reported in 
930: Ref.~\cite{prb70} (see also next section). The comparison with the
931: spectra for $J'=J$ reveals nearly identical excitations for 
932: $\vert\omega\vert\gtrsim J$. Thus, as expected, the different treatments 
933: of exchange interactions affect primarily the low-frequency excitations 
934: in the metallic wide band.    
935: 
936: 
937: 
938: \begin{figure}[t!]%11
939:   \begin{center}
940: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{NRG1J0.ps}
941:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig12.ps}
942:   \end{center}
943:   \vskip-2mm
944: \caption{\label{NRG12J0}
945: Spectral distribution of wide band in intermediate phase, calculated 
946: within NRG approach for $U=2.4$~eV, $J=U/4$, $J'=0$.
947: Green curve: $T=3$~meV; red curve: $T=34$~meV;  
948: dashed blue curve: bare density of states. 
949: Black dotted curve: spectrum for $J'=J$, $T=3$~meV from 
950: Fig.~\ref{NRG12}. 
951: }\end{figure}
952: 
953: 
954: 
955: \section{7 \ Comparison with previous QMC/DMFT results}
956: 
957: In this and the following sections we compare the ED and NRG results 
958: for $J'=0$ with available QMC/DMFT data in order to illustrate 
959: the consistency between these impurity treatments and to explore 
960: further the role of finite-size effects in the ED approach.
961: 
962: \begin{figure}[t!]%12
963:   \begin{center}
964: %  \includegraphics[width=5.5cm,height=8cm,angle=-90]{QMC.ED.ps}
965: %  \includegraphics[width=5.5cm,height=8cm,angle=-90]{QMCnew.ps}
966: %  \includegraphics[width=5.5cm,height=8cm,angle=-90]{QMC.NRG.ps}
967:   \includegraphics[width=5.9cm,height=8cm,angle=-90]{fig13a.ps}
968:   \includegraphics[width=5.9cm,height=8cm,angle=-90]{fig13b.ps}
969:   \includegraphics[width=5.9cm,height=8cm,angle=-90]{fig13c.ps}
970:   \end{center}
971:   \vskip-2mm
972: \caption{\label{QMCprb}
973: Comparison of subband self-energies $\Sigma_i(i\omega_n)$ in intermediate 
974: non-Fermi-liquid phase for $U=2.1,\ 2.4,\ 2.7$~eV, $J=U/4$, $J'=0$,
975: $T=31$~meV, calculated via three different impurity solvers:
976: (a) ED results for $n_s=8$: red solid curves: narrow band; 
977: blue dashed curves: wide band; 
978: (b) QMC results from Fig.~10 of Ref.~\cite{prb70};
979: red solid curves: narrow band; blue dashed curves: wide band;  
980: (c) comparison of ED (x) and NRG self-energies of wide band.
981: }\end{figure}
982: 
983: As shown in Ref.~\cite{prb70}, in the absence of spin-flip and
984: pair-exchange terms, purely metallic and insulating phases exist
985: for $U<U_{c1}\approx2.1$~eV and  $U>U_{c2}\approx2.7$~eV, 
986: respectively. Fig.~\ref{QMCprb} shows a comparison of subband 
987: self-energies at three representative Coulomb energies within 
988: the intermediate `bad-metal' non-Fermi-liquid region. The ED 
989: results for $n_s=8$ are seen to be in excellent agreement with 
990: the NRG self-energies, confirming the validity of the effective 
991: one-band model in the intermediate phase for $J'=0$. Moreover, 
992: both schemes agree very well with the QMC/DMFT self-energies 
993: reported in Ref.~\cite{prb70} for $T=31$~meV. Minor differences 
994: between the QMC and ED/NRG results are found only in the steepest 
995: parts of $\Sigma_i$ close to the first Matsubara frequency. 
996: (In contrast to the QMC calculations which are carried out at 
997: discrete Matsubara frequencies, the NRG self-energy is available 
998: continuously as a function of frequency. The ED self-energies 
999: could in principle also be obtained at arbitrary $i\omega$ but 
1000: were calculated here at $i\omega_n$. This explains the slightly 
1001: different form of the curves plotted in panel (c). At 
1002: $i\omega_n$ the ED and NRG data nearly coincide.)  
1003: 
1004: \begin{figure}[t!]%13
1005:   \begin{center}
1006: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{QMCa.ps}
1007: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{QMCb.ps}
1008:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig14a.ps}
1009:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig14b.ps}
1010:   \end{center}
1011:   \vskip-2mm
1012: \caption{\label{QMCab}
1013: Subband self-energies $\Sigma_i(i\omega_n)$ in intermediate 
1014: non-Fermi-liquid phase for $T=31$~meV, $U=2.1,\ 2.4,\ 2.7$~eV,
1015: $J=U/4$, $J'=0$. Solid (red) curves: $n_s=8$; dashed (blue) curves:   
1016: $n_s=6$. 
1017: (a) narrow subband: $\omega_n\Sigma_1(i\omega_n)$; 
1018: (b) wide   subband: $\Sigma_2(i\omega_n)$.  
1019: }\end{figure}
1020: 
1021: The overall 
1022: trend obtained previously within the QMC/DMFT is fully confirmed 
1023: by the new ED and NRG calculations: The narrow band is insulating 
1024: throughout this range of Coulomb energies, whereas the wide band 
1025: changes gradually from metallic to insulating via progressive 
1026: non-Fermi-liquid behavior. Thus, for small $\omega_n$,   
1027: Im\,$\Sigma_2(i\omega_n)\rightarrow c(U)$, 
1028: where $c(U)\rightarrow-\infty$ at $U\approx2.7$~eV.   
1029: 
1030: Evidently, despite their intrinsic numerical uncertainties,
1031: the three complementary impurity solvers: ED, NRG and QMC provide 
1032: perfectly consistent descriptions of the electronic properties of 
1033: the orbital-selective phase for Ising-like exchange. 
1034: 
1035: Since the temperature $T=31$~meV in these results is relatively 
1036: high, finite-size effects tend to be less pronounced than in the cases 
1037: discussed in the previous sections at lower $T$. This is illustrated 
1038: in Fig.~\ref{QMCab} where we compare the above ED results for $n_s=8$ 
1039: with those for $n_s=6$. The self-energies of the narrow band are 
1040: nearly identical and satisfy 
1041: $\omega_n\Sigma_1(i\omega_n)\rightarrow {\rm const.}$ at low 
1042: frequencies. This is plausible since the additional zero-energy 
1043: bath level for $n_s=8$ carries very little weight in the insulating 
1044: state. Because of breakdown of Fermi-liquid behavior, the self-energy 
1045: of the wide band approaches a finite value in the limit    
1046: $\omega_n\rightarrow0$. The differences between the results for 
1047: $n_s=6$ and $n_s=8$ are very small as long as this band is either 
1048: nearly metallic, like at $U=2.1$~eV, or nearly insulating, like at 
1049: $U=2.7$~eV. Slightly larger differences are found only in the middle 
1050: of the bad-metallic region near $U=2.4$~eV. As discussed in the 
1051: preceding section, the spectral function then has a more complicated
1052: four-peak structure as a result of the narrow pseudogap at $E_F$,
1053: which cannot be adequately represented using only two bath levels. 
1054: This region could possibly be even more accurately described by using 
1055: four bath levels for the wide band and two levels for the insulating 
1056: narrow band, maintaining the total $n_s=8$.      
1057: 
1058:  \begin{figure}[t!]%14
1059:   \begin{center}
1060: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{GnuQMCNRG.ps}
1061:   \includegraphics[width=6.1cm,height=8cm,angle=-90]{fig15.ps}
1062:   \end{center}
1063:   \vskip-2mm
1064: \caption{\label{QMCNRG}
1065: Spectral distributions of subbands in intermediate phase, calculated 
1066: within QMC/DMFT for $U=2.4$~eV, $J=U/4$, $J'=0$, $T=31$~meV.
1067: Red curve: metallic wide band; magenta curve: insulating narrow band;
1068: blue and black curves: bare densities of states. From Fig.~11(b) 
1069: of Ref.~\cite{prb70}.
1070: }\end{figure}
1071:          
1072: We close this section by showing in Fig.~\ref{QMCNRG} the spectral 
1073: distributions of both subbands, as calculated within the QMC/DMFT and 
1074: the maximum entropy method \cite{prb70}. The spectrum of the bad-metallic 
1075: wide band can be compared with the corresponding NRG spectra plotted in 
1076: Fig.~\ref{NRG12J0}. Both distributions exhibit the marked four-peak
1077: structure induced by the pseudogap and the Hubbard bands. The 
1078: low-frequency region is seen to be in excellent agreement. Both methods
1079: coincide in that, at $T\approx30$~meV, the interacting density of 
1080: states at $E_F$ of the wide band is only about one third of the 
1081: noninteracting one. The slightly larger differences at higher energies,
1082: in particular, the position and relative weight of the Hubbard bands, 
1083: may be caused by the choice of fitting parameters in the maximum entropy 
1084: procedure, and by the less accurate nature of the NRG approach at high 
1085: energies. 
1086: %Since the ED and QMC self-energies shown in Fig.~\ref{QMCprb}
1087: %are in excellent agreement, the corresponding ED results for 
1088: %$G_i(i\omega_n)$ could also be used, via $G_i(\tau)$ and the maximum
1089: %entropy method, to derive spectral distributions similar to those 
1090: %shown in Fig.~\ref{QMCNRG}.   
1091: 
1092: 
1093: \section{8 \ Comparison with QMC/DMFT results for $\bf W_2=10W_1$}
1094: 
1095: The breakdown of Fermi-liquid behavior in the orbital-selective phase
1096: with Ising-like exchange was recently also studied by Biermann 
1097: {\it et al.}~\cite{biermann05} for the same two-band Hubbard model 
1098: as above, except for $W_1=0.2$~eV and $W_2=2$~eV. 
1099: The subbands were also assumed to be non-hybridizing and half-filled. 
1100: QMC/DMFT calculations were carried out for $T=1/120\ldots1/40$~eV, 
1101: $U=0.8$~eV, $J=0$ and $J=U/4$, with $J'=0$, i.e., in the absence of 
1102: spin-flip and pair exchange terms. For $J=0$ as well as $J=U/4$, the 
1103: self-energy of the narrow band was found to diverge at low frequencies, 
1104: demonstrating that this band is insulating since $W_1\ll U$. The wide 
1105: band is in the intermediate orbital-selective region between the purely 
1106: metallic and insulating phases since $U\ll W_2$. The self-energy of this 
1107: band showed a striking variation with the magnitude of $J$: 
1108: For $J=0$, $\Sigma_2(i\omega_n)\sim i\omega_n$ at low frequencies, as 
1109: expected for Fermi-liquid behavior. For  $J=U/4$, however, 
1110: Im\,$\Sigma_2(i\omega_n)\rightarrow{\rm const.}\approx -0.09$ nearly 
1111: independently of temperature for $T=1/120\ldots1/40$~eV, indicating 
1112: breakdown of Fermi-liquid properties. Accordingly, the pinning condition 
1113: $N_2(0)=A_2(0)$ was found to be satisfied for $J=0$, but not for $J=U/4$. 
1114: These results are fully consistent with the trend discussed in 
1115: Ref.~\cite{prb70} for $W_1=2$~eV and $W_2=4$~eV.   
1116: 
1117: To check the accuracy of our finite-$T$ ED approach we have applied 
1118: it to the case investigated in Ref.~\cite{biermann05}. To provide a 
1119: picture of the Mott transitions in this two-band system we show first 
1120: in Fig.~\ref{Paris.zu} the variation of $Z_i(U)$ for three different 
1121: treatments of Hund's exchange. In all cases the narrow band 
1122: becomes insulating at about $U_{c1}=0.3\ldots0.4$~eV. 
1123: However, the range and nature of the orbital-selective phase of the 
1124: wide band, $U_{c1} < U < U_{c2}$, depend sensitively on the magnitudes 
1125: of $J/U$ and $J'/U$. For $J=J'=0$ the upper transition occurs for 
1126: $U_{c2}$ slightly larger than $W_2$, with a pronounced hysteresis 
1127: loop indicative of first-order behavior. A very weak hysteresis 
1128: is found also for $J=J'=U/4$, with $U_{c2}\approx 1.5$~eV, similar 
1129: to the one in Fig.~\ref{ZU} near $U\approx3$~eV. Finally, for $J=U/4$ and 
1130: $J'=0$, $U_{c2}\approx 1.2$~eV without evidence of first-order behavior.
1131: (The latter result is consistent with $U_{c2}\approx 2.4$~eV in Fig.~2
1132: of Ref.~\cite{prl05}.)  Although $Z_2(U)\approx 0.2\ldots0.8$ at 
1133: $U=0.8$~eV, i.e., for $U_{c1}\ll U \ll U_{c2}$, the analysis of the 
1134: self-energy reveals fundamentally different electronic properties 
1135: of the wide band in this region.     
1136: 
1137: \begin{figure}[t!]%15  see Gnuz in  ED8/Jump  and   ED6/T=0.010
1138:   \begin{center}
1139: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{Pariszu.ps}
1140:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig16.ps}
1141:   \end{center}
1142:   \vskip-2mm
1143: \caption{\label{Paris.zu}
1144: $Z_i(U)$ as a function of $U$ for $W_1=0.2$~eV, $W_2=2.0$~eV, 
1145: $T=1/120$~eV, calculated within ED/DMFT, $n_s=6$, for 
1146: $J=J'=0$,  $J=J'=U/4$, and $J=U/4$, $J'=0$.    
1147: Red and blue curves: wide band, black curves: narrow band.
1148: The $Z_1$ for  $J=J'=U/4$ and $J=U/4$, $J'=0$ are indistinguishable. 
1149: The dots mark the three cases shown in Fig.~\ref{Paris}.
1150: }\end{figure}
1151:    
1152: This is illustrated in Fig.~\ref{Paris} which shows 
1153: Im\,$\Sigma_2(i\omega_n)$ for $n_s=8$ and $n_s=6$. The ED results for 
1154: $n_s=8$ are in excellent agreement with the QMC data \cite{biermann05}
1155: for $J=J'=0$ as well as $J=U/4$, $J'=0$, with the exception of small 
1156: finite-size effects at the lowest Matsubara frequencies and lowest 
1157: temperature. (Since $W_2=2$~eV, $T=1/120\ldots1/40$~eV in these 
1158: calculations corresponds to about $T=16\ldots50$~meV in the case 
1159: $W_2=4$~eV considered in the preceding sections.)  Also, the 
1160: variation with $T$ is in our results slightly more pronounced 
1161: than in the case of the QMC/DMFT.
1162: \begin{figure}[t!]%16
1163:   \begin{center}
1164: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{Parisa.ps}
1165: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{Parisb.ps}
1166: %  \includegraphics[width=5.0cm,height=8cm,angle=-90]{Parisc.ps}
1167:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig17a.ps}
1168:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig17b.ps}
1169:   \includegraphics[width=5.0cm,height=8cm,angle=-90]{fig17c.ps}
1170:   \end{center}
1171:   \vskip-2mm
1172: \caption{\label{Paris}
1173: (a) Imaginary part of self-energy of wide band 
1174: for $J=0$ and $J=U/4$ in intermediate 
1175: phase for $U=0.8$~eV, $J'=0$, calculated within ED/DMFT with $n_s=8$. 
1176: $W_1=0.2$~eV, $W_2=2.0$~eV.  Red solid, blue dashed, and black dotted 
1177: curves: $T=1/120$~eV, $T=1/80$~eV, and $T=1/40$~eV, respectively.
1178: The dots denote the results for $T=1/120$~eV obtained in 
1179: Ref.~\cite{biermann05}. 
1180: (b) Same as (a) except for $n_s=6$.  
1181: (c) Same as (a) except for $J'=J=U/4$.
1182: }\end{figure}
1183: 
1184: For $n_s=6$ the ED results exhibit larger finite-size effects, as shown 
1185: in Fig.~\ref{Paris}(b). The variation with temperature is in this case 
1186: also larger than for $n_s=8$. Evidently, the subtle features of spectral 
1187: functions in the intermediate phase are not so well represented by 
1188: including only two bath levels. Nevertheless, the qualitative 
1189: difference between Fermi-liquid behavior for $J=0$ and the clear 
1190: deviation from this behavior for $J=U/4$ are very well reproduced 
1191: by these $n_s=6$ ED/DMFT calculations.   
1192: 
1193: For completeness we show in Fig.~\ref{Paris}(c) the $n_s=8$ ED/DMFT 
1194: results for isotropic Hund's coupling, i.e., $J'=J=U/4$. This case 
1195: is not yet accessible within QMC/DMFT because of sign problems at low 
1196: temperatures. 
1197: The inclusion of spin-flip and pair-exchange terms is seen to restore
1198: the limiting behavior $\Sigma_2(i\omega_n)\rightarrow 0$ for $\omega_n
1199: \rightarrow 0$. Nevertheless, compared to the case $J'=J=0$ shown in (a),
1200: $\Sigma_2$ now increases much more rapidly at small finite frequencies, 
1201: similarly to the data shown in Figs.~\ref{OSMT} and \ref{NRG34}.  
1202: (As argued in Section 5, the shoulder near $\omega_n\approx0.06$ is  
1203: caused by finite-size effects.)
1204: Thus, instead of quasiparticles with weight $Z_2\approx0.8$ for $J=J'=0$,
1205: finite Hund's exchange with $J'=J=U/4$ supports a state of infinite 
1206: lifetime at $E_F$, but does not satisfy Fermi-liquid criteria at 
1207: $\omega>0$. This picture is consistent with the $T=0$ ED/DMFT results
1208: by Biermann {\it et al.}~\cite{biermann05}.
1209: The spectral function of the wide band therefore should exhibit 
1210: a narrow peak at $E_F$, similar to the one shown in Fig.~\ref{NRG12} 
1211: for $U=2.4$~eV. 
1212: 
1213: 
1214: \section{9 \ Comments on IPT/DMFT results}
1215: 
1216: In Ref.~\cite{prb70} we reported DMFT calculations for the two-band 
1217: Hubbard model within QMC and iterated perturbation theory \cite{ipt}. 
1218: Both approaches were shown to yield a consistent picture of the
1219: Mott transition, in the sense that both exhibit a single first-order
1220: transition at which the narrow band becomes insulating and the 
1221: wide band begins to show progressive bad-metallic behavior.
1222: This band becomes fully insulating at a higher Coulomb energy in a 
1223: non-first-order manner. As shown via ED/DMFT at $T=0$ in Ref.~\cite{koga2} 
1224: and at $T>0$ in Ref.~\cite{prl05}, the nature of this upper 
1225: transition depends in a critical way on the treatment of onsite 
1226: exchange interactions. In the absence of spin-flip and pair-exchange
1227: terms it is continuous, consistent with the QMC results, while for
1228: full Hund's coupling it becomes first-order.
1229: 
1230: The question then arises whether the IPT approach also supports this
1231: picture. In fact, as pointed out in Ref.~\cite{prl05}, it is surprising 
1232: that the IPT gives only one common first-order transition even though 
1233: spin-flip and pair-exchange terms are included. It can easily be shown,
1234: however, that to second-order in the Coulomb interaction the subband 
1235: self-energies with and without spin-flip and pair-exchange are the same, 
1236: except for a slightly different relation between $U$ and $J$. Thus, 
1237: $J'=J=0.25\,U$ gives the same result as $J'=0$ and $J\approx0.22\,U$. 
1238: Corrections to the simple second-order diagram via renormalization of subband
1239: energies to yield the correct atomic limit \cite{kajueter,lichtenstein}
1240: are also insensitive to the choice of $J'$ for the present two-band model. 
1241: 
1242: To resolve this puzzle we have repeated the IPT calculations reported 
1243: in Ref.~\cite{prb70} at even lower temperatures and found indeed a tiny 
1244: hysteresis loop also at the upper Mott transition, with a critical 
1245: temperature of approximately $T_{c2}\approx5$~meV, i.e., significantly 
1246: lower than the critical temperature of the lower Mott transition, 
1247: $T_{c1}\approx50$~meV. These findings suggest that present formulations 
1248: of IPT are too simple to deal with the full complexity of Hund's  
1249: exchange. Diagrams beyond second-order are required to distinguish
1250: more clearly between $J'=J$ and $J'=0$ treatments.      
1251: 
1252: 
1253: \section{10 \ Summary and outlook} 
1254: 
1255: Finite-temperature ED/DMFT studies are carried out in order to 
1256: explore the usefulness of this approach for multi-band systems.
1257: The important feature here is that onsite Coulomb and 
1258: exchange interactions are fully included. Since in the past ED
1259: has been used mainly for single-band systems, a particular aim 
1260: of this work is to illustrate the dependence of self-energies 
1261: on the cluster size and to test the range of applicability of 
1262: this method. As an example, we focus on the Hubbard model for
1263: two bands of different widths and investigate the metal insulator 
1264: transition as a function of Hund's exchange.   
1265: 
1266: The surprising and potentially very useful result of this work is
1267: that even a cluster size of $n_s=6$, with only two bath levels per 
1268: impurity orbital, provides a qualitatively correct picture in all 
1269: of the important phases of the $T/U$ phase diagram. Moreover, 
1270: finite-size effects are substantially reduced for $n_s=8$, i.e., using
1271: one extra bath level per impurity orbital. Thus, in the phase just 
1272: below the main first-order Mott transition, both subbands exhibit 
1273: clear metallic properties, albeit with strongly reduced 
1274: quasi-particle weights. The intermediate region between the purely
1275: metallic and insulating phases depends in a subtle manner on the 
1276: exchange interactions included in the ED calculation.  
1277: For full Hund's coupling coexisting metallic and insulating
1278: subbands are found, where the wide band exhibits infinite lifetime
1279: at $E_F$, but non-Fermi-liquid behavior at finite frequencies.
1280: For Ising-like exchange, this breakdown of Fermi-liquid behavior
1281: is enhanced, giving finite lifetime even at $E_F$. 
1282: 
1283: To explore the influence of finite-size effects in the two-band 
1284: ED/DMFT approach NRG calculations were performed for an  effective 
1285: one-band model suitable to describe the wide band in the intermediate 
1286: phase. Both for full Hund's coupling and Ising-like exchange,
1287: very good agreement with the ED results is found for 
1288: temperatures in the range $T=20\ldots30$~meV. At lower $T$, larger
1289: differences appear on a quantitative level at low frequencies.
1290: Nevertheless, the important qualitative differences between various 
1291: phases, especially the characteristic low-frequency variation of 
1292: the self-energy in the two types of non-Fermi-liquid regions for 
1293: isotropic and anisotropic exchange coupling, are fully reproduced 
1294: by the ED approach, both for $n_s=6$ and $n_s=8$.     
1295: 
1296: To test the accuracy of the two-band ED approach we also have applied
1297: it to models studied earlier within QMC/DMFT, neglecting spin-flip
1298: and pair-exchange. Nearly quantitative agreement is obtained for
1299: a cluster size $n_s=8$, but even the ED results for $n_s=6$ are
1300: found to be qualitatively reliable. 
1301: 
1302: The present ED approach utilizes full diagonalization of the 
1303: impurity Hamiltonian. Since at low temperatures only a limited 
1304: range of excited states is relevant for the local Green's functions 
1305: and self-energies it should be very useful to generalize finite-$T$ 
1306: Lanczos one-band methods \cite{jaklic,capone} in order to bridge the 
1307: gap between the present work and the true $T=0$ limit, and to apply 
1308: the finite-temperature ED/DMFT approach to realistic two-band and 
1309: three-band materials.
1310: 
1311: \bigskip
1312: {\bf Acknowledgements}: \ 
1313: One of us (A. L.) likes to thank A. I. Lichtenstein for parts of 
1314: the multi-band ED code. Some of the ED and NRG DMFT calculations 
1315: were carried out on the IBM supercomputer (JUMP) of the 
1316: Forschungszentrum J\"ulich.
1317: 
1318: \bigskip
1319: Email: \ a.liebsch@fz-juelich.de; \ t.costi@fz-juelich.de 
1320: 
1321: \begin{thebibliography}{99}
1322: 
1323: %\bibitem{dmft1}
1324: %   W. Metzner and D. Vollhardt, 
1325: %      Phys. Rev. Lett. {\bf 62}, 324 (1989);
1326: %   A. Georges and G. Kotliar, 
1327: %      Phys. Rev. B {\bf 45}, 6479 (1992);
1328: %   M. Jarrel,                 
1329: %      Phys. Rev. Lett. {\bf 69}, 168 (1992).
1330: 
1331: \bibitem{rmp}
1332:    For a review of early papers, see:
1333:    A. Georges, G. Kotliar, W. Krauth and M. J. Rozenberg, 
1334:    Rev. Mod. Phys. {\bf 68}, 13 (1996).  
1335: 
1336: \bibitem{reviews}
1337:    For recent reviews, see: K. Held, cond-mat/0511293;
1338:    A. Georges {\it et al.}, cond-mat/0403123;  
1339:    G. Kotliar and S. Savrasov, cond-mat/0208241.  
1340: 
1341: \bibitem{qmc} 
1342:    J. E. Hirsch and R. M. Fye, 
1343:       Phys. Rev. Lett. {\bf 56}, 2521 (1986).
1344: 
1345: \bibitem{held}  
1346:    For a discussion of this problem, see: 
1347:    K. Held and D. Vollhardt, 
1348:       Eur. Phys. J. B {\bf 5}, 473 (1998). 
1349: 
1350: \bibitem{arita}
1351:    R. Arita and K. Held,
1352:       Phys. Rev. B {\bf 96}, 201102  (2005); see also:
1353:    S. Sakai, R. Arita and H. Aoki,
1354:       Phys. Rev. B {\bf 70}, 172504  (2004).
1355: 
1356: \bibitem{koga3}
1357:    A. Koga, N. Kawakami, T. M. Rice, and  M. Sigrist,
1358:       Phys. Rev. B  {\bf 72}, 045128 (2005).
1359: 
1360: \bibitem{rubtsov}
1361:    See also: A. N. Rubtsov, V. V. Savkin, and A. I. Lichtenstein,  
1362:       Phys. Rev. B  {\bf 72}, 035122 (2005).
1363: 
1364: 
1365: \bibitem{ed}  
1366:    M. Caffarel and W. Krauth, 
1367:       Phys. Rev. Lett. {\bf 72}, 1545 (1994).
1368: 
1369: \bibitem{anisimov}
1370:    V. I. Anisimov, I. A. Nekrasov, D. E. Kondakov, T. M. Rice 
1371:       and M. Sigrist,
1372:    Euro. Phys. J. B {\bf 25}, 191 (2002). 
1373: 
1374: \bibitem{epl} 
1375:    A. Liebsch, 
1376:       Europhys. Lett. {\bf 63}, 97 (2003). 
1377: 
1378: \bibitem{prl}  
1379:    A. Liebsch, 
1380:       Phys. Rev. Lett. {\bf 91}, 226401 (2003).
1381: 
1382: \bibitem{koga1}
1383:    A. Koga, N. Kawakami, T. M. Rice, and  M. Sigrist,
1384:       Phys. Rev. Lett. {\bf 92}, 216402 (2004).
1385: 
1386: \bibitem{prb70}  
1387:    A. Liebsch, 
1388:       Phys. Rev. B  {\bf 70}, 165103 (2004).
1389: 
1390: \bibitem{koga2}
1391:    A. Koga, N. Kawakami, T. M. Rice, and  M. Sigrist,
1392:       Physica B     {\bf 359-361}, 1366   (2005).
1393: 
1394: \bibitem{song}
1395:    Yun Song and L.-J. Zou, 
1396:       Phys. Rev. B     {\bf 72}, 085114 (2005).
1397: 
1398: \bibitem{prl05}  
1399:    A. Liebsch, 
1400:       Phys. Rev. Lett. {\bf 95}, 116402 (2005).
1401: 
1402: \bibitem{inaba}
1403:    K. Inaba, A. Koga, S. Suga and N. Kawakami,
1404:       J. Phys. Soc. Jpn. {\bf 74}, 2393 (2005).
1405: 
1406: 
1407: \bibitem{medici}
1408:    L. de' Medici, A. Georges and S. Biermann,
1409:       Phys. Rev. B     {\bf 72}, 205124 (2005).
1410: 
1411: \bibitem{ferrero}
1412:    M. Ferrero, F. Becca, M. Fabrizio, and M. Capone,
1413:       Phys. Rev. B     {\bf 72}, 205126 (2006).
1414: 
1415: \bibitem{knecht}
1416:    C. Knecht, N. Bl\"umer, and P.G.J. van Dongen,
1417:       Phys. Rev. B     {\bf 72}, 081103(R) (2005).
1418: 
1419: %\bibitem{comment}
1420: %  A. Liebsch, cond-mat/0506138.  
1421: 
1422: \bibitem{biermann05}
1423:    S. Biermann, L. de' Medici, and A. Georges,
1424:       Phys. Rev. Lett. {\bf 95}, 206401 (2005).
1425: 
1426: \bibitem{jaklic} 
1427:      J. Jakli\v{c} and P. Prelov\v{s}ek, 
1428:         Adv. Phys. {\bf 49}, 1 (2000).
1429: 
1430: \bibitem{capone}
1431:    M. Capone, L. de' Medici, and A. Georges,
1432:       cond-mat/ 0512484.
1433: 
1434: \bibitem{maeno}
1435:    T. Maeno, T. M. Rice, and M. Sigrist, 
1436:       Phys. Today {\bf 54}, (1) 42 (2001).
1437: 
1438: \bibitem{nacoo}
1439:    K. Takada {\it et al.}, Nature {\bf 422}, 53 (2003); 
1440:    R. E. Schaak {\it et al.}, Nature {\bf 424}, 527 (2003);
1441:    M. L. Foo {\it et al.}, Phys. Rev. Lett. {\bf 92}, 247001 (2004). 
1442: 
1443: \bibitem{pruschke}
1444:    Th. Pruschke and R. Bulla, 
1445:      Eur. Phys. J. B {\bf 44}, 217 (2005).
1446: 
1447: \bibitem{costi.02} 
1448:      T. A. Costi and N. Manini, 
1449:       J. Low Temp. Phys. {\bf 126}, 835 (2002)
1450: 
1451: \bibitem{wilson.74} 
1452:       K. G. Wilson, 
1453:       Rev. Mod. Phys. {\bf 47}, 773 (1975).
1454: 
1455: \bibitem{costi.94} 
1456:        T. A. Costi, A. C. Hewson and V. Zlatic, 
1457:        J. Phys.: Condens. Matter {\bf 6}, 2519 (1994).
1458: 
1459: \bibitem{hofstetter.00} 
1460:       W. Hofstetter, 
1461:       Phys. Rev. Lett. {\bf 85}, 1508 (2000).
1462: 
1463: \bibitem{cragg.79} 
1464:     D. M. Cragg and P. Lloyd, J. Phys. C: Solid State 
1465:        Phys. {\bf 12}, L215 (1979).
1466: 
1467: \bibitem{koller.05} 
1468:      W. Koller,  A. C. Hewson and D. Meyer, 
1469:         Phys. Rev. B{\bf 72}, 045117 (2005).
1470: 
1471: \bibitem{bulla.98} 
1472:       R. Bulla, T. Pruschke and A. C. Hewson, 
1473:       J. Phys.: Condens. Matter {\bf 10 }, 8365 (1998).
1474: 
1475: \bibitem{bulla.01}
1476:       R. Bulla, T. A. Costi and D. Vollhardt, 
1477:         Phys. Rev. B{\bf 64}, 045103 (2001).
1478: 
1479: \bibitem{costi} 
1480:        T. A. Costi, to be published.
1481: 
1482: \bibitem{metallic}
1483:    A. Liebsch, to be published.
1484: 
1485: 
1486: \bibitem{kajueter}
1487:    H. Kajueter and G. Kotliar, 
1488:       Phys. Rev. Lett. {\bf 77}, 131  (1997).
1489: 
1490: \bibitem{ipt}
1491:    A. Georges and G. Kotliar, 
1492:       Phys. Rev. B {\bf 45}, 6479 (1992).
1493: 
1494: \bibitem{lichtenstein}
1495:    A. I. Lichtenstein and M. I. Katsnelson,
1496:       Phys. Rev. B     {\bf 57}, 6884 (1998).
1497: 
1498: \end{thebibliography}
1499: \end{document}
1500: 
1501: