Metal- transitions

Masatoshi Imada Institute for State , University of Tokyo, Roppongi, Minato-ku, Tokyo, 106, Japan

Atsushi Fujimori Department of Physics, University of Tokyo, Hongo, Bunkyo-ku, Tokyo, 113, Japan

Yoshinori Tokura Department of Applied Physics, University of Tokyo, Hongo, Bunkyo-ku, Tokyo, 113, Japan

Metal-insulator transitions are accompanied by huge resistivity changes, even over tens of orders of magnitude, and are widely observed in condensed-matter systems. This article presents the observations and current understanding of the metal-insulator transition with a pedagogical introduction to the subject. Especially important are the transitions driven by correlation effects associated with the -electron interaction. The insulating caused by the correlation effects is categorized as the . Near the transition point the metallic state shows fluctuations and orderings in the spin, charge, and orbital degrees of freedom. The properties of these metals are frequently quite different from those of ordinary metals, as measured by transport, optical, and magnetic probes. The review first describes theoretical approaches to the unusual metallic states and to the metal-insulator transition. The Fermi- theory treats the correlations that can be adiabatically connected with the noninteracting picture. Strong-coupling models that do not require Fermi-liquid behavior have also been developed. Much work has also been done on the scaling theory of the transition. A central issue for this review is the evaluation of these approaches in simple theoretical systems such as the and t-J models. Another key issue is strong competition among various orderings as in the interplay of spin and orbital fluctuations. Experimentally, the unusual properties of the metallic state near the insulating transition have been most extensively studied in d-electron systems. In particular, there is revived interest in transition-metal , motivated by the epoch-making findings of high-temperature in cuprates and colossal magnetoresistance in manganites. The article reviews the rich phenomena of anomalous metallicity, taking as examples Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Ru compounds. The diverse phenomena include strong spin and orbital fluctuations, mass renormalization effects, incoherence of charge dynamics, and phase transitions under control of key parameters such as band filling, bandwidth, and dimensionality. These parameters are experimentally varied by , pressure, chemical composition, and magnetic fields. Much of the observed behavior can be described by the current theory. Open questions and future problems are also extracted from comparison between experimental results and theoretical achievements. [S0034-6861(98)00103-2]

CONTENTS 4. Green’s functions, self-energy, and spectral functions 1060 D. Fermi-liquid theory and various mean-field I. Introduction 1040 approaches: Single-particle descriptions of A. General remarks 1040 correlated metals and insulators 1062 B. Remarks on theoretical descriptions 1. Fermi-liquid description 1063 (Introduction to Sec. II) 1044 2. Hartree-Fock approximation and RPA for C. Remarks on material systematics (Introduction phase transitions 1070 to Sec. III) 1046 3. Local-density approximation and its D. Remarks on experimental results of anomalous refinement 1074 metals (Introduction to Sec. IV) 1047 4. Hubbard approximation 1078 II. Theoretical Description 1048 5. Gutzwiller approximation 1079 A. Theoretical models for correlated metals and 6. Infinite-dimensional approach 1080 Mott insulators in d-electron systems 1048 7. Slave-particle approximation 1084 1. Electronic states of d-electron systems 1048 8. Self-consistent renormalization 2. Lattice fermion models 1050 approximation and two-particle self- B. Variety of metal-insulator transitions and consistent approximation 1087 correlated metals 1053 9. Renormalization-group study of magnetic C. Field-theoretical framework for interacting transitions in metals 1089 fermion systems 1057 E. Numerical studies of metal-insulator transitions 1. Coherent states 1058 for theoretical models 1092 2. Grassmann algebra 1058 1. Drude weight and transport properties 1093 3. Functional integrals, path integrals, and 2. Spectral function and 1097 statistical mechanics of many-fermion 3. Charge response 1099 systems 1059 4. Magnetic correlations 1100

Reviews of Modern Physics, Vol. 70, No. 4, October 1998 0034-6861/98/70(4)/1039(225)/$60.00 © 1998 The American Physical Society 1039 1040 Imada, Fujimori, and Tokura: Metal-insulator transitions

5. Approach from the insulator side 1103 1. Fe3O4 1208 6. Bosonic systems 1103 2. La1ϪxSrxFeO3 1211 F. Scaling theory of metal-insulator transitions 1104 3. La2ϪxSrxNiO4 1212 1. Hyperscaling scenario 1105 4. La1ϪxSr1ϩxMnO4 1216 2. Metal-insulator transition of a noninteracting F. Double exchange systems 1218 system by disorder 1106 1. R1ϪxAxMnO3 1218 3. Drude weight and charge compressibility 1107 2. La2Ϫ2xSr1ϩ2xMn2O7 1225 4. Scaling of physical quantities 1109 G. Systems with transitions between metal and 5. Filling-control transition 1109 nonmagnetic insulators 1228 6. Critical exponents of the filling-control 1. FeSi 1228 metal-insulator transition in one dimension 1110 2. VO2 1232 7. Critical exponents of the filling-control 3. Ti2O3 1233 metal-insulator transition in two and three 4. LaCoO3 1235 dimensions 1110 5. La1.17ϪxAxVS3.17 1239 8. Two universality classes of the filling-control H. 4d systems 1241 metal-insulator transition 1111 1. Sr2RuO4 1241 9. Suppression of coherence 1111 2. Ca1ϪxSrxRuO3 1243 10. Scaling of spin correlations 1113 V. Concluding Remarks 1245 11. Possible realization of scaling 1114 Acknowledgments 1247 12. Superfluid-insulator transition 1115 References 1247 G. Non-fermi-liquid description of metals 1116 1. Tomonaga- in one dimension 1116 2. Marginal Fermi liquid 1118 I. INTRODUCTION H. Orbital degeneracy and other complexities 1119 1. Jahn-Teller distortion, spin, and orbital A. General remarks fluctuations and orderings 1119 2. Ferromagnetic metal near a Mott insulator 1122 The first successful theoretical description of metals, 3. Charge ordering 1123 insulators, and transitions between them is based on III. Materials Systematics 1123 noninteracting or weakly interacting electron systems. A. Local electronic structure of transition-metal The theory makes a general distinction between metals oxides 1123 1. Configuration-interaction cluster model 1123 and insulators at zero temperature based on the filling of 2. Cluster-model analyses of photoemission the electronic bands: For insulators the highest filled spectra 1125 band is completely filled; for metals, it is partially filled. 3. Zaanen-Sawatzky-Allen classification scheme 1126 In other words, the Fermi level lies in a band gap in 4. Multiplet effects 1128 insulators while the level is inside a band for metals. In 5. Small or negative charge-transfer energies 1130 the noninteracting electron theory, the formation of B. Systematics in model parameters and charge band structure is totally due to the periodic lattice struc- gaps 1131 ture of in crystals. This basic distinction between 1. Ionic crystal model 1131 metals and insulators was proposed and established in 2. Spectroscopic methods 1132 the early years of quantum mechanics (Bethe, 1928; 3. First-principles methods 1135 C. Control of model parameters in materials 1139 Sommerfeld, 1928; Bloch, 1929). By the early 1930s, it 1. Bandwidth control 1139 was recognized that insulators with a small 2. Filling control 1141 between the highest filled band and lowest empty band 3. Dimensionality control 1142 would be due to thermal excitation of IV. Some Anomalous Metals 1144 the (Wilson, 1931a, 1931b; Fowler, 1933a, A. Bandwidth-control metal-insulator transition 1933b). More than fifteen years later the transistor was systems 1144 invented by Shockley, Brattain, and Bardeen. 1. V2O3 1144 Although this band picture was successful in many re- 2. NiS2ϪxSex 1151 spects, de Boer and Verwey (1937) reported that many 3. RNiO3 1154 transition-metal oxides with a partially filled d-electron 4. NiS 1157 band were nonetheless poor conductors and indeed of- 5. Ca1ϪxSrxVO3 1163 B. Filling-control metal-insulator transition systems 1165 ten insulators. A typical example in their report was NiO. Concerning their report, Peierls (1937) pointed out 1. R1ϪxAxTiO3 1165 2. R1ϪxAxVO3 1171 the importance of the electron-electron correlation: C. High-Tc cuprates 1173 Strong Coulomb repulsion between electrons could be 1. La2ϪxSrxCuO4 1173 the origin of the insulating behavior. According to Mott 2. Nd2ϪxCexCuO4 1184 (1937), Peierls noted 3. YBa Cu O 1187 2 3 7Ϫy ‘‘it is quite possible that the electrostatic interaction 4. Bi2Sr2CaCu2O8ϩ␦ 1195 D. Quasi-one-dimensional systems 1199 between the electrons prevents them from moving at 1. Cu-O chain and ladder compounds 1199 all. At low temperatures the majority of the electrons

2. BaVS3 1205 are in their proper places in the ions. The minority E. Charge-ordering systems 1208 which have happened to cross the potential barrier

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1041

Mott, 1990), in his original formulation Mott argued that the existence of the insulator did not depend on whether the system was magnetic or not. Slater (1951), on the other hand, ascribed the origin of the insulating behavior to magnetic ordering such as the antiferromagnetic long-range order. Because most Mott insulators have magnetic ordering at least at zero tem- perature, the insulator may appear due to a band gap generated by a superlattice structure of the magnetic pe- riodicity. In contrast, we have several examples in which spin excitation has a gap in the Mott insulator without magnetic order. One might argue that this is not com- patible with Slater’s band picture. However, in this case, FIG. 1. Metal-insulator phase diagram based on the Hubbard both charge and spin gaps exist similarly to the band model in the plane of U/t and filling n. The shaded area is in insulator. This could give an adiabatic continuity be- principle metallic but under the strong influence of the metal- tween the Mott insulator and the band insulator, which insulator transition, in which carriers are easily localized by we discuss in Sec. II.B. extrinsic forces such as randomness and electron-lattice cou- In addition to the Mott insulating phase itself, a more pling. Two routes for the MIT (metal-insulator transition) are difficult and challenging subject has been to describe shown: the FC-MIT (filling-control MIT) and the BC-MIT and understand metallic phases near the Mott insulator. (bandwidth-control MIT). In this regime fluctuations of spin, charge, and orbital correlations are strong and sometimes critically en- hanced toward the MIT, if the transition is continuous find therefore all the other atoms occupied, and in or weakly first order. The metallic phase with such order to get through the lattice have to spend a long strong fluctuations near the Mott insulator is now often time in ions already occupied by other electrons. called the anomalous metallic phase. A typical anoma- This needs a considerable addition of energy and so lous fluctuation is responsible for mass enhancement in is extremely improbable at low temperatures.’’ V2O3, where the specific-heat coefficient ␥ and the Pauli These observations launched the long and continuing paramagnetic susceptibility ␹ near the MIT show sub- history of the field of strongly correlated electrons, par- stantial enhancement from what would be expected ticularly the effort to understand how partially filled from the noninteracting band theory. To understand this bands could be insulators and, as the history developed, mass enhancement, the earlier pioneering work on the how an insulator could become a metal as controllable MIT by Hubbard (1964a, 1964b) known as the Hubbard parameters were varied. This transition illustrated in approximation was reexamined and treated with the Fig. 1 is called the metal-insulator transition (MIT). The Gutzwiller approximation by Brinkmann and Rice insulating phase and its fluctuations in metals are indeed (1970). the most outstanding and prominent features of strongly Fermi-liquid theory asserts that the ground state and correlated electrons and have long been central to re- low-energy excitations can be described by an adiabatic search in this field. switching on of the electron-electron interaction. Then, In the past sixty years, much progress has been made naively, the carrier number does not change in the adia- from both theoretical and experimental sides in under- batic process of introducing the electron correlation, as standing strongly correlated electrons and MITs. In the- is celebrated as the Luttinger theorem. Because the oretical approaches, Mott (1949, 1956, 1961, 1990) took Mott insulator is realized for a partially filled band, this the first important step towards understanding how adiabatic continuation forces the carrier density to re- electron-electron correlations could explain the insulat- main nonzero when one approaches the MIT point in ing state, and we call this state the Mott insulator.He the framework of Fermi-liquid theory. Then the only considered a lattice model with a single electronic or- way to approach the MIT in a continuous fashion is the bital on each site. Without electron-electron interac- divergence of the single- mass m* (or more tions, a single band would be formed from the overlap of strictly speaking the vanishing of the renormalization the atomic orbitals in this system, where the band be- factor Z) at the MIT point. Therefore mass enhance- comes full when two electrons, one with spin-up and the ment as a typical property of metals near the Mott insu- other with spin-down, occupy each site. However, two lator is a natural consequence of Fermi-liquid theory. electrons sitting on the same site would feel a large Cou- If the symmetries of spin and orbital degrees of free- lomb repulsion, which Mott argued would split the band dom are broken (either spontaneously as in the mag- in two: The lower band is formed from electrons that netic long-range ordered phase or externally as in the occupied an empty site and the upper one from elec- case of crystal-field splitting), the adiabatic continuity trons that occupied a site already taken by another elec- assumed in the Fermi-liquid theory is not satisfied any tron. With one electron per site, the lower band would more and there may be no observable mass enhance- be full, and the system an insulator. Although he dis- ment. In fact, a MIT with symmetry breaking of spin and cussed the magnetic state afterwards (see, for example, orbital degrees of freedom is realized by the vanishing of

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1042 Imada, Fujimori, and Tokura: Metal-insulator transitions the single-particle (carrier) number, as in the transition MIT in correlated metals has been most thorough and to a band insulator. For example, this takes place in the systematic in d-electron systems, namely, transition- transition from an antiferromagnetic metal to an antifer- metal compounds. Many examples will be reviewed in romagnetic Mott insulator, where the folding of the Bril- this article. In d-electron systems, orbital degeneracy is louin zone due to the superstructure of the magnetic an important and unavoidable source of complicated be- periodicity creates a completely filled lower band. Car- havior. For example, under the cubic crystal-field sym- riers are doped into small pockets of Fermi surface metry of the lattice, any of the threefold degenerate t2g whose Fermi volume vanishes at the MIT. bands, dxy , dyz , and dzx as well as twofold degenerate From this heuristic argument, one can see at least two eg bands, dx2Ϫy2 and d3z2Ϫr2, can be located near the distinct routes to the Mott insulator when one ap- Fermi level, depending on transition-metal ion, lattice proaches the MIT point from the metallic side, namely, structure, composition, dimensionality, and so on. In ad- the mass-diverging type and carrier-number-vanishing dition to strong spin fluctuations, effects of orbital fluc- type. The diversity of anomalous features of metallic tuations and orbital symmetry breaking play important phases near both types of MIT is a central subject of this roles in many d-electron systems, as discussed in Secs. review. Mass enhancement or carrier-number reduction II.H and IV. The orbital correlations are frequently as well as more complicated features have indeed been strongly coupled with spin correlations through the observed experimentally. The experiments were exam- usual relativistic spin-orbit coupling as well as through ined from various, more or less independently devel- orbital-dependent exchange interactions and quan- oped theoretical approaches, as detailed in Sec. II. In drupole interactions. An example of this orbital effect particular, anomalous features of correlated metals near known as the double-exchange mechanism is seen in Mn the Mott insulator appear more clearly when the MIT is oxides (Sec. IV.F), where strong Hund’s-rule coupling continuous. Theoretically, this continuous MIT has been between the eg and t2g orbitals triggers a transition be- a subject of recent intensive studies in which unusual tween the Mott insulator with antiferromagnetic order metallic properties are understood from various critical and the ferromagnetic metal. Colossal negative magne- fluctuations near the of the MIT. toresistance near the transition to this ferromagnetic A prototype of theoretical understanding for the tran- metal phase has been intensively studied recently. sition between the Mott insulator and metals was Another aspect of orbital degeneracy is the overlap or achieved by using simplified lattice fermion models, in the closeness of the d band and the p band of ligand particular, in the celebrated Hubbard model (Anderson, atoms which bridge the elements in transition-metal 1959; Hubbard, 1963, 1964a, 1964b; Kanamori, 1963). compounds. For example, as clarified in Secs. II.A and The Hubbard model considers only electrons in a single III.A, in the transition-metal oxides, the oxygen 2p␴ band. Its Hamiltonian in a second-quantized form is level becomes close to that of the partially filled 3d band given by near the Fermi level for heavier transition-metal ele- ments such as Ni and Cu. Then the charge gap of the ϭ ϩ Ϫ␮N, (1.1a) HH Ht HU Mott insulator cannot be accounted for solely with d † electrons, but p-electron degrees of freedom have also tϭϪt͚ ͑ci␴cj␴ϩH.c.͒, (1.1b) to be considered. In fact, when we could regard the H ͗ij͘ Hubbard model as a description of a d band only, the

1 1 charge excitation gap is formed between a singly occu- UϭU ͑ni Ϫ 2 ͒͑ni Ϫ 2 ͒, (1.1c) H ͚i ↑ ↓ pied d band (the so-called lower Hubbard band) and a doubly occupied (with spin up and down) d band (the and so-called upper Hubbard band). However, if the p␴ level becomes closer, the character of the minimum charge Nϵ ni␴ , (1.1d) excitation gap changes to that of a gap between a singly ͚i␴ occupied d band with fully occupied p band and a singly where the creation (annihilation) of the single-band occupied d band with a p␴ hole. This kind of insulator, † electron at site i with spin ␴ is denoted by ci␴(ci␴) with which was clarified by Zaanen, Sawatzky, and Allen † ni␴ being the number operator ni␴ϵci␴ci␴ . In this sim- (1985), is now called a charge-transfer (CT) insulator as plification, various realistic complexities are ignored, as contrasted to the former case, the Mott-Hubbard (MH) we shall see in Sec. II.A. However, at the same time, insulator, which we discuss in detail in Secs. II.A and low-energy and low-temperature properties are often III.A, and the distinction is indeed observed in high- well described even after this simplification since only a energy spectroscopy. Correspondingly, compounds that small number of bands (sometimes just one band) are have the MH insulating phase are called MH com- crossing the Fermi level and have to do with low-energy pounds while those with the CT insulating phase are excitations. The parameters of the simplified models in called CT compounds. The term ‘‘Mott insulator’’ is this case should be taken as effective values derived used in this review in a broad sense which covers both from renormalized bands near the Fermi level. types. Recent achievements in the field of strongly cor- One of the most drastic simplification in the Hubbard related electrons, especially in d-electron systems, have model is to consider only electrons in a single orbit, say brought us closer to understanding more complicated the s orbit. In contrast, the experimental study of the situations in which there is an interplay between orbital

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1043 and spin fluctuations and lattice degrees of freedom. Al- the standard Fermi-liquid behavior. A typical anoma- though we focus on the d-electron systems in this article, lous property is the temperature dependence of the re- other systems also provide many interesting aspects in sistivity, which is linearly proportional to the tempera- terms of strong-correlation effects. For example, organic ture T, in contrast with the T2 dependence expected in systems with ␲-electron conduction such as BEDT-TTF the Fermi-liquid theory. The optical conductivity compounds show two-dimensional (2D) anisotropy with roughly proportional to the inverse frequency is another an interesting interplay of MIT and transition to super- unusual aspect, because it contradicts the well-known conducting state (see, for example, Kanoda, 1997). An- Drude theory. These indicate that the charge dynamics other example is seen in f-electron systems, where, for have a strongly incoherent character. Another example instance, non-Fermi-liquid properties due to orbital de- is the so-called pseudogap behavior observed in the un- generacy effects have been extensively studied. derdoped region near the Mott insulating phase. The Most theoretical effort has been concentrated on sim- pseudogap is a frequency threshold for the strong exci- plified single-orbital systems such as the Hubbard and tation of spin and charge modes. We review basic fea- t-J models, in which the interplay of strong quantum tures of these unusual properties observed in a number fluctuations and correlation effects has been reexam- of experimental probes, such as photoemission, nuclear ined. In fact, the single-band model was originally de- magnetic resonance, neutron scattering, transport mea- rived, as we mentioned above, to discuss low-energy ex- surements, and optical measurements, in Sec. IV.C, citations of renormalized simple bands near the Fermi while we discuss various theoretical approaches to un- level even in the presence of a complicated structure of derstanding them in Sec. II. We concentrate on proper- high-energy bands far from the Fermi level. This trend ties in the normal state related to the MIT, while the was clearly motivated and greatly accelerated by the dis- properties in the superconducting phase are not re- covery of high-Tc cuprate superconductors (Bednorz and viewed in this article. Mu¨ ller, 1986). In the cuprate superconductors, the low- Strong incoherence actually turned out to be a com- energy electronic structure of the copper is rather mon property of most transition-metal compounds near well described only by the dx2Ϫy2 band separated from the point, as discussed in Sec. IV. We the d3z2Ϫr2 and t2g bands. Although the 2p␴ band is call these incoherent metals. The incoherent charge dy- close to the dx2Ϫy2 band, as the copper oxides are typical namics as well as other anomalies in spin, charge, and CT compounds, the 2p␴ band is strongly hybridized with orbital fluctuations all come from almost localized but the dx2Ϫy2 band and consequently forms a single anti- barely itinerant electrons in the critical region of transi- bonding band near the Fermi level. This again lends sup- tion between metals and the Mott insulator. These criti- port for a description by an effective single-band model cal regions are not easy to understand from either the (Anderson, 1987; Zhang and Rice, 1988). metal or the insulator side, but their critical properties Mother compounds of the cuprate superconductors can be analyzed by the techniques for quantum critical are typical Mott insulators with antiferromagnetic order phenomena discussed in Sec. II. where the Cu is in a d9 state with a half-filled The two important parameters in the Hubbard model dx2Ϫy2 orbital. A small amount of carrier doping to this are the electron correlation strength U/t and the band Mott insulating state drives the MIT and directly results filling n. The schematic metal-insulator phase diagram in the superconducting transition at low temperatures in terms of these parameters is presented in Fig. 1. In the for low carrier doping. Because 2D anisotropy is promi- case of a nondegenerate band, the nϭ0 and nϭ2 fillings nent in cuprates due to the layered perovskite structure, correspond to the band insulator. For the half-filled case strong quantum fluctuations with suppression of antifer- (nϭ1), it is believed that the change of U/t drives the romagnetic long-range order are the key aspects. One- insulator-to-metal transition (Mott transition) at a criti- and two-dimensional systems without orbital degeneracy cal value of U/t except in the case of perfect nesting, are the extreme cases to enhance this quantum fluctua- where the critical value Uc is zero. This transition at a tion, where high-Tc superconductivity was discovered. finite Uc is called a bandwidth control (BC)-MIT. Mott This has strongly motivated studies of the ground state argued that the BC-MIT becomes a first-order transition and low-temperature properties of simple low- when we consider the long-range Coulomb force not dimensional models such as the 2D Hubbard model be- contained in the Hubbard model (Mott, 1956, 1990). He yond the level of mean-field approximations, taking ac- reasoned supposing that the carrier density decreases count of strong fluctuations of spin and charge. It is with increasing U/t. Then screening of long-range Cou- widely recognized that reliable treatments of normal- lomb forces by other carriers becomes ineffective, which state properties with strong low-dimensional anisotropy results in the formation of an electron-hole bound pair near the MIT are imperative for an understanding of at a finite U. This causes the first-order transition to the this type of superconductivity, while low dimensionality insulating state. Here we note that the validity of Mott’s makes various mean-field treatments very questionable argument is still an open question because, as mentioned The importance of strong correlation to our understand- above, the Mott transition is not necessarily of the ing of cuprate superconductors was stressed in the early carrier-number-vanishing type but can be of the mass- stage by Anderson (1987). diverging type, in contrast to the transition to the band In fact, the normal state of the cuprate superconduct- insulator. We shall discuss this later in greater detail. ors shows many unusual properties which are far from Irrespective of Mott’s argument the coupling to lattice

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1044 Imada, Fujimori, and Tokura: Metal-insulator transitions degrees of freedom also favors the first-order transition high-temperature cuprate superconductors. Several at- by discontinuously increasing the transfer amplitude t in tempts have been made, along the lines of Slater’s ap- the metallic phase. The coupling of spin and orbital may proach, to account for the ground state of Mott insula- also favor the first-order transition in real materials. In tors such as La2CuO4 and other transition-metal oxides the region adjacent to the nϭ1 insulating line (a thick as well as metallic phases, as we shall see in Sec. II.D. line in Fig. 1), the spin-ordered phase is frequently ob- Some recent developments to treat the correlation ef- served, that is, an antiferromagnetic insulating or metal- fects in the framework of Fermi-liquid theory are re- lic phase. viewed in Sec. II.D.1 together with descriptions of the The filling at noninteger n usually leads to the metal- basic phenomenology of this theory. Effects of fluctua- lic phase. The phase of particular interest is that of met- tion in spin and charge responses can be taken into ac- als near the nϭ1 insulating line which is derived by fill- count on the level of the mean-field approximation by ing control (FC) from the parent Mott insulator (n the random-phase approximation (RPA). When the ϭ1). Since the discovery of high-temperature supercon- symmetry is broken, the Hartree-Fock (HF) theory is the ductors, the concept of ‘‘carrier doping’’ or FC in the simplest way to describe the ordered state as well as the parent Mott insulators has been widely recognized as . The basic content of these mean-field one of the important aspects of the MIT in the 3d elec- approximations is summarized in Sec. II.D.2. Theoreti- tron systems. In contrast to the BC-MIT, the above two cal description beyond the mean-field level is the subject mechanisms for a first-order transition are not effective of later sections, Secs. II.E and II.F. To understand the for the FC-MIT. The reason is that the electron-hole correlation effect in d-electron systems on a more quan- bound states cannot make an insulator by themselves titative level, one has to go beyond the tight-binding ap- due to the absence of carrier compensation. The cou- proximation implicit in the Hubbard Hamiltonian. pling to the lattice is also ineffective because usually it When the correlation effect is not important, the local- does not couple to the electron filling. Therefore the density approximation (LDA) is a useful tool for calcu- transition can easily be continuous for an FC-MIT. In a lating the band structure of metals in the ground state. relatively large-U/t region near the insulating phase of However, it is well known that the simple LDA is poor nϭ0, 1, and 2, however, the compounds occasionally at reproducing the Mott insulating state as well as suffer from the carrier localization effect arising from anomalous metallic states near the Mott insulator. Re- the static random potential and/or electron-lattice inter- finements of the LDA are then an important issue. action. In addition, at some fractional but commensurate These attempts include the local spin-density approxi- fillings such as nϭ1/8, 1/3, and 1/2, the compounds mation (LSDA), generalized (density) gradient approxi- sometimes undergo the charge-ordering phase transition. mation (GGA), the GW approximation, the so-called We shall see ample examples in Sec. IV.E, associated LDAϩU approximation, and self-interaction correction with the commensurate charge-density and spin-density (SIC), which are reviewed in Sec. II.D.3. At finite tem- waves. The latter phenomenon is related not only with peratures, a mode-mode coupling theory [the self- the short-ranged electron correlation as represented by consistent renormalization (SCR) approximation] was U/t but also with the inter-site Coulomb interaction. developed in metals to consistently take into account Keeping these phenomena in mind, we shall review in weak-amplitude spin fluctuations, which are ignored in Sec. III these electronic control parameters and how the Hartree-Fock calculation (Moriya, 1985). This they are varied in actual materials. theory may be viewed as a self-consistent one-loop ap- proximation in the weak-correlation approach. It is B. Remarks on theoretical descriptions (Introduction briefly reviewed in Sec. II.D.8. The basic assumption of to Sec. II) the approaches initiated by Slater is that the perturba- tion expansion or the Hartree-Fock approximation and Section II describes different theoretical approaches RPA in terms of spin correlations of electron quasipar- for strongly correlated electron systems and their MITs. ticles is valid at least at zero temperature and well de- A conventional way of describing the correlation effects scribes metals and the Mott insulator separately. On the is to assume the adiabatic continuity of the paramag- other hand, in the approaches that go back to the origi- netic metallic phase with the noninteracting system nal idea of Mott, several attempts have been made to based on the Fermi-liquid description. The Mott insulat- take account of strong-correlation effects through strong ing phase is interpreted as in the pioneering work of charge correlations in a nonperturbative way. These at- Slater as a consequence of a charge gap’s opening due to tempts include the Hubbard approximation (Hubbard, symmetry breaking of spin or orbit. In the other ap- 1964a, 1964b), the Gutzwiller approximation proaches, qualitatively different aspects which are (Gutzwiller, 1965; Brinkman and Rice, 1970), various claimed to be inaccessible from the Fermi-liquid-type slave-particle approximations (Barnes, 1976, 1977; Cole- approach are stressed. Since the proposals by Mott and man, 1984), and the infinite-dimensional approach Slater, these two viewpoints have played complementary (Kuramoto and Watanabe, 1987; Metzner and Voll- roles sometimes in understanding the correlation effects, hardt, 1989, Mu¨ ller-Hartmann, 1989). In general, these even though they provide contradicting scenarios in approaches do not seriously take account of the wave- many cases, as we discuss in Sec. II in detail. A typical number dependence (or spatial dependence) of the cor- example of the controversy is seen in recent debates on relation. In addition, they usually neglect effects of spin

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1045

fluctuations. One coherent-potential approximation that this attempt is immediately faced with the difficulty of does not assume magnetic order was developed by Hub- describing properties above the Ne´el temperature TN , bard (1964a, 1964b) and is summarized in Sec. II.D.4. because the insulating behavior is robust well above TN This approach first succeeded in substantiating the ap- in many cases, such as La2CuO4 and MnO. In fact, in pearance of the Mott insulating state along the lines of La2CuO4, the charge excitation gap of the order of 2 eV, the original idea by Mott. Another approach on a basi- clearly observed in optical measurements above TN ,is cally mean-field level is the Gutzwiller approximation. two orders of magnitude larger than the energy scale of This approximation, first applied to the MIT by Brink- TN . To understand this, a formalism to allow the forma- man and Rice (1970), is discussed in Sec. II.D.5. These tion of electron-hole bound states well above TN is nec- two approaches, by Hubbard and by Brinkman and essary. This is similar to the separation of the Bose con- Rice, have close connections to recently developed densation temperature and the energy scale of bound treatments in infinite dimensions and the slave-particle boson formation as in 4He. An efficient way of describ- approximation. On the metallic side, the Gutzwiller ap- ing phenomena below the electron-hole excitation en- proximation, the original slave-boson approximation, ergy, namely, the charge excitation gap, is to derive an and the solution in infinite dimensions give essentially effective Hamiltonian of Heisenberg spins, as is dis- the same results for the low-energy excitations, indicat- cussed in Sec. II.A. The formation of electron-hole ing the basic equivalence of these three approximations bound states well above TN is a clear indication of the for a description of the coherent part. In fact, the breakdown of the single-particle picture. Of course since Gutzwiller and slave-particle approximations treat only the low-energy collective excitations enhance the quan- the coherent part and neglect the incoherent part. On tum spin fluctuations even at the level of zero-point os- the insulating side, the Hubbard approximation and the cillations, quantitatively reliable description of the infinite-dimensional approach provide similar results, ground state cannot be achieved by the usual single- because the Hubbard approximation considers only the particle description with a Hartree-Fock-type approxi- incoherent excitations. Sections II.D.6 and II.D.7 discuss mation. the results of the infinite-dimensional and slave-particle The second serious problem is in describing the MIT. approximations, respectively. The infinite-dimensional The breakdown of the single-particle description of elec- approach is sometimes called the dynamic mean-field trons is most clearly observed when the MIT is continu- theory and treats the dynamic fluctuation correctly when ous because the metallic ground state must be recon- spatial fluctuations can be ignored. structed from the low-energy states of the insulator Aside from the above more or less biased approaches, when the insulator undergoes a continuous transition to numerical methods have been developed for the pur- a metal. There, critical fluctuations can only be treated pose of obtaining insights without approximations. A by taking account of relevant collective excitations and number of numerical studies are discussed in Sec. II.E. growth of short-ranged correlations. Continuous MITs To discuss this problem further, we should keep the are the subject of extensive work in recent theoretical following point in mind: When we consider the ground and experimental studies. state of metals and insulators separately far away from The critical fluctuation is not observable as a true di- the transition point, each of the ground states may in vergence if the MIT is of first order. Therefore, in this many cases be correctly expressed by adiabatic continu- case, whether interesting anomalous fluctuations in met- ations of the fixed-point solutions obtained from rather als appear or not relies on whether the first-order tran- simple Hartree-Fock approximations or perturbative ex- sition is weak or not. Experimental aspects of first-order pansions in terms of the correlation. However, an ex- transitions are considered in Sec. IV. ample is known in 1D interacting systems where the As is discussed in Sec. IV, many compounds, such as usual perturbation expansions and Hartree-Fock ap- some titanium oxides and vanadium oxides with the 3D proximations clearly break down and the correct ground perovskite structure, some of the sulfides and selenides, state is not reproduced from the weak-coupling ap- as well as 2D systems including high-Tc cuprates, appear proach. Its correct fixed point is believed to be the to show a continuous MIT. To understand critical fluc- Tomonaga-Luttinger liquid. The analysis of 2D systems tuations near the MIT, it is crucial to extract the major in terms of the Fermi-liquid fixed point is still controver- driving force of the transition. One of the driving forces sial. In any case, aside from a few exceptional cases such is the Anderson localization, and another is the correla- as the 1D systems and systems under strong magnetic tion effect toward the Mott insulator. One may also ar- fields, the single-particle description of electrons is gen- gue the role of the electron- interaction. In real- erally robust. However, even in the cases where the istic situations, the effect of randomness, namely, the fixed point is more or less reproduced by the single- Anderson localization, must be seriously considered particle picture of electrons, weak-coupling approaches when the system is sufficiently close to the transition based on the perturbation expansion and/or Hartree- point. However, in many cases such as high-Tc cuprates, Fock-type analysis have serious difficulties in describing several other transition-metal oxides, and 2D 3He, the correlated metals. strong-correlation effect appears to be the dominant The first serious problem is in describing excitations. driving force of the MIT. As a prototype, the nature of For example, if one wants to ascribe the insulating na- the transition from (to) the Mott insulator in clean sys- ture of the Mott insulator to antiferromagnetic ordering, tems, if clarified, provides a good starting point and is

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1046 Imada, Fujimori, and Tokura: Metal-insulator transitions helpful in establishing new concepts. Theoretical under- mates. Once we know the parameter values, especially standing of the MIT caused by pure strong-correlation how they change when the chemical composition of ma- effects is an important subject of this review. terials is changed, we are able to predict to some extent To understand the continuous MIT, the idea of scaling the physical properties of those materials using the pa- was first applied to the Anderson localization problem rameter values as an input to theoretical calculations. in the 1970s (Wegner, 1976; Abrahams, Anderson, Lic- The current status of our understanding of the elec- ciardello, and Ramakrishnan, 1979). As we discuss in tronic structures of transition-metal compounds is Sec. II.F, it is now established that the scaling concept heavily based on spectroscopic methods that have been and renormalization-group analysis are useful especially developed in the last decade. A very useful approach for in 2D. The scaling theory was also developed for under- strongly correlated transition-metal compounds has standing the transition between a Mott insulator and a been the configuration-interaction (CI) treatment of superfluid of single-component bosons (see Sec. II.F.12). metal-ligand cluster models, in which correlation and In the case of transitions between metals and Mott insu- hybridization effects within the local cluster are quite lators in pure systems, a scaling theory has also been accurately taken into account. On the other hand, first- formulated recently and several new features of the MIT principles calculations have also greatly contributed to clarified (Imada, 1994c, 1995b). We discuss this recent our understanding of electronic structures, especially to achievement in Sec. II.F. our estimates of the parameter values. In Sec. III.A, the The difficulty of quasiparticle descriptions and mean- local description of electronic structure based on the field approximations of electrons appear most seriously configuration-interaction picture of a metal-ligand clus- in non Fermi . One-dimensional interacting sys- ter model is described. Basic parameters that character- tems showing the Tomonaga-Luttinger liquid behavior ize the electronic structure, namely, the charge-transfer provide a prototype of non Fermi liquids due to strong- energy ⌬, the Coulomb repulsion U [see Eq. (1.1)], etc., correlation effects. Its basic properties are summarized are introduced together with the important concept of in Sec. II.G.1. To explain several anomalous properties Mott-Hubbard-type versus charge-transfer-type insula- of high-Tc cuprates, the marginal Fermi-liquid theory tors. was proposed on phenomenological grounds. It is dis- In Sec. III.B, methods used to derive those param- cussed in Sec. II.G.2. eters are described: They are spectroscopic methods, Section II.H is devoted to recent attempts to describe first-principles methods, and more classical methods more complicated situations. Transition-metal com- based on the ionic point-charge model. The spectro- pounds have rich structures of phases due to spin and scopic methods are largely based on the configuration- orbital fluctuations. Among them, the interplay of spin interaction cluster model. Systematic variations of the and orbital ordering is the subject of Sec. II.H.1. Due to model parameters have been deduced and employed to the interplay of these two fluctuations, several predict physical properties of the transition-metal com- d-electron systems have phases of ferromagnetic and an- pounds. The first-principles methods are based on the tiferromagnetic metals. In particular, Mn and Co com- local-density approximation. The LDA with certain con- pounds show the MIT accompanied by the appearance straints (constrained LDA method) is used to calculate of ferromagnetic metals due to strong Hund’s-rule cou- U and other parameters. Fits of first-principles band pling. This is discussed in Sec. II.H.2. Other complexities structures to the tight-binding model Hamiltonian give such as the effects of charge ordering are also discussed such parameters as the t and ⌬. in Sec. II.H.3. In Sec. III.C we describe how to control the electronic parameters of the model for real materials systems. Two C. Remarks on material systematics (Introduction fundamental parameters of a strongly correlated elec- to Sec. III) tron system are the one-electron bandwidth W (or hop- ping interaction t of the conduction particle) and the In order to study Mott transitions and associated phe- band filling n (or doping level). The conventional nomena in real materials, microscopic models such as method of bandwidth control (BC) is application of ex- the Hubbard model [Eq. (1.1)] and its extended versions ternal or internal (chemical) pressure. In Sec. III.C.1 we are necessary to formulate and to solve problems. Those show some examples of the metal-insulator phase dia- models include parameters that are to be determined so gram afforded by such bandwidth control. As a typical as to reproduce experimentally measured physical prop- example of chemical pressure, we describe the general erties of real systems as closely as possible. For example, relation between the ionic radii of the composing atomic the Hubbard model has the parameters U and t while element (or the so-called tolerance factor) and the the d-p model [Eq. (2.1)] also has ␧dϪ␧pϵ⌬. There are metal-oxygen-metal bond angle in a perovskite-type more parameters when the degeneracy of the d orbitals structure. The d-electron hopping interaction t is medi- has to be taken into account. In order to check the va- ated by the oxygen 2p state (supertransfer process) and lidity of theoretical ideas through comparison with ex- hence sensitive to the bond angle. This offers the oppor- periment, however, it is often desirable not only to ad- tunity to control W by variation of the perovskite toler- just the model parameters but also to have independent ance factor. The example of RNiO3, is described in that estimates of those parameters a priori. Section III de- section, where R is the rare-earth ion with varying ionic scribes such methods and the results of parameter esti- radius. The band filling or doping level can be controlled

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1047 by modifying the chemical composition, as described in IV.A.5, we describe the electron spectroscopy of Sec. III.C.2, such as introducing extra oxygen or its va- Ca1ϪxSrxVO3, in which the bandwidth control can be cancies and/or alloying the heterovalent ions other than similarly achieved by perovskite distortion. on the tradition metal site. The perovskites and their In Sec. IV.B we describe the features of the metallic analogs are again the most typical for such filling control state derived from the Ti3ϩ- and V3ϩ-based Mott- (FC) and the FC range in the actual pseudocubic and Hubbard insulators by FC-MIT. R1ϪxAxTiO3 (IV.B.1) layered perovskites of 3d transition-metal oxide is de- is a relatively new but prototypical Mott-Hubbard sys- picted. The last subsection (III.C.3) is devoted to a de- tem, in which both BC and FC have been proven to be scription of some homologous series compounds in possible. The FC-MIT on the V3ϩ-based compound is which electronic dimensionality can be controlled to also observed in La1ϪxSrxVO3 (IV.B.2) with perovskite some extent. The cuprate ladder-type compounds and structure, which has long been known but is yet to be the Ruddlesden-Popper series (the so-called layered investigated from the modernized view. perovskite) structures represent the dimensional cross- Concerning the high-temperature superconducting cu- over from 1D to 2D and from 2D to 3D, respectively. prates, there have been numerous publications on their These systematics are helpful in connecting the rather anomalous metallic properties. They are the most im- general theoretical descriptions in Sec. II to specific ex- portant example of the FC-MIT. Here, we focus on the perimental examples in Sec. IV. They are also useful for filling dependence of the normal-state properties in pro- classifying experimental results. totypical hole-doped superconductors (IV.C.1. La2ϪxSrxCuO4 and IV.C.3. YBa2Cu3O6ϩy) and electron- D. Remarks on experimental results of anomalous metals doped ones (IVC.2 R2ϪxCexCuO4). (Introduction to Sec. IV) In Sec. IV.D, we briefly describe the recent attempts to derive the metallic state via carrier doping or FC in Section IV is devoted to experimentally observed fea- the quasi-1D cuprate compounds as well as the ther- tures of correlated metals and related MITs in mally induced BC-MIT in BaVS3. In particular, the d-electron systems (mostly oxides), which are mostly de- search for the FC-MIT in the quasi-1D cuprates have rived from the Mott insulators by bandwidth control recently been fruitful for one of the spin-ladder com- (BC) or filling control (FC). As we mentioned above, pounds, (Sr,Ca)14Cu24O41, and 10 K superconductivity other systems, including the organic and f-electron sys- was found under pressure, a finding which is described tems, show similar MIT properties, but we do not cover in a little more detail in Sec. IV.D.1. them in this article. Concerning the work done before The part of the Coulomb repulsion between electrons the mid-1980s, comprehensive reviews on d-electron on different sites occasionally helps the periodic real- compounds have been given in the literature, e.g., the space ordering of the charge carriers in narrow-band book written by Mott (1990) himself and one in honor of systems, although even without the intersite Coulomb the 80th anniversary of Mott’s birth (Edwards and Rao, repulsion, this real-space ordering may be stabilized due 1985). The review given here would be biased in favor of to the kinetic-energy gain (see Sec. II.H.3). These phe- the late advance on the FC-MITs. nomena are called charge ordering in the d-electron ox- In Sec. IV.A, we try to update the experimental un- ides and are widely observed for systems with simple derstanding of the BC-MIT systems, V2O3 (IV.A.1), NiS fractional filling. A classic example is the Verwey tran- (IV.A.2), NiS2ϪxSex (IV.A.3) and RNiO3 (IV.A.4). The sition in magnetite, Fe3O4 (Sec. IV.E.1), with spinel first two compounds have long been known as the pro- structure. The resistivity jump at TVϷ120 K was as- totypical systems that show the Mott insulator-to-metal signed to the onset of the charge ordering, namely a transition by application of pressure or by chemical sub- real-space ordering of the Fe2ϩ and Fe3ϩ on the spinel B stitution. The MIT induced by some nonstoichiometry site in the ferrimagnetic state, where the spins on the A or by Ti doping should rather be considered as FC-MIT, sites (Fe3ϩ) and on the B sites (Fe2ϩ/Fe3ϩ) direct oppo- yet all the features of the MIT in V2O3 are comprehen- sitely. The magnetic phase-transition temperature is 858 sively described in this section. In the modern view of K, much higher than TV . Thus we can only consider the the Zaanen-Sawatzky-Allen scheme, the insulating charge degree of freedom decoupled from the spin de- states of NiS and NiS2ϪxSex should be classified as gree of freedom upon the charge-ordering transition. In charge-transfer insulators with the charge gap between contrast, versatile phenomena take place in the the filled chalcogen 2p state and Ni 3d upper Hubbard d-electron oxides arising from simultaneous or strongly band, in which the chalcogen 2p state bandwidth or the correlated changes of the charge ordering and spin or- hybridization between the two states can be changed by dering. One such example is the case of La1ϪxSrxFeO3 applying pressure or alloying S site with Se. A similar (Sec. IV.E.2) with xϭ2/3, in which the MIT at 220 K is BC-MIT has been recently recognized in a series of associated with simultaneous charge ordering and anti- RNiO3, R being a rare-earth element from La to Lu, ferromagnetic spin ordering. In the layered (nϭ1) per- and the BC acting by the bond distortion of the ortho- ovskite La2ϪxSrxNiO4 (Sec. IV.E.3) with xϭ1/3 or 1/2, rhombic perovskite. In this case, the nominal valence of the charge-ordered phase shows up at higher tempera- Ni is 3ϩ with 3d7 electron configuration, indicating the tures (220–240 K), accompanying the resistivity jump, insulating phase with Sϭ1/2 in contrast to the former and subsequently antiferromagnetic spin ordering takes Ni-based conventional Mott insulators with Sϭ1. In Sec. place. It is now believed that similar charge-ordering

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1048 Imada, Fujimori, and Tokura: Metal-insulator transitions

phenomena in quasi-2D systems are present in the lay- of the misfit compounds, La1.17ϪxPbxVS3.17 (Sec. ered perovskites of manganites (Sec. IV.E.4) and cu- IV.G.5), which are nonmagnetic in the insulating state prates (Sec. IV.C.1) with specific fillings. In particular, (xϭ0.17) where both electron doping and hole doping the cases of the superconducting cuprates are possible. Section IV.H is devoted to other interest- La2ϪxSrxCuO4 and La2ϪxBaxCuO4, with xϭ1/8, are ing 4d transition-metal oxides, for example Sr2RuO4 known as ‘‘the 1/8 problem:’’ The superconductivity dis- (Sec. IV.H.1). This transition-metal oxide, which is iso- appears singularly at this filling even though 1/8 is near structural with La2ϪxSrxCuO4, was recently found to be the optimum doping level (xϭ0.15) at which Tc is maxi- superconducting below 0.9 K and to show a highly an- mum. The phenomenon accompanies a structural isotropic Fermi-liquid behavior. change, as confirmed in La2ϪxBaxCuO4, and perhaps also spin ordering. II. THEORETICAL DESCRIPTION Sec. IV.F is devoted to a review of double-exchange systems, which are of revived interest due to the large A. Theoretical models for correlated metals and Mott magnetoresistance and other magnetic-field-induced insulators in d-electron systems phenomena. The ferromagnetic interaction between the 1. Electronic states of d-electron systems local d-electron spins is mediated by the transfer of the conduction d electron (double-exchange interaction): The atomic orbitals of transition-metal elements are The itinerant carriers are coupled with the local spin via constructed as eigenstates under the spherical potential the strong on-site exchange interaction (Hund’s-rule generated by the transition-metal ion. When the solid is coupling). The metallic phase is derived by hole doping formed, the atomic orbital forms bands due to the peri- from the parent insulators, e.g., LaMnO3 and LaCoO3. odic potential of atoms. The bandwidth is basically de- In the perovskite manganites (Sec. IV.F.1), the strong termined from the overlap of two d orbitals on two ad- Hund’s-rule coupling works between the eg electron and jacent transition metals each. The overlap comes from the t2g local spin (Sϭ3/2). The eg electrons are strongly the tunneling of two adjacent so-called virtual bound correlated, which gives rise to various competing inter- states of d orbitals. Because of the relatively small ra- actions with the double-exchange interaction. In particu- dius of the wave function as compared to the lattice con- lar, we shall see that competing interactions, such as stant in crystals, d-electron systems have in general electron-lattice (Jahn-Teller), superexchange, and smaller overlap and hence smaller bandwidths than al- charge-ordering interactions, cause a variety of kaline metals. In transition-metal compounds, the over- magnetic-field-induced metal-insulator phenomena in lap is often determined by indirect transfer between d the manganites with controlled bandwidth and filling. orbitals through ligand p orbitals. This means that the The attempt at dimension control in manganites is re- bandwidth is determined by the overlap (in other words, viewed, taking examples of layered perovskite structures hybridization) of the d wave function at a transition- (nϭ2; Sec IV.F.2). Hole-doped cobalt oxides with the metal atom and the p wave function at the adjacent perovskite structure (Sec. IV.G.4) are other examples of ligand atom if the ligand atoms make bridges between double-exchange ferromagnets, although the itinerancy two transition-metal atoms. Because of this indirect of the conduction electrons is considerably larger than in transfer through ligand atomic orbitals, the d bandwidth the manganites. becomes in general even narrower. Another origin of The systems associated with gaps for both charge and the relatively narrow bandwidth in transition-metal spin are conceptually difficult to distinguish from the compounds is that 4s and 4p bands are pushed well band insulators, as discussed above. However, some above the d band, where screening effects by 4s and 4p d-electron compounds are considered to be strongly cor- electrons do not work well. This makes the interaction related insulators with a spin gap and undergo the tran- relatively larger than the bandwidth. In any case, be- sition to correlated metals usually by increasing tem- cause of the narrow bandwidth, the tight-binding models perature or by applying pressure (BC-MIT). Some constructed from atomic Wannier orbitals provide a examples are given in Sec. IV.G. FeSi (Sec. IV.G.1) has good starting point. For historic and seminal discussions been pointed out to show similarities to a Kondo insu- on this point, readers are referred, to the textbook ed- lator. The nonmagnetic ground state of VO2 (Sec. ited by Rado and Suhl (1963), particularly the articles by IV.G.2) arises from the dimerization of V sites. Never- Herring (1963) and Anderson (1963a, 1963b). theless, the Cr-doped compound takes the structure of a The bands that are formed are under the strong influ- homogeneous stack but remains insulating, i.e., is a Mott ence of anisotropic crystal fields in . Because the insulator. Thus the spin-dimerized state in VO2 may be 3d orbital has the total angular momentum Lϭ2, it has viewed as a spin Peierls state and the compound under- fivefold degeneracy (Lzϭ2,1,0,Ϫ1,Ϫ2) for each spin goes an insulator-metal transition with increasing tem- and hence a total of tenfold degeneracy including spins. perature. LaCoO3 (Sec. IV.G.4) undergoes a spin-state This degeneracy is lifted by the anisotropic crystal field. transition and the ground state corresponds to the low- In transition-metal compounds, a transition-metal atom spin (Sϭ0) state. The thermally induced MIT in this is surrounded by ligand atoms to help in the formation compound has recently been revisited and some uncon- of a solid through the increase in cohesive energy by ventional features arising from electron correlation have covalent bonds of the two species. Because the ligand been found. We further consider the recent investigation atoms have a strong tendency towards negative valence,

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1049

FIG. 2. Crystal-field splitting of 3d orbitals under cubic, te- tragonal, and orthorhombic symmetries. The numbers cited near the levels are the degeneracy including spins. the crystal field of electrons in the direction of the ligand atom is higher than in other directions. Figure 2 shows an example of the crystal field splitting, where the cubic lattice symmetry leads to a higher energy level of four- fold degenerate eg (or d␥) orbital and sixfold degenerate lower orbitals, t2g (or d␧). When a transition-metal atom is surrounded by ligand atoms with an octahedron con- figuration, the eg orbital has anisotropy with larger am- plitude in the direction of the principle axes, namely, toward neighboring ligand atoms. The basis of these or- FIG. 3. Examples of configurations for transition-metal 3d or- bitals which are bridged by ligand p orbitals. bitals may be expanded by dx2Ϫy2 and d3z2Ϫr2 orbitals. On the other hand, the t2g orbital has anisotropy with larger amplitude of the wave function toward other di- transition-metal element is changed from Sc to Cu. This rections and may be represented by dxy , dyz , and dzx is mainly because the positive nuclear charge increases orbitals. For other lattice structures with other crystal with this change, which makes the chemical potential of symmetries, such as tetragonal or orthorhombic as in the d electrons lower and closer to the p orbital. In fact, in case of 2D perovskite structure, similar crystal field split- the high-Tc cuprates, the 2p␴ orbital has a level close to ting appears. In the case of tetrahedral surroundings of the 3dx2Ϫy2 orbital of Cu. This tendency, as well as the ligand ions, eg orbitals lie lower than t2g , in contrast to larger overlap of the eg wave function with the ligand p␴ cubic symmetry or octahedron surroundings. Readers orbital for geometric reasons, causes a strong hybridiza- are referred to Sec. III for more detailed discussions of tion of the eg and ligand p bands in the late transition each compound. metals. Therefore, to understand low-energy excitations In general, the relevant electronic orbitals for low- on a quantitative level, we have to consider these strong energy excitations transition-metal compounds with hybridization effects. In contrast, for light transition- light transition-metal elements are different from those metal oxides, the oxygen p level becomes far from the with heavy ones. In the compounds with light transition- 3d orbital and additionally the overlap of t2g and p or- metal elements such as Ti, V, Cr,...,only a few bands bitals is weak. Then the oxygen p band is not strongly formed from 3d orbitals are occupied by electrons per hybridized with 3d band at the Fermi level, and the for- 2Ϫ atom. Therefore the t2g orbital (more precisely, the t2g mal valence of oxygen is kept close to O . (More cor- band under the periodic potential) is the relevant band rectly speaking, the oxygen p band is nearly full. How- for low-energy excitations in the case of the above- ever, the real valence of oxygen itself may be larger than mentioned octahedron structure because the Fermi level Ϫ2 because the oxygen p band is hybridized with 3d, crosses bands mainly formed by t2g orbitals. By contrast, 4s, and 4p orbitals.) Consequently the contribution in transition-metal compounds with heavy transition- from the oxygen p band to the wave function at the metal elements such as Cu and Ni, the t2g band is fully Fermi level may be ignored in the first stage. When the occupied far below the Fermi level, and low-energy ex- ligand atoms are replaced with S and Se, in general, the citations are expressed within the eg band, which is level of the 2p band becomes higher and can be closer formed mainly from eg atomic orbitals. If degenerate t2g to the d band. This systematic change can be seen, for or eg orbitals are filled partially, it generally leads again example, in NiO, NiS, and NiSe, as we shall see in Secs. to degeneracy of the ground state, which frequently in- III.A and IV.A. duces the Jahn-Teller effect to lift the degeneracy. The effect is in general large Another important difference between light and when two electrons with up and down spins each occupy heavy transition-metal compounds is the level of ligand the same atomic d orbital of a transition-metal atom. p orbitals. For example, in the transition-metal oxides, The Coulomb repulsion energy of two electrons at the the levels of the relevant 3d orbitals and oxygen 2p␴ same atomic orbital is determined by the spatial exten- orbitals, illustrated in Fig. 3, become closer when the sion of the orbital. The Coulomb energy thus deter-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1050 Imada, Fujimori, and Tokura: Metal-insulator transitions

mined is relatively large for the d orbital as compared to tron number ͗ni␴͘ is controlled at ͗ni␴͘ϭ1/2. For the the small bandwidth. Of course, since the Coulomb in- nearest-neighbor Hubbard model on a hypercubic lat- teraction is long ranged, the interatomic Coulomb inter- tice, the band structure of the noninteracting part is de- action up to the screening radius also has to be consid- scribed as ered in realistic calculations. † tϭ͚ ␧0͑k͒ck␴ck␴ , (2.3a) H k␴ 2. Lattice fermion models

Much progress has been made in our theoretical un- ␧0͑k͒ϭϪ2t cos kv , (2.3b) ϭ͚x y derstanding of d-electron systems through the tight- v , ,¯ binding Hamiltonian. In this section we introduce sev- where we have taken the lattice constant to be unity. eral lattice fermion systems derived from the tight- Here, the Fourier transform of the electron operator is binding approximation, for use and further discussion in introduced as later sections. † ik•rj † Considering all the above aspects of electronic wave ck␴ϭ͚ e cj␴ . (2.4) functions in d-electron systems, we obtain several sim- j plified tight-binding models. The most celebrated and The spatial coordinate of the j site is denoted by rj .At simplified model is the Hubbard model (Anderson, half-filling under electron-hole symmetry, the Fermi 1959; Hubbard, 1963; Kanamori, 1963) defined in Eqs. level lies at ␮ϭ␧Fϭ0. Because the band is half-filled at (1.1a)–(1.1d). The kinetic-energy operator in (1.1b) ␮ϭ0, the appearance of the insulating phase is clearly Ht is obtained from the overlap of two atomic Wannier or- due to the correlation effect arising from the term U . H bitals, ␸i␴(r) on site i and ␸j␴(r) on site j as Although both the Mott insulating state and the MIT (metal-insulator transition) are reproduced in the Hub- 1 bard model, many aspects on quantitative levels with (tϭ dr ␸* ͑r͒ ٌ2␸ ͑r͒, (2.1 ͵ i␴ 2m j␴ rich structure have to be discussed by introducing more complex and realistic factors. To discuss the charge- where m is the electron mass and the Planck constant ប ordering effect, at least the near-neighbor interaction ef- is set to unity. The on-site term U in (1.1c) describes fect should be considered as the Coulomb repulsion of two electronsH at the same site, ϭ ϩ , (2.5a) as derived from H HH HV e2 ϭV n n , (2.5b) Uϭ dr drЈ ␸i*␴͑r͒␸i␴͑r͒ ␸i*Ϫ␴͑rЈ͒␸iϪ␴͑rЈ͒. V ͚ i j ͵ ͉rϪrЈ͉ H ͗ij͘ (2.2) Cleary this Hamiltonian neglects multiband effects, a niϭ ni␴ . (2.5c) ͚␴ simplification that is valid in the strict sense only when the atom has only one s orbital, as in hydrogen atoms. For light elements of transition metal, such as V and When this model is used as a model of d-electron sys- Ti, the Fermi level is on the t2g bands, which are three- tems, it implicitly assumes that orbital degeneracy is fold degenerate under the cubic crystal field with pos- lifted by the strong anisotropic crystal field so that rel- sible weak splitting of this degeneracy under the Jahn- evant low-energy excitations can be described by a Teller distortion in the perovskite structure. When the e single band near the Fermi level. It also assumes that the Fermi level is on the g bands, as in Ni and Cu com- pounds, the degeneracy is twofold in the absence of the ligand p band in transition-metal compounds is far from Jahn-Teller distortion. If the Jahn-Teller splitting is the relevant d band or that they are strongly hybridized weak, we have to consider explicitly three orbitals, d , to form an effective single band. This Hamiltonian also xy d , and d , for the t band and two orbitals, d 2 2 neglects the intersite Coulomb force. The screening ef- yz zx 2g x Ϫy and d3z2Ϫr2, for the eg band in addition to the spin de- fect makes the long-range part of the Coulomb force generacy. This leads to the degenerate Hubbard model exponentially weak, which justifies neglect of the Cou- denoted by lomb interaction far beyond the screening radius. How- ϭ ϩ ϩ ϩ , (2.6a) ever, ignoring the intersite interaction in the short- HDH HDt HDU HDV HDUJ ranged part of the Coulomb force is a simplification that ␯,␯Ј † sometimes results in failure to reproduce important fea- DtϭϪ ͚ k͑tij ci␴␯cj␴␯ЈϩH.c.͒, (2.6b) tures, such as the charge-ordering effect, as we shall dis- H ͗ij͘ cuss in Sec. II.H. In the literature, the electron hopping ␴,␯,␯Ј term t is often restricted to the sum over pairs of nearest-neighborH sites ͗ij͘. DUϭ ͚ ͑1Ϫ␦␯␯Ј␦␴␴Ј͒U␯␯Јni␴␯ni␴Ј␯Ј , (2.6c) H i␯␯Ј In spite of these tremendous simplifications, the Hub- ␴,␴Ј bard model can reproduce the Mott insulating phase with basically correct spin correlations and the transition ␯␯Ј DVϭ ͚ Vij ni␴␯nj␴Ј␯Ј , (2.6d) between Mott insulators and metals. The Mott insulat- H ␴,␴Ј ing phase appears at half-filling where the average elec- ␯,␯Ј͗ij͘

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1051

ϭϪ J 1Ϫ␦ ͒c† c c† c DUJ ͚ 0␯␯Ј͓͑ ␯␯Ј i␯␴ i␯␴Ј i␯Ј␴Ј i␯Ј␴ H ␯␯Ј i␴␴Ј Ϫ 1Ϫ␦ ͒ 1Ϫ␦ ͒c† c† c c ͑ ␯␯Ј ͑ ␴␴Ј i␯Ј␴Ј i␯Ј␴ i␯␴ i␯␴Ј͔ (2.6e) where ␯ and ␯Ј describe orbital degrees of freedom. DUJ is the contribution of the intrasite exchange inter- Haction with e2 J ϭ dr drЈ ␸␯͑r͒␸␯Ј͑r͒ ␸␯͑rЈ͒␸␯Ј͑rЈ͒, (2.7) 0␯␯Ј ͵ i i ͉rϪrЈ͉ i i where ␸ is taken to be real. The intersite exchange term is neglected here for simplicity. The term DUJ gener- ates a Hund’s-rule coupling, since two electronsH on dif- ferent atomic orbitals feel higher energy for opposite spins due to the first term. In this orbitally degenerate FIG. 4. Examples of various second-order processes of two- model, in addition to spin correlations, we have to con- site systems with the sites i and j. The orbitals are specified by sider orbital correlations that can lead to orbital long- ␯ and ␯Ј. The upper, middle, and lower states are the initial, intermediate, and final states in the second-order perturbation range order. The term DU represents the intra-orbital Coulomb energy U Has well as the interorbital one expansion in terms of t/U: (a) off-diagonal contribution to the ␯,␯ spin-exchange process for the case without orbital degeneracy; U with ␯Þ␯Ј for the on-site repulsion. The intersite ␯,␯Ј (b) diagonal contribution for the case without orbital degen- repulsion is given in . HDV eracy; (c) twofold orbital degeneracy yielding off-diagonal or- For the case with orbital degeneracy as in Eq. (2.6a), bital exchange process; (d) same situation as (c) yielding both the effective Hamiltonian may be derived from the or- orbital and spin exchange. bital exchange coupling in analogy with the spin ex- change coupling Eq. (2.12) (Castellani et al., 1978a, Therefore the Hamiltonian becomes highly anisotropic 1978b, 1978c; Kugel and Khomskii, 1982). However, the for ␶ operators. A general form of the exchange Hamil- orbital exchange coupling has important differences tonian in the strong-coupling limit in this case is given by from the spin exchange. One such difference is that the orbital exchange may have large anisotropy. z z z ϩϪ ϩ Ϫ Ϫ ϩ ϭ͚ ͓Si•Sj͕JsϩJs␶␶i ␶j ϩ2Js␶ ͑␶i ␶j ϩ␶i ␶j ͒ As in the derivation of the usual superexchange inter- H ͗ij͘ action for the single-orbital system, we can derive the ϩ2Jϩϩ͑␶ϩ␶ϩϩ␶Ϫ␶Ϫ͒ϩ2Jzx͑␶ϩϩ␶Ϫ͒ strong-coupling Hamiltonian by introducing pseudospin s␶ i j i j s␶ j j representation by ␶ for the orbital degrees of freedom zy ϩ Ϫ z z x x i Ϫ2iJs␶ ͑␶j Ϫ␶j ͒ϩK␶s␶i ϩK␶s␶i ͖ in addition to the Sϭ1/2 operator S for the spins (␶ has i i x z z ϩϪ ϩ Ϫ Ϫ ϩ zx z ϩ Ϫ the same representation as the spin-1/2 operators for ϩJ␶␶i ␶j ϩ2J␶ ͑␶i ␶j ϩ␶i ␶j ͒ϩ2J␶ ␶i ͑␶j ϩ␶j ͒ † doubly degenerate orbitals). More explicitly, ␶i Ϫ2iJzy␶z͑␶ϩϪ␶Ϫ͒ϩ2Jϩϩ͑␶ϩ␶ϩϩ␶Ϫ␶Ϫ͒ † Ϫ † z 1 † † ␶ i j j ␶ i j i j ϭc ci , ␶ ϭc ci and ␶ ϭ 2 (c ci Ϫc ci ), re- i␯ ␯Ј i i␯Ј ␯ i i␯ ␯ i␯Ј ␯Ј z z x x spectively, for doubly degenerate orbits. In the simplest ϩK␶␶i ϩK␶␶i ͔. (2.8) case of twofold degenerate orbitals with a single elec- Here, the parameters depend on the direction of ជi Ϫជj .It tron per site on average, the Hilbert space of the two- should be noted that XY-type or Ising-type anisotropy site problem is expanded by four states, as shown in Fig. for the spin exchange easily occurs, and this anisotropy 4, when the doubly occupied site is excluded due to a depends on the pair (i,j). When the system is away from strong intrasite Coulomb interaction. Here we consider the integer filling or the on-site interaction is not large the transfer of two eg orbitals between two sites located enough to make the system Mott insulating, the transfer along the x or y directions. The transfers between two 2 2 2 2 term has to be considered explicitly. As we saw above, x Ϫy orbitals, tx1 , between two 3z Ϫr orbitals, tx2 , 2 2 2 2 the transfer itself is also highly anisotropic, depending and between x Ϫy orbital and 3z Ϫr orbital, tx12 , re- on the orbital. For example, the transfer for the x2Ϫy2 spectively, satisfy tx1Ͼtx12Ͼtx2 while tz2ӷtz1 and tz12 orbital is in general large in the x and y directions but ϭ0 are expected along the z direction. In fact, because small in the z direction. The opposite is true in the 3z2 of the anisotropy of the d wave function, these are Ϫr2 orbital. This makes for complicated orbital- scaled by a single parameter as dependent band structure in contrast with the spin de- 3 1 ) pendence. Another complexity is that the orbital ex- t ϭ t , t ϭ t , t ϭϪt ϭϪ t , x x1 4 0 x2 4 0 x12 y12 4 0 change coupling of the bonds in the direction is large for dxy , dzx ,ordx2Ϫy2 orbitals, while it is small for the tz1ϭtz12ϭ0, and tz2ϭ1. exchange including dyz or d3z2Ϫr2 orbital. For the bonds

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1052 Imada, Fujimori, and Tokura: Metal-insulator transitions

in other directions, y and z, different combinations have configuration, where three t2g bands as well as the large exchange coupling. In addition, orbital ordering d3z2Ϫr2 band are fully occupied, whereas the dx2Ϫy2 strongly couples with magnetic ordering. The coupling band is half filled. This makes the single-band descrip- of orbital and spin degrees of freedom is apparently dif- tion more or less valid because the low-energy excitation ferent from the usual spin-orbit coupling because it is a can be described only through the isolated dx2Ϫy2 band. nonrelativistic effect. For d-electron systems, the usual This is the reason why the Hubbard model can be a spin-orbit interaction is relatively weak as compared to good starting point for discussing physics in the cuprate rare-earth compounds. In general, in systems with or- superconductors. However, another feature of heavy bital degeneracy, we have to consider couplings of or- transition-metal oxides is the strong hybridization effect bital occupancy to the Jahn-Teller distortion, quadru- of the d orbital and the oxygen p orbital. Because the pole interaction, and spin-orbit interaction. Orbital conduction network is constructed from an oxygen 2p degeneracy also yields intersite orbital exchange cou- ␴ orbital and a 3d 2 2 orbital in this case, strong cova- pling as in Eq. (2.8). These will lead to a rich structure of x Ϫy lency pushes the oxygen 2p orbital closer to the Fermi physical properties, as is discussed in Sec. II.H.l. ␴ For Mn and Co compounds, the d electrons may oc- level. The d-p model is the full description of these cupy high-spin states because of Hund’s-rule coupling. 3dx2Ϫy2 and 2p␴ orbitals in the copper oxides where the Hamiltonian takes the form In this case, the t2g bands are occupied by three spin- aligned electrons with additional electrons in the e g dpϭ dptϩ dpUϩ dpV , (2.11a) bands ferromagnetically coupled with t2g electrons due H H H H to Hund’s-rule coupling. This circumstance is sometimes † dptϭϪ ͚ tpd͑di␴pj␴ϩH.c.͒ϩ␧d͚ ndi modeled by the double-exchange model (Zener, 1951; H ͗ij͘␴ i Anderson and Hasegawa, 1955; de Gennes, 1960):

DEϭ tϩ Hund , (2.9a) ϩ␧p͚ npj , (2.11b) H H H j

† tϭϪt͚ ͑ci␴cj␴ϩH.c.͒, (2.9b) H ͗ij͘ dpUϭUdd͚ ndi ndi ϩUpp͚ npi npi , (2.11c) ␴ H i ↑ ↓ i ↑ ↓

ϭϪJ Sជ ជ Hund H͚ i•␴i , (2.9c) dpVϭVpd͚ npindj . (2.11d) H i H ͗ij͘ x y z When the oxygen p level, ␧ , is much lower than ␧ , where Sជ iϭ(Si ,Si ,Si ) is defined from the eg-electron ␴ p d the oxygen p orbital contributes only through virtual operator ci␴ as processes. The second-order perturbation in terms of ϩ x y † Si ϭSi ϩiSi ϭci ci , (2.10a) ␧ Ϫ␧ generates the original Hubbard model. This type ↑ ↓ d p Ϫ x y † of transition-metal oxide in which ␧dϪ␧p is assumed to Si ϭSi ϪiSi ϭci ci , (2.10b) ↓ ↑ be larger than Udd is called a Mott-Hubbard-type com- pound. Because of U Ͻ͉␧ Ϫ␧ ͉, the charge gap in the z 1 † † dd d p Si ϭ ͑ci ci Ϫci ci ͒, (2.10c) Mott insulating phase is mainly determined by U . 2 ↑ ↑ ↓ ↓ dd In contrast, if ͉␧dϪ␧p͉ is smaller than Udd , the charge whereas ␴ជ i represents the localized t2g spin operators. A excitation in the Mott insulating phase is mainly deter- strong Hund’s-rule coupling JH larger than t may lead to mined by the charge transfer type where an added hole a wide region of ferromagnetic metal, as is observed in in the Mott insulator mainly occupies the oxygen p␴ Mn and Co compounds. In Co compounds, a subtle bal- band. The difference between these two cases is sche- ance of low-spin and high-spin states is realized, as we matically illustrated in Fig. 5. The importance of the shall see later in Sec. IV.G.4. More generally, the inter- oxygen p␴ band in this class of material was first pointed play of the Hund’s-rule coupling and exchange couplings out by Fujimori and Minami (1984). This type of com- with strong-correlation effects as well as orbital and pound is called a charge-transfer-type or CT compound Jahn-Teller fluctuations can lead to complicated phase (Zaanen, Sawatzky, and Allen, 1985). diagrams with ferromagnetic and various types of anti- The character of low-energy charge excitations ferromagnetic phases in metals, as well as in the Mott changes from d in the Mott-Hubbard type to strongly insulator, as will be discussed in Secs. II.H.2 and IV.F.1. hybridized p and d in the charge-transfer type. How- In the compounds with heavy transition-metal ele- ever, it may be affected by an adiabatic change in the ments such as Ni and Cu, t2g bands are fully occupied parameter space of ⌬ϭ␧dϪ␧p and U, as long as the with additional eg electrons. In particular, in the high-Tc insulating phase survives at half-filling of d electrons. At cuprates, two-dimensional perovskite structure leads to least in the Mott insulating phase, these two types of highly 2D anisotropy with crystal field splitting of d␥ compounds show similar features described by the bands to the lower orbit, d3z2Ϫr2, and the upper one, Heisenberg model, as is discussed below. In the metallic dx2Ϫy2, due to the Jahn-Teller distortion for the case of phase, one may also expect a similarity when the strong a CuO6 octahedron. The Mott insulating phase of the hybridization of the d and ligand p band forms well 9 high-Tc cuprates as in La2CuO4 is realized in the Cu d separated bonding, nonbonding, and antibonding bands

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1053

for the d-p model. The exchange between two d orbit- als, 1 and 2, in the presence of the ligand p orbital 3, is indeed derived from the direct exchange between the d orbital 1 and the neighboring p orbital 3, which is hy- bridized with the other d orbital 2 or vice versa. This generates a ‘‘superexchange’’ between 1 and 2 in the order shown above (Anderson, 1959). Even for doped systems, a second-order perturbation in terms of t/U in the Hubbard model leads to the so-called t-J model (Chao, Spalek, and Oles, 1977; Hirsch, 1985a; Anderson, 1987),

† t-JϭϪ ͚ Pd͑tijci␴cj␴ϩH.c.͒PdϩJ͚ Si•Sj , H ͗ij͘ ͗ij͘ ␴ (2.13)

where Pd is a projection operator to exclude the double occupancy of particles at the same site. To derive Eq. (2.13), we neglect the so-called three-site term given by

† tЈϪJϰ ͚ ci␴͑Sl͒␴␴Јcj␴Ј (2.14) H ͗ilj͘ from the assumption that the basic physics may be the same. However, this is a controversial issue, since it ap- pears to make a substantial difference in spectral prop- erties (Eskes et al., 1994) and superconducting instability FIG. 5. Schematic illustration of energy levels for (a) a Mott- (Assaad, Imada, and Scalapino, 1997). It was stressed Hubbard insulator and (b) a charge-transfer insulator gener- that the t-J model (2.13) represents an effective Hamil- ated by the d-site interaction effect. tonian of the d-p model (2.11a) by extracting the singlet nature of a doped p hole and a localized d hole coupled and the Fermi level lies near one of them. However, it with it (Zhang and Rice, 1988). Of course, for highly has been argued that some low-energy excitations dis- doped system far away from the Mott insulator, it is play differences between the Mott-Hubbard and CT questionable whether Eq. (2.13) with doping- types (Emery, 1987; Varma, Schmitt-Rink, and Abra- independent J can be justified as the effective Hamil- hams, 1987). With a decrease in ⌬, the MIT may occur tonian. even at half-filling. This is probably the reason why Ni compounds become more metallic when one changes the ligand atom from oxygen, sulfur to selenium. B. Variety of metal-insulator transitions Low-energy excitations of the Mott insulating phase and correlated metals in transition-metal compounds are governed by the Kramers-Anderson superexchange interaction, in which In order to discuss various aspects of correlated met- only collective excitations of spin degrees of freedom als, insulators, and the MIT observed in d-electron sys- are vital (Kramers, 1934; Anderson, 1963a, 1963b). The tems, it is important first to classify and distinguish sev- second-order perturbation in terms of t/U in the Hub- eral different types. To understand anomalous features bard model, as in Figs. 4(b) and 4(c), or the fourth-order in recent experiments, we must keep in mind the impor- perturbation in terms of tpd /͉␧dϪ␧p͉ or tpd /Udd in the tant parameter, dimensionality. We should also keep in d-p model yield the spin-1/2 Heisenberg model at half mind that both spin and orbital degrees of freedom play band filling of d electrons: crucial roles in determining the character of the transi- tion. Metal-insulator transitions may be broadly classi- ϭJ S S , (2.12) fied according to the presence or the absence of symme- HHeis ͚ i• j ͗ij͘ try breaking in the component degrees of freedom on ϩ x y † Ϫ x y † where Si ϭSi ϩiSi ϭci ci , Si ϭSi ϪiSi ϭci ci and both the insulating and the metallic side, because differ- z 1 ↑ ↓ ↓ ↑ Si ϭ 2 (ni Ϫni ). Here J is given in the perturbation ex- ent types of broken-symmetry states cannot be adiabati- pansion as↑ Jϭ↓4t2/U for the Hubbard model and cally connected. Here we employ the term ‘‘component’’ 4 to represent both spin and orbital degrees of freedom in 8tpd d-electron systems. Symmetry breaking in Mott insula- Jϭ 2 ͉͑␧dϪ␧p͉ϩVpd͒ ͉͑␧dϪ␧p͉ϩUpp͒ tors as well as in metals is due to the multiplicity of the particle components. For example, for antiferromagnetic 4t4 pd order, the symmetry of multiple spin degrees of freedom ϩ 2 ͉͑␧dϪ␧p͉ϩVpd͒ Udd is broken.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1054 Imada, Fujimori, and Tokura: Metal-insulator transitions

The Mott insulator itself in the usual sense is realiz- serts zero entropy as the normal condition of the ground able only in a multicomponent system. This is because state. The total singlet ground state of the Mott insulator the Mott insulating state is defined when the band is at is observed in many spin-gapped insulators. Examples integer filling but only partially filled, which is possible can be seen in spin Peierls insulators such as CuGeO3, only in multicomponent systems. In this context, the Peierls insulators with dimerization such as in TTF- Mott insulator, in the sense discussed in the Introduc- TCNQ and DCNQ salts, one-dimensional spin-1 chains tion, is the insulator caused by intrasite Coulomb repul- as in NENP, ladder systems (for example sion, as in the Hubbard model. This definition is gener- SrnϪ1CunO2nϪ1), two-dimensional systems as in CaV4O9 alized in the last part of this section to allow intersite and the Kondo insulator. The appearance of spin gaps in Coulomb repulsion as the origin of generalized Mott in- these cases may be understood in a unified way from the sulator. For the moment, we restrict our discussion to geometric structure of the underlying lattice (see, for the usual ‘‘intrasite’’ Mott insulator, as in the Hubbard example, Hida, 1992; Katoh and Imada, 1993, 1994). model. When intrasite repulsion causes the Mott insula- Here, the spin gap ⌬s is defined as the difference be- tor, energetically degenerate multiple states coming tween ground-state energies for the singlet and the trip- from a multiplicity of components are only partially oc- let state: cupied by electrons because of strong mutual intrasite repulsion of electrons. This origin of an insulator does N N N N ⌬ ϵE ϩ1, Ϫ1 ϪE , , (2.15) not exist when the multiplicity of states on a site is lost. s gͩ 2 2 ͪ gͩ 2 2 ͪ Although a multiplicity of components is necessary to realize a Mott insulator, it does not uniquely fix the be- havior of the component degrees of freedom. Namely, where Eg(N,M) is the total ground-state energy of N the insulator itself may or may not be a broken- electrons with up spin and M electrons with down spin. symmetry state of the component. To be more specific, The charge gap ⌬c for the hole doping into the singlet consider, for example, the spin degree of freedom as the ground state is given by component. In the antiferromagnetic or the ferromag- netic state, the spin-component symmetry is broken and 1 N N N N hence the spin entropy is zero in the ground state, as it ⌬ ϵ E , ϪE Ϫ1, Ϫ1 . (2.16) c 2 g 2 2 g 2 2 should be. Because of the superexchange interaction, ͫ ͩ ͪ ͩ ͪͬ antiferromagnetic long-range order is the most widely observed symmetry-broken state in the Mott insulator; When the spin excitation has a gap and the orbital exci- we shall see many examples in Sec. IV. tation is not considered, this Mott insulator has both When continuous symmetry is broken, as in SU(2) spin and charge gaps. In this it is actually similar to a symmetry breaking of the isotropic Heisenberg model, band insulator, in which spin and charge have the same only the Goldstone mode of the component, such as the electron-hole excitation gap. Although the spin gap can spin wave, survives at low energies accompanied by a be much smaller than the charge gap due to collective gap in the charge excitation. We show in Sec. II.E.5 that spin excitations in the above correlated insulators, in this this low-energy excitation of the Goldstone mode is nei- category of spin gapped insulators, it is hard to draw a ther relevant nor singular at the MIT point. Here the definite boundary between the band insulator and the SU(2) symmetry is defined as the rotational symmetry of Mott insulator. Indeed, they may be adiabatically con- the Pauli matrix Sx , Sy , and Sz . nected by changes in the parameters when one concen- Another possible symmetry breaking may be caused trates only on the insulating phase apart from metals. by the orbital order. For example, because the t2g orbital There is another type of Mott insulator in which nei- is threefold degenerate in the cubic symmetry, the Mott ther symmetry breaking of the components nor a com- insulator may have either staggered orbital order (simi- ponent excitation gap exists. Two well-known examples lar to antiferromagnetic order in the case of spins) or of this category are the 1D Hubbard model at half-filling uniform orbital order (similar to ferromagnetic order) or and the 1D Heisenberg model in which the spin excita- even helical (spiral) and more complicated types of or- tion is gapless due to algebraic decay of antiferromag- der to lift the orbital degeneracy. In general, the orbital netic correlation. These examples indicate the absence order can be accompanied by static lattice anomalies, of magnetic order because of large quantum fluctuations such as staggered Jahn-Teller-type distortion as dis- inherent in low-dimensional systems. To realize this cussed in Sec. II.H.1. From the experimental point of type, low dimensionality appears to be necessary. view, detection of the orbital order is not easy when it In terms of a quantum phase transition, this type of does not strongly couple with the lattice anomaly. When gapless spin unbroken-symmetry insulator appears just orbital and magnetic order take place simultaneously or at the lower critical dimension dil of the Mott insulator take place only partially, interesting and rich physics is for component order, or, in other words, just at the mag- expected, as we discuss in Secs. II.H.1 and IV. netic critical point. As we discuss in detail in Sec. II.F, When the component order is not broken in the below the lower critical dimension of the relevant com- ground state, it implies that a highly quantum mechani- ponent order, dϽdil , insulators with a disordered com- cal state like a total singlet state should be the ground ponent appear while order appears for dϾdil . The di- state to satisfy the third thermodynamic law, which as- mension dϭdil is marginal with gapless spin and

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1055 unbroken-symmetry behavior. The lower critical dimen- tightly connected with the double-exchange mechanism, sion may depend on details of the system under consid- which is made possible by the strong Hund’s-rule cou- eration such as lattice structure and number of compo- pling. Some of theoretical and experimental aspects of nents. For example, if lattice dimerization increases in this class are discussed in Secs. II.H.2 and IV.F, respec- the Heisenberg model, the ground state changes from tively. the antiferromagnetic state to a spin-gapped singlet at a For the case dрdml , order in the diagonal component critical strength of dimerization in 2D (Katoh and is lost. This class is further divided into three cases. In Imada, 1993, 1994). A strongly dimerized lattice in 2D the first case, the component excitation has a gap due to corresponds to dϽdil while, in the case of a uniform formation of a bound state. A typical example is singlet square lattice, dil is below 2. Cooper pairing, where the ground state has an off- In metals near the Mott insulator, a similar classifica- diagonal order, namely, superconducting for dу2, while tion scheme is also useful. Here, we define two types of gap formation only leads to algebraic decay of the pair- component order. One is diagonal, where there is a den- ing correlation at dϭ1 because long-range order is de- sity wave such as the in antiferromag- stroyed by quantum fluctuations. A known example of netic and ferromagnetic phases or an orbital density this class in 1D is the Luther-Emery-type liquid (Luther wave. The other is off-diagonal, in which superconduc- and Emery, 1974; Emery, 1979). Although no example is tivity is the case. Above the lower critical dimension of known to date, the possibility is not excluded in prin- metal for diagonal component order, dml , a diagonal ciple that the superconducting ground state could domi- component order exists, as in antiferromagnetic metals nate or coexist with diagonal component-order in a part and ferromagnetic metals. By contrast, for dрdml , the of the region dϾdml near the Mott insulator. This class diagonal component order disappears due to quantum does not exclude the case of an anisotropic gap with fluctuations. dml may depend on the number of compo- nodes as in d-wave pairing, where the true excitation nents. In principle, the insulating state can be realized gap is lost. through a diagonal component order of an arbitrarily The second case for dрdml is the formation of a large incommensurate periodicity at the Fermi level, if the Fermi surface consistent with the Luttinger sum rule. nesting condition is satisfied. This is indeed the case with This case is possible, for example, when the Fermi-liquid the stripe phase in the Hartree-Fock approximation, as state is retained close to the Mott insulating state. Here, is discussed in Sec. II.D.2 and II.H.3. However, in the the Fermi liquid is defined as an adiabatically continued Hartree-Fock approximation dynamic quantum fluctua- state with the ground state of the noninteracting system, tions are ignored. Usually, in reality, the ordering takes in which the interaction is gradually switched on. Then place only at some specific commensurate filling for a the Fermi surface in momentum space is retained in the such large electron density of the order realizable in sol- Fermi liquid with the same volume as in a noninteract- ids. Incommensurate order is hardly realized due to ing system, which is the statement of the Luttinger theo- quantum fluctuations. When the quantum fluctuation is rem (Luttinger, 1963). The formation of a large ‘‘Fermi large enough, and the impurity-localization effect and surface’’ is not necessarily restricted to the Fermi liquid. the effect of anisotropic gap (in other words, discrete In the case of the Tomonaga-Luttinger liquid in 1D, the symmetry breaking as in the Ising-like anisotropic case) Fermi surface is characterized by singularity of the mo- are not serious, usually metal appears just adjacent to mentum distribution function, and there is a large Fermi the Mott insulator. volume consistent with the Luttinger theorem. If this metal has a diagonal component order, the The third case is the formation of small pockets of most plausible case is the same order with the same pe- Fermi surface. This is basically the same as what hap- riodicity, like the order in the Mott insulating state. In pens in a band insulator upon doping. Therefore no this case, a small pocket of the Fermi surface is expected clear distinction from the doping into the band insulator if the metal is obtained by carrier doping of the Mott is expected. insulator. This may be realized in the antiferromagnetic We summarize three cases of insulators and four cases metallic state near the Mott insulator. Another impor- of metals near the Mott insulator: tant case of this class is that of ferromagnetic metals, as (I-1) insulator with diagonal order of components observed in Mn and Co compounds such as (I-2) insulator with disordered component and compo- (La,Sr)MnO3. Here the symmetry breaking is quite dif- nent excitation gap ferent between the antiferromagnetic order in the Mott (I-3) insulator with disordered component and gapless insulator and the ferromagnetic order in the doped sys- component excitations tem. The appearance of ferromagnetic metals is theo- (M-1) metal with diagonal order of components and retically argued in connection with Nagaoka ferromag- small Fermi volume netism (Nagaoka, 1966), where the is (M-2) metal or superconducting state with the bound- proven when a hole is doped into the Hubbard model at state formation dominated by off-diagonal com- half filling in a special limit of infinite U. Ferromag- ponent order (frequently with component excita- netism is also proven in some special band structures tion gap) with a flat dispersion at the Fermi level. In the above (M-3) metal with disordered component and ‘‘large’’ argument the orbital degrees of freedom are ignored. In Fermi volume consistent with the Luttinger theo- some cases, the formation of the ferromagnetic order is rem

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1056 Imada, Fujimori, and Tokura: Metal-insulator transitions

TABLE I. Rough and simplified sketch for types of metal-insulator transitions in overall phase diagram observed in various compounds. In parentheses, section numbers where discussions may be found in Sec. IV are given.

Insulator I-1 I-2 Metal Component order Component gap

M-1 \ V 2ϪyO 3(IV.A.1) La 1ϪxSr xCoO 3 diagonal Ni 2ϪyS 2ϪxSe x(IV.A.2) (IV.G.4) component-order R1ϪxAxMnO3(IV.F.1) M-2 high-Tc cuprates (Sr,Ca) 14Cu 24O 41 bound state (IV.C) (IV.D.1) (superconductor) BEDT-TTF compounds

M-3 La 1ϪxSr xTiO 3 component disorder (IV.B.1)

(large Fermi RNiO3 La 1ϪxSr xVO 3 volume) (IV.A.3) (IV.B.2) M-4 Kondo component insulator disorder (IV.G) (small Fermi volume)

(M-4) metal with disordered component and ‘‘small’’ ent spin orderings can be realized between insulating Fermi volume as in a doped band insulator and metallic phases in this category. A good example is In d-electron systems experimentally and also in the- the double-exchange system, such as La1ϪxSrxMnO3, oretical models, a variety of MITs are observed. Here where the Mott insulating state at xϭ0 has a type of we classify them in the form ͓I-i M-j͔ (iϭ1,2,3, j antiferromagnetic order with orbital and Jahn-Teller or- ϭ1,2,3,4) where a transition from↔ an insulator of the derings, whereas the metallic phase has ferromagnetic type I-i to a metal of the type M-j takes place. We list order. As for the orbital order, this compound has the below both cases of filling control (FC-MIT) and band- character of ͓I-1 M-3͔, because it is speculated that width control (BC-MIT) and summarize examples of the the orbital order↔ is lost upon metallization. compounds in Table I. ͓I-1 M-2͔: This is the case of high-Tc cuprates at ͓I-1 M-1͔: A typical example is the transition be- low temperatures.↔ Recent studies appear to support that tween↔ an antiferromagnetic insulator and an antiferro- the superconducting phase directly undergoes transition magnetic metal. Examples are seen in V2O3Ϫy (BC-MIT to a Mott insulator (Fukuzumi et al., 1996) by the route or FC-MIT) and in NiS2ϪxSex (BC-MIT). Although the of FC-MIT. No normal metallic phase is seen between Hartree-Fock approximation usually gives this class of superconductor and insulator at zero temperature in the behavior for the MIT, it is important to note that in cuprates. Another example of this type of transition is systems with strong 2D anisotropy, no example of this that of BEDT-TTF compounds (BC-MIT; for a review, class is known to date. This lends support to the idea see Kanoda, 1997). that d is larger than two for the antiferromagnetic or- ͓I-1 M-3͔: This class can be observed for d Ͼd ml ↔ ml der, at least in single orbital systems. Quantum Monte Ͼdil . As we have already discussed, in 3D, Carlo results in 2D for a single band system also support La1ϪxSrxTiO3 (FC-MIT) and RNiO3 may be examples, the absence of antiferromagnetic order in metals. Even although the possible existence of a tiny in 3D, all the examples of antiferromagnetic metals antiferromagnetic-metal region cannot be ruled out. known to date are found when orbital degeneracy exists. When the effect of orbital degeneracies becomes impor- For doped spin-1/2 Mott insulators, antiferromagnetic tant, even for doped spin-1/2 insulators, rather compli- order quickly disappears by metallization upon doping cated behavior may appear, as in (Y,Ca)TiO3 (FC-MIT) or bandwidth widening, and no antiferromagnetic metals where a wide region of orbital ordered insulator is are scarcely found, as is known in La1ϪxSrxTiO3 (FC- speculated. In 2D, the second layer registered phase of 3 MIT) and RNiO3 (BC-MIT), although they have orbital He adsorbed on graphite may also be interpreted as a degeneracy. Therefore dml for antiferromagnetic order Mott insulator (Greywall, 1990; Lusher et al., 1993; without orbital degeneracy can be larger than three, al- Imada, 1995b). Solidification of 3D 3He has been dis- though we do not have a reliable theory. With increasing cussed before in analogy with the Mott transition in the component degeneracy due to orbital degrees of free- Gutzwiller approximation (Vollhardt, 1984). However, 3 dom, dml may become smaller than three. Quite differ- the 3D solid of He should not be categorized as a Mott

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1057

insulator, because commensurability with the underlying for the high-Tc cuprates, ascribed to the more general periodic potential is crucial for a Mott insulator while it anomalous character of metals near the Mott insulator, is absent in 3D 3He. In contrast, the second layer of as discussed in Secs. II.E, II.F, and II.G. We discuss ex- adsorbed 3He on graphite indeed undergoes a transition perimental details of this category in Sec. IV.D.1. between a quantum liquid and a Mott insulator. In this ͓I-2 M-4͔: This is basically the same as the transi- ↔ case, the Mott insulator is formed on the van der Waals tion between a band insulator and a normal metal. periodic potential of the triangular lattice of the 3He Kondo insulators are also supposed to belong to this solid in the first layer. The second-layer Mott insulator class. For d-electron systems, spin excitation gaps or appears to be realized at a commensurate density on a low-spin configurations appear in the insulating phase triangular lattice. The anomalous quantum liquid state in, for example, FeSi, TiO2,Ti2O3, and LaCoO3. Re- of 3He observed upon doping can basically be classified cently, it was found that several other systems also show as belonging to this class, ͓I-1 M-3͔, though the tran- spin-gapped excitations in the insulating phase. Ex- sition is of weakly first order (Imada,↔ 1995b). An advan- amples are spin Peierls compounds such as CuGeO3 tage of the 3He system (FC-MIT) is that one can inves- (Hase et al., 1993), NaV2O5 (Isobe and Ueda, 1996), tigate a prototype system without orbital degeneracies, spin-1 chain compounds such as Y2BaNiO5 (Buttrey impurity scattering, interlayer coupling, or long-range et al., 1990), ladder compounds as already discussed Coulomb interactions. Although the Anderson localiza- above, CaV4O9 (Taniguchi, et al., 1995) and La ϩxVS (Takeda et al., 1994). Carrier doping of tion effect appears to be serious, La1.17ϩxVS3.17 also be- 1.17 3.17 longs to this class (Nishikawa, Yasui, and Sato, 1994; these compounds has not been successful so far, except Takeda et al., 1994). In this compound there exists a in the ladder compounds, and it is not clear whether spin subtlety related to dimerization and spin gap formation excitation becomes gapless or not in these cases upon in the insulating phase, and hence it can also be charac- metallization. ͓I-3 M-3͔: Experimentally this class has not been terized as ͓I-2 M-4͔. The 2D Hubbard model clarified ↔ by quantum Monte↔ Carlo calculations appears to belong observed in the strict sense. In theoretical models, ex- to a theoretical model of this class as we shall see in Sec. amples are 1D systems such as the 1D Hubbard model. II.E. In general, a new universality class of transition is Although it is not Fermi liquid, the singularity of the expected in this category, as we discuss in Sec. II.F. momentum distribution clearly specifies the formation of ‘‘large Fermi volume’’ in the metallic region. ͓I-2 M-1͔: An example is seen in La1ϪxSrxCoO3, ↔ When the electron density is commensurate to the where the low-spin state of LaCoO3 at low temperatures becomes a ferromagnetic metal under doping. number of lattice sites, charge ordering may occur. This ͓I-2 M-2͔: Extensive effort has been made to real- is observed in the cases of Fe3O4,Ti4O7, and several ize this↔ class of transition experimentally, as in doping of perovskite oxides (see Sec. IV). This class of charge- dimerized, spin-1, or ladder systems (Hiroi and Takano, ordering transition has often been called (for example, 1995; McCarron et al., 1988; Siegrist et al., 1988). This is by Mott, 1990) the Wigner transition or the Verwey based on a proposal of the pairing mechanism for doped transition. However, this type of charge ordering usually spin-gapped systems and theoretical results where the takes place only at simple commensurate filling. In the spin gap remains finite and pairing correlation becomes case of d-electron systems, the Wigner crystal in the dominant when carriers are doped into various spin- usual sense of an electron cannot be realized be- gapped 1D systems such as dimerized systems (Imada, cause the electron density is too great. The commensu- 1991, 1993c), spin-1 systems (Imada, 1993c), the frus- rability of the electron filling is crucial for realizing trated t-J model (Ogata et al., 1991), ladder systems charge ordering. Although the intersite Coulomb repul- (Dagotto, Riera, and Scalapina, 1992; Rice et al., 1993, sion helps the charge ordering, the gain in kinetic energy Tsunetsugu et al., 1994; Dagotto and Rice, 1996). These could be enough to stabilize the charge-ordered state. In apparently different systems are continuously connected this sense, a charge-ordered insulator may also be in the parameter space of models (Katoh and Imada, viewed as a generalized Mott insulator where the elec- tron density is kept constant at a commensurate value 1995). Although a ladder system LaCuO2.5 was metal- lized upon doping, the metallic state appears to be real- and it is incompressible. The charge-ordering transition ized from the 3D network by changing the geometry of is discussed in Sec. II.H.3. Another mechanism for the ladder network. Therefore it is questionable whether charge ordering is the Peierls transition when the nest- it can be classified as a doped ladder system. One clear ing condition is well satisfied. In this case, the periodicity origin of the difficulty in realizing superconductivity is can be incommensurate to the lattice periodicity. the localization effect, which must be serious with 1D anisotropy. Doping of a spin-gapped Mott insulator in C. Field-theoretical framework for interacting 2D would therefore be desirable. The recent discovery fermion systems of a superconducting phase in a spin ladder system (Sr,Ca)14Cu24O41 by Uehara et al. (1996) may belong to In this section, a basic formulation for treating inter- this category. At the moment, it is not clear enough acting fermions based on the path-integral formalism is whether this superconductivity is realized by carrier summarized for use in later sections. To show how the doping of a quasi-1D spin-gapped insulator via the formulation is constructed, we take the Hubbard model mechanism described above or by a mechanism like that as an example. Since this section discusses only the basic

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1058 Imada, Fujimori, and Tokura: Metal-insulator transitions formalism needed to understand the content of later sec- where we have used Eq. (2.18). tions, readers are referred to other references (for ex- ample, Negele and Orland, 1988) for a more complete and rigorous description. 2. Grassmann algebra 1. Coherent states The basic algebra of the Grassmann variable ͕␺␣͖ is Fermion Fock space can be represented by a basis of given by Eq. (2.18). In addition, several other relations Slater determinants. However, Fock space can alterna- are summarized here. From the adjoint state of (2.20), tively be represented by a basis of coherent states, as we show below. The coherent states are defined as eigen- ¯ ¯ states of the fermion annihilation operator. This is ex- ͗␺͉ϭ͗0͉exp Ϫ͚ c␣␺␣ ϭ͗0͉͟ ͑1ϩ␺␣c␣͒, (2.23) ͫ ␣ ͬ ␣ pressed as ¯ c␣͉␺͘ϭ␺␣͉␺͘, (2.17) we can define the conjugation of ␺␣ as ␺␣ . From this definition, it is clear that the conjugation operation fol- where c␣ is the fermion annihilation operator with the quantum number ␣, whereas ͉␺͘ is a coherent state with lows: the eigenvalue ␺␣ . When one operates more than one ¯ annihilation operator to ͉␺͘, one gets the commutation ␺␣ϭ␺␣ , (2.24a) relation for ␺␣ from that of c␣ : ¯ ␺␣␺␥ϩ␺␥␺␣ϭ0. (2.18) ␭␺␣ϭ␭*␺␣ , (2.24b) This means that ␺ ’s are not c numbers. Actually, this ␣ ¯ ¯ ¯ commutation relation constitutes the Grassmann alge- ␺␣␺␤␺␥¯ϭ¯␺␥␺␤␺␣ (2.24c) bra. In order to demonstrate that the basic Grassmann al- where ␭ is a complex number with its complex conjugate gebra given below is indeed useful and necessary to rep- ␭*. † resent the Fock space, we first introduce the explicit The operation of c␣ can be associated with the deriva- form of the coherent state. Because ␺␣’s are not c num- tive of Grassmann variables as bers, we are forced to enlarge the original Fock space by including Grassmann variables ␺␣ as coefficients. Any † † † † vector ␾͘ in the generalized Fock space is given by the c␣͉␺͘ϭc␣͑1Ϫ␺␣c␣͒ ͑1Ϫ␺␤c␤͉͒0͘ ͉ ␤͟Þ␣ linear combination of vectors in the original Fock space with Grassmann variables as coefficients: † ץ † † ϭc␣ ͟ ͑1Ϫ␺␤c␤͉͒0͘ϭϪ ͑1Ϫ␺␣c␣͒ ␣␺ץ ␣Þ␤ ͉␾͘ϭ ␺␣͉␸␣͘, (2.19) ͚␣ ץ ϫ ͑1Ϫ␺ c† ͉͒0 ϭϪ ͉␺ , (2.25) where ␺␣ is a Grassmann variable with ͕͉␸␣͖͘ being the ͟ ␤ ␤ ͘ ͘ ␣␺ץ ␣Þ␤ complete set of vectors in the original Fock space. In the generalized Fock space, the fermion coherent state is where the derivative of ␺ is defined in a natural way as given as ␣ ץ † † ¯ ¯ ¯ ͉␺͘ϭexp Ϫ͚ ␺␣c␣ ͉0͘ϭ͟ ͑1Ϫ␺␣c␣͉͒0͘. (2.20) ͓a0ϩa1␺␣ϩa2␺␣ϩa3␺␣␺␣͔ϭa1Ϫa3␺␣ (2.26) ␣␺ץ ␣ ͬ ␣ ͫ For the coherent state ͉␺͘ defined in Eq. (2.20), it is for c number constants a0 , a1 , a2 , and a3 . Note that necessary to require the commutation relations ¯ ¯ ¯ ␺␣)(␺␣␺␣)ϭϪ␺␣ . Equationץ/ץ)␺␣)(␺␣␺␣)ϭϪץ/ץ) † † ␺␣c␤ϩc␤␺␣ϭ0, (2.21a) (2.25) is indeed easily shown from the definitions (2.20) and (2.26). One can verify the conjugation of (2.25): ␺␣c␤ϩc␤␺␣ϭ0. (2.21b) † † ץ Since ␺␣c␣ and ␺␤c␤ commute from Eqs. (2.21a) and (2.21b), we can show that ͗␺͉c␣ϭ ͗␺͉. (2.27) ¯ ␣␺ץ † † c␣͉␺͘ϭ ͑1Ϫ␺␤c␤͒c␣͑1Ϫ␺␣c␣͉͒0͘ ␤͟Þ␣ The overlap of two coherent states is shown from Eqs. (2.20) and (2.27) as † ϭ ͟ ͑1Ϫ␺␤c␤͒␺␣͉0͘ ␤Þ␣ ¯ † ͗␺͉␺Ј͘ϭ͗0͉ ͑1ϩ␺␣c␣͒͑1Ϫ␺␣Јc␣͉͒0͘ ͟␣ † † ϭ ͑1Ϫ␺␤c␤͒␺␣͑1Ϫ␺␣c␣͉͒0͘ϭ␺␣͉␺͘, ␤͟Þ␣ ¯ ¯ ϭ͟ ͑1ϩ␺␣␺␣Ј ͒ϭexp ͚ ␺␣␺␣Ј . (2.28) (2.22) ␣ ͫ ␣ ͬ

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1059

The matrix element of a normal-ordered operator The generalization of this proof to many-particle states † † H(c␣ ,c␣) between two coherent states immediately fol- ͉␺͘ϭ⌸␣(1Ϫ␺␣c␣)͉0͘ is straightforward. lows from Eq. (2.28) as From Eqs. (2.20), (2.23), and (2.33), the trace of an operator A is † ¯ ¯ ͗␺͉H͑c␣ ,c␣͉͒␺Ј͘ϭexp ͚ ␺␣␺␣Ј H͑␺␣ ,␺␣͒. (2.29) Tr Aϭ ␸ A ␸ ͫ ␣ ͬ ͚ ͗ i͉ ͉ i͘ i

Through Eq. (2.29), operators are represented only by ¯ ¯ the Grassmann variables. ϭ ͟ d␺␣d␺␣exp Ϫ͚ ␺␣␺␣ ͵ ␣ ͫ ␣ ͬ To calculate matrix elements of operators as well as trace summations, it is useful to introduce a definite in- ϫ ͗␸i͉␺͗͘␺͉A͉␸i͘ tegral in terms of ␺␣ . Any regular function of ␺␣ may ͚i be expanded as ¯ ¯ f͑␺ ͒ϭa ϩa ␺ (2.30) ϭ ͟ d␺␣ d␺␣exp Ϫ͚ ␺␣␺␣ ␣ 0 1 ␣ ͵ ␣ ͫ ␣ ͬ 2 because of ␺␣ϭ0. Therefore it is sufficient to define ϫ ͗Ϫ␺͉A͉␸i͗͘␸i͉␺͘ ͐d␺␣1 and ͐d␺␣␺␣ . When we assume the same prop- ͚i erties as those of ordinary integrals from Ϫϱ to ϩϱ over functions which vanish at infinity, it is natural to ¯ ¯ define the definite integral so as to satisfy vanishing in- ϭ ͟ d␺␣ d␺␣exp Ϫ͚ ␺␣␺␣ ͗Ϫ␺͉A͉␺͘. ͵ ␣ ͫ ␣ ͬ tegral of a complete differential form: (2.34) A Gaussian integral of a pair of conjugate Grassmann ץ d␺ f͑␺ ͒ϭ0. (2.31) ͵ ␣ ␣ variables is ␣␺ץ

¯ ¯ ¯ ␣␺ ) ␺ , we have ¯ Ϫ␺␣a␺ץ/ץ)Then, because of 1ϭ ␣ ␣ ͵ d␺␣ d␺␣ e ϭ ͵ d␺␣ d␺␣͑1Ϫ␺␣a␺␣͒ϭa. (2.35) d␺ 1ϭ d¯␺ 1ϭ0. (2.32a) ͵ ␣ ͵ ␣ Similarly, a Gaussian integral in matrix form is proven to be The definite integral of ␺␣ can be taken as a constant. n Only for convenience, to avoid annoying coefficients in ¯ ¯ ¯ ¯ d␺␣ d␺␣exp͓Ϫ␺␣A␣␤␺␤ϩ␨␣␺␣ϩ␨␣␺␣͔ later matrix elements, we take this constant to be unity: ͵ ␣͟ϭ1

ϭ A ¯ AϪ1 ¯ ¯ ͑det ͒exp͓␨␣ ␣␤␨␤͔, (2.36) ͵ d␺␣␺␣ϭ ͵ d␺␣␺␣ϭ1. (2.32b) ¯ From the definition of the integrals, we obtain the where ␨␣ and ␨␣ are another pair of Grassmann vari- useful closure relation ables. Using Eqs. (2.29), (2.34), and (2.36), we can cal- culate the partition function and other functional inte- grals in the path-integral formalism. ¯ ¯ ͟ d␺␣ d␺␣exp Ϫ͚ ␺␣␺␣ ͉␺͗͘␺͉ϭ1. (2.33) ͵ ␣ ͫ ␣ ͬ

If we restrict the Hilbert space to single-particle states 3. Functional integrals, path integrals, and statistical with particle number 0 or 1, Eq. (2.33) is proven as mechanics of many-fermion systems In the case of the Hubbard model (1.1a), the matrix ¯ ¯ ͑2.33͒ϭ d␺␣ d␺␣exp͓Ϫ␺␣␺␣͔ element of the Hamiltonian HH in the coherent state ͵ representation reads from Eq. (2.29) as † ¯ ϫ͑1Ϫ␺␣c ͉͒0͗͘0͉͑1ϩ␺␣c␣͒ U ␣ ͓¯␺,␺͔ϭ ϩ Ϫ ␮ϩ , (2.37a) HH Ht HU ͩ 2 ͪ N ¯ ¯ † ϭ d␺␣d␺␣͓͑1ϩ␺␣␺␣͉͒0͗͘0͉Ϫ␺␣c␣͉0͗͘0͉ ͵ ¯ ¯ tϭϪt͚ ͑␺i␴␺j␴ϩ␺j␴␺i␴͒, (2.37b) H ͗ij͘ ¯ ¯ † ϩ͉0͗͘0͉␺␣c␣ϩ␺␣␺␣c␣͉0͗͘0͉c␣͔ ¯ ¯ † UϭU͚ ␺i ␺i ␺i ␺i , (2.37c) ϭ͓͉0͗͘0͉ϩc␣͉0͉͗͘0͉c␣͔ϭ1. H i ↑ ↓ ↓ ↑

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1060 Imada, Fujimori, and Tokura: Metal-insulator transitions

␤ ¯ Ϫi͑k•rϪ␻n␶͒ ϭ͚ ␺i␴␺i␴ , (2.37d) ␺͑k͒ϭ dr d␶ e ␺͑x͒, (2.42a) N i␴ ͵ ͵0

c† c which is obtained by replacing i␴ and i␴ with Grass- 1 mann variables ¯␺ and ␺ , respectively, after proper ␺͑x͒ϭ ei͑k•rϪ␻n␶͒␺͑k͒, (2.42b) i␴ i␴ ␤V ͚ normal ordering. The partition function for a many- k,␻n particle system is where x and k are Euclidean four-dimensional vectors Ϫ␤ ZϭTr e H xϵ(r,␶) and kϵ(k,␻n), with ␻n being the Matsubara 1 frequency ␻nϵ2␲T(nϩ 2 ), nϭ0,Ϯ1,Ϯ2,... . Hereaf- ¯ ¯ Ϫ␤ ϭ ͵ d␺␣ d␺␣exp Ϫ͚ ␺␣␺␣ ͗Ϫ␺͉e H͉␺͘. ter the volume V and the number of sites Ns will have ͫ ␣ ͬ the same meaning for lattice systems. The Fourier- (2.38) transformed form of the action is

Ϫ␤ When one divides e H into L imaginary-time slices ¯ ¯ ¯ (Trotter slices) by keeping the inverse temperature S͓␺,␺͔ϭ͚ Ϫi␻n ͵ dk ␺͑k͒␺͑k͒ϩ ͓␺͑k͒,␺͑k͔͒ . (⌬␶)Lϭ␤ constant and inserts the closure relation Eq. n ͫ H ͬ (2.33) in each interval of slices, using Eqs. (2.28) and (2.43) (2.33), one obtains the standard path-integral formalism in the limit L ϱ in the form 4. Green’s functions, self-energy, and spectral functions → The ␶-dependent thermal Green’s function for sys- L tems with translational symmetry is defined by Zϭ d¯ ͑l͒ d ͑l͒ ϪS ¯ lim ͵ ͟ ␺␣ ␺␣ exp† ͓␺,␺͔‡, (2.39a) L ͫ lϭ1 ͬ ϱ → ͑k,␶͒ϭϪ͵ dr eϪik•r͗Tc͑r,␶͒c†͑0,0͒͘, (2.44) L L G ¯ ¯ S͓␺,␺͔ϭ͚ ͚ lm͓␺,␺͔, (2.39b) lϭ1 mϭ1 S with the time-ordered product T and the thermody- namic average ͗¯͘. The spectral representation of ␦ Ϫ␦ (k,␶)is ¯ ¯͑l͒ ͑m͒ lm lϪ1,m G lm͓␺,␺͔ϭ⌬␶͚ ␺␣ ␺␣ S ␣ ͫ ͫ ⌬␶ ͬ ͑k,␶͒ϵ e␤„FϪEm͑N͒…ϩ„Em͑N͒ϪEn͑NϪ1͒ϩ␮…␶ G ͚mn ϩ [¯␺(l) ,␺(m)]␦ (2.39c) H ␣ ␣ lϪ1,mͬ † ϫ͗⌽m͑N͉͒ck͉⌽n͑NϪ1͒͘ with ␺(0)ϭϪ␺(L) . By defining the functional integra- ␣ ␣ ϫ ⌽ ͑NϪ1͉͒c ͉⌽ ͑N͒ (2.45) tion ͗ n k m ͘ † L for ␶Ͼ0, and for ␶Ͻ0 the same but with ck and ¯ ¯ ͑l͒ ͑l͒ ck exchanged. Here, F is the free energy ͓␺,␺͔ϭ lim ͟ ͟ d␺␣ d␺␣ , (2.40) Ϫ␤ D L ϱ lϭ1 ␣ FϵϪ(1/␤)ln Tr e H and ͕Em(N)͖ are eigenvalues of → a complete set of eigenstates ͕⌽m(N)͖ in an N-electron one obtains the partition function system with the chemical potential ␮. Although we do not explicitly take the summation over N in the grand canonical ensemble with a fixed ␮, it is hereafter implic- ¯ Zϭ ͓¯␺,␺͔eϪS[␺,␺], (2.41a) itly assumed. Because the spectral representation of ͵ D proves to satisfy (k,␶ϩ␤)ϭϪ (k,␶), it can be ex-G G G with the action panded in the Fourier series as

␤ 1 Ϫi␻n␶ ץ ¯ ¯ ͑k,␶͒ϭ ͚ ͑k,␻n͒e , (2.46) S͓␺,␺͔ϭ d␶͚ ␺␣͑␶͒ ␺␣͑␶͒ G ␤ n G ␶ץ ␣ 0͵

␤ ␤ ¯ i␻n␶ ϩ d␶ ͓␺␣͑␶͒,␺␣͑␶͔͒ (2.41b) ͑k,␻n͒ϭ d␶ ͑k,␶͒e . (2.47) ͵0 H G ͵0 G under the constraint ␺␣(0)ϭϪ␺␣(␤). The thermal Green’s function of the free-fermion sys- The Fourier transforms of the Grassmann variables tem is obtained with the help of the Grassmann repre- are sentation as

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1061

␤ 1 † 0͑␰,␶1 ;␩,␶2͒ϵϪ Tr T exp Ϫ d␶ 0 c␰͑␶1͒c␩͑␶2͒ G Z ͫ ͫ ͵0 H ͬ ͬ 0

͐ ͓¯␺,␺͔exp Ϫ͐d␶ d␶ ͓¯␺,␺͔ ␺ ͑␶ ͒¯␺ ͑␶ ͒ D † 3 4 S␶3 ,␶4 ‡ ␰ 1 ␩ 2 ϭϪ␦␰␩ ͐ ͓¯␺,␺͔exp Ϫ d␶ d␶ ͓¯␺,␺͔ D ͫ ͵ 3 4 S␶3␶4 ͬ

2 ͐ ͓¯␺,␺͔exp͓Ϫ͐d␶ d␶ ͓¯␺,␺͔ϩ͐d␶J*͑␶͒␺͑␶͒ϩ¯␺͑␶͒J͑␶͔͒ D 3 4 S␶3␶4 ץ ϭϪ␦␰␩ ¯ ¯ ͒ J͑␶ץ͒ J*͑␶ץ 1 2 ͕͐ ͓␺,␺͔exp Ϫ͐d␶3 d␶4 ␶ ␶ ͓␺,␺͔ ͖ ͯ † ‡ D S 3 4 JϭJ*ϭ0

2ץ ¯ Ϫ1 Ϫ1 ϭϪ␦␰␩ exp d␶3 d␶4 J*͑␶3͒† ͓␺,␺͔‡␶ ␶ J͑␶4͒ ϭϪ␦␰␩͓ ͔␶ ␶ , (2.48) J͑␶ ͒ ͫ ͵ S 3 4 ͬͯ S 1 2ץ͒ J*͑␶ץ 1 2 JϭJ*ϭ0

where is the matrix whose (l,m) component is S 2 ¯ A͑k,␻͒ϭ ͉͗⌽n͑NϪ1͉͒ck͉⌽g͑N͉͒͘ l,m͓␺,␺͔. After the Fourier transformation, the ther- ͚n malS Green’s function for the noninteracting Hamil- tonian 0 is easily obtained from Eqs. (2.43) and (2.48) ϫ␦„␻ϩEn͑NϪ1͒ϪEg͑N͒… as H † 2 ϩ ͉͗⌽n͑Nϩ1͉͒ck͉⌽g͑N͉͒͘ ͚n 1 0͑k,␻n͒ϭ , (2.49) ϫ␦ ␻ϪE ͑Nϩ1͒ϩE ͑N͒ , (2.54) G i␻nϪ␧0͑k͒ „ n g … where g specifies the ground state. From a comparison where ␧0(k) is the dispersion of H0 as in Eq. (2.3b). of the spectral representation of GR(k,␻) and the ana- The real-time-displaced retarded Green’s function is lytic continuation, we find that defined by ϱ A͑k,␻Ј͒ GR͑k,␻͒ϭ lim d␻Ј GR r,t͒ϭϪi␪ t͒ c r,t͒,c† 0,0͒ , (2.50) ͵ Ϫ ϩ ͑ ͑ ͕͗ ͑ ͑ ͖͘ ␦ ϩ0 Ϫϱ ␻ ␻Ј i␦ → with the anticommutator ͕A,B͖ϵABϩBA, the Heavi- ϭ lim „i͑␻ϩi␦͒… (2.55) ␦ ϩ0 G side step function ␪, and the Heisenberg representation → i t Ϫi t A(t)ϭe H Ae H . The causal Green’s function is simi- is satisfied, where GR is analytic in the upper half plane larly defined by of complex ␻. A(k,␻) is the imaginary part of the single-particle GC͑r,t͒ϭϪi͗Tc͑r,t͒c†͑0,0͒͘. (2.51) Green’s function and therefore contains full information about the temporal and spatial evolution of a single elec- The thermal Green’s function is analytically continued tron or a single hole in the interacting many-electron GR to a complex variable z, system. Using the Fourier transform (k,␻) of the re- tarded Green’s function (2.50), we find the spectral func- tion ϱ A͑k,␻Ј͒ ͑k,Ϫiz͒ϭ d␻Ј, (2.52) ͵ zϪ␻Ј 1 G Ϫϱ A͑k,␻͒ϭϪ Im GR͑k,␻͒. (2.56) ␲ with the spectral function This directly follows from Eq. (2.55). The k-integrated spectral function or the density of states is defined by A͑k,␻͒ϵ e␤„FϪEm͑N͒…͑e␤͑␻Ϫ␮͒ϩ1͒ n͚,m 2 ␳͑␻͒ϵ͚ A͑k,␻͒. (2.57) ϫ͉͗⌽n͑NϪ1͉͒ck͉⌽m͑N͉͒͘ k

ϫ␦„␻ϪEm͑N͒ϩEn͑NϪ1͒…. (2.53) If we integrate A(k,␻) with respect to ␻ up to the chemical potential ␮, we obtain the electron momentum At zero temperature the spectral function is reduced to distribution function

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1062 Imada, Fujimori, and Tokura: Metal-insulator transitions

␮ Therefore Re ⌺(k,␻) and ϪIm ⌺(k,␻) give, respec- n͑k͒ϭ d␻ A͑k,␻͒. (2.58) tively, the energy shift and the broadening of the one- ͵Ϫϱ electron eigenvalue ␻ϭ␧0(k) due to the interaction. If we integrate A(k,␻) above and below the chemical Because the Green’s function GR(k,t) is a linear re- potential, then sponse function to an external perturbation at tϭ0, the imaginary part and the real part of its Fourier transform ϱ GR(k,␻) should satisfy causality, i.e., the Kramers- d␻ A͑k,␻͒ϭ1, (2.59) Kronig relation: ͵Ϫϱ per energy band. ϱ GR R 1 Im ͑k,␻Ј͒ In a noninteracting system, we derive Re G ͑k,␻͒ϭ d␻Ј, (2.64) ␲ P͵Ϫϱ ␻Ϫ␻Ј 1 R ϱ GR G ͑k,␻͒ϭ , R 1 Re ͑k,␻Ј͒ ␻Ϫ␧0͑k͒Ϫi␦ Im G ͑k,␻͒ϭϪ d␻Ј. (2.65) ␲ P͵Ϫϱ ␻Ϫ␻Ј That is, GR(k,␻) has to be an analytic function of ␻ in C 1 G ͑k,␻͒ϭ . the complete upper ␻ half-plane. ␻Ϫ␧0͑k͒ϩi␦ sgn„␧0͑k͒Ϫ␧F… Then, ⌺(k,␻) should also be analytic in the same ␻ The Green’s function of an interacting electron sys- region and satisfy the Kramers-Kronig relation: tem GR or is related to that of the noninteracting RG system G (k,␻)ϭ1/͓␻Ϫ␧ (k)Ϫi␦͔,or (k,␻) 1 ϱ Im ⌺͑k,␻Ј͒ϪIm ⌺͑k,ϱ͒ 0 0 G0 n ϭ1/ i␻ Ϫ␧ (k) where ␧ (k) is the energy of a nonin- Re ⌺͑k,␻͒ϭ d␻Ј ͓ n 0 ͔ 0 ␲ P͵Ϫϱ ␻Ϫ␻Ј teracting Bloch electron, through the Dyson equation (2.66)

͑k,␻n͒ϭ 0͑k,␻n͒ϩ 0͑k,␻n͒⌺͑k,␻n͒ ͑k,␻n͒. 1 ϱ Re ⌺͑k,␻Ј͒ G G G G (2.60) Im ⌺͑k,␻͒ϪIm ⌺͑k,ϱ͒ϭϪ d␻Ј. ␲ P͵Ϫϱ ␻Ϫ␻Ј Here ͚(k,␻n) is the self-energy into which all of the (2.67) interaction effects are squeezed. The solution of the Equations (2.53) and (2.54) for ␻Ͻ␮ and ␻Ͼ␮ can Dyson equation formally reads be measured by photoemission and inverse photoemis- sion spectroscopy, respectively, in the angle-resolved (k- 1 ͑k,␻ ͒ϭ . (2.61) resolved) mode using single-crystal samples. Here ␻ is n i␻ Ϫ␧ ͑k͒Ϫ⌺͑k,␻ ͒ G n 0 n related to the kinetic energy ␧kin of the emitted (ab- After analytic continuation of the Matsubara frequency sorbed) electron through ␻ϭ␧kinϪh␯, where h␯ is the i␻ to the real ␻, the Green’s function is energy of the absorbed (emitted) photon. ␳(␻), given by n Eq. (2.57), is measured by the angle-integrated mode of 1 photoemission and inverse photoemission experiments. GR͑k,␻͒ϭ . (2.62) Here it should be noted that the photoemission and in- ␻Ϫ␧0͑k͒Ϫ⌺͑k,␻͒ verse photoemission spectra are not identical to the

The pole ␻ϭ␧0(k) for the noninteracting system is spectral function A(k,␻) due to several experimental conditions, e.g., the spectral function is modulated by modified to ␻ϭ␧0(k)ϩ͚(k,␻). For simplicity, we consider the case in which there is dipole matrix elements for the optical transitions to only one energy band in the Brillouin zone. The self- yield measured spectra; finite mean free paths of photo- electrons violate the conservation of momentum per- energy ⌺(k,␻) gives the change of ␧0(k) caused by electron-electron interaction, electron-phonon interac- pendicular to the sample surface, KЌ , making it difficult tion, and/or any kind of interaction that cannot be incor- to determine (K)Ќ , etc. [see the text book on photoelec- tron spectroscopy by Hu¨ fner (1995)]. The latter problem porated in the noninteracting energy spectrum ␧0(k). In this article, we implicitly assume that the self-energy cor- does not exist for quasi-1D or 2D materials, in which the rection is due to electron-electron interaction. In a non- dispersions of energy bands in one or two directions is interacting system, the spectral function has a ␦-function negligibly small and hence the nonconservation of (K)Ќ does not cause any inconvenience. peak at ␻ϭ␧0(k): A(k,␻)ϭ␦„␻Ϫ␧0(k)…. With interac- tion,

1 D. Fermi-liquid theory and various mean-field approaches: A͑k,␻͒ϭϪ Im GR͑k,␻͒ Single-particle descriptions of correlated metals ␲ and insulators 1 ϪIm ⌺͑k,␻͒ ϭ 2 2 . To understand the fundamental significance of strong- ␲ ͓␻Ϫ␧0͑k͒ϪRe ⌺͑k,␻͔͒ ϩ͓ϪIm ⌺͑k,␻͔͒ correlation effects, as well as the difficulties in treating (2.63) them, it is helpful first to recollect basic and established

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1063 single-particle theories. This section discusses several Fermi level, we may restrict our consideration to small ␻ single-particle theories of metals. In particular, the and k measured from the Fermi surface. Then the real achievements to date of weak-correlation approaches and imaginary parts of the self-energy can be expanded and mean-field methods, as well as their limitations and as difficulties, are surveyed in this section in order to un- Re ⌺͑k,␻͒ϭRe ⌺͑kϭk ,␻ϭ␮͒Ϫb ͑␻Ϫ␮͒ derstand the motivation for the strong-correlation ap- F kF proaches presented in later sections. 2 ϪO„͑␻Ϫ␮͒ …ϩa͑kF͒͑kϪkF͒ 1. Fermi-liquid description 2 ϩO„͑kϪkF͒ …, (2.68a) In Fermi-liquid theory, adiabatic continuation of the Im ⌺͑k,␻͒ϭϪ⌫͑␻Ϫ␮͒2ϩO ͑␻Ϫ␮͒3 correct ground state in the interacting system from that „ … 2 3 in its noninteracting counterpart is assumed (Landau, ϩ␨͑kF͒͑kϪkF͒ ϩO„͑kϪkF͒ …. 1957a, 1957b, 1958; Anderson, 1984). In this subsection, (2.68b) we mainly discuss basic statements of the Fermi-liquid theory. The reader is also referred to references for fur- Here we have neglected possible logarithmic correc- ther detail (Pines and Nozie`res, 1989). In Fermi-liquid tions. The fact that the interaction term generates only theory, the Green’s function G defined by Eq. (2.61) or quadratic or higher-order terms for Im ͚ in terms of ␻ is (2.62) has basically the same structure of poles as in a the central assumption of Fermi-liquid theory. In the noninteracting system. The interaction term of the Fermi liquid, the quasiparticle decays through excita- Hamiltonian may be considered in the perturbation ex- tions of electron-hole pairs across the chemical poten- pansion. The Green’s function can be expressed in a tial. The lifetime of the quasiparticle is proportional to self-consistent fashion using the self-energy introduced the number of available electron-hole pairs and is there- in Sec. II.C.4, which includes terms in the perturbation fore proportional to (␻Ϫ␮)2 in the ␻ ␮ limit: expansion summed to infinite order. If the Fermi-liquid Im ⌺(k,␻)ϰϪ(␻Ϫ␮)2. Then the real part of→ the self- description is valid, the self-energy ⌺ remains finite and energy should be proportional to Ϫ(␻Ϫ␮)ϩconst in G is characterized by a pole at ␻ϭ␧0(k)ϩ⌺(k,␻). Be- the vicinity of ␻ϭ␮. The renormalized Green’s function cause Fermi-liquid theory is meaningful only near the has the form

Z kF GR͑k,␻͒Ӎ , ␻Ϫ␧*͑kϪk ͒ϩi Z ⌫͑␻Ϫ␮͒2ϪZ ␨͑k ͒͑kϪk ͒2 (2.69) F „ kF kF F F …

for small ␻Ϫ␮ and kϪkF where The renormalization of the quasiparticle density of states, as compared to that of the noninteracting system Z ϭ1/͑1ϩb ͒ (2.70) kF kF at the Fermi level, is kץ kץ is the renormalization factor, whereas N*͑␮͒ ϭ dk dk (2.73a) ␧ץ ␧* Ͳ ͵kϭkץ ␧*͑kϪkF͒ӍZk ͓␧0͑k͒ϩa͑kF͒͑kϪkF͔͒ (2.71) ␳ ͑␮͒ ͵kϭk F 0 F F 0 is the renormalized dispersion. Re ⌺(kϭkF ,␻ϭ␮) can m* m␻ mk ϭ ϭ (2.73b) be absorbed into a constant shift of ␧0(k), which should m m • m keep the constraint of fixed density. The necessary con- b b b dition for the validity of the Fermi-liquid description is where *Ϫ ӷ Ϫ Ϫ kץ ␧ ␮ Im ͚(k,␻) for small ␻ ␮ and k kF , which is indeed satisfied in this case if Z Ͼ0. The renormaliza- m ϭ dk, (2.73c) ␧ץ ͵ kF b tion factor Z scales basically all the other quantities. kϭkF 0 kF From this Green’s function, other physical quantities are m ϭm /Z , (2.73d) ␻ b kF derived. If the spectral function A(k,␻) and the renor- and malization factor Zk(␻) [or Re ⌺(k,␻)] are known, Z (␻)Ϫ1A(k,␻) gives the energy distribution function Ϫ1 k mkϭ„vF0ϩa͑kF͒… vF0mb , (2.73e) of with momentum k, in which each qua- k) ͉ andץ/ ␧ץ)with the bare Fermi velocity v ϭ siparticle carries weight 1 instead of Z of the spectral F0 0 kϭkF function. Hence the density of states of quasiparticles the bare band mass mb . We employ the notation mb N*(␻) is given by instead of m0 to specify the band mass of the noninter- acting Bloch electrons, although the noninteracting band Ϫ1 dispersion is denoted by ␧0 . Here, m␻ and mk are often N*͑␻͒ϭ͚ Zk͑␻͒ A͑k,␻͒. (2.72) k called the ␻ mass and the k mass, respectively, with the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1064 Imada, Fujimori, and Tokura: Metal-insulator transitions total effective mass m*ϭm m /m . The ratio of the k ␻ b Zk„␧*͑k͒… A͑k,␻͒ӍϪ specific-heat coefficient ␥ to that of noninteracting sys- ␲ tem, ␥b , is also scaled by the effective mass enhance- ment m*/mb : Zk„␧*͑k͒…Im ⌺„k,␧*͑k͒… ϫ 2 2 , ␥ m* ͓␻Ϫ␧*͑k͔͒ ϩ͓Zk„␧*͑k͒…Im ⌺͑k,␻͔͒ ϭ . (2.74) ␥b mb (2.82) 2 where Z is generalized to Z (␻)ϵ͓1 ␮ is scaled as kF kץ/nץ ( The charge compressibility ␬ϭ(1/n .␻͔Ϫ1(Ͻ1) to allow ␻ dependenceץ/(Re ⌺(k,␻ ץϪ 1 m* 1 Equation (2.82) indicates that the quasiparticle peak at ␬ϭ 2 s , (2.75) n mb 1ϩF0 ␻ϭ␧*(k) has approximately a Lorentzian line shape s with a width of Zk ␧*(k) Im ⌺ k,␧*(k) . It also indi- where a Landau parameter F0 is introduced phenom- „ … „ … enologically as the spin-independent and spatially iso- cates that the quasiparticle weight is Zk„␧*(k)… instead tropic part of the quasiparticle interaction. Similarly, the of one as in the noninteracting case. uniform magnetic susceptibility has the scaling form In the transport process, it is known that the mass m, which appears in the conductivity formula, ␴ϰne2␶/m m* 1 and the Drude weight Dϰn/m, is not scaled by the ef- ␹sϭ a , (2.76) mb 1ϩF0 fective mass m* but by the bare mass m0 for systems where the spin antisymmetric and spatially isotropic part with Galilean invariance. This is because the electron- a electron interaction conserves the total momentum and of the quasiparticle interaction is described by F0 . The renormalization factor Z can be interpreted as hence the current does not change under an adiabatic kF switching of the interaction. However, the transport of follows: From Eqs. (2.54), (2.55), and (2.69), Z is re- kF electrons in solid is associated with Umklapp scattering written as and impurity scattering. In particular, the Drude weight Z ϭ ⌽ ͑Nϩ1͒ c† ⌽ ͑N͒ 2, (2.77) in perfect crystals may be strongly affected by Umklapp kF ͉͗ g ͉ k ,␴͉ g ͉͘ F scattering, which may be regarded as the origin of the where Z is interpreted as the overlap between the true kF Mott transition in the perturbational framework. This ground state of the (Nϩ1)-particle system and the wave effect of the discrete lattice, not seen in the Galilean function for a bare single particle at kF is added to the invariant system, may cause a renormalization of the true N-particle ground state. Z is also given as the kF transport mass from the bare value to mD* , which may in jump of the momentum distribution function ͗n(k)͘ general be different from m* and mb . When the elec- † ϵ͗ck␴ck␴͘ at Tϭ0. The spectral representation tron correlation is large, electron-electron scattering is 2 the dominant decoherence process of the electron wave ͉͗⌽g͑N͉͒ck͉⌽n͑Nϩ1͉͒͘ GC͑k,␻͒ϭ lim ͚ function. This decoherence process is believed to govern ␦ ϩ0 n ␻ϪEnϩi␦ → the electron relaxation time ␶, although Umklapp scat- † 2 tering is necessary to cause inelastic relaxation. The ͉͗⌽g͑N͉͒ck͉⌽n͑NϪ1͉͒͘ ϩ (2.78) quasiparticle-quasiparticle scattering time is given by ␻ϩEnϪi␦ 1/␶ ϭZk Im ⌺. Then, in the usual Fermi liquid, the ␻ 2 leads to dependence Zk Im ⌺ϰ(␻Ϫ␮) may be transformed to temperature dependence by replacing ␻Ϫ␮ with T be- 1 ϱ ͗n͑k͒͘ϭ lim ͵ d␻ GC͑k,␻͒ei␻t. (2.79) cause the thermal excitation has a characteristic energy 2␲i t ϩ0 Ϫϱ scale of T from the Fermi surface. Therefore the resis- → Using the renormalized form tivity at low temperatures in strongly correlated systems in general has the temperature dependence Z C k 2 G ͑k,␻͒ϭ , (2.80) ␳ϭAT ϩ␳0 , (2.83) ␻Ϫ␧*͑kϪkF͒ϩi␥ sgn͑␧*Ϫ␧F͒ where ␳ is the residual resistivity determined from the the jump of ͗n(k)͘ at k , is given by 0 F impurity scattering process. The temperature depen- n͑k͒ Ϫ n͑k͒ ͗ ͘kϭkFϪ␦ ͗ ͘kϭkFϩ␦ dence (2.83) is widely observed in many d and f electron compounds at low temperatures. We discuss some i ϱ ϭϪ lim ͵ d␻ ei␻t examples, NiS2ϪxSex , NiS, RNiO3,Ca1ϪxSrxVO3, 2␲ t ϩ0 Ϫϱ R1ϪxSxTiO3, R1ϪxAxMnO3, etc., in Sec. IV. However, → we also discuss how T2 dependence is replaced with Zk Zk ϫ Ϫ other temperature dependences in cases such as the ͩ ␻Ϫ␧*͑kϪk ͒Ϫi␥ ␻Ϫ␧*͑kϪk ͒ϩi␥ͪ high-T cuprates and NiSSe. The origin of these unusual F F c properties is one of the important issues to be discussed ϭZ k . (2.81) in this article. On empirical grounds, it has been sug- The spectral function defined in Eq. (2.63) becomes gested that A seems to be proportional to ␥2 in many

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1065

␹ ϭ1Ϫ͑X ϪX ͒, (2.88) ␹b ↑↑ ↑↓ ␬ ϭ1Ϫ͑X ϩX ͒, (2.89) ␬b ↑↑ ↑↓ with

⌺␴͑k,␻ϭ␮͒ץ X ͑k͒ϭ , (2.90) ͯ Hץ ↑↑ ␴ H␴ϭ0 FIG. 6. Diagrammatic representation of the four-point vertex ⌺␴͑k,␻ϭ␮͒ץ function ⌫(k1 ,k2 ,k3 ,k4). X ͑k͒ϭ . (2.91) ͯ Hץ ↓↑ Ϫ␴ HϪ␴ϭ0 heavy-fermion compounds, that is, A/␥2 seems to be a From Eqs. (2.84) and (2.85), X can be expressed as ↑↓ universal constant in many cases (Kadowaki and Woods, 1986). This constant is also similar in some perovskite Ti X ͑k͒ϭ͚ A͑kЈ,␻ϭ␮͒⌫ ͑k,kЈ;kЈ,k͒. (2.92) ↑↓ k ↑↓ and V compounds (Ca1ϪxSrxVO3, R1ϪxAxTiO3 etc.; see Ј Secs. IV.A.5 and IV.B.1). However, this empirical rule is In this notation, the mass enhancement is basically as- not satisfied near antiferromagnetic transition, as in cribed to the enhancement of ͚k ⌫ (kϭ0,kЈ;kЈ,kϭ0). 2 2 Ј ↑↓ NiS2ϪxSex (Sec. IV.A.2), as well as in the high-Tc cu- Up to the order of T and ␻ ,Im͚(k,␻) is calculated as prates. ␻2ϩ͑␲T͒2 To derive thermodynamic as well as dynamic proper- Im ͑k,␻͒ϭϪ ␲ A͑kϪq,␻ϭ␮͒ ͚ 2 ͚ ties based on the Fermi-liquid theory, it is necessary to kЈ,q include self-consistently the vertex correction in the self- ϫA͑kЈ,␻ϭ␮͒A͑kЈϩq,␻ϭ␮͒ energy. For this purpose, the Ward-Takahashi identity provides a useful relation between the self-energy and ϫ ⌫2 ͑k,kЈ;kЈϩq,kϪq͒ the vertex function under spin-rotational invariance ͫ ↑↓ (Koyama and Tachiki, 1986; Yamada and Yosida, 1986; 1 Kohno and Yamada, 1991). The Ward-Takahashi iden- 2 ϩ ⌫ A͑k,kЈ;kЈϩq,kϪq͒ (2.93) tity leads to 2 ↑↑ ͬ with the antisymmetrized vertex function ⌺␴͑k,i␻͒ץ ⌺␴͑k,i␻͒ץ ϭ Ϫ A͑kЈ,␻ϭ␮͒ ␮ ͚ ⌫ A͑k1 ,k2 ;k3 ,k4͒ץ i␻͒͑ץ kЈ,␴Ј ↑↑ ϫ ϭ⌫ ͑k1 ,k2 ;k3 ,k4͒Ϫ⌫ ͑k1 ,k2 ;k4 ,k3͒. (2.94) ⌫␴␴Ј͑k,kЈ;kЈ,k͒ (2.84) ↑↑ ↑↑ and Using the damping factor

(⌺␴͑k,i␻͒ ˜⌫͑k͒ϭϪZ͑k͒Im ͑k,␻͒, (2.95ץ ⌺␴͑k,i␻͒ץ ϭ Ϫ A͑kЈ,␻ϭ␮͒ ͚ i H ͚ ␴ kЈ ץ ␻͒ ͑ץ we can express the coherent part of the conductivity as ϫ⌫␴␴͑k,kЈ;kЈ,k͒ (2.85) k2 i with H ϭg␮ H␴/2 under a magnetic field H where the 2 ␴ B ␴coh͑␻͒ϳe ͚ A*͑k,␻ϭ␮͒Jk Jk k ⌺ ͑K k͒ ˜ four-point vertex function ⌫␴␴Ј is defined in Fig. 6 and i i• ␻ϩi⌫͑k͒ where all of ␻ are taken zero. We note the identity (2.96) H . Below we neglect theץ/ ⌺ץH ϩץ/ ⌺ץ␮ϭץ/ ⌺ץ ␴ ␴ ␴ ␴ Ϫ␴ with the sum ͚ over the internal product of k and all wave-number dependence of the self-energy for the time i the reciprocal lattice vectors K coming from Umklapp being. Because i scattering. It should be noted that the electron-electron ⌺ Ϫ1 interaction itself does not generate a nonzero resistivityץ Z ϭ 1Ϫ (2.86) -␻ͪ because of momentum conservation, and Umklapp scatץ ͩ k tering is needed to yield a momentum relaxation process is satisfied from Eqs. (2.68a), (2.68b), and (2.70), we ob- (Yamada and Yosida, 1986). Yamada and Yosida suc- tain ceeded in deriving the momentum conservation by tak- ing account of proper vertex and self-energy corrections ⌺␴͑k,␻͒ץ ␥ ϭ1Ϫ͚ . (2.87) utilizing the Ward-Takahashi identity. This guarantees ͯ ␻ץ b k␴␥ ␻ϭ␮ zero resistivity in the absence of Umklapp scattering. Similarly, the enhancement of the uniform magnetic sus- Because the inverse of the quasiparticle spectral weight Ϫ1 ˜ ceptibility and charge susceptibility as compared to the A* , the current Jk , and ⌫(k) are all renormalized by noninteracting band values ␹b and ␬b satisfies Zk(␻), they are canceled in ␴(␻ϭ0). Therefore ␴(␻

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1066 Imada, Fujimori, and Tokura: Metal-insulator transitions

to spin fluctuations, this enhancement must be due to ⌳. Thus neglecting k and ␻ dependence of ⌳(k,kϩQ;i␻), we obtain

␹͑qϩQ͒ϳ⌳͑qϩQ͒␹b͑qϩQ͒, (2.104) where we employed simplified notation ⌳(q)ϳ⌳(k,k ϩq;i␻) and where ␹b is the susceptibility without the vertex correction. From Eq. (2.103), if we neglect k and kЈ dependence of ⌫ (kϩq,kЈ;k,kЈϩq) as denoted by ↑↓ FIG. 7. Diagrammatic representation of the three-point vertex ⌫ (q), we obtain ↑↓ function ⌳(k1 ,k2 ,k3 ,k4). 2 ␹͑Qϩq͒ϳ␹b͑Qϩq͒⌫ ͑Qϩq͒. (2.105) ↑↓ When the strong peak structure of ␹(Qϩq,␻) can be Ϫ ϭ0) is basically scaled by ͓Im ͚(k,␻)͔ 1, which may be approximated as then related to ␥ and ␹ through ⌫ . 1 It was argued that if the charge↑↓ fluctuation could be ␹͑Qϩq,␻͒ϳ 2 2 ␹b͑Qϩq,0͒, (2.106) assumed to be suppressed as ␬ 0, as compared with Ϫi␻ϩDs͑K ϩq ͒ → spin fluctuations, we might take we have X Ӎ1ϪX , (2.97) ␹͑Qϩq,␻ϭ0͒ ↑↑ ↑↓ ϳ␹b͑Qϩq,␻ϭ0͒⌫ ͑Qϩq͒ which leads to ␹b͑Qϩq,␻ϭ0͒ ↑↓ ␹ 1 ϭ2 X ϭ2 A͑kЈ,␻ϭ␮͒⌫ ͑k,kЈ;kЈ,k͒, ϳ . (2.107) ͚ ͚ D K2ϩq2 ␹b k ↑↓ k,kЈ ↑↓ s͑ ͒ (2.98) This type of susceptibility is justified for an order param- ␥ eter that is not a conserved quantity of the Hamiltonian ϭ ͚ A͑kЈ,␻ϭ␮͒⌫␴␴Ј͑k,kЈ;kЈ,k͒ and that is coupled to metallic and gapless Stoner exci- ␥b k,kЈ tations. Because of the coupling to the Stoner excitation ␴Ј Eq. (2.106) contains an ␻-linear term in the denomina- 1 ␹ tor, which comes from the overdamped nature of the ϭ ϩ2͚ A͑kЈ,␻ϭ␮͒⌫ ͑k,kЈ;kЈ,k͒. (2.99) 2 ␹b k,kЈ ↑↑ spin fluctuation. The antiferromagnetic transition at Q Þ0 may indeed belong to this class, as we discuss in Sec. Therefore the Wilson ratio is given by II.D.9. Then the NMR relaxation rate is given as ␹ ␥ b Im ␹ k,␻ϭ0 RWϭ Ϫ1 ͑ ͒ 2 ␹b ␥ ͑T1T͒ ϳ lim ͚ f͑k͒ ϳ͚ f͑k͒⌫ ͑k͒, ␻ 0 k ␻ k ↑↓ → 2 (2.108) ϭ . 4␹b while from Eqs. (2.93), (2.95), and (2.96), neglecting 1ϩ ͚k,k A͑kЈ,␻ϭ␮͒⌫ ͑k,kЈ;kЈ,k͒ ␹ Ј ↑↑ ⌫ , we obtain ↑↑ (2.100) 1 If the Wilson ratio is close to two, it gives ␴͑␻͒ϰ , (2.109) Ϫi␻ϩ˜⌫ A͑kЈ,␻ϭ␮͒⌫ ͑k,kЈ;kЈ,k͒Ӎ0. (2.101) ͚ ˜ 2 2 2 kkЈ ↑↑ ⌫ϳIm ϰ͓␻ ϩ͑␲T͒ ͔ ⌫ ͑k͒. (2.110) ͚ ͚k ↑↓ The static susceptibility ␹(q) is expressed as Therefore, when we assume weak k dependence of ␹b , ϱ d␻ the basic scaling of these two quantities is determined ␹͑q͒ϭϪ͚ ⌳͑k,kϩq;0͒G͑k,i␻͒G͑kϩq,i␻͒ k ͵Ϫϱ 2␲i from (2.102) kdϪ1 2 dϪ4 with the three-point vertex function ⌳ illustrated in Fig. ⌶ϵ ͓⌫ ͑k͔͒ ϰ dk 2 2 2 ϰK . (2.111) ͚k ↑↓ ͵ ͑K ϩk ͒ 7. The ⌳ and four-point vertex functions ⌫ are related as This leads to ⌳͑k,kϩq,0͒ϭ1ϪT ⌫ ͑kϩq,kЈ;k,kЈϩq͒ Ϫ1 dϪ4 ͚ ͚ ͑T1T͒ ϰAϰK (2.112) n kЈ ↑↓ for the resistivity coefficient A defined by ␳ϭAT2. Here ϫG ͑kЈϩq,i␻n͒G ͑kЈ,i␻n͒. (2.103) ↑ ↓ we have neglected q dependences of the nuclear form Therefore, if ␹(q) is strongly peaked around kϭQ due factor f(q) by assuming weak q dependence. As we see

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1067 from Eqs. (2.99) and (2.107) the specific-heat coefficient assumed above through vanishing charge compressibility ␥ and the local susceptibility ␹ for the singular part fol- ␬ϳ0, is not justified in metals near the Mott insulator. low the scaling Let us consider the resistivity and the frequency- dependent conductivity by considering the dependence 1/K for 1D qdϪ1 on the renormalization factor Z explicitly. As we dis- ␥ϰ␹ϰ dq ϰ Ϫln K for 2D (2.113) cussed above, the optical conductivity given by Eqs. ͵ K2ϩq2 ͭϪͱK for 3D. (2.109) and (2.96) can be rewritten from the fact that A*Ϫ1, J , and ˜⌫(k) are all scaled by the renormaliza- When the susceptibility can be described by a Curie- k tion factor Z(␻): Weiss law ˜ C ⌫͑␻͒nZ͑␻͒ ␹͑Q͒ϳ (2.114) Re ␴͑␻͒ϰ (2.123) Tϩ⌰ ␻2ϩ˜⌫͑␻͒2 where C is the Curie-Weiss constant and ⌰ the Curie- with Weiss temperature, then Eq. (2.107) implies Kϰ(T 1/2 ϩ⌰) , leading to ˜⌫͑␻͒ϳZ͑␻͒Im ⌺ϰ͵ dz Z͑z͒Im ␹͑q,z͒ Ϫ1 d/2 Ϫ2 ͑T1T͒ ϰAϰ͑Tϩ⌰͒ . (2.115) ϫIm ␹b͑Q,␻Ϫz͓͒nB͑z͒ϪnB͑zϪ␻͔͒ Similarly, we obtain for the singular part Z͑␻͒ Ϫln͑Tϩ⌰͒ for 2D ϰ G͑␻͒ (2.124) ␥ϰ␹ϰ (2.116) nDs ͭ ϪͱTϩ⌰ for 3D. in the Born approximation at TӶ␻. Here ␹b is the sus- We note that, aside from the singular part given in Eq. ceptibility of the noninteracting band electrons while the ␤z (2.116), the leading term is a constant for 3D. These Bose distribution nB is defined by nB(z)ϭ1/(e Ϫ1); comprise the basic framework of the antiferromagnetic G(␻) is proportional to ␻2 in a simple Fermi liquid, spin-fluctuation theory based on the Fermi-liquid while it may have ␻ dependence G(␻)ϰ␻d/2 in a wide theory, as we discuss in detail in Secs. II.D.8 and II.D.9. range of ␻ if the spin-fluctuation effect is important. In three dimensions, at the critical point ⌰ϭ0 of the This dependence is easily understood when we replace magnetic transition, it leads to the T dependence of ˜⌫ discussed above with ␻ depen- dence. Note that the temperature dependence of ˜⌫ is ͑T T͒Ϫ1ϰTϪ1/2, (2.117) 1 obtained from Eqs. (2.110), (2.111), (2.114), and (2.107) ␳ϰT3/2, (2.118) as ˜⌫ϳT2KdϪ4ϳTd/2 at TϾ␻ and ⌰ϳ0. Then the ␻- integrated conductivity, that is, the Drude weight D ␥ϰ␹ϰconstϪͱT (2.119) ϱ ϭ͐0 ␴coh(␻)d␻, can be sensitively dependent on the while in two dimensions at ⌰ϭ0 it leads to doping concentration near the Mott insulator only through the renormalization factor Z in the region with- Ϫ1 T1 ϰconst, (2.120) out magnetic order when a singular k-dependent mass m is absent. This is because the spin-diffusion constant ␳ϰT, (2.121) k Ds is not critically dependent on the doping concentra- and tion and n must remain finite near the Mott insulator in the Fermi-liquid theory due to the Luttinger theorem. In Ϫ T ␥ϰ␹ϰ ln , (2.122) fact, because the spin-diffusion constant Ds should be as is known in the self-consistent renormalization theory connected with the overdamped mode of the spin wave, (Ueda, 1977; Moriya, 1985; Moriya, Takahashi, and it should remain finite near the antiferromagnetic Mott insulator. Therefore, the Drude weight decreases to van- Ueda, 1990). In particular, at the antiferromagnetic tran- ish toward the Mott transition only when ZϪ1 is criti- sition point in some 3D metals such as ␤-Mn, cally enhanced at small ␻. In this case, the total Drude NiS Se , and Ni Co S this behavior is indeed ob- 2Ϫx x 1Ϫx x 2 weight derived from Eq. (2.123) depends on the doping served (Watanabe, Moˆ ri, and Mitsui, 1976; Katayama, concentration ␦ mainly in the form DϰZ. This necessar- Akimoto, and Asayama, 1977; Miyasaka et al., 1997). In ily leads to critical enhancement of ␥ through Eqs. 2D, it was claimed that Eqs. (2.120) and (2.121) repro- (2.73a)–(2.73e) and (2.74) provided that the k mass is duce the anomalous behavior in the high-Tc cuprates, as not critically suppressed. At least for the paramagnetic we discuss in Sec. IV.C.1. metal region, the reduction of D near the Mott insulator As is evident from the above arguments, spin- transition is necessarily accompanied by enhancement of fluctuation theory alone fails to reproduce physical ␥ in the Fermi-liquid description if mk does not have properties near the Mott transition point if the MIT is of singular k dependence. As we discuss in Sec. IV, such the continuous type. This is because the above treatment behavior in D and ␥ contrasts with that observed in of spin fluctuations takes into account only the coherent high-Tc cuprates because D decreases while ␥ is not en- part of the charge excitations. We also discuss in Secs. hanced. In the region where antiferromagnetic order ex- II.E.3 and II.F why the neglect of charge fluctuation, ists, of course, D can be reduced to zero without the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1068 Imada, Fujimori, and Tokura: Metal-insulator transitions

the sum rule, Eq. (2.58). If the line shape of the spectral function ␳(␻) did not change with the increase of elec- tron concentration ⌬n, the spectral weight that could be transferred from above to below the chemical potential would be ␳(␮)⌬␮ЈϭZN*(␮)⌬␮Ј [Eq. (2.126)], which is less than ⌬nϭN*(␮)⌬␮Ј by the renormalization fac- tor Z, and would violate the sum rule. The rest of the spectral weight (1ϪZ)⌬n is necessarily transferred from well above the chemical potential to well below it through a change in the line shape of ␳(␻). It is there- fore concluded that, if there is electron correlation, i.e., Ϫ1 -␻͉␻ϭ␮͔ Ͻ1, a change in the band fillץ/Re ⌺ ץif Zϭ͓1Ϫ ing necessarily leads to a change in the line shape of the spectral function. FIG. 8. Schematic representation of the chemical potential Here it should be noted that carrier doping in real shift ⌬␮ induced by a change ⌬n in the electron concentration compounds such as transition-metal oxides has to be n and spectral weight transfer near the chemical potential in carried out under the condition of charge neutrality be- an isotropic Fermi liquid. N*(␻) is the quasiparticle density cause of the presence of the long-range Coulomb inter- and ␳(␻) is the (spectral) density of states. Note that ␳(␻) action in real systems. That is, in the case of perovskite- ϭZN*(␻) near the chemical potential. type transition-metal oxides, in order to dope the system with electrons (holes), one must usually replace A-site enhancement of ␥ because small pockets of the Fermi metal ions by ions with higher (lower) valencies or oxy- surface can be arbitrarily small. This, however, is not the gen vacancies (excess oxygens). Therefore the ‘‘particle’’ case in the high-Tc cuprates. In contrast to the Drude in the definition of the chemical potential, the derivative weight, the resistivity depends mainly on ␦ through Im ⌺ of the Gibbs free energy with respect to the particle while the dependence on Z is canceled. Therefore criti- number, is not the bare electron or hole but an electron cal enhancement of A has to be ascribed to Im ⌺.Itwas or hole whose long-range Coulomb interaction is argued that if Z were not reduced and the k dependence screened by the positive background charges of the sub- of ⌫ could be neglected, the Kadowaki-Woods rela- stituted ions or oxygen defects. This implies that micro- tion↑↓ would result from ␥ϰ⌫ and Aϰ⌫2 . scopic models which neglect the long-range Coulomb in- Aside from all the above↑↓ phenomenological↑↓ and teraction, such as the Hubbard model, may be relevant rather sketchy arguments, as far as the authors know, no for analysis of the chemical potential shift and the theoretical studies have been attempted in this Fermi- charge susceptibility, although the long-range Coulomb liquid description with a self-consistent treatment of spin interaction should influence the dynamic (spectroscopic) and charge fluctuations to discuss the paramagnetic re- behaviors through ⌺(k,␻). gion near the Mott insulator. This remains for further Below we discuss in some detail a phenomenological studies. In particular, the k dependence of the self- way of analyzing experimentally observed line shapes of energy has to be considered seriously near a Mott insu- spectral weight. We discuss the case in which the self- energy is local and hence a(k)ϭ0, and Z ϭZ indepen- lator with magnetic order, while all the above arguments kF basically neglect this dependence. In Sec. II.F, we dis- dent of the direction of kF . In this case, ⌬*ϵ1/Z⌫, cuss how the singular k dependence of the self-energy called the quasiparticle coherence energy, defines the should be considered to satisfy the scaling relation of the energy scale below which (͉␻Ϫ␮͉Ӷ⌬*) the quasiparti- Mott transition. cle is well defined. The quasiparticle dispersion is given The shift of the electron chemical potential ␮ as a by ␻Ϫ␮ϭZ͓␧0(k)Ϫ␮͔: The band is uniformly nar- function of electron concentration n defines the charge rowed by the renormalization factor Z(Ͻ1), meaning Ϫ1 n, as discussed that the density of states (per unit energy) of theץ/␮ץsusceptibility ␹c through ␹c ϵ above. This quantity can be measured by photoemission quasiparticles, or equivalently the effective mass m* of spectroscopy as a shift of the Fermi level as a function of the conduction electrons, is enhanced by the factor band filling n. Experimentally, the energies of photo- 1/Z(Ͼ1) compared with the noninteracting case. Thus electrons are usually referenced to the electron chemical the mass enhancement factor m*/mb , where mb is the potential of the spectrometer, which is in electrical con- noninteracting conduction-band mass at the chemical tact with the sample. A chemical potential shift would potential, is equivalent to 1/Z. therefore cause a uniform shift of all the core-level and The quasiparticle peak carries spectral weight Z, and valence-band spectra if there were no interaction be- the remaining spectral weight 1ϪZ is distributed as an tween electrons. ‘‘incoherent’’ background mostly at higher energies, i.e., Now we consider the spectral weight transfer caused at larger ͉␻Ϫ␮͉, as schematically shown in Fig. 9. It by the change ⌬n, referring to Fig. 8 (Ino et al., 1997a, should be noted that the k-integrated spectral function 1997b). As the electron concentration increases by ⌬n, ␳(␻)at␻ϳ␮is not affected by the local self-energy cor- spectral weight ⌬n should be transferred from above to rection, i.e., ␳(␮)ϭNb(␮), where Nb(␻) is the density below the chemical potential, a condition imposed by of states of the noninteracting band, because the mass

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1069

(1990) and more recently by Langer et al. (1995) and Deisz et al. (1996). The density of states at the chemical potential ␳(␮)is modified from the noninteracting value Nb(␮)as

␳͑␮͒/Nb͑␮͒ϭmk/mb, (2.125) which means that k dependence of the self-energy is necessary to have ␳(␮)ÞNb(␮) (Greef, Glyde, and Clements, 1992). The ratio between the density of states FIG. 9. Spectral function ␳(␻) for a constant renormalization ␳(␮) and the density of states of quasiparticles N*(␮) factor Z. N (␻) is the band density of states, m is the band b b ϭ(m*/mb) Nb(␮) is given by Z (Allen et al., 1985), mass, and mk is the k mass. ␳͑␮͒ϭZN*͑␮͒. (2.126) enhancement factor 1/Z is canceled out by the quasipar- That is, the density of states at the chemical potential is ticle spectral weight Z. Therefore one can claim that if reduced from the quasiparticle density of states by the ␳(␮) is different from Nb(␮) there must be k depen- quasiparticle residue Z. This is a special case of the re- dence in the self-energy, i.e., the self-energy should be lationship between the spectral function and the energy nonlocal. This is the case not only for a Fermi liquid but distribution function of quasiparticles, Eq. (2.72). also for a non-Fermi liquid. Therefore, as shown in Fig. In the limit of large dimensionality or large lattice 9, if Z is k independent throughout the quasiparticle connectivity, the k dependence of the self-energy is lost band, the quasiparticle band in the spectral function because the fluctuation of the neighboring sites is aver- ␳(␻) is narrowed by a factor Z with the peak height aged out and the central atom feels only the average unchanged, resulting in the loss of spectral weight 1 field from the neighboring sites without fluctuation. This ϪZ from the coherent quasiparticle band region. The is described in greater detail in Sec. II.D.6. In that case, lost spectral weight 1ϪZ is distributed as incoherent the spectral weight at the chemical potential, ␳(␮), is not weight at higher ͉␻Ϫ␮͉. influenced by electron correlation. It is also expected Although the original Fermi-liquid theory assumes a that the k dependence of the self-energy will become structureless incoherent part and neglects its contribu- less important for a large orbital degeneracy (Kuramoto tion to low-energy excitations, recent studies described and Watanabe, 1987). below in Secs. II.D.4, II.D.6, II.E, II.F, and II.G have In the definition of the k dependence of the self- clarified the importance of the growing weight of inco- energy, one must define the reference mean-field state herent contributions. Based on the Fermi-liquid descrip- to which the self-energy correction is applied. Such a tion, a phenomenological treatment of coherent and in- mean-field state is either the band structure calculated coherent parts on an equal footing has been given using the local-(spin-)density approximation (LDA or (Kawabata, 1975; Kuramoto and Miyake, 1990; Miyake LSDA) or that calculated using the Hartree-Fock ap- and Narikiyo, 1994a, 1994b; Narikiyo and Miyake, 1994; proximation. In the Hartree-Fock approximation, the Narikiyo, 1996; Okuno, Narikiyo, and Miyake, 1997). nonlocality of the exchange interaction is properly dealt This approach may be justified when the incoherent part with and the nonlocality of the electron-electron inter- described by ‘‘upper’’ and ‘‘lower’’ Hubbard bands give action is already included on the mean-field level (and serious renormalization to the quasiparticle coherent not beyond the mean-field level) in the one-electron en- part. ergy ␧0(k), so that ⌺(k,␮)ϳ0. In the L(S)DA, on the In discussing Eq. (2.72) above, we ignored the k de- other hand, the exchange potential is approximated by a pendence of the self-energy for simplicity. However, the local potential. That is, the potential at a spatial point r effect of electron correlation is not restricted to each is a function of the electron (and spin) densities at this r, atomic site but is generally extended over many atomic i.e., V␴(r)ϭf␴͓n (r),n (r)͔. Therefore ⌺(k,␮) may be ↑ ↓ sites, necessitating a k-dependent self-energy. The de- substantially different from zero, meaning that ⌺(k,␮) gree of nonlocality of the self-energy generally increases will have to be taken into account if one analyzes pho- as the range of interaction increases. Nevertheless, even toemission spectra starting from the LDA band struc- the on-site Coulomb interaction included in the Hub- ture. bard model leads to some k dependence in the self- For the free-electron gas, the LDA gives ␧0(k) almost energy through the intersite hybridization term. In par- identical to the noninteracting dispersion ␧0(k) 2 2 ticular, in cases where there is a nesting feature in the ϭប k /2me , while the Hartree-Fock approximation Fermi surface, the self-energy should show an anomaly yields for k around the nesting wave vector. Because of this, ប2k2 2e2 1 1Ϫx2 1ϩx the k dependence of the self-energy is more pronounced ␧0͑k͒ϭ Ϫ kF ϩ ln , for low (one or two) dimensions, since the nesting con- 2me ␲ ͩ 2 4x ͯ 1Ϫx ͯͪ dition is more easily satisfied. The effect of antiferro- (2.127) magnetic fluctuation on the spectral function arising where xϵk/kF (Ashcroft and Mermin, 1976). There- from Fermi-surface nesting was discussed for the 2D fore, if one starts from the LDA, ⌺(kF ,␮)Ϫ⌺(0,␮) 2 Hubbard model originally by Kampf and Schrieffer ϭ2e kF /␲, meaning that the occupied bandwidth is in-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1070 Imada, Fujimori, and Tokura: Metal-insulator transitions creased by ϳ2e2k /␲ compared to the LDA band F † structure due to the effect of nonlocal exchange. In fact, Kϭ͚ ͚ ci␴͕ ijϪ͓␮ϩ͑2Ϫy͒U/4͔␦ij͖cj␴ , H ␴ ij K the screening effect, which is completely neglected in the (2.128b) Hartree-Fock approximation and can be treated in the U RPA, will reduce the ⌺(kF ,␮)Ϫ⌺(0,␮) to some extent, † † 2 Uϭ ͚ ͓y͑ci ci ϩci ci Ϫ1͒ but such a screening effect would be weak when the H 4 i ↑ ↑ ↓ ↓ electron density is low, as is the case in most of the 1/3 † † metallic transition-metal oxides. Since k ϭ(3n/8␲) ϩ2͑2Ϫy͒ci ci ci ci ͔, (2.128c) F ↓ ↑ ¯ ¯ ↑ ↓ ϳ1/2r, where r is the average distance between nearest where ij is Ϫt for the pair (i,j) summed in Eq. (2.37b). 2 ¯ K electrons, ⌺(kF ,␮)Ϫ⌺(0,␮)ϳe /␲r. That is, the total Here y is an arbitrary constant and we have omitted a 2 ¯ constant term U/4. For this Hamiltonian, the partition bandwidth is increased by ϳ⌺(kF ,␮)Ϫ⌺(0,␮)ϳe /r due to the exchange contribution from the long-range function (2.41a) can be expressed in the path integral of Coulomb interaction. If we consider that the long-range a noninteracting system through the Stratonovich- Hubbard (SH) transformation. Several formally differ- Coulomb interaction is screened by the electric polariza- ent but equivalent representations through different SH tion of constituent ions, represented by the optical di- transformations are possible. In a simple and ordinary electric constant ␧ , which is ϳ3 – 5 in oxides, the band ϱ mean-field approximation, one can introduce only ˜ by 2 ¯ ⌬s widening becomes ϳe /␧ϱrϳ0.5 eV (Morikawa et al., taking yϭ0 to discuss magnetic ordering. Below we 1996). Equation (2.127) means that for small kF the sec- show a general choice of SH transformations in which ond term (exchange energy ϰk ) dominates the first F both ⌬˜ and ⌬˜ are introduced to represent possible 2 ¯ c s term (kinetic energy ϰkF), i.e., for large r. Therefore charge and spin fluctuations separately. We use the iden- the nonlocal exchange is important in systems with low tity carrier density. In ordinary metals and alloys, the carrier 2Ϫy density is high enough (kF is large) that the kinetic en- ˜ ˜ ˜ † 1ϭ ⌬si* ⌬si exp Ϫ ͚ ͑2⌬si*ϪUci ci ͒ ergy dominates Eq. (2.127) and the nonlocal exchange ͵ D D ͫ 2U i ↑ ↓ contribution is negligible, explaining why the LDA works well in ordinary metals and alloys. The increase of ˜ † ϫ͑2⌬siϪUci ci ͒ , (2.129a) the bandwidth due to nonlocal exchange is also found ↓ ↑ ͬ using a tight-binding Hamiltonian if one treats the inter- and site Coulomb interaction in the Hartree-Fock approxi- mation: It contributes through ϪV ͗c†c ͘ to the transfer y ij i j ˜ ˜ † † integral t , where V is the intersite Coulomb energy, 1ϭ d⌬ciexp Ϫ ͚ „4⌬ciϪiU͑ci ci ϩci ci Ϫ1͒… ij ij ͵ ͫ 4U i ↑ ↑ ↓ ↓ thereby effectively increasing ͉tij͉. ˜ † † 2. Hartree-Fock approximation and RPA ϫ„4⌬ciϪiU͑ci c ϩci ci Ϫ1͒… , (2.129b) ↑ ↑ ↓ ↓ ͬ for phase transitions for complex c-number ⌬˜ and real variable ⌬˜ . Then The Hubbard Hamiltonian (1.1a) is rewritten as si ci we can rewrite the partition function in the coherent- ϭ ϩ , (2.128a) state representation using the Grassmann variables, H HK HU

ץ ␤ ¯ ˜ ˜ ˜ ¯ ¯ Zϭ ͓␺,␺͔ ⌬si͑␶͒ ⌬si*͑␶͒ ⌬ci͑␶͒exp Ϫ d␶ ͚ ␺i␴ ␺i␴Ϫ͚ ␺i␴͑ ijϪ␮␦ij͒␺j␴ ␶ ij Kץ D D D D ͩ ͵0 ͫ i ͵

2 4y ˜ ˜ ˜ 2 ˜ ¯ ˜ ¯ ϩ ͑2Ϫy͒ ⌬si*͑␶͒⌬si͑␶͒ϩ ⌬ci͑␶͒ Ϫ͑2Ϫy͒ ͓⌬si*͑␶͒␺i ␺i ϩ⌬si͑␶͒␺i ␺i ͔ U ͚i U ͚i ͚i ↓ ↑ ↑ ↓

˜ ¯ ¯ Ϫ2iy͚ ⌬ci͑␶͒͑␺i ␺i ϩ␺i ␺i Ϫ1͒ . (2.130) i ↑ ↑ ↓ ↓ ͬͪ

˜ ˜ account of magnetic symmetry breaking, as we shall see The variables ⌬s and ⌬c are called Stratonovich fields or Stratonovich variables. The normalization constants in below. The partition function may also be expressed ˜ ˜ ˜ Eqs. (2.129a) and (2.129b) are absorbed in ⌬c and only through ⌬si without introducing ⌬ci by using Eqs. ˜ ˜ D ⌬s* ⌬s . Although this is an identical transformation, a (2.129a) and (2.37a)–(2.37d). These different choices DStratonovichD transformation of this type readily takes should give the same result if the integration over the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1071

SH variables ⌬ is rigorously performed. However, ap- ¯ Z ͑⌬˜ ͒ϵ ⌿ ⌿¯ eϪ⌿Q⌿ proximations like the mean-field approximation lead in Q ͵ D D general to different results. Ϫ1 In the Fourier-transformed form, the partition func- ϭdet Qϭexp͓ln det Q0ϩln det͑1ϩQ0 Q1͔͒ tion is further rewritten by using the spinor notation ϭexp͓ϪTr ln͑ϪG0͒ϩTr ln͑1ϪG0Q1͔͒, ⌿¯ ϵ͑¯␺ ,¯␺ ͒, (2.131a) (2.134b) ↑ ↓ ␺ 2͑2Ϫy͒ 4y ⌿ϵ ↑ , (2.131b) S ϵ ⌬˜ *⌬˜ ϩ ⌬˜ 2ϩ2iy⌬˜ , (2.134c) ͩ␺ ͪ ⌬ U s s U c c ↓ in the form where Tr ln Aϭln det A is used. In Eq. (2.132c), the ma- trix Q is decomposed to ¯ ˜ * ˜ ˜ ϪS Zϭ ͵ ͓⌿,⌿͔ ⌬s ⌬s ⌬c e , (2.132a) D D D D QϭQ0ϩQ1 , (2.135a) 2͑2Ϫy͒ 4y Q k k ¯ ˜ ˜ ˜ 2 ˜ ͑ , Ј͒,0 Sϭ⌿Q⌿ϩ ⌬s*⌬sϩ ⌬c ϩ2iy⌬c , Q ϵ ↑↑ , (2.135b) U U 0 ͩ 0, Q ͑k,kЈ͒ͪ (2.132b) ↓↓ QЈ ͑k,kЈ͒ Q ͑k,kЈ͒ Q ϩQЈ Q Q ϵ ↑↑ ↑↓ . (2.135c) Qϭ ↑↑ ↑↑ ↑↓ , (2.132c) 1 ͩ Q ͑k,kЈ͒ QЈ ͑k,kЈ͒ͪ ͩ Q Q ϩQЈ ͪ ↓↑ ↓↓ ↓↑ ↓↓ ↓↓ The noninteracting Green’s function is defined as with G0 0 Q k k ϭ Ϫi ϩ Ϫ ↑ ͑ , Ј͒ ␦kkЈ͑ ␻n ␧0͑k͒ ␮͒, (2.133a) G0ϭ , (2.136a) ↑↑ ͩ 0 G ͪ 0 ↓ Q ͑k,kЈ͒ϭ␦kk ͑Ϫi␻nϩ␧0͑k͒Ϫ␮͒, (2.133b) ↓↓ Ј Ϫ1 G0␴͑k,kЈ͒ϭ␦kkЈ„i␻nϪ␰0͑k͒… , (2.136b) ˜ Q ͑k,kЈ͒ϭϪ͑2Ϫy͒⌬s͑kϪkЈ͒, (2.133c) ↑↓ ␰0ϵ␧0͑k͒Ϫ␮. (2.136c) ˜ Q ͑k,kЈ͒ϭϪ͑2Ϫy͒⌬s*͑kϪkЈ͒, (2.133d) ˜ ˜ ↓↑ ⌬ or ⌬*-dependent terms of S have the form ˜ QЈ ͑k,kЈ͒ϭQЈ ͑k,kЈ͒ϭϪ2iy⌬c͑kϪkЈ͒, (2.133e) 2͑2Ϫy͒ 4y ↑↑ ↓↓ ⌬˜ *͑k͒⌬˜ ͑k͒ϩ ⌬˜ *͑k͒⌬˜ ͑k͒ U ͚ s s U ͚ c c where kϵ(k,i␻n). In Eqs. (2.132a), (2.132b), and k k ␻nϾ0 (2.132c), the products ⌿¯ Q⌿ and ⌬˜ *⌬˜ are abbreviated 4y forms of the matrix products ˜ 2 ˜ ϩ ⌬c͑0͒ ϩ2iy⌬c͑0͒ϪTr ln͓1ϪG0Q1͔. (2.137) 1 U ⌿¯ Q⌿ϵ ⌿¯ ͑k,␻ ͒Q͑k,␻ ;kЈ,␻ ͒ V ͚ ͚ n n nЈ The mean-field (Hartree-Fock) approximation of the ͑␤ ͒ k,kЈ n,nЈ magnetic ordering is based on several approximations. ϫ⌿͑kЈ,␻nЈ͒ The first important assumption is that only static fluctua- ˜ ˜ ˜ and tions of ⌬s or ⌬c are important, so that ⌬c,s(k,␻n)at ␻Þ0 may be ignored. This static approximation dis- 1 n ˜ ˜ ˜ ˜ cards dynamic fluctuations though they are crucial in un- ⌬*⌬ϵ ͚ ͚ ⌬*͑k,␻n͒⌬͑k,␻n͒, ␤V k n derstanding anomalous features of metals near the Mott insulator. Another approximation in the Hartree-Fock respectively. The noninteracting dispersion ␧ (k) is the 0 ˜ Fourier transform of . theory is that it determines the static values ⌬c(␻nϭ0) K ˜ ˜ ˜ and ⌬s(␻nϭ0) by the saddle-point approximation. After Fourier transformation, ⌬c(k)ϭ⌬c*(Ϫk) is sat- More precisely, it is obtained by replacing the functional isfied. Therefore the functional integration over ⌬˜ (k) c ϪS ˜ ˜ may be written as integral of e over ⌬ and ⌬* with the saddle-point estimation of eϪS at ⌬(0) and ⌬*(0) obtained from ϭ0. So far, we have discussed possible spin and*⌬ץ/Sץ ⌬˜ ϵ d⌬˜ ͑k͒ϭd⌬ ͑kϭ0,␻ ϭ0͒ c ͟ c c n ˜ ˜ D k charge fluctuations by introducing ⌬c and ⌬s . Hereafter, ˜ for simplicity, we take ⌬c(0)ϵy⌬c(0) and ⌬s(0)ϵ(2 ϫ d˜ * d˜ ˜ ⌬c ͑k,␻n͒ ⌬c͑k,␻n͒. Ϫy)⌬s(K) with the staggered antiferromagnetic order- ͟k ␻͟Ͼ0 n ing wave vector kϭK. The condition of extremum is After integrating the Grassmann variables ⌿ and ⌿¯ , explicitly written as we are led to Q1ץ S 2⌬sץ ϭ ϩTr ͑1ϪG Q ͒Ϫ1G ϭ0 0͒ͬ͑*⌬ץ U ͫ 0 1 0 *⌬ץ Z͑⌬˜ ͒ϵ ⌬˜ * ⌬˜ ⌬˜ Z ͑⌬˜ ͒eϪS⌬, (2.134a) s s ͵ D s D s D c Q (2.138a)

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1072 Imada, Fujimori, and Tokura: Metal-insulator transitions and Qץ ⌬S 4ץ c Ϫ1 1 ϭ ϩTr͑1ϪG0Q1͒ G0 ϩ2iϭ0. (2.138b) c͑0͒⌬ץ c* U⌬ץ Equation (2.138a) is nothing but the self-consistent equation for magnetic ordering at the wave vector K, which may be rewritten as

⌬s͑0͒ 1 ⌬s͑0͒ ϭ ͚ 2 (2.139) U ␤V k,␻n ͑Ϫi␻nϩ␧0Ϫ␮͒͑Ϫi␻nϪ␧0Ϫ␮͒Ϫ͉⌬s͑0͉͒

or 2 2 E͑k͒ϭϮͱ␧0͑k͒ ϩ͉⌬s͑0͉͒ Ϫ␮. (2.141)

⌬s͑0͒ 1 ⌬s͑0͒ ϭ U ␤V ͚ 2 2 k,␻n ͑Ϫi␻nϩͱ␧0͑k͒ ϩ͉⌬s͑0͉͒ Ϫ␮͒ The saddle-point solution of ⌬c(0) for Eq. (2.138b) gives 1 ϫ . (2.140) 2 2 ͑i␻nϪͱ␧0͑k͒ ϩ͉⌬s͑0͉͒ Ϫ␮͒ Ui ⌬c͑0͒ϭ ͑͗n͘Ϫ1͒ (2.142) where we have assumed ␧0(kϩK)ϭϪ␧0(k)asinthe 4 case of the antiferromagnetic order of the nearest- neighbor Hubbard model. The extension to more gen- eral cases is straightforward. The quasiparticle excitation where ͗n͘ is the average electron density. The single- from this broken-symmetry state is given by the disper- particle Green’s function in the Hartree-Fock approxi- sion mation is

Ϫ1 Ϫ1 G␴␴Ј͑k,␻n ;kЈ,␻nЈ͒ϵϪQ ϭ 2 ͓Ϫi␻nϩ␧0͑k͒Ϫ␮͔͓Ϫi␻nϪ␧0͑k͒Ϫ␮͔Ϫ͉⌬s͑0͉͒

͑Ϫi␻nϪ␧0Ϫ␮͒␦k,kЈ␦n,nЈ , ⌬s͑0͒␦kЈ,kϩK␦n,nЈ ϫ , ͩ ⌬s͑0͒␦kЈ,kϩK␦n,nЈ , ͑Ϫi␻nϩ␧0Ϫ␮͒␦k,kЈ␦n,nЈͪ

where 2 U ¯ 2 S␥ϭ⌫Q␥⌫ϩ ͉⌬s͑0͉͒ Ϫ ͑nϪ1͒͑͗n͘Ϫ1͒ (2.145a) G G U 4 Gϵ ↑↑ ↓↑ . (2.143) ͩ G G ͪ ↑↓ ↓↓ After the Bogoliubov transformation Q␥ 0 Q ϭ ↑↑ , (2.145b) ␥ ͩ 0 Q ͪ ␥ ␥ ͑k͒ ␺ ͑k͒ ↓↓ ⌫ϭ ↑ ϭU ↑ (2.144a) ͩ ␥ ͪ ͩ ␺ ͑k͒ͪ ͑k͒ ↓ ↓ with Q␥␴␴ϭ␦kkЈ„Ϫi␻nϪ␮ϩ␴E͑k͒…. (2.145c) u͑k͒, v͑k͒ Uϭ , (2.144b) The Green’s function for the quasiparticle is therefore ͩ Ϫv͑k͒, u͑k͒ͪ and G␥␴␴Ј͑k,kЈ͒ϵϪ͗␥␴͑k͒␥␴͑kЈ͒͘ 1 ␧0͑k͒ u͑k͒2ϭ 1ϩ , (2.144c) 1 2 ͩ E͑k͒ ͪ ϭ␦␴␴Ј␦kkЈ . (2.146) i␻nϩ␮Ϫ␴E͑k͒ 1 ␧0͑k͒ v͑k͒2ϭ 1Ϫ , (2.144d) 2 E k ͩ ͑ ͒ ͪ We next discuss how the random-phase approxima- ˜ the action (2.132b) in the Hartree-Fock approximation tion is given by the Gaussian approximation for ⌬s . The is diagonalized, using the quasiparticles ␥, into the form dynamic spin susceptibility defined by

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1073

1 ␹͑q,␻ ͒ϭ ͗TSϩ͑q,␻ ͒SϪ͑q,␻ ͒͘ (2.147) m 4 m m

1 ˜ ˜ ϭ ⌬s* ⌬s 4␤VZ ͵ D D

2ץ ϪS⌬ ˜ ϫe ZQ͑⌬͒ (2.148) ˜ ˜ s͑q,␻m͒⌬ץs*͑q,␻m͒⌬ץ 1 2 ϭ ͗⌬˜*͑q,␻ ͒⌬˜ ͑q,␻ ͒͘Ϫ␤V 2␤VU ͩU s m s m ͪ (2.149) is obtained from Eqs. (2.134a)–(2.134c) after partial in- tegration. In the paramagnetic phase, up to second order in Q1 for Eq. (2.134), we obtain ˜ ZQ͑⌬͒Ӎexp͓ϪTr ln͑ϪG0͒ϪTr G0Q1

ϪTr G0Q1G0Q1͔. (2.150)

The term linearly proportional to Q1 (the second term) vanishes and the last term is

˜ ˜ Tr G0Q1G0Q1ϭ8 ␹b͑q,␻m͒⌬s*͑q,␻m͒⌬s͑q,␻m͒, q͚,m (2.151) FIG. 10. Hartree-Fock phase diagram of three-dimensional where ␹ (q,␻ ) is the spin susceptibility at Uϭ0. b m nearest-neighbor Hubbard model in the plane of U and a half Therefore, combined with the term S in Eq. (2.134), ⌬ of filling ͗n͘ (Penn, 1966). Here the ordinate C /EЉ represents ˜ ˜ 0 ͗⌬s*⌬s͘ is estimated as U/2t in our notation. ANTI, antiferromagnetic; PARA, para- U 1 magnetic; FERRO, ferromagnetic; S. FERRI, spiral ferrimag- ˜ ˜ netic; and S.S.SDW, spiral spin-density-wave phases. ͗⌬s*͑q,␻m͒⌬s͑q,␻m͒͘ϭ ␤V , 2 1Ϫ2U␹b͑q,␻m͒ (2.152) Korenman, Murray, and Prange, 1977a, 1977b, 1977c; which leads to Schulz, 1990c), in which case the partition function Z

␹b͑q,␻m͒ given in Eq. (2.130) is rewritten as ␹͑q,␻m͒ϭ . (2.153) 1Ϫ2U␹b͑q,␻m͒ This is nothing but the RPA result. The Hartree-Fock phase diagram calculated by using the above procedure is shown in Fig. 10 for the 3D nearest-neighbor Hubbard model (Penn, 1966) and in Fig. 11 for the 2D Hubbard model (Hirsch, 1985b). Here it is assumed that the or- dering vector K is either antiferromagnetic [Kϭ(␲,␲) or (␲,␲,␲)] or ferromagnetic [Kϭ(0,0) or (0,0,0)] with the lattice constant unity. The characteristic feature is a rather wide region of antiferromagnetic and ferromag- netic metals near the Mott insulator, nϭ1. We discuss more complicated types of spin or charge orderings in Sec. II.H.3 and IV. So far we have restricted our choice of the Stratonovich-Hubbard transformation with ⌬s corre- † † † sponding to the decoupling of ci ci ci ci into ci ci and c† c as in Eq. (2.129a). However,↑ ↓ this↓ ↑ transformation↑ ↓ FIG. 11. Hartree-Fock phase diagram of two-dimensional i i nearest-neighbor Hubbard model in the plane of U/t and fill- makes↓ ↑ possible symmetry breaking only in the xy plane ing ͗n͘ϵ␳ (Hirsch, 1985a, 1985b). P, A, and F represent para- of the spin space. To retain the SU(2) symmetry even magnetic, antiferromagnetic, and ferromagnetic phases, re- ˜ជ after the SH transformation, a vector SH field ⌬s may be spectively. The possibility of incommensurate order is not introduced (C. Herring, 1952; Capellmann, 1974, 1979; considered here.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1074 Imada, Fujimori, and Tokura: Metal-insulator transitions

for particle number, energy, and momentum. Some ¯ ˜ជ ˜ជ ˜ Zϭ ͓␺,␺͔ ⌬si͑␶͒ ⌬si*͑␶͒ ⌬ci͑␶͒ higher-order diagrams were also considered in addition ͵ D D D D to the Hartree-Fock diagram so as to retain the above ץ ␤ ¯ conservation laws (Bickers and Scalapino, 1989; Bickers ϫexp Ϫ d␶ ͚ ␺i␴ ␺i␴ and White, 1991), which may be viewed as an attempt to ␶ץ i ͩ 0͵ ͫ include perturbatively fluctuation effects that are ig- ¯ nored in the Hartree-Fock approximation. This is called Ϫ ␺i␴͑ ijϪ␮␦ij͒␺j␴ ͚ij K the fluctuation exchange (FLEX) approximation.

2͑2Ϫy͒ ˜ជ ˜ជ ϩ ⌬si*͑␶͒•⌬si͑␶͒ 3. Local-density approximation and its refinement 3U ͚i In a number of ways, as we see in this article, theoret- ͑2Ϫy͒ ˜ជ ¯ ical understanding of strongly correlated systems has Ϫ ͚ ͑⌬si*•␺i␴S␴␴Ј␺i␴Ј 3 i been made possible through studies on effective lattice Hamiltonians derived from the tight-binding approxima- 4y ˜ជ ¯ ˜ 2 tion for a small number of bands near the Fermi level. ϩ⌬si͑␶͒•␺i␴S␴Ј␴␺i␴Ј͒ϩ ͚ ⌬ci͑␶͒ U i Although the simplification to effective tight-binding Hamiltonians is useful to elucidate many universal fea- ˜ ¯ ¯ tures of correlation effects, detailed experimental data Ϫ2iy͚ ⌬ci͑␶͒͑␺i ␺i ϩ␺i ␺i Ϫ1͒ (2.154) i ↑ ↑ ↓ ↓ ͪͬ such as those from photoemission often have to be ana- lyzed from a more original Hamiltonian which contains ˜ជ where ⌬s•S␴␴Ј is an internal product with more or less all the electronic orbitals in the continuum space under the periodic potential of the real lattice. ͑S␴␴ ͒xϭ␦␴ ␦␴ ϩi␦␴ ␦␴ , (2.155a) Ј ↑ Ј↓ ↓ ↑ This is because high-energy spectroscopy reflects the in- ͑S␴␴Ј͒yϭ␦␴ ␦␴Ј Ϫi␦␴ ␦␴ , (2.155b) dividual character of the compounds. Electronic struc- ↑ ↓ ↓ ↑ ture calculations starting from atoms and electrons with 1 their mutual Coulomb interactions in the continuum ͑S␴␴ ͒zϭ ͑␦␴ ␦␴ Ϫ␦␴ ␦␴ ͒. (2.155c) Ј 2 ↑ Ј↑ ↓ Ј↓ space are sometimes called first-principles calculations or ab initio calculations. The expected role of these first- ˜ជ Here we note that ⌬ci is real while ⌬si is a complex principles calculations is not restricted to quantitative ˜ជ three-dimensional vector. In particular, ⌬si is rewritten analyses of high-energy spectroscopy and the structural as stability of lattices but is also extended to derive or jus- tify starting effective Hamiltonians of simplified lattice ˜ជ i␪i ⌬siϭ⌬s0e nជi (2.156) fermion models for each compound and to bridge the gap between high- and low-energy physics. For example, where ⌬s0 is a scalar for the amplitude of this Hartree- band-structure calculations provide a generally accepted Fock field while ␪i is the phase and nជ is the unit vector way of determining parameters of effective lattice mod- ˜ជ to specify the direction of ⌬s in the 3D spin space. In the els if carefully examined. It is also expected that the phase of the ground state will be correctly reproduced magnetically ordered phase, ⌬s0 has a finite amplitude, so that the excitation of this amplitude mode has a finite without introducing any parameter fitting. In fact, we describe below several treatments that have been suc- gap. In contrast, ␪i and nជ i represent the gapless excita- tion mode. Therefore low-energy excitations are ex- cessful in reproducing correct ground states of transition-metal compounds. ជ hausted by the excitations through ␪ and n. Even when As it stands, ‘‘first-principles’’ calculation does not there is no symmetry breaking, provided that the corre- mean a rigorous, practical way of determining electronic lation length is sufficiently long, the low-energy excita- structure, contrary to what one might imagine from its tion can be described only through ␪ and nជ . This is in- terminology. Basically, three types of approximations deed the case for low-dimensional systems below the have been employed to make the ‘‘first principles’’ cal- mean-field transition temperature TMF where quantum culation tractable in practice. fluctuations destroy the symmetry breaking while the One is the configuration-interaction (CI) method, in amplitude itself is developed below TMF . The low- which small clusters under the atomic potential with all energy action described by ␪ and nជ frequently contains the relevant electronic orbitals are solved by the diago- topological terms that cause singular and nonlinear ex- nalization of a Hamiltonian matrix. This method is exact citations in the continuum limit (for example, for 1D; but applicable only for small clusters, typically a few unit Haldane, 1983a, 1983b). Except in the 1D case, the role cells. Therefore it frequently underestimates the effect of these topological terms, such as the Berry phase term, of electronic kinetic energy and coherence. We further has not been fully clarified yet. We discuss the low- discuss details of the configuration-interaction method energy effective Hamiltonian for this case later. in Sec. III.A.1. This method has a computational simi- The Hartree-Fock approximation is a self-consistent larity to the exact diagonalization method reviewed in approach which satisfies microscopic conservation laws Sec. II.E.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1075

The second type of approximation is the Hartree- 1964; Kohn and Sham, 1965). A general and exact state- Fock approximation, which is one of the best ways to ment of the density-functional theory is that the true approximate a true many-body fermion state by a single ground-state energy is obtained by two steps: First, one Slater determinant. By taking a single Slater determi- finds the many-body state ⌿͓n(r)͔ that minimizes the nant as a variational state of the many-body Hamil- expectation value of the total many-body Hamiltonian, E n n tonian g͓ (r)͔, under given single-particle density (r). Sec- ond one minimizes Eg͓n(r)͔ as a functional of n(r). An ប2 e2 important point is the existence of the energy functional ϭϪ ⌬iϩ Vlat͑ri͒ϩ , E ͓n(r)͔. Of course these two steps are difficult to per- H ͚i 2m ͚i ͚ ͉r Ϫr ͉ g e ͗ij͘ i j form in practice. A basic underlying assumption in prac- (2.157) tical applications of the Kohn-Sham theory is that one determines the single-particle state ␺l(ri) from the Eg͓n(r)͔ can be calculated by starting from single- minimization of energy by solving the self-consistent particle states given by a single Slater determinant that equation reproduces n(r), with this single-particle state being ob- tained as a solution of the Schro¨ dinger equation ប2 e2 Ϫ ⌬ϩV ͑r͒ϩ drЈ ͉␺ ͑rЈ͉͒2 ␺ ͑r͒ ប2 lat ͚ ͵ Ϫ lЈ l 2m l ͉r rЈ͉ Ϫ ⌬ϩV͑r͒ ␺ ͑r͒ϭ␧ ␺ ͑r͒ (2.162) ͫ Ј ͬ ͩ 2m ͪ l l l 2 e with V(r) chosen to give Ϫ Ј dr ␺*͑rЈ͒␺ ͑rЈ͒ ␺ ͑r͒ ͚ ͫ͵ ͉rϪrЈ͉ lЈ l ͬ lЈ lЈ 2 n͑r͒ϭ͚ ͉␺l͑r͉͒ . (2.163) ϭ␧l␺l͑r͒. (2.158) l Here, l and lЈ specify quantum numbers including or- We discuss this procedure in greater detail below. In the bital degrees of freedom, and the sum ͚Ј is taken only original many-body problem, the ground-state energy is over lЈ, which has the same spin as that of l. The atomic obtained through minimization of the energy expecta- periodic potential of the lattice is denoted as an external tion value one by Vlat . The electron-electron Coulomb interaction E͓n͑r͔͒ϭ͗⌿͓n͑r͔͉͒ Kϩ extϩ U͉⌿͓n͑r͔͒͘ is considered through the third term, the Hartree term, H H H (2.164) and the fourth (last) term, the exchange term. The ex- with respect to variations of both ⌿͓n(r)͔ and the den- change term can be rewritten as sity functional n(r), where ⌿͓n(r)͔ is a many-body l wave function ⌿(r ,r ,...,r ) that satisfies Vex͑r͒␺l͑r͒, (2.159) 1 2 N

2 2 l l e dr dr ,...,dr ͉⌿͓n͑r͔͉͒ ϭn͑r ͒. (2.165) V ͑r͒ϭ drЈ␳ ͑r,rЈ͒ (2.160) ͵ 2 3 N 1 ex ͵ ex ͉rϪrЈ͉ The kinetic energy, interaction energy, and external with atomic potential are denoted as , , and , re- HK HU Hext ␺*͑r͒␺ ͑rЈ͚͒Ј ␺*͑rЈ͒␺ ͑r͒ spectively. After minimization with respect to ⌿ with l l l lЈ lЈ lЈ ␳ex͑r,rЈ͒ϭϪ 2 (2.161) fixed n(r), the minimum in terms of n(r) is obtained ͉␺l͑r͉͒ from where the exchange interaction is expressed as if it were ␦Eg͓n͑r͔͒ a static Coulomb term. A computational difficulty of the ϭ0, (2.166) Hartree-Fock approximation is that the exchange poten- ␦n͑r͒ l tial Vex(r) depends on the solution ␺l(r) itself, and dif- where Eg͓n(r)͔ is the minimized value with respect to ferent exchange potentials must be taken for each quan- ⌿. An important point here is that, in this step, the tum number l while the number of states l to be ground-state energy is calculated only from the energy considered increases linearly with the system size. It functional Eg͓n(r)͔ without any detailed knowledge of should be noted that the Hartree-Fock approximation the many-body wave function. Of course, the functional does not take into account correlations between elec- form of Eg is not easy to obtain. trons with different spins and hence an up-spin electron When we can assume that the first minimization pro- behaves independently of the down-spin electrons if cedure to obtain Eg͓n(r)͔ may be performed within the there is no symmetry breaking of spins such as the spin- single Slater determinant under an appropriate fictitious density wave state as in Sec. II.D.2. To take into account external potential V(r), the minimization may be per- short-ranged correlation effects of opposite-spin elec- formed from the minimization of the energy functional, trons, which is important in strongly correlated systems, one has to go beyond the Hartree-Fock approximation. E͓n͑r͔͒ϭEK͓n͑r͔͒ϩ dr Vlat͑r͒n͑r͒ The third approximation is the local-density approxi- ͵ mation, which gives a basic starting point for widely e2 n͑r͒n͑rЈ͒ used band-structure calculations. The local-density ap- ϩ drdrЈ ϩ drV ͑r͒n͑r͒, ͵ Ϫ ͵ ex proximation is based on the density-functional theory of 2 ͉r rЈ͉ Hohenberg, Kohn, and Sham (Hohenberg and Kohn, (2.167)

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1076 Imada, Fujimori, and Tokura: Metal-insulator transitions with der, 1980) of the ground-state energy for a uniform elec- tron gas with density n(r) (Perdew and Zunger, 1981). e2 From the difference between the Monte Carlo result V ͑r͒ϵ drЈ␳ ͑r,rЈ͒ , (2.168) ex ͵ ex ͉rϪrЈ͉ and the corresponding energy for the first three terms of Eq. (2.167), Vex is estimated. 2 The practical difficulties of applying the density- ͉␺l͑r͉͒ l ␳ex͑r,rЈ͒ϵ ␳ex͑r,rЈ͒. (2.169) functional procedure to strongly correlated systems can ͚l n͑r͒ be appreciated from the difficulty of calculating Vex(r), Here, the first term of Eq. (2.167), EK , is the kinetic- into which all the many-body effects are squeezed. The n energy term written as a functional of (r). For ex- assumption that Vex(r) is determined from the local ample, EK is given from Eq. (2.162) as density n(r) is in general not satisfied in strongly corre- lated systems. Failure of the LDA calculation is widely observed in EKϭ͚ ͓␧lϪ͗␺l͉V͑r͉͒␺l͔͘. (2.170) l the Mott insulating phase of many transition-metal com- The sum over l in Eq. (2.170) is taken to fill the lowest N pounds as well as in phases of correlated metals near the Mott insulator. MnO, NiO, NiS, YBa2Cu3O6, and levels of ␧l for an N-particle system. La2CuO4 are well known examples of the LDA’s inabil- However, the assumption that Eg͓n(r)͔ may be ob- tained within the single Slater determinant is in general ity to reproduce the insulating (or semiconducting) not justified in correlated systems. Therefore we leave phase (Terakura, Oguchi, Williams, and Ku¨ ller, 1984; Pickett, 1989; Singh and Pickett, 1991). the form of Vex(r) unknown and just define it as the difference between the true ground-state energy and the To improve the LDA, several different approaches contribution from the first three terms in Eq. (2.167). have been taken. The local spin-density approximation (LSDA) explicitly introduces the spin-dependent elec- The point is that Vex(r) is still represented by a func- tional of n(r). We also note, however, that there exists tron densities n (r) and n (r) separately, to allow for ↑ ↓ no proof that V(r) always exists, though its existence is possible spin-density waves or antiferromagnetic states necessary in this procedure, and in this approach it is (von Barth and Hedin, 1972; Perdew and Zunger, 1981). assumed to exist. The minimization of Eq. (2.167) in Instead of Eq. (2.167), the exchange-correlation energy terms of the functional n(r) gives the condition ͐drVex(r)n(r) is replaced with ͐dr͓n (r) ↑ ϩn (r)͔Vex͓n (r),n (r)͔ in the LSDA, where Vex is taken↓ as a functional↑ ↓ of both n (r) and n (r). Corre- ␦EK n͑rЈ͒ Ϫ ϭV ͑r͒ϩe2 drЈ spondingly, the Kohn-Sham equations↑ (2.162),↓ (2.163), ␦n͑r͒ lat ͵ ͉rϪrЈ͉ (2.170), and (2.171), which need to be solved self- ␦ consistently, are replaced with spin-dependent forms, ϩ drV ͑r͒n͑r͒. (2.171) ␦n͑r͒ ͵ ex ប2 Ϫ ⌬ϩV ͑r͒ ␺ ͑r͒ϭ␧ ␺ ͑r͒, (2.173) The ground-state energy is obtained by solving the ͩ 2m ␴ ͪ l␴ l␴ l␴ Kohn-Sham equations (2.162), (2.163), and (2.171) self- consistently, if we know the functional dependence of V (r)on n(r). In general, the exchange potential 2 ex ͕ ͖ n␴͑r͒ϭ ͉␺l␴͑r͉͒ , (2.174) ͚l Vex(r) depends on n(rЈ) for rÞrЈ in a nonlocal fashion. The local-density approximation (LDA) assumes that Vex(r) is determined only by n(rЈ)atrЈϭr. This as- n͑r͒ϭ n␴͑r͒, (2.175) sumption may be justified when effects of spatial varia- ͚␴ tions of the electron density may be neglected in the exchange potential, so that Vex(r) is calculated from the uniform electron gas with density n(r). This condition EKϭ͚ ͑␧l␴Ϫ͗␺l␴͉V␴͑r͉͒␺l␴͒͘, (2.176) for justification is explicitly given as l␴

␦ n͑r͒ EK n͑rЈٌ͒ Ӷ1 (2.172) Ϫ ϭV ͑r͒ϩe2 drЈ k ͑r͒n͑r͒ lat ͵ Ϫ F ␦n␴͑r͒ ͉r rЈ͉ provided that the Fermi wave number kF can be de- ␦ scribed as a local variable under the same assumption of ϩ drV ͑r͒n͑r͒. (2.177) ␦n ͑r͒ ͵ ex slow modulation of the electron density. ␴ Because we have left Vex(r) just as the difference of In the LSDA, Vex(r) is assumed to be determined only the true ground-state energy and the contribution from from the local spin densities n (rЈ) and n (rЈ)atrЈϭr, ↑ ↓ the first three terms in Eq. (2.167), it has to be calculated and usually Vex is given from the energy difference be- from other explicit calculations taking account of the tween the quantum Monte Carlo result and the single- correlation effects honestly. This is usually done by us- particle solution as in the LDA. For example, the LSDA ing the quantum Monte Carlo result (Ceperley and Al- succeeds in reproducing the presence of the band gap at

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1077 the Fermi level for MnO by allowing antiferromagnetic many-body systems by introducing the self-energy cor- order (Terakura, Oguchi, Williams, and Ku¨ ller, 1984). It rection to the single-particle Green’s function obtained also succeeds in reproducing the antiferromagnetic insu- from LDA-type calculations. This attempt is called the lating state of LaMO3 (MϭCr,Mn,Fe,Ni) (Sarma et al., ‘‘GW’’ approximation (Hedin and Lundqvist, 1969). It 1995). was applied to NiO (Aryasetiawan, 1992), although sys- However, the LSDA fails to reproduce the antiferro- tematic studies on transition-metal compounds are not magnetic ground state of many strongly correlated com- available so far (for a review, see Aryasetiawan and pounds such as NiS, La2CuO4, YBa2Cu3O6, LaTiO3, Gunnarsson, 1997). Effects of local three-body scatter- and LaVO3. Even in MnO, strongly insulating behavior ing on the self-energy correction have also been exam- observed well above the Ne´el temperature implies that ined (Igarashi, 1983). the charge-gap structure is not a direct consequence of To counter the general tendency to underestimate the antiferromagnetic order, although antiferromagnetic or- band gap of the Mott insulator in LDA-type calcula- der is needed to open a gap in the LSDA calculation. tions, a combination of the LDA and a Hartree-Fock- The LDA band-structure calculation makes a drastic type approximation called the LDAϩU method was approximation in that the exchange-correlation energy proposed (Anisimov, Zaanen, and Andersen, 1991). The Vex(r) is determined only from the local electron den- Coulomb interaction for double occupancy of the same sity n(r). When the electron density n(r) has a strong site as in the Hubbard model requires a large energy U spatial dependence, this approximation breaks down. if the eigenstate is localized as compared to extended This is indeed the case for strongly fluctuating spin and states. This is not correctly taken into account in the charge degrees of freedom in correlated metals. To im- LDA calculation, which estimates the exchange interac- prove this local-density approximation, the gradient ex- tion using information from the uniform electron gas, pansion of n(r) may be considered to take into account which is appropriate only for extended wave functions. the nonlocal dependence of Vex(r)onn(rЈ), where rЈ In the LDAϩU approach the contribution of the inter- Þr. Langreth and Mehl (1983), Perdew and Wang action proportional to U is added to the LDA energy in (1986), Perdew (1986), and Becke (1988) developed a an ad hoc way when the orbital is supposed to be local- way of including the gradient expansion of n(r) in the ized. The method thus combines arbitrariness in specify- exchange energy by keeping the constraint of negative ing the localized orbital with a drastic approximation in density of the exchange hole everywhere in space with a the treatment of the U term (for example, the Hartree- total deficit of exactly one electron. This is called the Fock approximation was used.) The LDAϩU method generalized gradient approximation (GGA) method. It has the same difficulty as the other many-body ap- was applied to reproduce the bcc ferromagnetic phase of proaches in the tight-binding Hamiltonian because it Fe by Bagno, Jepsen, and Gunnarson (1989). Recently does not necessarily provide a reliable way of treating U this method was applied to the transition-metal com- term, while, in a sense, all the important effects of strong pounds FeO, CoO, FeF2, and CoF2 with partial im- correlation are relegated to this term. Another problem provements of the LDA results. It was also applied to of this method is that the estimate of U may depend on transition-metal oxides with 3D perovskite structure, the choice of the basis (Pickett et al., 1996). Different LaVO3, LaTiO3, and YVO3 (Sawada, Hamada, Tera- choices of local atomic orbitals such as the linear muffin- kura, and Asada, 1996). It showed improvements in re- tin orbital (LMTO) and linear combination of atomic producing the band gap in LaVO3, but the ground state orbitals (LCAO) give different U because the definition was not correct for YVO3 and LaTiO3. of the orbital becomes ambiguous whenever there is A serious problem in LDA and LSDA calculations is strong hybridization. that the self-interaction is not canceled between the con- To take account of frequency-dependent fluctuations tributions from the Coulomb term and the exchange in- ignored in the LDA method, Anisimov et al. combined teraction term because of the drastic approximation in the LDA method with the dynamic mean-field approxi- the exchange-interaction term. This produces a physi- mation (the dϭϱ method described in Sec. II.D.6; cally unrealistic self-interaction, which makes the ap- Anisimov et al., 1997). They treated strongly correlated proximation poor for strongly correlated systems. Per- bands by the dynamic mean-field approximation and de- dew and Zunger (1981) proposed imposing a constraint scribed other itinerant bands by the LDA with a linear to remove the self-interaction. This method of self- muffin-tin orbital basis. Recently, the exchange- interaction correction (SIC) was applied to the Hubbard correlation energy functional was calculated as the sum model as well as to models of transition-metal oxides of the exact exchange energy and a correlation energy with quantitative improvement of the band gap and the given from the RPA without introducing the local- magnetic moment for MnO, FeO, CoO, NiO, and CuO density approximation (Kotani, 1997). (Svane and Gunnarsson, 1988a, 1988b, 1990). As we have seen above, in this decade substantial In principle, the Kohn-Sham equation provides the progress has been achieved in the band-structure calcu- procedure to calculate exactly the ground state at the lation of transition-metal compounds. However, it is still Fermi level. However, it does not guarantee that excita- not sufficient to reproduce the Mott insulator correctly tions will be well reproduced. A possible way of improv- in strongly correlated materials such as La2CuO4 and ing estimates for the excitation spectrum is to combine YBa2Cu3O6. A more serious problem lies in the descrip- the Kohn-Sham equation with a standard technique for tion of the metallic state near the Mott insulator, where

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1078 Imada, Fujimori, and Tokura: Metal-insulator transitions spin and charge fluctuations play crucial roles even in where the spatial (or wave-number) dependence of the the ground state, while the band-structure calculation self-energy is neglected. The probability to have the has difficulty in including such fluctuation effects. For ˜ (l) value ⌬c is denoted by Pl . Originally, Pl was deter- example, the continuous MIT should be described by mined from the Gaussian approximation obtained from the continuous reduction of coherent excitation when exp͓ϪS ͔ with S in Eq. (2.134c). However, in the Hub- approaching the insulator, while the available ap- ⌬ ⌬ bard approximation, as we describe below, P is taken proaches mentioned in this subsection fail to describe l rather differently, in order to consider nonlinear fluctua- important contributions from the increasing weight of incoherent part. tions ignored in the Gaussian approximation. We shall see in Sec. II.D.6 that ⌺ is not dependent on wave num- 4. Hubbard approximation ber in infinite dimensions, although, in general, it is wave-number dependent in finite dimensions. Although To gain insight into the original suggestion by Mott, ⌺(␶) is exactly calculated in infinite dimensions in Sec. Hubbard (1963) developed the coherent-potential ap- II.D.6, the Hubbard approximation calculates it only ap- proximation (CPA), which is combined with a Hartree- proximately as follows: The partition function (2.178a) is Fock-type approximation to address the problem of the metal-insulator transition. The CPA is a technique de- rewritten in the same manner as Eq. (2.134), veloped to treat the effects of random potentials (see, for example, Lloyd and Best, 1975), which Hubbard had ZlϭPlexp͓ϪTr ln͑ϪG0͒ϩTr ln͑1ϪG0Q1l͔͒ already developed before its widespread application (2.179) (see also Kawabata, 1972). In the Hubbard approxima- tion, the Stratonovich-Hubbard transformation, using with ˜ only the charge fluctuations ⌬ci(␶) in Eq. (2.130), is em- ˜ ͑l͒ ployed to obtain a static approximation represented by ⌬c 0 the Stratonovich-Hubbard field; the static field is re- Q1lϭϪ4i . (2.180) ͩ 0 ⌬˜ ͑l͒ͪ garded as a random potential in the CPA. Below some c details of the Hubbard approximation are given. This may be rewritten as ˜ In the CPA, the ␶ dependence of ⌬c(r,␶) is neglected in the same way as in the Hartree-Fock approximation. Z ϭP exp͕ϪTr͓ln͑ϪG͒Ϫln 1ϪG͑Q Ϫ⌺͒ ͔͖, After integrating out the Grassmann variables, one re- l l „ 1l … (2.181) ˜ places ⌬c(r,␶) with the ␶-independent constants ⌬c(r), which is called the static approximation. In contrast to where the Green’s function G has the form the Hartree-Fock approximation, ⌬c(r) is assumed to ˜ (l) Ϫ1 Ϫ1 take more than one value ⌬c (lϭ1,2,...) in a random G ͑k,␻n͒ϭG0 ͑k,␻n͒Ϫ⌺͑␻n͒, (2.182) way from site to site. In fact, in the Hubbard approxi- ˜ with ␻-dependent self-energy ⌺. Now the on-site mation, two values 0 and Ui/4 are assumed for ⌬c . These two values are obtained from the saddle-point so- Green’s function Gloc(␻n) is defined as lution (2.142) if we take either ͗n͘ϭ1or͗n͘ϭ2. These two values can intuitively be understood because an electron at a doubly occupied site feels a static interac- Gloc͑␻n͒ϭ G͑k,␻n͒. (2.183) ͚k tion potential U while its potential is absent for a singly occupied site if all the other electrons are frozen so that Then, to include locally tractable and averaged effects in dynamic effects can be neglected. (As we shall see later the second term of Eq. (2.181), we rewrite it as in this subsection, in the CPA, dynamic effects are par- tially taken into account in another way by introducing ZlϭPlexp͓ϪTr͕ln͑ϪG͒Ϫln͓1ϪGloc͑Q1lϪ⌺͔͒ afterwards an ␻-dependent self-energy ⌺ as determined variationally in the single-particle excitation spectrum.) Ϫln͓1Ϫ͑GϪG ͒ ͔͖͔, (2.184) loc Tl As in Eq. (2.130) the partition function is first assumed where is a t matrix, to have the form Tl ϭ͑Q Ϫ⌺͓͒1ϪG ͑Q Ϫ⌺͔͒Ϫ1. (2.185) Tl 1l loc 1l Zϭ͚ Zl , (2.178a) l The CPA is regarded as the lowest-order expansion of the t matrix, since it neglects the last term in (2.184). Since it determines ⌺ in a variational way, the condition ץ ¯ ¯ ZlϭPl͟ ␺␴ ␺␴exp Ϫ dx␺␴͑x͒ ␺␴͑x͒ for determining is Z ϭ ⌺ 0. This leads toץ/ ץ ␶ ⌺ץ ͵ ͫ ␴ ͵ D D

Gץ Ϫ dx ¯␺ ͑x͒͑ Ϫ␮͒␺ ͑x͒ 2 loc ␴ ␴ ͚ ZlTr GlocϪ l ϭ0, (2.186) ͮ ⌺ ͪTץ ͩͭ K l ͵ 4 ˜ ͑l͒ ˜ ͑l͒2 from which we obtain the self-consistent equation for ϩ4i⌬c ͑n ͑x͒ϩn ͑x͒Ϫ1͒Ϫ ⌬c , (2.178b) ↑ ↓ U ͬ the self-energy,

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1079

Ϫ1 Ϫ1 ⌺ϭ͚ ZlQ1l͕1ϪGloc͑Q1lϪ⌺͖͒ ͚ Zl͕1ϪGloc͑Q1lϪ⌺͖͒ . (2.187) l Ͳ l

In the case of the Hubbard approximation for the nearest-neighbor Hubbard model at half filling, Q ͉⌿͘ϭ ͑1Ϫgnj nj ͉͒⌽͘, (2.188) 11 ͟j ↑ ↓ ϭ0, Q12ϭU, P1ϭ1/2, and P2ϭ1/2, where the number of l’s is taken as two. When ⌺ is determined by solving where ⌽ is a single-particle state represented by a Slater Eq. (2.187), Z is obtained from Eq. (2.184) by putting determinant and g is a variational parameter to take ϭ0. The procedure of the Hubbard approximation is into account the reduction of double occupancy of elec- Tl summarized as follows: trons at the same site. The free-fermion determinant was (1) Take a variational form of G and ⌺. originally taken as the Slater determinant. It may be (2) Calculate Gloc using Eq. (2.183) and Zl using Eq. generalized to cases with other types of single Slater de- (2.184). terminant where some symmetry is broken, as in the (3) Calculate ⌺ from Eq. (2.187) by substituting the spin-density wave state or in the superconducting state. present form of ⌺, Gloc , and Zl in the right-hand side. The ground-state energy (4) Using the form of ⌺ obtained in step (3), calculate G from Eq. (2.182). Egϭ͗⌿͉ ͉⌿͘/͗⌿͉⌿͘ (2.189) H (5) From the form of G and ⌺ determined in steps (3) is estimated from optimization of the parameter g and and (4), repeat the process from step (2) again until the the choice of a single Slater determinant. This estimate iteration converges. satisfies the variational principle and hence gives the up- Because an electron feels the static potential of 0 or U per bound of the ground-state energy. in the static approximation, it is rather obvious that it Then an additional approximation is introduced to es- succeeds in splitting the band into the so-called upper timate the ground state. When one assumes the number and lower Hubbard bands for the Mott insulating phase, D as shown in the bottom of Fig. 13 for large U, while it of doubly occupied sites , the interaction energy is exactly given by ͗⌿͉ ͉⌿͘/͗⌿͉⌿͘ϭDU. However, the merges to a singly connected band for small U, as shown HU kinetic energy ͗⌿͉ ͉⌿͘/͗⌿͉⌿͘ is difficult to estimate in the top panel of Fig. 13. Ht We shall see in Sec. II.D.6 that the solution in infinite analytically. By neglecting the spin and charge short- dimensions is similar for the insulating phase, i.e., for ranged correlation, one can calculate the energy as a large U. This is because dynamic charge fluctuations ne- function of D and g for a given system size N and num- ¯ glected in the Hubbard approximation are irrelevant in ber of up spins N and down spins N . With ␧␴ ↑ ↓ the insulating phase. We shall also see in Sec. II.D.6 that ϭ͗⌽͉ t͉⌽͘, the optimization in terms of g at half filling H ⌺ can be determined rigorously in infinite dimensions by gives introducing dynamic mean fields instead of static charge- Eg Stratonovich variables. In the metallic phase, neglect of ϭq͑¯␧ ϩ¯␧ ͒ϩUd, (2.190) the dynamic charge fluctuations in the Hubbard approxi- N ↑ ↓ mation is not at all justified. where dϵD/N and qϭ8d(1Ϫ2d). From minimization A related problem in the Hubbard approximation is with respect to d, we obtain that spin fluctuations are not taken into account because E it starts from the Stratonovich-Hubbard transformation, g ¯ ¯ 2 ϭϪ͉͑␧ ϩ␧ ͉͒͑1ϪU/Uc͒ (2.191) in which there is decomposition only in the charge den- N ↑ ↓ sity. In Sec. II.D.8, we discuss a similar type of CPA with U ϭ8͉¯␧ ϩ¯␧ ͉ at UϽU , while at UϾU , the in- developed for the case of decomposition into spin den- c c c sulating state↑ with↓ gϭ1 is stabilized with E ϭ0. As the sity fluctuations. A crucial drawback of the Hubbard ap- g MIT point UϭU is approached, the renormalization proximation in the metallic phase is that the Fermi vol- c factor Zϭqϭm/m* is renormalized to zero as qϭ1 ume is small for small doping away from the Mott Ϫ(U/U )2. For details of the derivation, readers are re- insulator, which does not satisfy the Luttinger require- c ferred to Vollhardt (1984). In the Gutzwiller approxima- ment of large Fermi volume for the Fermi-liquid de- tion, the uniform magnetic susceptibility ␹ , the charge scription. s susceptibility ␹c , and the effective mass m* follow Ϫ1 5. Gutzwiller approximation ␹sϰ͑1ϪU/Uc͒ , (2.192) Ϫ1 The Gutzwiller approximation (Gutzwiller, 1965) was m*/mbϰ͑1ϪU/Uc͒ , (2.193) first applied to the MIT problem by Brinkman and Rice ␹ ϰ͑1ϪU/U ͒, (2.194) (1970). This approach is based on two stages of approxi- c c mation. We discuss the case of the Hubbard model [Eqs. in the metallic region UϽUc . The MIT scenario of the (1.1a)–(1.1d)] here. First, it employs a variational wave Gutzwiller approximation is described by mass diver- function of the form gence within the Fermi liquid. The Gutzwiller approxi-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1080 Imada, Fujimori, and Tokura: Metal-insulator transitions mation ignores the wave-number dependence of the suppressed and hence the mean-field approximation be- self-energy as well as the incoherent part of the excita- comes valid. In fermionic systems with metallic spin and tions. charge excitations, however, the situation is much more Away from half filling, for the doping concentration ␦ complicated because dynamic quantum fluctuations are 1 (i.e., n ϵN /Nϭn ϵN /Nϭ 2 Ϫ␦), Eg is given by important even in large dimensions. In infinite dimen- ↑ ↑ ↓ ↓ sions, the number of spatial dimension is infinite, so spa- Eg ϭq͑¯␧ ϩ¯␧ ͒ϩUd (2.195) tial fluctuations are suppressed, while the number of N ↑ ↓ temporal dimension is always one with an inherent with quantum dynamics. Therefore the mean-field theory ͑1/2ϩ␦Ϫd͒͑ͱϪ2␦ϩdϩͱd͒2 must be constructed in a dynamic way to describe me- qϭ 2 . (2.196) tallic states in large dimensions. In the dynamic mean- 1/4Ϫ␦ field theory, the spatial degrees of freedom are replaced This is minimized at dϭr␦ with rϭ͓(␥ϩ2)yϪ1͔/2 and with a single-site problem, while dynamic fluctuations yϭ͓Ϫ(␥ϩ2)ϩ2ͱ␥2ϩ4␥ϩ1͔/͓␥(␥ϩ4)͔ where ␥ must be fully taken into account. In the path-integral ϭ4(U/UcϪ1). At UϾUc , we obtain approach, the ‘‘mean field’’ determined self-consistently should be dynamically fluctuating in the temporal direc- E g ¯ ¯ 2 3 tion. ϭ͑␧ ϩ␧ ͒͑rϩ1ϩy͒„4␦ϩ8͑1ϩr͒␦ …ϪUr␦ϩO͑␦ ͒ N ↑ ↓ To understand how the dynamical fluctuation is ex- (2.197) actly taken into account in infinite dimensions, it is im- portant to notice that the self-energy of the single- for ␦Ӷ1. Therefore the FC-MIT takes place at UϾUc at particle Green’s function has no wave-number ␦ϭ0 with a change in ␦.AtUϾUc,qhas the form q 2 dependence and hence is site diagonal (Mu¨ ller- ϭq1␦ϩq2␦ with Hartman, 1989). By using the site-diagonal self-energy, Ϫ1/2 q1ϰ͑UϪUc͒ (2.198) one can obtain a set of self-consistent equations for the exact solution of the Green’s function in infinite dimen- near UϭUc . The charge susceptibility sions. The MIT has been extensively studied with the 1 Ϫ1 -help of numerical or perturbational solutions of the self 2͒␦ץ/ 2Eץ͑ ␹ ϭ c ͫN g ͬ consistent equations. Below we first define the Hubbard ¯ ¯ model in infinite dimensions. Next the self-energy of the remains finite at ␦Þ0 and lim␦ ϩ0␹cϭp/͓16(␧ ϩ␧ )͔ 2 2 2 → ↑ ↓ Green’s function is shown to be site diagonal. The self- with pϭ␥ /͓(1ϩ␥) Ϫͱ␥ ϩ4␥ϩ1͔ for UϾUc .AtU consistent equations are derived and several different ϭUc, the charge susceptibility has an exponent given by 2/3 ways of solving them are discussed. Consequences ob- ␹cϰ␦ . The above mean-field exponents are not the tained from the results are then discussed. For more de- same as the exponent in 1D and 2D, as is discussed in tailed discussions of the infinite-dimensional approach, Sec. II.E and Sec. II.F. readers are referred to the review article of Georges, The second step of the Gutzwiller approximation does Kotliar, Krauth, and Rozenberg (1996). not satisfy the variational principle. Alternatively, the In infinite dimensions, the Hubbard model defined in energy of this Gutzwiller wave function may be esti- Eqs. (1.1a)–(1.1d) with nearest-neighbor hopping char- mated numerically from Eq. (2.189) by Monte Carlo acterized by the dispersion in Eqs. (2.3a)–(2.3b) has the sampling (Yokoyama and Shiba, 1987a, 1987b) without same form. The dispersion of Eq. (2.3b) in d dimensions introducing the second step of the Gutzwiller approxi- is explicitly written as mation. This Monte Carlo method satisfies the varia- d tional principle and gives the upper bound of the ␧0͑k͒ϭϪ2t cos kj (2.199) ground-state energy. Variational Monte Carlo results j͚ϭ1 correctly predict the absence of MIT in the 1D Hubbard model at half filling, that is, the insulating state for all for a simple hypercubic lattice. For d ϱ, we have to → UϾ0. take a proper limit to yield a nontrivial model (Metzner As we see below in Sec. II.D.6, the Gutzwiller ap- and Vollhardt, 1989), because the on-site interaction proximation reproduces to some extent the coherent term (1.1c) is still well defined in dϭϱ while the kinetic- part of the spectral weight, while it fails to treat the energy term t has to be rescaled. This is easily under- H incoherent part, in contrast to the Hubbard approxima- stood from the density of states ␳(E) in the noninteract- tion. ing limit Uϭ0, which has the form 1 6. Infinite-dimensional approach 2 ␳͑E͒ϭ lim 1/2 exp͓Ϫ͑E/2tͱd͒ ͔. (2.200) d ϱ 2t͑␲d͒ In a simple model of phase transitions such as the → Ising model for the ferromagnetic transition, the univer- in the case of a hypercubic lattice. This density of states sality class and critical exponents are correctly repro- is derived from the equivalence between ␳(E) and the duced by the mean-field theory when the system is probability distribution function of the x coordinate of a above the upper critical dimension dc . For sufficiently random walker in 2D after d steps of length 2t each with large dimensional systems, fluctuations are in general a uniform and random choice of direction in the 2D

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1081

limit d ϱ, we see that the intersite terms vanish as → compared to ⌺ii . This heuristic argument can be ex- tended to all the higher-order terms in U, which guar- antees that all the intersite terms of the self-energy can be neglected in dϭϱ. We rewrite the functional integral of the Hubbard model introduced in Eqs. (2.41a) and (2.41b) with FIG. 12. The lowest-order diagrams of the intersite self- (2.37a)–(2.37d) as energy. The interaction is denoted by a black dotted line, while G0 is given by a full curve. The sites are denoted by i and j. Zϭ ⌿¯ ⌿ exp͓ϪS͔, (2.202a) ͵ D D ץ ␤ plane (x,y). Equation (2.200) is simply the solution of ¯ Sϭ d␶ ͚ ␺i␴͑␶͒ ␦ij ϩtijϪ␮␦ij ␺j␴͑␶͒ ϩSU , ͬ ͪ ␶ץ ͩ ͫ the diffusion equation for d ϱ as it should be in the ͵0 i,j,␴ case of a long time limit of the→ random walk. To take the (2.202b) limit with a finite density of states, one must scale t as ␤ tϭt*/2ͱd with a finite t*. Therefore the bare t is scaled S ϭ S ϭ d U ͚ Ui ͚ ͵ ␶ Ui͑␶͒, (2.202c) to zero as d increases. i i 0 H As we see in Eq. (2.200), the density of states for a ¯ ¯ Ui͑␶͒ϭU␺i ͑␶͒␺i ͑␶͒␺i ͑␶͒␺i ͑␶͒. (2.202d) simple hypercubic lattice has a Gaussian distribution of H ↑ ↓ ↓ ↑ width t*. A drawback of this density of states is that the Here ⌿ is a vector which depends upon site i (or wave MIT does not take place at finite U in the strict sense, number k), imaginary time ␶ (or the Matsubara fre- because the exponential but finite tail of the density of quency ␻n), and spin ␴. The matrix product, such as the states blocks charge-gap formation at any finite U.To ⌿¯ A⌿ used below, follows the same notation as that get around this difficulty, van Dongen (1991) considered from Eqs. (2.131a)–(2.131b) through (2.134a)–(2.134c). replacing the simple hypercubic lattice by the Bethe lat- The nointeracting Green’s function at Uϭ0 is defined as tice, where the density of states has a semicircular form: Ϫ1 ץ 2 G0ijϵ ␦ij ϩtijϩ␮␦ij , (2.203) ͪ ␶ץ ͩ (␳͑E͒ϭ ͱ1Ϫ͑E/t*͒2. (2.201 ␲t* which is transformed in Fourier space as In an explicit calculation of the Green’s function and the self-energy described below, we use only the density 1 G0͑k,␻͒ϭ . (2.204) of states as input, while the calculation does not depend i␻nϪ␧0͑k͒ϩ␮ on more detailed knowledge of the actual lattice struc- When we use G , the action has the form ture. This indicates that, in dϭϱ, lattice structure, con- 0 SϭS ϩS , (2.205a) nectivity for the hopping transfer, and a measure of the G0 U distance are all irrelevant. In other words, spatial corre- lations are completely ignored in dϭϱ. S ϵϪTr ⌿¯ GϪ1⌿, (2.205b) G0 0 When t is scaled as tϰ1/ͱd, one can show that the ␤ where Trϵ͐ d␶ ͚i . The Green’s function G at UÞ0is intersite self-energy of the Green’s function ⌺ij(␻)(i 0 Þj) is a higher order of the 1/d expansion than on-site formally given by introducing the self-energy ⌺ as self-energy ⌺ (␻) (Mu¨ ller-Hartman, 1989). In the dia- Ϫ1 Ϫ1 ii G ϭG0 Ϫ⌺, (2.206) grammatic expansion of the self-energy as in Fig. 12, the from which the action is rewritten as lowest order is a part of the on-site term ⌺ii , which contains one bare Green’s function G . In contrast, the 0ii SϭSGϪS⌺ϩSU , (2.207a) lowest-order term of the intersite self-energy is second ¯ ˆ order in U with three bare Green’s functions, as in Fig. S⌺ϭTr ⌿⌺⌿. (2.207b) 12. The intersite bare Green’s function G is scaled as 0ij Here, in the real-space and imaginary-time representa- ͉RiϪRj͉ G0ijϳt , tion, ⌺ˆ is a diagonal matrix given by ␴I with the identity I dϭ where ͉RiϪRj͉ is the Manhattan distance of the i and j matrix , since ⌺ is site diagonal in ϱ. Then the par- tition function may be given as sites. Then the lowest-order term of ⌺ij for the nearest- neighbor pair (i,j) is at most scaled by t3. Therefore the self-energy which contains the i site has contributions Zϭ ⌿¯ ⌿ exp͓ϪS ϩS ϪS ͔ ͵ D D G ⌺ U from the on-site term ⌺iiϳO(1), the nearest-neighbor 3 3 term ⌺ijϳ(number of nearest neighbors)ϫt ϳdt , ϭZG͗exp͓S⌺ϪSU͔͘G (2.208a) and the next-nearest-neighbor term ⌺ijϳ(number of 6 2 6 with next-nearest neighbors )ϫt ϳd t ¯ . Because of t ϳ1/ͱd, these are reduced to ⌺on-siteϳO(1), ¯ ZGϵ ⌿ ⌿ exp͓ϪSG͔, (2.208b) ⌺nearest-neighborϳ1/ͱd, ⌺next-nearest-neighborϳ1/d etc. In the ͵ D D

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1082 Imada, Fujimori, and Tokura: Metal-insulator transitions

1 1 ͗ ͘ ϵ ⌿¯ ⌿ exp͓ϪS ͔ . (2.208c) G ͑␻ ͒ϭ ¯ G ͵ G ¯ ii n ͚ Ϫ Ϫ ZG D D k i␻n ␧0͑k͒ ⌺͑␻n͒ Because ⌺ and U are site diagonal, for an N-site system, ␳͑⑀͒ the average over G may be replaced with the average ϭ d⑀ . (2.215) ͵ Ϫ Ϫ i␻n ⑀ ⌺ over the on-site Green’s function GiiϵGloc after inte- ¯ gration over ␺j and ␺j with jÞi first, as (4) From Gii(␻n) and ⌺(␻n) obtained in step (3), Z exp͓S ϪS ͔ ϭZ exp͓S ϪS ͔ , calculate G˜ (␶) from Eq. (2.214). G͗ ⌺ U ͘G Gloc͗ ⌺i Ui ͘Gloc (2.209) (5) Use G˜ (␶) obtained in step (4) as the input to solve where Eq. (2.213) and repeat the process from step (1) again until the iteration converges so that self-consistency is ␤ ¯ satisfied. S⌺iϭ d␶Јd␶␺i͑␶͒⌺͑␶Ϫ␶Ј͒␺i͑␶Ј͒. (2.210) ͵0 From this iteration, the single-site problem with dy- namical fluctuations is solved exactly. The lattice struc- Here, ͗¯͘G is defined by a local average, loc ture enters the procedure only through the density of states in Eq. (2.215). The basic idea of this self- ͗ ͘ ϭ ¯␺ ␺ ͗exp͓¯␺ GϪ1␺ ͔ ͘/Z , ¯ Gloc ͵ D i D i i ii i ¯ Gloc consistent procedure was first proposed by Kuramoto (2.211a) and Watanabe (1987) for the periodic Anderson model and extended to the Falikov-Kimball model by Brandt Z ϭ ¯␺ ␺ exp͓¯␺ GϪ1␺ ͔, (2.211b) and Mielsch (1989). This self-consistent procedure may Gloc ͵ D i D i i ii i be compared with the coherent-potential approximation and we have used the identity employed in the Hubbard approximation and the self- consistent renormalization approximation. While they ¯ Ϫ1 ¯ exp͓␺iGii ␺i͔ϭ d␺jd␺jexp͓ϪS͔. are similar, an important difference of the dϭϱ ap- ͵͟jÞi proach from the Hubbard approximation and the self- The on-site Green’s function is related to its Fourier consistent renormalization approximation is that the transform as mean field is dynamic and the dynamic fluctuations are 1 taken into account exactly in ␻ dependence of ⌺. Ϫi␻n␶ Gii͑␶͒ϭ G͑k,i␻n͒e . (2.212) It is difficult to solve Eq. (2.213) in an analytical way. N␤ ͚␻ ͚k n Several different numerical ways for solving Eq. (2.213) Then from Eqs. (2.208a)–(2.208c) and (2.209), the par- have been proposed. One is the quantum Monte Carlo tition function has the form method (Jarrell, 1992; Rozenberg, Zhang, and Kotliar, 1992), in which a single-site problem with dynamically ¯ Zϭ ␺i ␺i ˜ ͵ D D fluctuating G is solved numerically in the path-integral formalism following the algorithm for the impurity ␤ ¯ ˜ Ϫ1 Anderson model developed by Hirsch and Fye (1986). ϫexp Ϫ d␶Јd␶͕Ϫ␺i͑␶͒G ͑␶Ϫ␶Ј͒␺i͑␶Ј͒ ͫ ͵0 Although this is a numerically exact procedure, within statistical errors, it is rather difficult to apply in the low-

ϩ Ui͖ (2.213) temperature region. Another approach is the perturba- H ͬ tion expansion in which Eq. (2.213) is solved perturba- with tively in terms of U (Georges and Kotliar, 1992). This is called the iterative perturbation method. Among several ˜ Ϫ1 Ϫ1 G ͑␶͒ϵGii ͑␶͒Ϫ⌺͑␶͒. (2.214) different approaches, the most precise way of estimating physical quantities in the low-energy scale seems to be ˜ Ϫ1 In dynamic mean-field theory, G plays a role corre- the numerical renormalization-group method (Sakai and sponding to the Weiss mean field in the usual mean-field Kuramoto, 1994). The MIT of the Hubbard model in d theory. Now the problem is reduced to solving a single- ϭϱ has been extensively studied using these site problem (2.213), where G and ⌺ are exactly ob- procedures.1 Other strongly correlated models, such as tained in a self-consistent fashion as follows: the degenerate Hubbard models, the periodic Anderson (1) Calculate the Green’s function Gii from the single- site partition function (2.213) by assuming a trial form of model, the d-p model, and the double-exchange model have also been studied (Jarrell, Akhlaghpour, and Prus- G˜ . This is possible because Eq. (2.213) is a single-site problem. We discuss several different ways of solving Eq. (2.213) later. 1Studies of dϭϱ include those of Georges and Kotliar, 1992; (2) From the obtained result Gii in step (1) and the ˜ Georges and Krauth, 1992; Jarrell, 1992; Rozenberg, Zhang, assumed G, obtain ⌺(␶) from Eq. (2.214). and Kotliar, 1992; Jarrell and Pruschke, 1993; Pruschke, Cox, (3) From the Fourier transform ⌺(␻n) derived from and Jarrell, 1993; Zhang, Rozenberg, and Kotliar, 1993; Caf- ⌺(␶) in step (2), calculate the on-site Green’s function farell and Krauth, 1994; Jarrell and Pruschke, 1994; Sakai and from Kuramoto, 1994; Moeller et al., 1995.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1083

bility ␹s is enhanced but approaches a constant ϳ1/J at the transition, in contrast with the divergence in the Gutzwiller approximation. Here J is the energy scale of the exchange interaction. This difference comes from the fact that ␹s has contributions from the incoherent part, namely, upper and lower Hubbard bands, while the incoherent contribution is ignored in the Gutzwiller ap- proximation. Even for the filling-control MIT, results in dϭϱ show behavior similar to that of the Gutzwiller approximation where ZϪ1, m*, and ␥ are also enhanced while ␬ remains finite. As we shall see in Secs. II.E and II.F, ␬ appears to diverge at the transition point in 2D in contrast to the dϭϱ result. Comparisons of the spectral weight and the density of state with experimental data on Ca1ϪxSrxVO3, La1ϪxSrxTiO3, and other compounds are made in later sections. A controversy exists concerning the BC-MIT. The coherence peak at the Fermi level seems to disap- pear when the metal-insulator transition in Ca1ϪxSrxVO3 and other series of compounds such as FIG. 13. Density of states ␳(␻)ϭϪIm G by the dϭϱ ap- VO2, LaTiO3, and YTiO3 takes place with the incoher- proach at Tϭ0 for the half-filled Hubbard model at U/t*ϭ1, ent upper and lower Hubbard band kept more or less 2, 2.5, 3, and 4 from top to bottom. The calculation is done by unchanged, yet the optical conductivity shows that the iterative perturbation in terms of U. The bottom one charge gap closes continuously at the transition. The d (U/t*ϭ4) is an insulator. From Georges, Kotliar, Krauth, and ϭϱ results are consistent with the former but appear to Rozenberg, 1996. contradict the optical data, as will be discussed in Sec. IV. Although the mass enhancement in dϭϱ is similar chke, 1993; Si and Kotliar, 1993; Furukawa, 1995a, to that in some experimental systems such as 1995b, 1995c; Kajueter and Kotliar, 1996; Rozenberg, La1ϪxSrxTiO3, in the dϭϱ result, the Wilson ratio RW 1996a; Mutou and Hirashima, 1996; Mutou, Takahashi, ϵ␹s /␥ vanishes at the transition while experimentally it and Hirashima, 1997) remains constant around 2 in La1ϪxSrxTiO3, as is dis- Here basic results for the dϭϱ Hubbard model are cussed in Sec. IV.B.1. However, quantitatively this dis- summarized below. We first discuss the case of the BC- crepancy is not remarkable. MIT. In the metallic phase, the Fermi-liquid state with The solution in dϭϱ seems to combine two ap- Luttinger sum rule is realized in the solution in dϭϱ if proaches, the Hubbard approximation and the one considers only the paramagnetic phase (Georges Gutzwiller approximation, in a natural way. In the insu- and Kotliar, 1992). A remarkable property of the solu- lating phase, the Hubbard approximation provides simi- tion in dϭϱ is seen in the density of states ␳(␻) as illus- lar results, whereas, in the metallic phase, many of prop- trated in Fig. 13. At half filling it has basically the same erties have similarities with the Gutzwiller structure as the Hubbard approximation in the insulat- approximation This merging of the Gutzwiller and Hub- ing region UϾUc , where the Fermi level lies in the bard approximations was pointed out in an early work of Hubbard gap sandwiched between the upper and lower Kawabata (1975). He showed that the inelasticity of Hubbard bands. The reason why the Hubbard approxi- quasiparticle scattering is crucial for realization of a mation can treat the dynamics more or less properly has sharp Fermi surface, which is not considered in the Hub- been discussed in Sec. II.D.4. In the metallic phase U bard approximation. The dϭϱ approach further makes ϽUc but near Uc , a resonance peak appears around the it possible to discuss dynamics and excitations such as Fermi level in addition to the upper and lower Hubbard spectral weight. Thus one of the mean-field pictures of bands. The peak height is kept constant in the metallic the MIT, when combined with the dϭϱ results, has be- region, while the width of the peak becomes smaller and come clearer. smaller as the MIT point UϭUc is approached. The The original solution in dϭϱ contains only a para- peak at the Fermi level eventually disappears with a magnetic phase with complete ignorance of the antifer- vanishing width at UϭUc . This is associated with a van- romagnetic order. However, by introducing two sublat- ishing renormalization factor Z and hence a diverging tices A and B, one can reproduce the antiferromagnetic effective mass m*. In the metallic phase this critical di- order even in dϭϱ (Jarrell, 1992; Jarrell and Pruschke, vergence of m* is essentially the same as what happens 1993). The solution in this case indicates an antiferro- in the Gutzwiller approximation (Brinkman and Rice, magnetic order which persists not only in the insulating 1970), that is, ZϪ1, m* and the specific-heat coefficient phase but also in a wide region of metals similarly to the Ϫ1 ␥ are enhanced in proportion to ͉UϪUc͉ while the simple Hartree-Fock-RPA results. Therefore the MIT compressibility ␬ decreases until it vanishes at the tran- observed by excluding magnetic order is usually inter- sition point. The Drude weight is also proportional to rupted and masked by the appearance of the magnetic ͉UϪUc͉ in the dϭϱ approach. The magnetic suscepti- transition in reality at dϭϱ. It was pointed out that the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1084 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 15. U dependence in the dϭϱ Hubband model. (a) An- tiferromagnetic transition temperature as a function of U: ᭺, dϭϱ Hubbard model; छ, dϭ3 at half-filling. (b) U depen- ϭ FIG. 14. Phase diagram of the d ϱ Hubbard model at half dence of the square of the local moment ␮2. From Jarrell and * * filling in the plane of temperature T/t and interaction U/t in Pruschke, 1993. the case where the magnetic order is entirely suppressed by frustration. The solid curve is the first-order transition line pare the results in dϭϱ with those in realistic low- with the critical point at the black square. The region sur- dimensional systems including 3D, the region of the an- rounded by dotted curve shows the coexistence of metal and tiferromagnetic ordered phase may extend beyond those insulator, while the transition at Tϭ0 is a continuous transi- for dϭ2 and 3. In the case without orbital degeneracy, tion. From Georges, Kotliar, Krauth, and Rozenberg, 1996. the antiferromagnetic order can be remarkably sup- pressed due to quantum fluctuations in finite dimen- original MIT in the paramagnetic phase could be more sions, while it requires some other origin, such as frus- or less recovered by introducing frustrations because the tration, to suppress magnetic order in dϭϱ. In the limit magnetic order could be suppressed by the frustration dϭϱ, each exchange coupling is scaled to vanish, and (Georges and Krauth, 1993; Rozenberg, Kotliar, and hence we do not expect spin-glass-like order. However, Zhang, 1994). in this case we have nonzero entropy at Tϭ0. Another The phase diagram of metals and insulators in the more important difference between low-dimensional plane of U and temperature T shown in Fig. 14 was systems and dϭϱ occurs in the paramagnetic metallic argued to be similar to that of V2O3 under pressure or phase due to difference in the universality class of the doping (Rozenberg, Kotliar, and Zhang, 1994). Al- MIT as well as the absence of magnetic short-range cor- though the first-order transition appears to take place in relations in dϭϱ. For example the Drude weight D ap- dϭϱ at finite temperatures even with the transfer t be- pears to scale linearly with ␦ for the FC-MIT at dϭϱ, p ing fixed, in the experimental situation observed in the while it can be suppressed as Dϰ␦ , pϾ1 in the scaling theory. The dϭϱ approach may be justified only above case of V2O3, the change in the lattice parameter at the transition may also be an origin of the strong first-order the upper critical dimension dc , as we shall discuss in transition where discontinuous reduction of the lattice Sec. II.F. Although some attempts have been made to constant may stabilize the metallic state through the in- calculate 1/d corrections to compare with the Gutzwiller crease in the bandwidth. In the dϭϱ approach, the first- results in finite dimensions (Schiller and Ingersent, order transition at TÞ0 is replaced with a continuous 1995), as it stands, the convergence properties of the 1/d transition at Tϭ0 (Moeller et al., 1995). expansions are not clear. To make connections to finite- At half filling, when a two-sublattice structure is intro- dimensional results, it would be desirable to develop a systematic approach in terms of the 1/d expansion. In duced, the antiferromagnetic transition temperature TN as a function of U shows a peak at a moderate U (see the Mott insulating phase, the Heisenberg spin model Fig. 15), whose structure is similar to the SCR result as has been treated by an approach similar to that in this well as to the quantum Monte Carlo results in 3D (Scal- section (Kuramoto and Fukushima, 1998). The expan- ettar et al., 1989), although temperature and energy sion can be calculated up to the order of 1/d (or up to scales have an ambiguity in dϭϱ due to the rescaling of the order of 1/zn where zn is the number of nearest t to t*. It is clear that 2D systems show quite different neighbors), because the polarization function remains site diagonal up to this order in the quantum spin model. behavior because TN should always be zero. Away from This approximation goes beyond the one-loop level and half filling, TN still shows similar behavior to RPA re- sults with a wide region of antiferromagnetic metal. A corresponds to the spherical approximation known in serious problem in dϭϱ with antiferromagnetic order is spin systems. that the short-ranged spin fluctuation inherent in lower- 7. Slave-particle approximation dimensional systems is completely ignored. As we shall see in Sec. II.F the scenarios of the MIT in In strongly correlated systems, the strong-coupling dϭϱ have some similarity, such as the mass divergence, limit is frequently a good starting point for understand- to the low-dimensional case while they are rather differ- ing the basic physics. The strong-coupling limit is de- ent from the predictions of the scaling theory in many fined in the single-band model as the limit where double respects, such as the critical exponents themselves and occupation of the same orbital is strictly prohibited. The the role of antiferromagnetic correlations. If we com- t-J model (2.13) offers an example of this limit. To

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1085 implement this local constraint, the slave-particle and method has been developed. As an example, we discuss ⌬ ϭ͗b b†͘ (2.220) the case of the slave-boson method (Barnes, 1976, 1977; ij i j Coleman, 1984; Baskaran, Zou, and Anderson, 1987; Af- to decouple the transfer term proportional to t, whereas fleck and Marston, 1988; Kotliar and Liu, 1988; Su- Dijϭ͗fi fj Ϫfi fj ͘ (2.221) zumura, Hasegawa, and Fukuyama, 1988) where the ↑ ↓ ↓ ↑ electron operator ci␴ is decoupled as and ␹ij for decoupling of the exchange term propor- † tional to J. By using the mean-field approximation ci␴ϭfi␴bi . (2.216) † ͗bi bj͘ϰ␦ to replace the local constraint with an aver- Here fi␴ is a fermion operator called a spinon, which aged one, we can reduce the transfer term to carries spin degrees of freedom, while a boson operator b† called a holon represents a charged hole. The local † i tϭϪt␦͚ fi␴fj␴ . (2.222) constraint prohibiting double occupancy is represented H ␴ ͗ij͘ as This leads to an approximation that is similar to the † † Gutzwiller result but not strictly identical, in which the fi␴fi␴ϩbi biϭ1 (2.217) ͚␴ bandwidth becomes narrow in proportion to the doping concentration ␦ and hence the mass is enhanced criti- at all the sites i. cally. It should be noted that the holon degrees of free- Another way of decoupling is to assign fermion statis- dom are frozen in the approximation b†b ϭconstant tics for charge degrees of freedom and boson statistics ͗ i j͘ and a Fermi liquid of spinons is realized. When we fur- for spins. Aside from minor differences, this reduces to ther decouple the exchange term by using the RVB or- the spin-wave theory at half filling where the spins are der parameter ␹ , the bandwidth becomes proportional represented by Schwinger bosons. This formalism is ij to ⌬ ϩ␹ , which is proportional to J even in the limit called the slave-fermion method (Arovas and Auerbach, ij ij of 0. When the charge degrees of freedom are re- 1988; Yoshioka, 1989a, 1989b). The advantage of the ␦ tained→ in a dynamical manner, with the assumption that slave-fermion method is that it is easier to implement antiferromagnetic order at half filling than with the ␹ij is frozen below the mean-field temperature for ␹ij , slave-boson method. Away from half filling, however, the transfer term may be rewritten as the mean-field solution of the slave-fermion method † generally predicts a wide region of incommensurate an- tϭϪt͚ ␹ijbi bj . (2.223) H ␴ tiferromagnetic order, in contrast with the rapid disap- ͗ij͘ pearance of magnetic order believed to occur in 2D. A Then, if ␹ is assumed to be a mean-field constant ¯␹,it related formalism was also developed in the spin- ij leads to Bose condensation, b Þ0, below the transition fermion model in a rotating reference frame of antifer- ͗ i͘ temperature when a weak interaction of the bosons is romagnetic correlations (Ku¨ bert and Muramatsu, 1996). assumed. On the other hand, if the Bose condensation of We do not discuss the details of the slave-fermion ap- b takes place, by neglecting the strict local constraint, proximation here. i the Fermi-liquid state of f is realized again with finite In the slave-boson method, the t-J model defined in i mean-field order parameters for ␹ij and ͗bi͘. This gives Eq. (2.13) is represented using these operators as 2 a bandwidth proportional to t͗bi͘ ϩ(J/2)͗␹ij͘. When † † J † † we neglect J, we obtain a bandwidth proportional to ϭϪt b f f b Ϫ f f f f 2 ͚ i i␴ j␴ j ͚ ͓ i␴ j␴ j␴Ј i␴ b ϰ␦ again, as in the Gutzwiller approximation. H ͗i,j͘ 2 ͗ij͘ Ј ͗ i͘ ␴ ␴␴Ј Therefore, at Tϭ0, this mean-field framework always leads to the same level of approximations as the ϩf† f† f f . (2.218) i␴ j␴Ј j␴ j␴Ј͔ Gutzwiller approximation for the MIT. As discussed in When the local constraint (2.217) is rigorously enforced, Secs. II.D.6 and II.F, this approximation of course gives (2.218) is equivalent to the original t-J model. However, only mean-field exponents with no information on the in practical treatments, as we shall see below, this local incoherent part of the charge excitation. constraint is in many cases replaced by a global, aver- In the large-N limit, where the SU(2) symmetry of the aged one. In addition to this approximation, various original spin-1/2 models is generalized to SU(N) with types of mean-field decoupling have been attempted to arbitrary integer N, this mean-field solution becomes ex- discuss the phase diagram of this Hamiltonian, including act in any dimension and the incoherent part disappears flux phase, commensurate flux phase, dimer phase, and in the leading order of 1/N (see the reviews of Newns the so-called resonating valence bond (RVB) state. As a and Read, 1987; Kotliar, 1993; see also Jichu, Matsuura, typical example, here we discuss the case of RVB decou- and Kuroda, 1989). pling. The uniform RVB state introduces three order Mean-field solutions of the slave-particle formalism parameters for mean-field decoupling: give various phases such as the dimer phase, flux phase, and RVB phase. When the dimer and flux phases are † suppressed, the metallic state at Tϭ0 is the Fermi liq- ␹ijϭJ͚ ͗fi␴fj␴͘ (2.219) ␴ uid, and the MIT near ␦ 0 in this framework is char- →

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1086 Imada, Fujimori, and Tokura: Metal-insulator transitions acterized by Zϰ␦ and Dϰ␦ where Z is the renormaliza- above, this approach gives only a simple mean-field re- tion factor and D is the Drude weight. The specific-heat sult similar to that of a Gutzwiller-type approximation, coefficient ␥ is given by ␥ϰ1/(␦ϩ␣J/t) with a constant ␣ due to Bose condensation of bi . In the limit T 0, this of order unity (Grilli and Kotliar, 1990). The effective treatment ignores the incoherent part of the excitations→ mass m* is enhanced but remains finite, m*ϳ(t/J)mb in although the incoherence becomes important near the the limit ␦ 0. Mott transition even at Tϭ0. This incoherence is con- → Above the Bose condensation temperature of bi , sidered in the above scenario only partially via gauge- when using the Hamiltonian (2.218) as a starting point, field fluctuations at finite temperatures. It is also difficult one must treat the dynamic fluctuation of ␹ij with in this approach to explain the small amplitude of RH strongly coupled spinons and holons if one wishes to go observed in the high-Tc cuprates at high temperatures beyond the mean-field approximation. A gauge-field ap- (see Secs. II.E.1 and IV.C). Diverging compressibility at proach (Ioffe and Larkin, 1989; Ioffe and Kotliar, 1990; Tϭ0 near ␦ϭ0 (see Secs. II.E.3 and IV.C.1) is also dif- Nagaosa and Lee, 1990; Lee and Nagaosa, 1992) can in ficult to explain in this scheme. All these problems are of principle treat this fluctuation, where the local constraint fundamental importance and contribute to the failure of and the fluctuation of ␹ij are represented by a gauge this approach in treating the Mott transition in finite field coupled to both spinon and holon. This may be dimensions in a manner similar to the Gutzwiller ap- implemented through the fluctuation of ␪ which is simi- proximation. lar to Eq. (2.156) but actually uses the Stratonovich- One of many serious difficulties of the slave-particle Hubbard field to decouple Eq. (2.218), where the cou- model is that it is not clear whether the gauge-field cou- pling of holon to spinon is represented by chirality pling is weak enough to justify the original spin-charge fluctuations of ␪. Recently this approach was extended separation. Without a weak enough coupling, spinons to reproduce the SU(2) symmetry at half filling. The ␲- flux state at half filling is connected to the RVB phase and holons recombine to the original electrons, which upon doping (Wen and Lee, 1996). would invalidate the model. Because this approach is In the gauge-field approach, it has been claimed that based on a very strong statement, namely, an explicit the linear-T resistivity is reproduced from the dynamics spin-charge separation, it is highly desirable to find di- of holes (slave bosons) strongly coupled to the gauge rect evidence for or against it. From a theoretical point field. So far, in this gauge-field theory, the T-linear re- of view, the 2D Hubbard model at low concentration sistivity has been ascribed to the incoherence of slave has been extensively investigated and, although it is still bosons due to low boson concentration together with controversial (see for example, Narikiyo, 1997; Castel- the speculated retraceable nature of bosons under the lani et al., 1998), the conclusion appears to be that the coupling to the dynamic gauge field. However, the gauge Fermi-liquid state is stable and hence the spin-charge fluctuation has been considered mostly within the separation does not take place. In general, in 2D, the Gaussian approximation without a self-consistent treat- Fermi liquid is destroyed only when there is a singular ment. Self-consistent treatment of holons and spinons interaction between quasiparticles. coupled by the gauge field has not been attempted so When we take into account the J term, below the far, and it is not clear whether the above claims will be mean-field temperature of Dij , the spin degrees of free- confirmed by a self-consistent treatment. In fact, the en- dom may have a pseudogap structure on the mean-field ergy scale of the dynamic fluctuation of spinons is only level. It was claimed that this may be relevant to the J, which yields incoherent charge dynamics only on the pseudogap anomaly observed in the high-Tc cuprates same energy scale through the spinon-holon coupling by (Kotliar and Liu, 1988; Suzumura, Hasegawa, and Fuku- the gauge field. This appears to contradict the experi- yama, 1988). The basic mechanism could be closely re- mental and numerical observations to be described in lated to superconducting pairing due to antiferromag- Secs. II.E.1 and IV.C, where ␴(␻)ϰ1/␻ extends to the netic fluctuations. However, the formation of order of the bare bandwidth in the absence of the char- pseudogaps above the superconducting transition tem- acteristic crossover energy scale. The energy scale for T the appearance of incoherence is of the order of the perature c , expected at low doping in this framework, Mott gap and much larger than J near the Mott insula- is rather difficult to explain from antiferromagnetic spin- tor. In Secs. II.E.1 and II.F, we shall see that the inco- fluctuation theory, while it may have an interesting con- herence may be ascribed to different origins near the nection to the pairing mechanism for doped spin-gap MIT. Mott insulators. Pseudogap formation may be just a con- The description of the MIT ␦ 0 by the slave-particle sequence of the preformed singlet pair (fluctuations) of approach including the gauge-field→ method is rather electrons due to some pairing force, for example, and is primitive in many other respects: First, it does not prop- not necessarily the consequence of this particular for- erly take account of the growth of (short-ranged) anti- malism of spin-charge separation. Indeed, on general ferromagnetic correlation, which plays an important role grounds, it is natural to have a separation of Tc and the in the incoherent dynamics. [The RPA level of antifer- pseudogap formation temperature TPG at low doping, as romagnetic correlation was additionally considered we shall discuss in Secs. II.F and IV.C. The critical tem- within a uniform RVB treatment (Tanamoto, Kohno, perature must go to zero as ␦ 0 because the coherence → and Fukuyama, 1992)]. Secondly, at Tϭ0, as mentioned temperature Tcoh should go to zero when the system

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1087

approaches the continuous MIT point, while TPG may coupling theory developed for weak ferromagnetic or remain finite due to a finite pairing energy even close to antiferromagnetic systems with small-amplitude spin the MI transition point. fluctuations within the Gaussian approximation. In addi- Although the slave-particle approach provides in prin- tion to the RPA diagrams of bubbles, within the one- ciple an attractive way to describe the problem if the loop approximation, the polarization function contains local constraint is reliably implemented and the fluctua- diagrams whose bubbles are coupled to the spin fluctua- tion around the starting mean-field solution is properly tions with different wave numbers. This generates mode estimated, its practical realization is rather difficult and couplings, which are treated self-consistently. This is ba- the approximations employed so far are unfortunately sically a self-consistent treatment of the one-loop ap- not well controlled. As in all the mean-field theories, the proximation. The original mode-coupling treatment takes account of dynamic fluctuations but is justified way of decoupling the original Hamiltonian by assuming only for weakly correlated systems at the one-loop level order parameters in the mean-field treatment of the con- and at the level of minor vertex corrections. On the straint (2.217) is not unique and we can argue many pos- other hand, our main interest in this review is strongly sibilities. Mean-field results give a criterion for local sta- correlated systems. Therefore we shall use the terminol- bility of the solution while it is hard to show global ogy of the SCR approximation for more generalized stability. It is also hard to estimate fluctuation effects in treatments, formulated later, with the static approxima- a controlled way. Generally speaking, in low- tion of the Stratonovich-Hubbard variable, to cover dimensional systems, this is a serious problem because both small and large amplitudes (Moriya, 1985). This of large quantum fluctuations. In fact, if the dimension- may be viewed as an extrapolation of the original SCR ality is below the upper critical dimension, phase transi- to the strong-correlation regime but within the static ap- tions must be described beyond the mean-field theories. proximation of the Stratonovich-Hubbard variables. As The slave-particle approximation does not reproduce we shall see below, in principle, the mode-coupling con- the large dynamic exponent zϭ4 observed in 2D nu- tribution to the additionally introduced self-energy can merical studies, discussed in Secs. II.E and II.F, because be implemented so as to include the original SCR treat- of the neglect of strong wave-number dependence. ment as a special case, although it has not been done Therefore it does not reproduce important physical seriously. Though there exist several different ways of quantities associated with the MIT in 2D. For qualita- formulating the SCR theory, the level of approximations tively correct descriptions of the problem, the slave- used is similar. A typical formulation of the SCR theory particle approach is useful for limited use on a mean- (Moriya and Hasegawa, 1980) is reviewed below. We field level, but it is important to consider the overall also briefly review a more or less equivalent consistency of the obtained results with those of other renormalization-group approach by Millis (1993) in Sec. theoretical approaches and with experimental indica- II.D.9. tions. To calculate the partition function in the path-integral formalism with the Stratonovich-Hubbard transforma- tion, we need to calculate Eq. (2.132a). Like the 8. Self-consistent renormalization approximation Hartree-Fock formulation for the strong-correlation re- and two-particle self-consistent approximation gime, the SCR approximation for the strong-correlation When electron correlation effects become stronger, in regime employs a static approximation in which dynamic ˜ ˜ general, spin fluctuations have to be considered seri- fluctuations of ⌬s and ⌬c are ignored. This static ap- ously. In the Hubbard approximation and Gutzwiller ap- proximation becomes a serious drawback in treating proximation, as well as in the infinite-dimensional ap- anomalous metals with fluctuations related to correlated proach, wave-number- and frequency-dependent spin insulators. fluctuations are not taken into account. Although the Instead of the saddle-point approximation of a single RPA, discussed in Sec. II.D.2, provides a formalism to mode of ⌬˜ in the Hartree-Fock approximation, the SCR deal with spin fluctuations perturbatively, its validity is theory employs a mode-coupling theory for Gaussian restricted to cases of sufficiently weak spin fluctuations, fluctuations combined with a coherent potential ap- as is easily understood from its derivation. It describes proximation (CPA) which regards the wave-number- magnetic transitions only at the mean-field level with dependent static Stratonovich fields ⌬˜ as random po- uncontrolled treatment of fluctuations. To consider s tentials. The framework of the CPA is quite similar to wave-number-dependent spin fluctuation of any ampli- that for the Hubbard approximation discussed in Sec. tude, large or small, and especially to improve the RPA ˜ and the Hartree-Fock approximations at finite tempera- II.C.4. Instead of the ⌬c considered in the Hubbard ap- ˜ tures, the self-consistent renormalization (SCR) ap- proximation, the fields ⌬s are mainly taken into account proximation was developed by Moriya (1985). In the in SCR. The first goal is to calculate the Green’s func- SCR theory, as we see below, wave-number-dependent tion spin fluctuations of large amplitude can be taken into Ϫ1 G ͑k,kЈ͒ϵϪQ ͑k,kЈ͒. (2.224) account more completely, while dynamic fluctuations ␴␴Ј ␴␴Ј are more or less neglected because of the static approxi- Instead of using Eq. (2.143) for the quasiparticle mation employed to treat the strong-correlation regime. Green’s function in the Hartree-Fock approximation, we In the original terminology of SCR, it was a mode- consider the self-energy correction in the form

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1088 Imada, Fujimori, and Tokura: Metal-insulator transitions

ϪQ␥ Ϫ⌺1 ,0 ther approximation such as the Gaussian approximation GϪ1ϭϪQ ϭ ↑↑ , ˜ ␥ ␥ ͩ 0, ϪQ Ϫ⌺ ͪ of ⌬s is taken to cut off the t-matrix expansion (Moriya ␥ 2 ↓↓ (2.225) and Hasegawa, 1980). The formal procedure for solving the problem in this approximation is rather similar to with Q given in Eq. (2.145c). The self-energy is de- ␥␴␴ that explained in the Hubbard approximation (Sec. scribed as ⌺ and ⌺ . In the absence of symmetry 1 2 II.D.4). The result in the paramagnetic phase is breaking, namely, in the paramagnetic phase, the self- energy correction may be written in the original single- 1 2 ⌺͑␻ ͒ϵ⌺ ϭ⌺ ϭϪ ⌺2͑␻ ͒Ϫ2␲ particle Green’s function (2.143) in the form n 1 2 ͫ n ͩN␤ͪ

⌺1 0 GϪ1ϭϪQ Ϫ . (2.226) ˜ ˜ 0 ͩ 0 ⌺ ͪ ϫ͚ ͗⌬s͑k͒⌬s͑Ϫk͒͘ Gloc͑␻n͒. (2.232) 2 kÞ0 ͬ The self-energy is determined by the CPA combined It should be noted that, in contrast to the infinite- with some approximations such as the Gaussian approxi- dimensional approach, the CPA self-consistent equation ˜ mation for the static fluctuation ⌬s(k,␻nϭ0). The co- (2.227) cannot be exactly solved because of its explicit herent potential approximation neglects the wave- wave-number dependence. Since ͚kÞ0⌬s(k)⌬s(Ϫk) be- number dependence of the self-energy and considers comes ␤ linear at low temperatures, Re ͚ϰT is obtained. only ⌺1(kϭ0)ϭ⌺2(kϭ0) in the absence of magnetic From the Kramers-Kronig relation, this implies Im ͚ fields. In other words, the CPA condition is to take into ϰT2, which is consistent with the Fermi-liquid descrip- ˜ account the effect of Q1 (or effects of dynamic ⌬s)by tion. the local on-site self-energy on average. This may be When ⌺ϭ0, the static approximation with only a ˜ achieved by solution of the self-consistent equation single mode of ⌬s retained leads directly to the Hartree- 1 Ϫ1 Fock approximation, while the Gaussian approximation Q k Ϫ G Q Ϫ ˜ 1͑ ͒ 1 ͚ ͑k,␻͒ ͑ 1͑k͒ ⌺͒ for ⌬s leads to the RPA result. Therefore, at Tϭ0, the ͳ ͫ ͩ N k ͪ ͬ ʹ ⌺ϭ Ϫ1 SCR approximation is reduced to the Hartree-Fock- 1 RPA result. The most important contribution of the 1Ϫ ͚ G͑k,␻͒ ͑Q1͑k͒Ϫ⌺͒ . ͳ ͫ ͩ N k ͪ ͬ ʹ SCR approximation is an improvement in the prediction (2.227) of temperature dependence for various physical quanti- ties in weak ferromagnetic or antiferromagnetic metals. as in Eq. (2.187). Here ͗¯͘ is the average over the dis- ˜ One of them is the Curie-Weiss-like temperature depen- tribution of ⌬s corresponding to ͚lZl¯/͚lZl in Eq. (2.187). In this case, the probability P is taken as dence of the susceptibility derived in metallic magne- l tism. In contrast to the critical behavior of the Mott in- exp͓ϪS⌬͔ defined in Eq. (2.134). After one has determined the wave-number- sulating systems, in which the local magnetic moment is independent but frequency-dependent ⌺(␻), a system- temperature independent, the Curie-Weiss-like suscepti- atic t-matrix expansion is in principle possible to allow bility in this case is mainly ascribed to temperature de- for corrections coming from the wave-number depen- pendence of the squared local moment, which becomes dence. The effective action for ⌬ obtained from Eqs. T-linear at the critical coupling constant for TϾTcϭ0. (2.134a)–(2.134c), The Curie-Weiss-like susceptibility obtained in this framework is not so surprising because in the SCR ap- Seff͑⌬͒ϭ͓ϪTr ln͑ϪG0͒ϩTr ln͑1ϪG0Q1͔͒, proximation we reproduce the critical behavior of the (2.228) Gaussian treatment, as we discuss in Sec. II.D.9. For is rewritten as analyses of experimental results in weakly correlated materials, readers are referred to the review article of S ͑⌬͒ϭTr ln͑ϪGϪ1ϩQ ͒ϭTr ln͑ϪGϪ1ϩQ Ϫ⌺͒ eff 0 1 1 Moriya (1985). The basic predictions of this treatment for transport, magnetic, and thermodynamic properties, ϭϪTr͓ln͑ϪG͒Ϫln„1ϪG͑Q1Ϫ⌺͒…͔ especially temperature dependences, have already been ϭϪTr͓ln͑ϪG͒Ϫln͕1ϪGloc͑Q1Ϫ⌺͖͒ described in Sec. II.D.1, where the basic temperature Ϫln͕1Ϫ͑GϪG ͒ ͖͔, (2.229) dependences are derived from rather simple dimen- loc T sional analyses. Experimentally, an example that follows where the t matrix is defined as the basic predictions of the SCR approximation near the Ϫ1 magnetic transition is discussed in Sec. IV.A. Several ϭ͑Q1Ϫ⌺͓͒1ϪGloc͑Q1Ϫ⌺͔͒ (2.230) T f-electron systems also appear to follow this prediction, with the local Green’s function while some systems such as CeCu6ϪxAux show different 1 criticality at the antiferromagnetic transition (von G ͑␻ ͒ϭ G͑k,␻ ͒. (2.231) Lo¨ hneysen, 1996), where the specific heat C, the resis- loc n N ͚ n k tivity ␳, and the uniform magnetic susceptibility ␹ follow The CPA condition (2.227) is solved in some approxima- C/TϰϪln(T), ␳ϭ␳0ϩAЈT, and ␹ϭ␹0(1Ϫ␣ͱT), re- tion, while charge fluctuations are neglected. Then a fur- spectively. Other f-electron systems such as UxY1ϪxPd3

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1089

and CeCu2Si2 also show a variety of anomalous features, II.F in detail, such mean-field predictions are no longer which we do not discuss in detail here. justified in low dimensions. The mean-field prediction The most serious drawback of the SCR approximation also ceases to be satisfied near the Mott transition point in treating strongly correlated systems is the combina- because the nesting condition has to be considered. Low tion of the Gaussian approximation with the static ap- dimensionality and the proximity of the Mott insulating proximation. First, dynamic spin fluctuations beyond the phase are two typical cases in which quantum fluctua- Gaussian level (one-loop level) are neglected. The pre- tions are crucially important in deriving various features scription for the vertex correction is also insufficient. beyond the level of the SCR approximation, as is de- Second, since the static approximation employed in the tailed in Sec. II.F. strong-correlation regime is justified only when the tem- Recently, an approximation at the same level as the perature is much higher than all the characteristic spin SCR but extended to allow explicit charge fluctuations fluctuation energies, various interesting quantum fluc- simultaneously and to allow an incommensurate period- tuations may be ignored. Intrinsically quantum- icity of order has been developed as the two-particle mechanical phenomena have to be treated beyond the self-consistent approach (Vilk, Chen, and Tremblay, ´ static approximation. 1994; Dare, Vilk, and Tremblay, 1996). In this approach, The metal-to-Mott-insulator transition, one of the the same functional forms as for the RPA for ␹s and ␹c main subjects in this article, is a good example in which are assumed, but with renormalized interactions Us and the above drawbacks become serious. MITs are essen- Uc . These parameters Us and Uc are determined self- tially quantum phase transitions at Tϭ0 controlled by consistently to satisfy the constraint on a relation to ͗n n ͘. This constitutes a conserving one-loop approxi- quantum fluctuations such as the band-width and the ↑ ↓ filling. There quantum dynamics plays a crucial role. In mation at the two-particle level, in contrast to the one- fact, near the Mott insulator, the characteristic spin- particle level in the fluctuation exchange (FLEX) ap- fluctuation energy remains at a finite value even in the proximation. FLEX is regarded as the same level of insulating phase, whereas the characteristic charge exci- approximation as the original weak-coupling SCR ap- tation energy has to vanish toward the MIT point if the proximation without the vertex correction. transition is of continuous type. Therefore the tempera- In the strong-coupling limit, that is, in localized spin ture range at which the static approximation is justified systems, it appears that the SCR approximation in this lies outside the interesting temperature range of charge subsection can be extended to reproduce the spherical dynamics. In addition, the existing framework of SCR approximation of spin systems. Kuramoto and Fuku- on the one-loop level does not properly take into ac- shima (1997) clarified the level of approximation needed count the contribution from incoherent excitations or to reproduce the spherical approximation in more detail. the reduction of the renormalization factor, which is a dominant effect near the continuous MIT. 9. Renormalization-group study of magnetic transitions in metals Another point to be discussed with SCR is the treat- ment of magnetic transitions at finite temperatures as In this subsection, renormalization-group studies of well as at zero. Self-consistent renormalization leads to a the Gaussian fixed point of magnetic transitions in the mean-field prediction for critical exponents near the metallic phase, developed by Hertz (1976) and Millis transition point, because it is based on the Gaussian ap- (1993), are reviewed. As is mentioned in Sec. II.D.8, this proximation in the paramagnetic region. This is only treatment of the Gaussian fixed point essentially leads to valid when the system is above the upper critical dimen- equivalent results to the SCR approximation. We start sion of the Tϭ0 quantum transition. In terms of critical with the path-integral formalism with the Stratonovich- phenomena, a proper treatment beyond the mean-field Hubbard transformation introduced in Sec. II.D.2. level becomes necessary in low dimensions. Recently, an When we neglect the charge fluctuation part by taking equivalent approach to the weak-coupling SCR approxi- yϭ0 in Eq. (2.132b), the ⌬s-dependent part of the ac- mation was formulated by Hertz (1976) and Millis tion derived in Eq. (2.137) is rewritten as (1993), using renormalization-group (RG) analysis, as 4 we review in the next subsection (Sec. II.D.9). This RG ˜ ˜ Sϭ ͚ ⌬s*͑k͒⌬s͑k͒ϪTr ln͓1ϪG0Q1͔. (2.233) approach is also justified only above the upper critical U k dimension. A basic point there is that the dynamic ex- ˜ ponent z is 2 for an antiferromagnetic transition in the Since Q1 is linear in ⌬s , as defined in Eqs. (2.135c) and metallic phase while zϭ3 for a ferromagnetic transition (2.133c)–(2.133e), we are able to expand S in terms of ˜ ˜ in metals. These dynamic exponents are derived from ⌬s and ⌬s* . Hereafter, for simplicity, we employ the no- ˜ the low-energy action for an overdamped spin-wave due tation ⌬s for 2⌬s . Up to the fourth order of ⌬s ,itis to the continuum of the Stoner excitation coupled to the written as spin wave. From the renormalization-group study, it was suggested that the upper critical dimension appears to Sϭ͚ v2͑k͒⌬s*͑k͒⌬s͑k͒ϩ ͚ v4͑k1 ,k2 ,k3 ,k4͒ be dϭ3 for antiferromagnetic transitions provided that k k1¯k4 the Fermi surface does not satisfy the nesting condition, 4 while it is controversial for the ferromagnetic case (see ϫ␦ ͚ ki ⌬s*͑k1͒⌬s*͑k2͒⌬s͑k3͒⌬s͑k4͒ϩ¯ (2.234) the next subsection). As we discuss in Secs. II.D.9 and ͩ iϭ1 ͪ

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1090 Imada, Fujimori, and Tokura: Metal-insulator transitions where and finite. We discuss the possible breakdown of this 1 assumption in Sec. II.F. The difference between the fer- v ͑k͒ϭ Ϫ␹ ͑k͒ (2.235) romagnetic and antiferromagnetic cases arises from the 2 U b fact that the order parameter, that is, the magnetization, while commutes with the Hamiltonian and hence is conserved 1 for the ferromagnetic case while this is not so for the ␹ ͑k͒ϭϪ G ͑q͒G ͑kϩq͒ staggered magnetization in the antiferromagnetic case. b ␤ ͚ 0 0 q Because of the conserved magnetization, the inverse Ϫ1 f͑␧q͒Ϫf͑␧kϩq͒ lifetime ␶ of the ferromagnetic fluctuation becomes k ϭϪ͚ (2.236) linear for small k and hence the lifetime diverges in the q ␧qϪ␧kϩqϩ␻ limit k 0. The ␻n-linear dependence of the action is is the magnetic susceptibility of the noninteracting sys- the consequence→ of the overdamped nature of the fluc- tem. The coefficient v4 is given by the fourth-order loop tuation due to coupling to the continuum Stoner excita- and can be expressed using G0 . The Gaussian coeffi- tions of metal. cient v (k) is expanded in terms of k and ␻ by using 2 n Next we consider the scaling properties of these ac- 2 1 k ␲ ͉␻n͉ tions. When we take the scale transformation kЈϭkb ␹ ͑k,i␻ ͒ϭ␳͑␧ ͒ 1Ϫ Ϫ ϩ z b n F ͫ 3 ͩ 2k ͪ 2 ͉k͉v ¯ͬ and ␻Јϭ␻b , the Gaussian terms F F (2.237) 2 2 for the case of ferromagnetic transitions with isotropic S2Fϭ ͑⌬ϩk ϩ ͉␻n͉/͉k͉͉͒⌬s͑k͉͒ (2.240) ͚k Fermi sphere. It should be noted that this expansion is possible only for dϾ3. In dϭ1 and dϭ2, the suscepti- for the ferromagnetic case and bility is singular at ͉k͉ϭ2kF even for an isotropic Fermi surface and cannot be expanded regularly (Kittel, 1968). 2 2 S2AFϭ͚ ͓⌬ϩk ϩ ͉␻n͉␶͔͉⌬s͑k͉͒ (2.241) A crucial assumption of the Gaussian treatment de- k scribed in this subsection is that the dispersion in v2 has for the antiferromagnetic case are transformed as nonzero coefficient for the quadratic k dependence and this is the leading contribution at small k. Recently, the 2 Ϫ2 1Ϫz 2 S2ЈFϭ͚ ͓⌬ϩkЈ b ϩ ͉͑␻nЈ͉/͉kЈ͉͒ b ͔͉⌬s͑k͉͒ nonanalyticity of ␹b in k for dр3 in the leading order kЈ was considered more seriously (Vojta et al., 1996; Kirk- (2.242) patrick and Belitz, 1997), which led to the conclusion and that, for the ferromagnetic case, the mean-field descrip- tion is incorrect at dϭ2 and 3, in contrast with the re- 2 Ϫ2 Ϫz 2 sults obtained from Eq. (2.237). S2ЈAFϭ͚ ͓⌬ϩkЈ b ϩ ͉͑␻nЈ͉␶͒ b ͔͉⌬s͑k͉͒ , kЈ If the above assumption for the regular expansion in (2.243) (2.237) is allowed, we obtain respectively. The action can be kept in the same form ͉␻n͉ after this scaling if the dynamic exponent z is given by v ͑k͒ϭ⌬ϩk2ϩ , (2.238) 2 ͉k͉ zϭ3 for the ferromagnetic case and zϭ2 for the anti- ferromagnetic case with the following scaling taken si- where we have assumed nonzero and finite coefficients 2 multaneously: for ⌬, k , and ͉␻n͉/͉k͉, which can be taken to be unity in appropriate units. Here ⌬ is the parameter measuring ⌬Јϭ⌬b2 (2.244) ϭ the distance from the critical point (⌬ 0 at the transi- and tion point) and should not be confused with the order Ϫ͑dϩzϩ2͒/2 parameter ⌬s . In the antiferromagnetic case, v2(k)is ⌬sЈϭ⌬sb . (2.245) similarly expanded around the ordering wave number Q The k dependence of the fourth-order coefficient v is through ␹ (Qϩk,i␻ ) for small k and ␻ if the nesting 4 b n n not important for critical properties and may be re- condition is not satisfied and the spatial dimension d placed with the value uϵv (kϭ0) because v (kϭ0) satisfies dу3. When the nesting condition is satisfied, as 4 4 Ͼ0 is assumed to be finite to make the transition from in the case of transitions to the Mott insulating state, this paramagnetic phase ͗⌬ ͘ϭ0 to the magnetic phase type of simple expansion is not possible, because v in s n ͗⌬ ͘Þ0 possible at ⌬ϭ0, as in the usual Landau free- the expansion (2.234) has higher-order singularities for s energy expansion for ⌬ . Then the fourth-order term is larger n. We do not discuss this case here. The Mott s scaled as transition is the subject of Sec. II.F, and we discuss this problem in Sec. II.F.11. After taking proper units, we 4 4Ϫ͑dϩz͒ have S4ϭb u ͚ ␦ ͚ kiЈ ͩ iϭ1 ͪ kЈ kЈ 2 1¯ 4 v2͑k͒ϭ⌬ϩk ϩ͉␻n͉␶ (2.239) ϫ⌬Ј*͑kЈ͒⌬Ј*͑kЈ͒⌬Ј͑kЈ͒⌬Ј͑kЈ͒. (2.246) for the antiferromagnetic case, where we have again as- s 1 s 2 s 3 s 4 2 sumed that coefficients for ⌬, k , and ͉␻n͉ are nonzero Therefore, the scaling of u should be

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1091

uЈϭub4Ϫ͑dϩz͒ (2.247) easily calculated up to first order in u, and rescalings of ⌬, T, and u are determined to keep the same form for F to keep the scale-invariant form for S4 . This means that after the scale transformation b. The resultant the quartic term is scaled to zero for dϩzϾ4 after re- renormalization-group equations up to first order in u peating the renormalization. This is the reason why the are mean-field approximation is justified for dϩzϾ4 be- cause the term S becomes irrelevant. [Strictly speaking, d⌬ 4 ϭ2⌬ϩ⌶͑T,⌬͒u, (2.254) as usual, the term S4 is dangerously irrelevant. See, for d ln b example, Ma (1976)]. In this case, in the paramagnetic phase, the Gaussian approximation retaining only the dT ϭzT, (2.255) quadratic term S2 is justified in describing the quantum d ln b phase transition at Tϭ0. We note, however, that if the dispersion itself in Eq. (2.240) or (2.241) is given by du ϭ͓4Ϫ͑dϩz͔͒u (2.256) higher-order k␨ such as k4 instead of k2, the above cri- d ln b terion dϩzϾ4 for the justification of the mean-field treatment must be replaced with dϩzϾ2␨. This is in- with deed the case for the Mott transition, discussed in Sec. d 2 ⌳ d k zϪ2 II.F, where hyperscaling is justified. ⌶͑T,⌬͒ϭ2͑nϩ2͒ d k ͫ ␲ ͵0 ͑2␲͒ If only the Gaussian part of S2F or S2AF is considered, Ϫ␤F ϪS the partition function Zϭe ϵ͐ ⌬* ⌬ e is given zϪ2 D s D s k 1 by ϫcoth 2T 1ϩ͑⌬ϩk2͒2 ⌳ d ⌫ d k k d␧ ␧ ␧/⌫k Ϫ1 1 FϭV d coth tan 2 , d␧ ␧ ␧ ͵0 ͑2␲͒ ͵0 ␲ 2T ͫ⌬ϩk ͬ ϩKd coth 2 2 , (2.248) ͵0 ␲ 2T ␧ ϩ͑⌬ϩ1͒ ͬ where we introduce a cutoff ⌳ for the momentum and ⌫ (2.257) for the energy, with ⌫ ϭ⌫ for antiferromagnets and k where ⌫kϭ⌫k for ferromagnets. When we first integrate out Ϫ d/2 the rapidly varying part of a thin frequency-momentum Kdϭ͑2␲͒ /͑dϪ2͒!! shell defined by the region ⌳уkу⌳/b and ⌫kу␧ z у⌫k /b , we obtain for even d and d ⌳/b d k ⌫ /bz d␧ ␧ ␧/⌫ Ϫ ͑dϩ1͒/2 k Ϫ1 k Kdϭ2͑2␲͒ /͑dϪ2͒!! FϭV d coth tan 2 ͵0 ͑2␲͒ ͵0 ␲ 2T ⌬ϩk for odd d. The solution of these equations is

ϩF1ln b, (2.249) z TϭT0b , (2.258) ⌫⌳ d␧ ␧ ␧/⌫ d Ϫ1 ⌳ F1ϭV⌳ Kd coth tan 2 uϭu b4Ϫ͑dϩz͒, (2.259) ͵0 ␲ 2T ⌬ϩ⌳ 0

Vz ⌳ ddk ⌫ 1 ln b k Ϫ1 ⌬ϭb2 ⌬ ϩu dxe͓2Ϫ͑dϩz͔͒xF(T ezx,⌬ e2x) . ϩ d coth tan 2 . (2.250) 0 0 ͵ 0 0 ␲ ͵0 ͑2␲͒ 2T ⌬ϩk ͩ 0 ͪ (2.260) After the rescaling of k to kЈϭbk and ␧ to ␧Јϭ␧bz to recover the original range of k and ␧, we need the re- From the solution for u, it is clear that the Gaussian scaling treatment is justified in the paramagnetic phase when 4 Ϫ(dϩz)Ͻ0. Two regimes exist when the scaling stops ⌬ϭ⌬Ј/b2, (2.251) at ⌬ϳ1. One is the quantum regime, where TӶ1 is sat- TϭTЈ/bz (2.252) isfied when the scaling stops at ⌬ϳ1. The other is the classical regime, where Tӷ1 for ⌬ϳ1. The boundary of to keep the same form for F as these two regimes is obtained by putting Tϭ0 and ⌬ϭ1 d ⌳ d k ⌫k d␧ into Eq. (2.260) to get b and substituting the resultant b Ϫ͑dϩz͒ Ј Ј Ј FϭVЈb d into Eq. (2.258). The condition for the occurrence of the ͵0 ͑2␲͒ ͵0 ␲ quantum regime is

␧Ј Ϫ ␧Ј/⌫kЈ ϫcoth tan 1 . (2.253) 1ӷT /rz/2, (2.261) 2TЈ ⌬ЈϩkЈ2 0

Then the renormalization-group equation of ⌬ and u for uF͑Tϭ0,⌬0͒ rϭ⌬0ϩ , (2.262) S2 is d⌬/d ln b ϭ2⌬ and dT/d ln b ϭzT. The non- zϩdϪ2 Gaussian term S4 can be treated perturbatively, when we treat only Gaussian behavior. The free energy can be where T0 and ⌬0 are the bare values of the temperature

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1092 Imada, Fujimori, and Tokura: Metal-insulator transitions and the control parameter. In the quantum regime, as is ϳC/(Tϩ␪), although the coefficients C and ␪ may be usual in the Gaussian model, the correlation length is renormalized from the bare values. This is related to the given by Curie-Weiss law derived in the SCR approximation, where treatment by the Gaussian fixed point is justified ␰ϭrϪ1/2. (2.263) around ⌬ϳ0 even at low temperatures Tϳ0. In this re- When T is scaled large, the renormalization-group equa- gion, as pointed out by the SCR approximation, the tem- tion should be rewritten with a new variable vϭuT be- perature dependence of the local moment ϳT plays an cause ⌶(T)ϳCT for Tӷ1, and the right-hand side of important role in deriving the Curie-Weiss behavior. Eq. (2.254) is a function of the single variable uT. Then Consequences of the Gaussian treatment for other the renormalization-group equation reads quantities such as the susceptibility, spin correlations, and resistivity are discussed from intuitive picture in d⌬ ϭ2⌬ϩCv, (2.264) Secs. II.D.1 and II.D.8. In the strong-correlation regime, d ln b the amplitude of the order parameter grows faster than dv the weak-correlation regime above the transition tem- ϭ͑4Ϫd͒v (2.265) perature, where the low-energy excitation is described d ln b through nជ i and ␪i by Eq. (2.156). In this region, the up to linear order in v. The initial condition for these Gaussian treatment described in this subsection is no equations is obtained from a solution of the original longer justified, even in the disordered phase. The equations (2.254)–(2.256) when the renormalization Gaussian treatment also becomes invalid when the dis- stops at Tϭ1, as persion of the relevant gapless excitation becomes flatter in low-dimensional systems. In this case we have to con- ⌬¯ ϭTϪ2/z͓rϩBu T͑dϩzϪ2͒/z͔, (2.266) 0 0 0 sider the scaling analysis. We discuss in Sec. II.F when ¯ ͑dϩzϪ4͒/z and how the Gaussian treatment breaks down. vϭu0T0 , (2.267) where E. Numerical studies of metal-insulator transitions 1 1 for theoretical models Bϭ dT T͑2ϪdϪ2z͒/z͓⌶͑T͒Ϫ⌶͑0͔͒/͑nϩ2͒. 2z ͵ 0 Algorithms for numerical computation of strongly (2.268) correlated systems have recently been developed and The renormalization-group equations (2.264) and applied extensively. Several comprehensive articles in (2.265) are solved using the initial conditions (2.266) and the literature are available on these algorithms and their (2.267). The Ginsburg criterion, which is the condition applications (Scalapino, 1990; von der Linden, 1992; Ha- for justification of Gaussian treatment, is given by vӶ1 tano and Suzuki, 1993; Imada, 1993d, 1995d; Dagotto, when the scaling stops at ⌬ϭ1. This criterion is 1994). Two different algorithms have been widely stud- uT1ϩ1/z ied and applied, so far. One is the exact diagonalization 1ϩ1/z Ӷ1, (2.269) of the Hamiltonian matrix including the Lanczos rϩ͑BϩC͒uT method. Using this method, one can obtain physical which is always justified at Tϭ0 and is violated only in a quantities with high accuracy but only for small clusters. narrow critical region near the critical point rϩ(B Tractable system size is strictly limited because the nec- ϩC)uT1ϩ1/zϭ0 for TÞ0. We note that this derivation essary memory size increases exponentially with system of the Ginsburg criterion is justified only when dϩz size. For instance, for the Hubbard model, the largest Ϫ4Ͼ0 and u remains small after renormalization. tractable number of sites for the ground state is around The free energy can be computed from the solution of 20. The exact diagonalization method has close ties with the renormalization-group equation. The specific heat the configuration-interaction method, which is the sub- coefficient ␥ϭC/T at low T is calculated as ject of Sec. III.A.1. 2 Rigorous treatments of finite-temperature properties ␥ϰ␥0ϩO͑T ͒ (2.270) as well as dynamic properties are possible only when the in the quantum regime with a constant ␥0 , while Hamiltonian matrix is fully diagonalized. However, full diagonalization by, for example, the Householder ␥ϰln T (2.271) method is possible only for even smaller-sized system for zϭd and than the Lanczos method for the ground state only. To relax the severe limitation of tractable size for treating ␥ϰ␥ ϩ␣T1/z (2.272) 0 finite-temperature and dynamic properties, a combina- for zϽd in the classical regime. These are indeed the tion of diagonalization and statistical sampling was in- results obtained in the SCR approximation, which is es- troduced a decade ago (Imada and Takahashi, 1986). In sentially a treatment by the Gaussian fixed point with this method, the trace summation is replaced by a sam- self-consistent determination of the coefficients of the pling of randomly generated basis states, while the Ϫ␶ i t Gaussian term. When this description is justified, it is imaginary-time or real-time evolution, e H or e H ,is natural to expect that ␹(q) follows the criticality of the computed exactly as in the power method. In this quan- mean-field theory and hence the Curie-Weiss law ␹(q) tum transfer Monte Carlo method and the quantum mo-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1093 lecular dynamics method, the tractable system size is ex- matrix by properly considering the boundary condition tended to sizes comparable to the Lanczos method. A of the system. It was recently applied, as well, to a lattice very similar method was also applied recently to the t-J of stripe such as 6ϫn and 8ϫn square lattices to discuss model by replacing the power method with Lanczos di- 2D systems (White and Scalapino, 1996). The recently agonalization (Jaklicˇ and Prelovsˇek, 1994). developed constrained path-integral method also ap- The quantum Monte Carlo method provides another pears to give high accuracy in any dimension, although it way of calculating strong-correlation effects for rela- relies on a variational estimate (Zhang, Carlson, and tively larger systems. A frequently used algorithm of Gubernatis, 1997). quantum Monte Carlo calculation is based on the path- There is an enormous literature on the numerical in- integral formalism combined with the Stratonovich- vestigation of strongly correlated systems, especially on Hubbard transformation. This takes a similar starting models for high-Tc cuprates. Because extensive reviews point to that of the method described in Secs. II.C and of numerical results on strongly interacting lattice fer- II.D. The Stratonovich-Hubbard variable employed in mion models are available elsewhere (Scalapino, 1990; the Monte Carlo calculation is usually chosen to break von der Linden, 1992; Imada, 1993d, 1995d; Dagotto, up the interaction into two diagonal operators. A crucial 1994), here we give only a brief summary of results di- difference from the Hartree-Fock-type treatment is that rectly related to the metal-insulator transition and the dynamic and spatial fluctuations of the Stratonovich- anomalous metallic states near it. We do not discuss in Hubbard variables are taken into account numerically in detail the superconducting properties studied by nu- an honest fashion in contrast to saddle-point estimates, merical methods. The reader is referred to other reviews static approximations, or Gaussian approximations em- (Scalapino, 1995, and the references cited above). ployed in the mean-field-type treatments. Therefore, in Near the MIT point, various quantities exhibit critical terms of interaction effects, the quantum Monte Carlo fluctuations. Here we discuss calculated results for the study provides a controlled method of calculation. A Hubbard model, the d-p model, and the t-J model. drawback of the numerical methods is that they can Most of the effort in numerical studies is concentrated in treat only finite-size systems. Thus careful analyses of 1D and 2D systems without orbital degeneracy, partly finite-size effects are important in numerical studies, es- due to the interest in these systems in relation to the pecially for low-energy excitations. high-Tc cuprates and partly due to the feasibility of es- There exist various types of algorithms for quantum timating finite-size effects within present computer capa- Monte Carlo simulations. For the algorithm for fermion bilities. Another important motivation for investigating systems with the Stratonovich-Hubbard transformation low-dimensional systems is that they show a variety of described above, readers are referred to the literature interesting and anomalous quantum fluctuations. To un- (von der Linden, 1992; Hatano and Suzuki, 1993; Imada, derstand real complexity in d-electron systems with or- 1993d, 1995d). In addition to this algorithm, several dif- bital degeneracy, an algorithm for the degenerate Hub- ferent methods are known. The so-called world-line al- bard model, treating the MIT and orbital and spin- gorithm is a powerful method for nonfrustrated quan- fluctuation effects, was developed and applied by tum spin systems as well as for fermion models in one Motome and Imada (1997, 1998). dimension. Recently for the world-line method, two use- ful algorithms have been introduced. One is the cluster 1. Drude weight and transport properties updating algorithm or the loop algorithm (Evertz, Lana, As we discuss in Sec. II.F.3, the Drude weight and the and Marcu, 1993; Evertz and Marcu, 1993), which was charge compressibility are the most basic and relevant developed based on the original cluster algorithm for quantities for discussing the Mott transition. The Drude classical systems (Swendsen and Wang, 1987). By using weight D for periodic systems is defined from the the cluster algorithm, one can update the configuration frequency-dependent conductivity ␴(␻)as of a large cluster at once, so that the sample can be more efficiently updated with shorter autocorrelation time. ␴͑␻͒ϭD␦͑␻͒ϩ␴reg͑␻͒ (2.273) This is especially powerful for example, when the critical at Tϭ0 where ␴reg describes the contribution from in- slowing down is serious near the critical point. The other coherent excitations and must be smooth as a function useful algorithm is for continuous time simulation, of ␻. where the path integral can be simulated without taking The Drude weight has been calculated by the exact a discrete time slice, so that the procedure to take the diagonalization method for the 1D and 2D Hubbard continuum limit of the breakup in the time slice can be models as well as for the t-J models.2 Although the ex- avoided (Beard and Wiese, 1996). In 1D systems, some trapolation to the thermodynamic limit is difficult be- other efficient methods of numerical calculation have cause of small system sizes, it appears to show that the also been developed. The density-matrix renormalization-group (DMRG) method developed by White (1993) is a typical example, where the numerical 2Exact diagonalization studies include those of Moreo and renormalization-group method originally developed by Dagotto, 1990; Sega and Prelovsˇek, 1990; Stephen and Horsch, Wilson (1975) is successfully extended to a study of the 1990, 1992; Fye et al., 1991; Wagner, Hanke, and Scalapino, ground state of 1D lattice models using a 1991; Dagotto, Moreo, Ortolani, Riera, and Scalapino, 1991; renormalization-group scheme applied to the density Tohyama and Maekawa, 1992; Tsunetsugu and Imada, 1997.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1094 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 16. Diagonalization result for the total kinetic energy K in (a) and Drude weight D in (b) as a function of electron filling ͗n͘ in the ground state for the 4ϫ4 cluster of the Hub- bard model: triangles, U/tϭ4; squares, U/tϭ8; pentagons, U/tϭ20; hexagons, U/tϭ100. The curves without symbols are for Uϭ0. From Dagotto et al., 1992.

Drude weight decreases continuously and vanishes at half filling. The vanishing Drude weight clearly shows the appearance of the Mott insulator at half filling. The Hubbard model and the t-J model show similar behav- iors, as can be seen in Figs. 16 and 17. In 1D systems, exactly solvable models follow Dϰ␦, as we show in Sec. II.F. In constrast, 2D systems seem to follow Dϰ␦2 (Tsunetsugu and Imada, 1998). FIG. 18. Optical conductivity ␴(␻)ofthet-Jmodel with one Experimental results on the cuprate superconductors hole at Tϭ0: (a) two-dimensional 4ϫ4 cluster for three and several other transition-metal oxides such as choices of J/tϭ0.25, 0.5, and 0.75; (b) one-dimensional 16-site cluster at J/tϭ0.5 for the Sϭ1/2 ground state (solid curve) and La Sr TiO appear to show similar continuous reduc- 1Ϫx x 3 the Sϭ15/16 ferromagnetic state (dashed curve). Note that the tion of the Drude weight with decreasing doping con- Drude peak appears at a finite frequency ␻ϳ0.1 due to the centration (see Sec. IV). In the strict sense, the Drude open boundary condition. The spectra are broadened with weight is defined as the singularity of the conductivity at width 0.1. The inset in (a) shows the case with a higher reso- lution (the broadening is 0.02). The inset in (b) shows the weight integrated from 0 to ␻. From Stephan and Horsch, 1990.

␻ϭ0 and Tϭ0, but experimental results always suffer from finite-temperature broadening as well as lifetime effects from the impurity scattering. This issue is further discussed in Secs. II.F and IV. The frequency-dependent conductivity ␴(␻) has also been calculated in the Hubbard and t-J models (Inoue and Maekawa, 1990; Moreo and Dagotto, 1990; Sega and Prelovsˇek, 1990; Stephen and Horsch, 1990). It was pointed out that the 2D system under doping has large weight in the incoherent part ␴reg(␻) inside the original charge gap. Upon doping, in addition to the growth of the Drude-like part, the weight is progressively trans- ferred from the higher ␻ region above the charge gap of the insulator to the region inside the gap as shown in FIG. 17. The same as Fig. 16 as a function of hole doping Fig. 18. This is similar to what happens in the cuprate concentration ␦ϵxϭ1Ϫ͗n͘ for the 4ϫ4 cluster of the t-J superconductors (see, for example, Uchida et al., 1991) model: squares, J/tϭ0.1; solid triangles, J/tϭ0.4; open tri- as discussed in Sec. IV.C. It was also pointed out that angles, J/tϭ1. From Dagotto et al., 1992. the long tail of ␴(␻) at large ␻ seen in the numerical

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1095

FIG. 20. Temperature dependence of resistivity ␳ for 4ϫ4 FIG. 19. Optical conductivity at finite temperatures for 4ϫ4 cluster of the t-J model for various doping concentrations: cluster of the t-J model at J/tϭ0.3 and ␦ϭ3/16 where ␳0 dashed curves, J/tϭ0; solid curves, J/tϭ0.3. The same nota- 2 ϭប/e , with e being the electronic charge and nhϵ␦ the dop- tion as Fig. 19. From Jaklicˇ and Prelovsˇek, 1995a. ing concentration. From Jaklicˇ and Prelovsˇek, 1995a. the marginal Fermi-liquid hypothesis (for the marginal results is similar to the experimental results (Azrak Fermi liquid, see Sec. II.G.2, and for experimental as- et al., 1994). In contrast with 2D systems, the incoherent pects of the cuprates, see Sec. IV.C). The dc resistivity part of the charge excitation is not clearly seen in 1D has indeed been calculated at relatively high tempera- systems under doping, as shown in Fig. 18 (Stephen and tures (Jaklicˇ and Prelovsˇek, 1995a, 1995b; Tsunetsugu Horsch, 1990). The transferred weight from the region and Imada, 1997). What should be noted here is that the above the gap is exhausted only in the coherent Drude linear-T resistivity RϰT characteristic in the high- part. Recently, it was suggested that the integrable temperature limit with totally incoherent dynamics model in 1D has a finite Drude weight with the ␦- (Ohata and Kubo, 1970; Rice and Zhang, 1989) seems to function peak at ␻ϭ0in␴(␻), even at finite tempera- be retained even in the temperature region below the tures (Zotos and Prelovsˇek, 1996). It should be noted superexchange interaction for the planar t-J model (see that the overall feature of ␴(␻) of the single ladder sys- Fig. 20). This seems to be related to the suppression of tem shows an incoherent tail similar to the 2D case, in coherence due to the anomalous character of the Mott contrast to the naive expectation of similarity to 1D sys- transition (see Secs. II.F.9 and II.G.2). Broad incoherent tems (Tsunetsugu and Imada, 1997). tails roughly scaled by 1/␻ are seen not only in the high- For the t-J model in 2D, in the temperature range of Tc cuprates but also in many transition-metal com- the superexchange interaction and lower, ␴(␻) shown in pounds near the Mott transition point. We discuss ex- Fig. 19 and calculated by Jaklicˇ and Prelovsˇek (1995a) amples of this unusual feature in Sec. IV, for example in appears to follow Secs. IV.B and IV.D. The optical conductivity of the Mn 1ϪeϪ␤␻ and Co compounds discussed in Secs. IV.F and IV.G ␴͑␻͒ϭ C͑␻͒ (2.274) shows even more incoherent ␻ dependence. ␻ The Hall coefficient is another interesting subject of and the current correlation function the transport properties. The frequency-dependent Hall coefficient RH(B,␻) in a magnetic field B is defined by ϱ ␴ /␶ i␻t 0 C͑␻͒ϵ dt e ͗j͑0͒j͑t͒͘Ӎ 2 2 (2.275) 1 ␴x,y͑␻͒ ͵Ϫϱ ␻ ϩ͑1/␶͒ RH͑B,␻͒ϭ , B ␴x,x͑␻͒␴y,y͑␻͒Ϫ␴x,y͑␻͒␴y,x͑␻͒ with temperature-independent ␴0 and ␶. The scale of 1/␶ (2.276) seems to be larger than or comparable to the transfer t (Jaklicˇ and Prelovsˇek, 1995a). This form implicitly as- where sumes that the carrier dynamics are totally incoherent. ie2 ␴ ͑␻͒ϭ D ͑␻͒ (2.277) We discuss this problem in Secs. II.F.9 and II.G.2, as ␣,␥ ␻ϩi␩ ␣,␥ well as a related problem in Sec. II.E.3. Jaklicˇ and Pre- lovsˇek claimed the similarity of this result to the experi- with mentally observed transport properties in the normal 1 ͗K͘ ϱ state of high-T cuprates. From the above form, ␴(␻) D ͑␻͒ϭ Ϫ ␦ Ϫ dt eϪi␻t͗j ͑t͒;j ͑0͒͘ c ␣,␥ Ld d ␣,␥ ͵ ␣ ␥ has a rather long tail proportional to 1/␻ in the range ͩ 0 ͪ 0Ͻ␻Ͻ1/␶, while the dc conductivity is proportional to (2.278) ␤, consistent with experimental results on cuprates and for a linear system of size L and the infinitesimal con-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1096 Imada, Fujimori, and Tokura: Metal-insulator transitions vergence factor ␩ (Shastry, Shraiman, and Singh, 1993; Assaad and Imada, 1995). The canonical correlation ͗P;Q͘ is defined by ␤ d␭ Ϫ␤ ␭ Ϫ␭ ͗P;Q͘ϵ Tr͓e He HPe HQ͔. (2.279) ͵0 ␤ Here the averaged kinetic energy is denoted by ͗K͘ and the current operator is given by

† j ϭϪi ͓tជ ជ ͑Aជ ͒c cជ ជ ␣ ͚ i,a␣ ជi,␴ i,ϩa␣ ,␴ ជi,␴

Ϫtជ ជ ជ ͑Aជ ͒cជ ជ cជ ͔ (2.280) iϩa␣ ,Ϫa␣ iϩa␣ ,␴ i,␴ and

2␲i ជi ϩaជ ជ ជ ជ tជi ,aជ͑A͒ϭϪt exp A•dl , (2.281) ͩ ⌽0 ͵ជi ͪ with the flux quantum ⌽0 and the vector potential Aជ . Here, a hypercubic lattice structure is assumed for sim- plicity. Since the effect of the lifetime ␶ of the charge is washed out in the frequency ␻ӷ1/␶, RH at large ␻ rep- resents a simpler and more intrinsic property of strong correlations than the zero-frequency Hall coefficient. In FIG. 21. Temperature dependence of high-frequency Hall co- fact it is a better probe to measure the carrier density, as efficient R* and equal-time spin structure factor S(␲,␲)at␦ pointed out by Shastry et al. (1993). At ␻ ϱ, one can H → ϭ0.05 for 6ϫ6 cluster of the 2D Hubbard model. Solid line show with no symbols corresponds to the U/tϭ0 result. From As- d saad and Imada, 1995. 4L i͗͑jx ,jy͒͘ R*ϵR ͑␻ ϱ͒ϭϪ (2.282) H H → e2B ͗K͘2 with the commutator (a,b)ϭabϪba. Recently the ␻- is gradually released with the growth of antiferromag- dependent Hall coefficient was measured experimentally netic correlations and up- and down-spin electrons start (Kaplan et al., 1996). forming large Fermi surface consistently with the Lut- The temperature dependence of the Hall coefficient tinger volume. This state with large carrier number leads RH* as well as its frequency dependence has been calcu- to a small amplitude of the Hall coefficient. The sign of lated by several authors (Shastry, Shraimann, and Singh, RH itself may depend on details of the Fermi surface. If 1993; Assaad and Imada, 1995; Tsunetsugu and Imada, the curvature of the Fermi surface is hole-like, RH* may 1997). In the strong-correlation regime of 2D systems, it be positive. shows that, at small hole doping, RH* is electronlike Another interesting crossover happens when a spin (RH*Ͻ0) with small amplitude at TϾU, while it be- gap or pseudogap due to singlet pairing is formed. Al- comes hole-like (RH*Ͼ0) with large amplitude at JϽT though the tractable system size is limited to small clus- ϽU where U and J are the energy scales of the repul- ters, this was numerically studied by explicitly forming a sive interaction and the superexchange interaction, re- spin gap in a ladder structure of the lattice. Below the spectively. When T is smaller than the energy scale of spin-gap temperature, RH* becomes large and generally J, RH* again decreases and becomes negative (electron- positive (Tsunetsugu and Imada, 1997), and the system like; see Fig. 21). is transformed essentially to single-component bosons These two crossovers are explained by the following near the Mott insulator, although coexistence of singlet simple physical picture: At TϾU, the Hall coefficient pairs and quasiparticles with dynamic fluctuations may approaches the value of noninteracting electrons be- show aspects different from simple bosons. Because the cause the interaction effect may be neglected. In the re- hard-core bosons near the Mott insulator are mapped to gion JϽTϽU, the interaction effect becomes important, the case of dilute bosons near the vacuum by the while the spin degrees of freedom are still degenerate, electron-hole transformation, the single-component so that the particles behave essentially as spinless fermi- bosons near the Mott insulator behave as dilute holes. ons. In terms of the spinless fermions, the band is com- This is the basic reason for hole-like RH* with large am- pletely filled for the filling corresponding to the Mott plitude. insulator. Therefore a system with small hole doping is All the above seem to be consistent with the mea- indeed expected to behave as hole-like with small car- sured temperature dependence of RH in the high-Tc cu- rier number (that is, RH has large amplitude). When the prates if the RH at large and small frequencies are quali- temperature is further lowered to TϽJ, the spin entropy tatively similar. As is discussed in Secs. IV.C.1 and

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1097

IV.C.3, the experimentally measured RH in the under- Hanke, and W. von der Linden, 1995; Preuss et al., doped region showed strong temperature dependence 1997). Figure 22 shows an example of the doping con- from small amplitude at higher temperatures to large centration dependence of ␳(␻). The density of states at and positive values basically scaled by the inverse hole half filling clearly shows the opening of the charge gap, concentration 1/␦ at low temperatures. The crossover as we discuss below. The chemical potential shifts imme- diately upon doping to the edge of the ‘‘Hubbard gap.’’ occurred at the temperature Tcr where the uniform magnetic susceptibility decreased remarkably with de- The chemical potential slowly shifts upon further doping creasing temperature (Nishikawa et al., 1993; Hwang within a large density of states observed near the gap et al., 1994; Nakano et al., 1994). All of these properties edge. This shift is accompanied by a drastic and continu- can be explained by gradual and progressive pseudogap ous reconstruction from ‘‘upper’’ and ‘‘lower’’ Hubbard band structure to a single merged band structure. When formation (preformed singlet pair fluctuation) below the the doping concentration is low, in the 2D Hubbard or crossover temperature T , which decreases from the cr t-J models, A(k,␻) shows dispersive quasiparticle-like value ϳ500ϳ600 K at small doping to0Kintheover- structure with a bandwidth comparable to that of J in doped region (␦ϳ0.2ϳ0.3). addition to broad incoherent backgrounds, as in Fig. 23. In the case of ladder systems, with the lattice structure This basic and overall structure seems to be robust in of two chains coupled by rungs, however, the particular some interval of doping range. The dispersive part has a geometry of the ladder leads to a large negative value of flat band structure with stronger damping around k RH* below the temperature of spin-gap formation, al- ϭ(␲,0) and three other equivalent points (Bulut, Scala- though it has been shown in a small cluster study that a pino, and White, 1994b; Dagotto, Nazarenko, and Bon- substantial interladder coupling generally drives RH* to a insegni, 1994). The dispersions there appear to be flatter large positive value (Tsunetsugu and Imada, 1997). In than that expected from the usual van Hove singularity. most of the above cases, the frequency dependence of This flatter dispersion could be related to other unusual RH(␻) has not yet been examined systematically. How- properties observed numerically, such as incoherent ever, the calculated ␻ dependence available so far ap- charge transport (Sec. II.E.1), singular charge compress- pears to show similar behavior between small and large ibility (Sec. II.E.3), and incoherent spin dynamics (Sec. ␻. It is not clear whether RH(␻) has a strong ␻ depen- II.E.4). This idea is discussed in Sec. II.F.11. Recently, it dence at the frequency scale of the inverse relaxation was shown that the dispersion around (␲,0) follows Ϫ1 time ␶ of carriers. quartic dispersion (ϳk4) in agreement with the picture in Sec. II.F.11 (Imada, Assaad, et al., 1998). Although 2. Spectral function and density of states the numerical results of clusters have suggested the ex- Two other quantities important to our understanding istence of a quasiparticle band of width ϳJ, this ‘‘band’’ of the MIT and anomalous metallic states are the spec- itself could be dominantly incoherent near the Fermi tral function A(k,␻) and the density of states ␳(␻) de- level. Because these numerical results are results on sim- fined from Eqs. (2.50), (2.56), and (2.57) as plified models, more complicated features can also be observed in realistic situations. Effects of impurity levels ␳͑␻͒ϭ͚ A͑k,␻͒, (2.283) and the long-range Coulumb interaction are among k these complexities. We compare these results on theo- retical models with photoemission results on transition- 1 A͑k,␻͒ϭϪ Im GR͑k,␻͒, (2.284) metal oxides in Secs. IV.C.3, IV.C.4 and IV.D.1. In par- ␲ ticular, in a recent comparison of 1D and 2D systems by ϱ Kim et al. (1996), it was suggested that weakly dispersive GR ϭϪi dt ei␻t t c t c† ͑k,␻͒ ͵ ␪͑ ͕͒͗ k␴͑ ͒, k␴͑0͖͒͘, bands with incoherent character in 2D be contrasted Ϫϱ with a clearer dispersion in 1D, as is discussed in Sec. (2.285) IV.D.1. This difference between 1D and 2D is related where the thermodynamic average ͗¯͘ and the Heisen- with the difference in the Drude weight discussed in Sec. Ϫi t i t berg representation O(t)ϭe H Oe H are introduced. II.E.1. In the ground state, A(k,␻) may be rewritten as Eq. Angle-resolved photoemission spectra of a model (2.54). Experimentally, A(k,␻) is measured by angle- compound of the parent insulator of high-Tc supercon- resolved photoemission and inverse photoemission ex- ductors, Sr2CuCl2O2 (Wells et al., 1995), can be com- periments as described in Sec. II.C.4. The angle- pared with the Monte Carlo simulations of the Hubbard integrated intensity gives ␳(␻). Numerically, A(k,␻) as well as of the t-J model (Dagotto, 1994), since the and ␳(␻) have been calculated either by the exact diago- single-band Hubbard model and the t-J model are effec- nalization method or by quantum Monte Carlo calcula- tive Hamiltonians for the CuO2 plane. (The top of the tions combined with the maximum-entropy method. occupied band is derived from the d9 d9Lគ local-singlet Numerical results on A(k,␻) and ␳(␻) for the Hub- spectral weight and the bottom of the→ unoccupied band bard model and the t-J model show drastic but continu- forms the d9 d10 spectral weight). Figure 24 shows that ous change upon doping (Dagotto, Moreo, et al., 1991, in going away→ from kϭ(0,0) a peak disperses towards 1992; Stephen and Horsch, 1991; Tohyama and lower binding energies; along the (␲,␲) direction, the Maekawa, 1992; Bulut, Scalapino, and White, 1994a; peak reaches closest to EF at ϳ(␲/2,␲/2) and then is Moreo, Haas, Sandvik, and Dagotto, 1995; Preuss, shifted back toward higher binding energies, in good

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1098 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 23. The spectral weight A(k,␻) by quantum Monte Carlo calculations of the 8ϫ8 Hubbard model at Uϭ8t. Dark areas correspond to large spectral weight, white areas to small spectral weight. The solid lines in (a) and (c) are tight-binding fits of the data while in (b) they are the quantum Monte Carlo result at ͗n͘ϭ1.0. Preuss et al., 1997.

agreement with the prediction of the Monte Carlo simu- lations using the t-J model and the Hubbard model. Along the (␲,0) direction, on the other hand, the peak does not disperse so close to EF , although the t-J and Hubbard-model calculations predict a peak position nearly equal to the ϳ(␲/2,␲/2) point. That the occupied band has maxima along the (␲,0)-(0,␲) line as in the Monte Carlo study is a generic feature of the single- band Hubbard model and is also the case in the weak- coupling spin-density wave picture (Bulut et al., 1994b). The discrepancy between theory and experiment around the (␲,0) point remains unexplained. Around (␲,0) the peak is rather broad due to large incoherent character, leading to uncertainty that might be the origin of the FIG. 22. Doping concentration and temperature dependence discrepancy. Another possibility is that a more realistic of the density of states ␳(␻) for 8ϫ8 Hubbard model at U/t ϭ8 for (a) ␦ϭ0, (b) ␦ϭ0.13, and (c) ␦ϭ0.3. The ordinate model, such as the t-tЈ-J model or d-p model including N(␻) is the density of states ␳(␻) in our notation. From Bulut, apex oxygens, will be necessary (see, for example, Naza- Scalapino, and White, 1994a. renko et al., 1995).

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1099

FIG. 24. Quasiparticle dispersions in angle-resolved photo- FIG. 25. The renormalized interaction Urn derived so as to emission spectra of Sr2CuCl2O2 (Wells et al., 1995) compared reproduce the numerically observed charge gap by the with t-J model calculations given by solid curve (left panel) Hartree-Fock calculation is plotted against the bare U for the and crosses with an interpolation of crosses by dotted curve 2D Hubbard model (from N. Furukawa and M. Imada, unpub- (right panel). From Bulut, Scalapino, and White, 1994a, 1994b. lished). This indicates that the Hartree-Fock gap becomes cor- rect for small U. 3. Charge response In connection with the vanishing Drude weight, a gradually and continuously vanish when one controls the charge excitation gap opens at half filling. This is directly doping concentration from the metallic phase to the observed in the density of states at half filling (White, MIT point ␦ϭ0. The scaling plot of ␹c even shows a 1992). A direct estimate of the charge gap is also pos- singular divergence in the form sible from Eq. (2.16). Both the quantum Monte Carlo ␹ ϰ1/͉␦͉ (2.287) finite-temperature algorithm applied by Moreo, Scala- c pino, and Dagotto (1991) and the projector algorithm for ␦Þ0, as illustrated in Fig. 26 (Furukawa and Imada, for Tϭ0 applied by Furukawa and Imada (1991b, 1991c, 1992, 1993). The singular divergence in the form of 1992) to the nearest-neighbor Hubbard model at U/t (2.287) is observed not only in the nearest-neighbor ϭ4 show the charge gap ⌬cϳ0.6. Recently a more Hubbard model (1.1a) but also in a model extended by elaborate algorithm for estimating zero-temperature adding a finite next-nearest-neighbor transfer tЈ. This properties in the Mott insulating phase was developed indicates that the origin of this scaling has nothing to do using stabilized matrix calculations (Assaad and Imada, either with electron-hole symmetry, perfect nesting, or 1996a). The application of this algorithm to the Hubbard with the Van Hove singularity because they are absent model at U/tϭ4 showed ⌬c /tϭ0.66Ϯ0.015 (Assaad and in a model with nonzero tЈ. Similar behavior is also ob- Imada, 1996b). The amplitude of the charge gap exam- served in the 2D t-J model (Jaklicˇ and Prelovsˇek, 1996; Ϫ1 ined as a function of U in the 2D Hubbard model agreed Kohno, 1997). In 1D systems the same scaling ␹cϰ␦ is with the prediction of the Hartree-Fock approximation derived in exactly solvable cases, as we shall see in Sec. for small U but deviated for larger U, as is expected. II.F.6. There we shall discuss how the 1D and 2D sys- Figure 25 shows how the charge gap scaled as a function tems differ in terms of the quantum phase transition de- of U. In Fig. 25, Uren is the interaction with which the Hartree-Fock equation of the Hubbard model gives the same charge gap as the numerically observed value in the quantum Monte Carlo. The Hartree-Fock result pre- dicts that ⌬c is scaled as ⌬cϳt exp͓Ϫ2␲ͱt/U͔, while Hirsch (1985b) suggested a slight modification, namely, replacing the prefactor t with ͱtU. The charge compressibility ␬ is essentially the same as the charge susceptibility ␹c and is defined as 2 (␮ (2.286ץ/nץ␹cϭn ␬ϭ where n is the electron density and ␮ is the chemical potential. The charge susceptibility vanishes in the Mott insulating phase because of incompressibility. However, quantum Monte Carlo results on 2D systems show that doped systems can be very compressible near the MIT FIG. 26. Chemical potential vs ␦2 for the 2D Hubbard model on the metallic side (Otsuka, 1990; Furukawa and of various sizes at Tϭ0: open symbols, without next-nearest- Imada, 1991b, 1992, 1993). In 2D systems, in contrast to neighbor transfer tЈ; filled symbols, with tЈϭ0.2. From Fu- the Drude weight, the charge susceptibility ␹c does not rukawa and Imada, 1993.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1100 Imada, Fujimori, and Tokura: Metal-insulator transitions spite their seemingly identical behavior. This is due to 4. Magnetic correlations the difference in universality class. Experimental aspects Effects of spin fluctuations are another intensively of ␹ and the chemical-potential shift upon doping of a c studied issue. In low-dimensional systems, the antiferro- 2D MIT are discussed in Secs. IV.C.1 and IV.G.5 and magnetic correlation is under the strong influence of interpretation in terms of Fermi-liquid theory is given in quantum fluctuations. The one-dimensional spin-1/2 Sec. II.D.1. Heisenberg model and the Hubbard model at half filling When the charge susceptibility is singularly divergent show a power-law decay of the antiferromagnetic corre- as a power law of ␦ for ␦ ϩ0, the single-particle de- lation at Tϭ0, ͗S(r) S(0)͘ϰ1/r. For spin-1/2 systems, scription of low-energy excitations→ is made possible only • the spin-wave theory in 2D predicts the existence of an- by the divergence of the effective mass of a relevant tiferromagnetic order (Anderson, 1952; Kubo, 1952). particle. This is in contrast with the usual transition be- Although the existence of long-range order is proven for tween metals and band insulators, in which the number a spin Sу1 on a square lattice (Neves and Perez, 1986; of carriers goes to zero as the effective mass of the qua- Kubo and Kishi, 1988), at the moment, no rigorous siparticle remains finite. This numerical study by Fu- proof for long-range order exists at Sϭ1/2, even for rukawa and Imada prompted subsequent theoretical nearest-neighbor coupling on a square lattice. Quantum studies of continuous MITs based on the scaling theory, Monte Carlo results on the square-lattice Heisenberg the main subject of Sec. II.F. model appear to show antiferromagnetic order at Tϭ0 When the charge susceptibility is scaled by Eq. (Okabe and Kikuchi, 1988; Reger and Young, 1988; (2.287), it may be regarded as tending strongly toward Makivic´ and Ding, 1991). In the 2D Hubbard model, phase separation as ␦ 0 although real phase separation → quantum Monte Carlo results also suggest the existence is not achieved. In fact, a continuous transition to a of antiferromagnetic long-range order in the Mott insu- phase-separated state may be probed by the singularity lating phase, ␦ϭ0 (White et al., 1989). The existence of ␹c ϱ. The growth of charge-density fluctuations, ob- long-range order is also inferred in a parameter region → served when ␹c is enhanced is reminiscent of the pro- of the d-p model (Dopf, Wagner, et al., 1992). posal of phase separation by Visscher (1974) and Emery, In the 2D Hubbard model, long-range order appears Kivelson, and Lin (1990). The possibility of phase sepa- to be lost immediately upon doping (Furukawa and ration in the 2D t-J model was examined recently by Imada, 1991b, 1992). The Fourier transform of the Hellberg and Manousakis (1997), who claimed to find a equal-time spin-spin correlation function phase separation for any J/t at close to half filling. This 1 result contradicts the observation of Kohno mentioned S͑k͒ϭ eϪik•r͗S͑r͒ S͑0͒͘dr (2.288) 3 ͵ • above. The numerical methods of these two studies are similar, while Hellberg et al. derived the opposite con- shows a short-ranged incommensurate peak at kϭQ clusion basically by relying on a single data point near away from the antiferromagnetic Bragg point k half filling. Further studies in the t-J model are clearly ϭ(␲,␲) when the doping proceeds as we see in Fig. 27 needed, including a reexamination of the convergence to (Imada and Hatsugai, 1989; Moreo, Scalapino, et al., the ground state for this region. The quantum Monte 1990; Furukawa and Imada, 1992). Similar incommensu- rate peaks are observed in the 2D t-J model (Moreo and Carlo results in the Hubbard model imply that ␹c di- verges only for ␦ ϩ0, and there the transition to a Dagotto, 1990) and in the d-p model (Dopf, Mura- Mott insulator takes→ place without a real phase separa- matsu, and Hanke, 1992). Figure 28 indicates that this tion. If a phase separation took place at finite doping peak value S(Q) becomes finite for ␦Þ0, which means concentration, the quantum Monte Carlo method could the disappearance of long-range order in the whole me- tallic region, ␦Þ0. The scaling of S(Q) appears to follow indeed detect it, as we shall see later in the case of the Ising-like boson t-J model in Sec. II.E.12. We of course S͑Q͒ϰ1/␦, (2.289) have to note here the limitation of the numerical study which can be interpreted as the antiferromagnetic corre- for the ultimate limit of ␦ 0, because the numerical Ϫ1/2 → lation length ␰m scaled by ␰mϰ␦ (Furukawa and results can be obtained only at finite ␦ in finite-size sys- Imada, 1993). Here it should be noted that the existence tems. In any case, the overall numerical results indicate of an antiferromagnetic correlation length in the metal- that rapid growth of the charge compressibility and the lic phase is nontrivial because the correlation at an as- tendency towards phase separation do exist near half ymptotically long distance should follow a power-law filling. It is interesting to regard the critical enhancement decay in the usual spin gapless metals and, naively, this of ␹c at small ␦ as a general tendency not of static but of should lead to divergence of the correlation length. The dynamic fluctuation toward the phase separation. In Monte Carlo results in 2D suggest that the spin correla- fact, static phase separation is suppressed due to the tion of the period given by the wave number Q is con- long-range Coulomb force, ignored in this discussion, stant as in the Heisenberg model for the distance 1Ӷr ␥ and the above controversy is not so serious in real ma- Ӷ␰m while it decays as a power law ϰ1/r with ␥Ͼ2in terials. The dynamic phase separation provides an alter- the region rӷ␰m . In the 1D doped Hubbard model, it native view to Eq. (2.287). We discuss this issue further has been established that the incommensurate spin cor- 1ϩK␳ in Secs. II.H.3, IV.C., and IV.E. relation decays as 1/r for 1ӶrӶ␰m while as 1/r for

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1101

FIG. 28. Inverse of equal-time spin structure factor SϪ1(Q,t ϭ0) at the peak point qϭQ as a function of doping ␦ for the 2D Hubbard model at U/tϭ4 with various sizes. From Fu- rukawa and Imada, 1992.

stance in such d-electron systems with orbital degeneracy as V3Se4,V5Se8, and V3S4. Even in 3D, so far, antiferromagnetic metals are observed only in cases with strong orbital degeneracy. From both theoretical and experimental viewpoints, the stability of antiferro- FIG. 27. Incommensurate peak in the momentum space of magnetic and metallic phases near the Mott insulator for equal-time magnetic structure factor S(q,tϭ0) for the 2D single-band systems, even in 3D, calls for further study. Hubbard model of a 10ϫ10 cluster at Tϭ0: (a) ␦ϭ0.18; (b) ␦ϭ0.26; (c) ␦ϭ0.42; and (d) ␦ϭ0.5 at U/tϭ4. From Furukawa It is not clear whether the lower critical dimension of and Imada, 1992. in metals, dml , is even higher than three for systems without orbital degeneracy, that is, two-component systems. This uncertainty is due in part Ϫ1 rӷ␰m with ␰mϰ␦ (Imada, Furukawa, and Rice, 1992; to the difficulty of finding an effectively single-band sys- Iino and Imada, 1995). K␳ is the exponent of the tem in 3D. Tomonaga-Luttinger liquid given in Sec. II.G.1. This The temperature dependence of the antiferromag- nontrivial way of defining ␰m in 1D and 2D is further netic correlation length ␰m has also been studied exten- discussed in Sec. II.F.10. We note that the immediate sively by numerical approaches. In the undoped case, ␰m destruction of antiferromagnetic order, which is numeri- continuously and rapidly increases until T 0 because → cally observed, may be a characteristic feature in single- the antiferromagnetic transition takes place at Tϭ0in band 2D systems. 2D. Chakraverty, Halperin, and Nelson (1989) pointed Experimentally, so far we have no definite example of out the relation of the quantum Heisenberg model in 2D an antiferromagnetic metal near the Mott insulator in to the O(3) nonlinear sigma model in 3D and suggested highly 2D single-band systems, in agreement with the the existence of two regions, namely, where ␰mϰ1/T at above numerical studies. As is well known and discussed higher temperatures (the quantum critical regime) and in Sec. IV.C, the cuprate superconductors in general lose where ␰m increases exponentially at lower temperatures antiferromagnetic order upon small doping without any in the renormalized classical regime. This basic picture indication of antiferromagnetic metals. Antiferromag- has been more or less confirmed numerically (for ex- netic order survives up to ␦ϭ0.02, while insulating be- ample, Makivic´ and Ding, 1991). havior persists until ␦ϭ0.05 for La2Ϫ␦ Sr␦CuO4. The At finite doping, ␰m(T) was also computed for the 2D MIT at small but finite ␦ is clearly due to Anderson Hubbard model (Imada, 1994a). It displayed a sharp localization under random potentials. contrast with the undoped case. Both ␰m and the peak In 3D systems, however, antiferromagnetic metals value of the equal-time spin structure factor S(Q) may exist, although no numerical study is available so started growing at high temperatures, TϾJ. With de- far. They have been found adjacent to the Mott insula- creasing temperature, however, they showed rapid satu- tor in limited examples, such as NiS2ϪxSex with pyrite ration at rather high temperatures, as in Fig. 29. For structure and V2O3Ϫy , as will be discussed in Sec. IV.A. example, at ␦ϳ0.15, ␰m more or less saturated at TϳJ However, it should also be noted that these two materi- and did not grow below it. This means that ␰m quickly als have orbital degeneracies with spins larger than 1/2 approaches the intrinsic correlation length ϳ1/ͱ␦ given in the insulator, which may stabilize the antiferromag- from Eq. (2.289). The temperature below which ␰m be- netic metals. In the weak-correlation regime, other ex- comes more or less temperature independent is denoted amples of antiferromagnetic metals are known, for in- as Tscr . Tscr rapidly increases from 0 at ␦ϭ0toTscrϳJ at

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1102 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 29. Temperature dependence of S(Q,tϭ0) for various fillings nϭ1Ϫ␦ of the 2D Hubbard model at U/tϭ4. From Imada, 1994a.

␦ϳ0.15ϳ0.2. A similar and consistent result was also obtained in the 2D t-J model by Jaklicˇ and Prelovsˇek (1995b), who applied the finite-temperature Lanczos method mentioned above and found that Im ␹(q,␻)/␻ has a peak around (␲,␲) and grows with decreasing tem- perature, while the width in ␻ and q is temperature in- dependent at TϽJ for the doping ␦ϳ0.15; see Fig. 30. This peak structure is associated with the ‘‘incoherent part’’ of the spin correlation, while the ‘‘coherent part’’ FIG. 30. Dynamic spin susceptibility Im ␹(q,␻)/␻ fora4ϫ4 has only a low weight at lower frequencies when the cluster of the t-J model at ␦ϭ3/16: solid line, T/tϭ0.1; dashed doping concentration becomes low. This coherent part line, T/tϭ0.2; dashed-dotted line, T/tϭ0.3; dotted line, T/t may give rise to the Fermi-surface effect at low tempera- ϭ0.5. From Jaklicˇ and Prelovsˇek, 1995b. tures, which is associated with a power-law decay of the spin correlation at long distances in metals. These numerical results are consistently interpreted ␻ space. The essential features of the numerical results by the following picture, as discussed by Imada (1994a) may be written, approximately and phenomenologically, and Jaklicˇ and Prelovsˇek (1995b): Although it is not for ␻Ͼ0as completely correct, it is useful to analyze the spin corre- C lation as the sum of incoherent and coherent parts. As S ͑q,␻͒ϳ , (2.292) inc ␻2ϩ͓D ͑q2ϩK2͔͒2 the doping concentration decreases, the weight of the s Ϫ1 2Ϫz d coherent part decreases to zero and the incoherent with Kϵ␰m , Dsϰ␰m , Cϰ␰m . Here we take the di- weight becomes dominant. The coherent part comes mension dϭ2. The important point is that ␰m has to be from an asymptotically long-time and long-distance part temperature independent at TϽTscr to be consistent z beyond ␶m(ϵ␰m) and ␰m , where the correlation decays with numerical results. The coherent part Scoh may have as a power law at Tϭ0. Here z is the dynamic exponent strong temperature dependence even below Tscr .In discussed in detail in Sec. II.F. The incoherent part fact, Scoh represents the long-ranged power-law decay of arises from the short-time and short-distance part (␶ the spin correlation and reflects the Fermi-surface effect Ͻ␰z and rϽ␰ ). As ␰ ϱ with decreasing ␦, it is clear m m m→ at low temperatures. Therefore Scoh grows in general as that the incoherent spin correlation dominates and the incommensurate and rather sharp non-Lorentzian peaks coherent weight vanishes. The spin structure factor below the coherence temperature TcohϵTF . The total ϱ structure factor S is schematically illustrated in Fig. 31 as Ϫi␻t iq r S͑q,␻͒ϵ dr dt e e • ͗S͑r,t͒•S͑0,0͒͘ (2.290) the sum of these two parts. Although neutron scattering ͵Ϫϱ may detect this structure coming from Scoh at low tem- peratures, the weight of S becomes smaller at lower may be given as the sum of the coherent and incoherent coh doping. Therefore q-integrated or ␻-integrated values contributions, are generally dominated by Sinc at low doping. Because S͑q,␻͒ӍSinc͑q,␻͒ϩScoh͑q,␻͒. (2.291) K and Ds are temperature independent below Tscr , Im ␹͑q,␻͒ϭ͑1ϪeϪ␤␻͒S͑q,␻͒ (2.293) This phenomenological way of writing S(q,␻) with co- herent and incoherent contributions was also suggested for ␻Ͼ0 has a temperature dependence coming mainly by Narikiyo and Miyake (1994). The incoherent contri- from the Bose factor (1ϪeϪ␤␻). We note that Im ␹ and Ϫ1 Ϫz bution Sinc has typical widths ␰m in q space and ␰m in S at ␻Ͻ0 are determined from the requirement

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1103

eigenvectors in the Hilbert space of N particles. The two-body correlation function is also independent of ␮ within the charge gap. However, the single-particle Green’s function defined in Eq. (2.44) as ͑r,␶͒ϭ͗Tc͑r,␶͒c†͑0,0͒͘ (2.295) G depends on ␮ because it propagates a particle created above the gap in the Nϩ1-particle sector for the interval of ␶Ͼ0. This Green’s function yields a factor of eϪ␶(EnϪ␮) in Eq. (2.295) for a created particle with en- ergy En . Then the ␮ dependence of (r,␶) is simply e␮␶. Therefore, when (r,␶)at␮ϭG0 is calculated, (r,␶)at␮Þ0 is givenG as (r,␶,␮Þ0)ϭ (r,␶,␮ FIG. 31. Total typical and schematic structure of S(q,␻) Gϭ0)e␮␶. This means that all the G(r,␶) for ␮ withinG the around the antiferromagnetic peak point. The incoherent con- charge gap can be obtained by calculatingG (r,␶) once at tribution is illustrated as a bold gray curve and the total am- ␮ϭ0. Because there may be numericalG instability for plitude is given by a black solid curve. large ␶ due to the factor e␮␶, a numerical stabilization technique for the matrix calculation is necessary (As- Im ␹(q,Ϫ␻)ϭϪIm ␹(q,␻), leading to the result that saad and Imada, 1996a). ͐dq Im ␹(q,␻) is universally scaled as In any case, the calculated Green’s function provides the localization length ␰l defined by Ϫ␤␻ dq Im ␹͑q,␻͒ϰ͑1Ϫe ͒ (2.294) ϱ ͵ ͑r,␻ϭ␮͒ϵ ͑r,␶,␮͒d␶ϳeϪr/␰l. (2.296) Ϫz G ͵0 G at ␻Ͻ␰m when the incoherent contribution is domi- nant. It was pointed out (Imada, 1994a) that this ex- Here ␰l may be regarded as the localization length of the plains the neutron and NMR results in high-Tc cuprates wave function of the virtually created state at the chemi- where ͐dq Im ␹(q,␻) is reported to be universally cal potential. The localization length ␰l has to diverge at scaled by ␻/T (Hayden et al., 1991a; Keimer et al., 1992; the MIT. The scaling of ␰l near ␮c calculated by the Sternlieb et al., 1993) while the NMR relaxation rate above method for the two-dimensional Hubbard model T1ϳconst at TϾTcoh (Imai et al., 1993). The coherence shows temperature T is estimated to be ϳ100ϳ300 K at the coh ␰ ϳ ␮Ϫ␮ Ϫ␯, optimal doping ␦ϳ0.15. More detailed analysis of the l ͉ c͉ antiferromagnetic correlation based on scaling theory is ␯ϭ0.26Ϯ0.05, (2.297) given in Sec. II.F.10. Experimental aspects of high-Tc cuprates are discussed in Secs. IV.C.1 and IV.C.3. as in Fig. 32 (Assaad and Imada, 1996a, 1996b). In Sec. II.F, we compare this exponent ␯ϳ1/4 with the exponent 5. Approach from the insulator side of the compressibility analysis given in Eq. (2.287) in the metallic side, assuming hyperscaling, and find them to be So far we have discussed the Mott transition from the consistent. The confirmation of the large dynamic expo- metallic side. Recently, a method of probing the transi- nent has prompted subsequent studies on the t-U-W tion from the insulator side was developed (Assaad and model in which the two-particle process defined in Eq. Imada, 1996a, 1996b). If the transition is continuous, it (2.347) is introduced explicitly and added to the Hub- can be shown that critical exponents in both sides should bard model because of its greater relevance than that of be the same. Therefore, any information gained about the irrelevant single-particle process characterized by z the transition from the insulator side also helps to clarify ϭ4 (Assaad, Imada, and Scalapino, 1996, 1997; Assaad its character on the metallic side. A large advantage of and Imada, 1997; see also Sec. II.F.9). The (t-U-W) observing the transition from the insulator side is that model has made it possible to study directly the quan- we can avoid the negative-sign problem, which has tum transition between an antiferromagnetic Mott insu- plagued the quantum Monte Carlo method (Loh et al., lator and a d-wave superconductor and thereby to 1990; Furukawa and Imada, 1991a). For the case of clarify the character of the obtained d-wave supercon- electron-hole symmetry, as in the nearest-neighbor Hub- ductor at Tϭ0. Antiferromagnetic correlation is ex- bard model on a bipartite lattice, the negative sign does tremely compatible in the superconducting phase with not appear at half filling (Hirsch, 1983). divergence of the staggered susceptibility. Finite tem- The insulator undergoes a transition to a metal when perature fluctuations were also studied as well. the chemical potential ␮ approaches the critical point ␮c from the region of the charge gap. The ground-state wave function ⌽ in the insulating phase should not ͉ g͘ 6. Bosonic systems change when ␮ changes because the number of particles is fixed at N within the gap, while the Hamiltonian con- Transitions between Mott insulators and superfluids tains ␮ only through the term ␮N irrespective of the in bosonic systems have been the subject of recent inten-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1104 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 33. Schematic phase diagram of a metal and a Mott in- sulator in the plane of ␮/t and t/U. Route A shows filling- control transition, FC-MIT (generic transition), while route B is the bandwidth-control transition, BC-MIT (multicritical transition). Dotted curves illustration density contour lines.

ment was speculated to be the persistence of the Ising- like component order for ␦р␦c .

F. Scaling theory of metal-insulator transitions

The Mott transition defined as the transition between a metal and a Mott insulator, has two types, the filling- control transition (FC-MIT) and the bandwidth-control FIG. 32. Localization length ␰l vs ͉␮Ϫ␮c͉ for the Hubbard model at U/tϭ4 for (a) two-dimensional systems, (b) one- transition (BC-MIT), as explained in Sec. I and illus- dimensional systems. Monte Carlo data are illustrated by solid trated in Fig. 33. Both of these transitions are controlled circles. From Assaad and Imada, 1996b. by quantum fluctuations rather than by temperatures. The FC-MIT has as control parameters the electron con- centration or the chemical potential. The BC-MIT has as sive studies. When disorder drives the transition in a control parameter the ratio of the bandwidth to the bosonic systems, a superfluid undergoes a transition to a interaction, t/U. The quantum fluctuation is enhanced Bose glass phase. Studies of the critical exponents of the for large t/U, which stabilizes the metallic phase, transition from superfluid to Bose glass were motivated namely, the quantum liquid state of electrons, while by the proposal of the scaling theory (Fisher et al., 1989). large U/t makes the localized state more stable. Because The results of quantum Monte Carlo calculations appear these control parameters are intrinsically quantum me- to support the scaling theory, as discussed in Sec. II.F.12 chanical, the MIT is the subject of research in the more (So”rensen et al., 1992; Wallin et al., 1994; Zhang et al., general field of quantum phase transitions. The transi- 1995). tion itself can take place either through continuous In the transition between a superfluid and a Mott in- phase transitions or through a first-order transition ac- sulator, an interesting case is that of multicomponent companied by hysteresis. Critical fluctuations of metallic bosons. In analogy with the fermion t-J model (2.13), states can be observed more easily near the continuous the boson t-J model of two-component hard-core transition point. In this section, we look more closely at bosons was investigated by Imada (1994b). In the case of continuous transitions. Ising-exchange for J, Monte Carlo results clearly show a In Sec. II.D, various theoretical approaches to the phase separation into the Mott insulating phase at ␦ Mott transition and correlated metals are described, in- ϭ0 and another component-ordered phase at a finite ␦c cluding the Gutzwiller approximation, the Hubbard ap- near the Mott insulator (Motome and Imada, 1996). A proximation, the infinite-dimensional approach, slave- remarkable fact is that the component correlation length particle methods, and spin-fluctuation theories. diverges at the continuous transition point ␦ϭ␦c when Although their physical consequences vary and some- the filling is changed from ␦Ͼ␦ to ␦ ␦ while phase times conflict with each other, these methods share a c → c separation takes place for ␦Ͻ␦c . Strong mass enhance- common limitation in that they all rely on the mean-field ment and vanishing superfluid density are observed with approximation. The level of mean-field approximations ␦ ␦c . This appears to come from a mechanism of mass used ranges from dynamically correct treatments that enhancement→ at finite ␦ which is different from the neglect spatial correlations, as in the dϭϱ approach, to mechanism of the mass divergence in the 2D Hubbard simple static approximations on the Hartree-Fock level. model discussed above. The origin of the mass enhance- In Sec. II.D.9, we have seen that the magnetic transition

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1105 in metals independent of the MIT seems to be described The universality class obtained in 2D is characterized by by mean-field fixed points for antiferromagnetic transi- an unusually large dynamic exponent zϭ4. An impor- tions at dу3 and for ferromagnetic transitions at dу2 tant consequence of this new universality class is the or dу3 under certain assumptions. However, in this sec- suppression of coherence as compared to that in the tion, we discuss some rather different features of the transition between band insulators and metals. We dis- cuss in Sec. II.F.9 how and why the coherence of charge Mott transition. and spin dynamics is suppressed for the Mott transition. In general, mean-field approximations are justified if We also discuss the instability of the large-z state to the dimension is high enough. However, in the Mott superconducting pairing. The success of the hyperscaling transition of fermionic models, the upper critical dimen- description is a nontrivial consequence for large z be- sion beyond which the mean-field approximation is jus- cause, generally speaking, the upper critical dimension tified is not so straightforwardly estimated as in simple becomes low for large z. A possible microscopic origin magnetic transitions of the Ising model. As we saw in for hyperscaling with zϭ4 in 2D is discussed in Sec. Sec. II.D.6, the infinite-dimensional approach provides II.F.11. In Sec. II.F.12, superfluid-insulator transitions an example where the mean-field approximation be- are discussed. comes exact at dϭϱ by taking account of dynamic fluc- tuations correctly. On the other hand, other mean-field 1. Hyperscaling scenario approximations do not take account of on-site dynamic Continuous MITs were first analyzed in terms of the fluctuations and incoherent excitations correctly, leading scaling concept in the Anderson localization problem in to failure of the method even in the limit of large dimen- the 70s. Scaling has proven to be useful in establishing sions, because incoherent excitations become over- the marginal nature of 2D systems for localization by whelming near the transition point. disorder (Wegner, 1976; Abrahams et al., 1979). The un- In low-dimensional systems, temporal as well as spa- derlying assumption in this approach is the hyperscaling tial fluctuations from both thermal and quantum origins hypothesis. In this subsection, we introduce the hyper- become important. A generally accepted view is that the scaling description of the MIT caused by the electron mean-field theories for phase transitions become invalid correlation from more general point of view following below the upper critical dimension due to large spatial recent developments (Continentino, 1992, 1994; Imada, fluctuations. 1994c, 1995a, 1995b). Because the MIT is controlled by Below the upper critical dimension, where the mean- quantum fluctuations, we may use control parameters field description breaks down, phase transitions are be- such as the electron chemical potential ␮ or the band- lieved in general to be described by scaling theory, with width t and measure the distance from the critical point hyperscaling assumed. In this subsection we review the by ⌬. The control parameter can either be the chemical scaling theory of the Mott transition and examine its potential to control the filling or the bandwidth (or the validity based on recent numerical work. Its conse- interaction), as can be seen in Fig. 33. The scaling theory quences and experimental relevance are also discussed. assumes the existence of a single characteristic length Among various Mott transitions categorized in Sec. scale ␰, which diverges as ͉⌬͉ 0, and a single character- II.B, ͓I-1 M-1͔, ͓I-1 M-2͔, ͓I-2 M-2͔, and → istic frequency scale ⍀, which vanishes as ͉⌬͉ 0. The ͓I-2 M-4↔͔ are basically↔ characterized↔ by vanishing correlation length ␰ is assumed to follow → carrier↔ number as the MIT is approached. In this case, the universality class of the MIT may be characterized as ␰ϳ͉⌬͉Ϫ␯ (2.298) the same as that for a transition between a band insula- as ⌬ 0, which defines the correlation length exponent tor and a standard metal, as will be discussed in Sec. ␯. The→ frequency scale ⍀ is determined from the quan- II.F.8. On the other hand, for the case of ͓I-1 M-3͔, tum dynamics of the system independently of the length an unusual type of MIT can take place with an↔ anoma- scale in general. The dynamical exponent z determines lous metallic phase. In this case the MIT is characterized how ⍀ vanishes as a power of ⌬, by diverging single-particle mass (Imada, 1993a). As we shall see below and also in Sec. II.E, critical properties ⍀ϳ␰Ϫzϳ͉⌬͉z␯. (2.299) of the MIT are observed numerically in a wide region In critical phenomena, hyperscaling is believed to hold around the critical point. The existence of such a wide between the upper critical dimension d and the lower critical region also leads to a wide region of an anoma- c critical dimension d (see, for example, Ma, 1976). lous metallic state near the MIT with strongly incoher- l Above d , the mean-field description is justified and the ent charge dynamics. In this sense, the scaling argument c hyperscaling description breaks down, whereas below d for the MIT discussed in this section is not limited to a l the transition itself disappears. Here, d and d should narrow region but appears to be relevant for a wide re- c l not be confused with d and d defined in Sec. II.B. d gion of unusual metallic behavior observed in transition- ml il c and d are defined as the critical dimensions of the MIT metal compounds. l and not of the magnetic transition. Hyperscaling asserts In 2D systems, numerical analyses support the as- the homogeneity in the singular part of the free-energy sumption of hyperscaling and the mean-field description density f in terms of an arbitrary length-scale transfor- breaks down. This observation may be related to the s mation parameter b: strong and singular wave-number dependence of the Ϫ͑dϩz͒ 1/␯ ␯͑dϩz͒ renormalized charge excitations near the Mott insulator. fs͑⌬͒ϳb fs͑b ⌬͒ϳ⌬ . (2.300)

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1106 Imada, Fujimori, and Tokura: Metal-insulator transitions

This scaling is derived, in the absence of anomalous di- Since the order parameter exponent is shown to be non- mensions, from dimensional analysis of the free-energy critical (McKane and Stone, 1981) for a transition con- density defined by trolled by disorder, Eq. (2.305) leads to the relation yh 1 ϭd. The order-parameter susceptibility also has the fϭϪ lim lim ln Z, (2.301) scaling form ␤ ϱ N ϱ ␤N 2fץ 1 → → which is proportional to [frequency]ϫ[length]Ϫd. The ␹͑⌬͒ϭ ϭb2yhϪd␹͑b1/␯⌬,byhh͒. (2.306) ͯ h2ץ T 1/␯ hϭ0 scale transformation of the argument ϰb ⌬ in fs is made dimensionless using the single characteristic From the fluctuation dissipation theorem (Kubo, 1957), length scale (2.298). When the inverse temperature ␤ we see that Eq. (2.306) is equal to the density-density and the linear dimension of the system size L are both correlation function large but finite, a finite-size scaling function is ex- ␮ץ/nץpected to hold as F i C͑k,␻͒ϭ (2.307) Ϫi␻ϩD k2 f ͑⌬͒ϳ⌬␯͑dϩz͒ ͑␰/L,␰z/␤͒ (2.302) c s F in the combination of dimensionless arguments. at ␻ϭ0. Because of the noncriticality of the compress- Here, we confirm that the hyperscaling scenario is sat- ibility at a transition driven by disorder, we have isfied for the transition between a metal and a band in- Ϫ1 Ϫ2 C͑k,␻ϭ0͒ϰDc k . (2.308) sulator in any number of dimensions 1рdрϱ. In a non- interacting fermion system, the control parameter ⌬ is Therefore C, which follows the same scaling form as Eq. the chemical potential that controls the filling, and the (2.306), ground-state energy has the form C͑k,␻ϭ0͒ϭb2yhϪdC͑bk,␻ϭ0͒, (2.309) k 2 dϩ2 F k kF leads to scaling of the diffusion coefficient D as E ϳ dk kdϪ1 ϳ c g ͵ 2m 2͑dϩ2͒m 0 2ϩdϪ2yh Dc͑k,␻ϭ0͒ϭb Dc͑bk,␻ϭ0͒. (2.310) ͑2m͒d/2 ϭ ⌬͑dϩ2͒/2, (2.303) Combined with dϭyh , this is indeed consistent with the dϩ2 original scaling function of the conductivity ␴ϰDc intro- where ⌬ is the chemical potential measured from the duced by Wegner (1976) in the form bottom of the band edge with a quadratic dispersion, ␴͑k,␻͒ϭb2Ϫd␴͑bk,bz␻͒. (2.311) ␧(k )ϭk2 /2m. Comparison of Eqs. (2.303) and (2.300) F F This demonstrates that dϭ2 is marginal. This scaling clearly shows that the hyperscaling scenario is satisfied form has inspired subsequent studies by perturbation ex- with ␯ϭ1/2 and zϭ2. The characteristic length ␰ and the pansion in terms of ␧ /␶ in the weakly localized regime. characteristic energy ⍀ are nothing but the Fermi wave- F The scaling form (2.310) is supported by numerical length kϪ1 and the Fermi energy E ϭ␮ ϭ⌬, respec- F F F analysis (MacKinnon and Kramer, 1993). tively. The reason why hyperscaling holds in any dimen- In the case of the Anderson transition, the compress- sion in this case is rather trivial, namely, the ibility ␬, ␥, and the density of states are all nonsingular Hamiltonian is quadratic and it is always at the Gaussian at the critical point, in contrast with the Mott transition. fixed point. This is because a finite density of states exists on both sides of the mobility edge. As we shall see later, ␥ and ␬ ␯(dϪz) 2. Metal-insulator transition of a noninteracting system are scaled by ⌬ , leading to zϭyhϭd. by disorder So far we have discussed the noninteracting case. To The idea of scaling for the MIT in the form first intro- treat the Anderson transition with an interaction term, duced (Wegner, 1976, 1979; Abrahams et al., 1979) for we can write the original action in the path-integral for- the Anderson localization problem is certainly equiva- malism. After making the Stratonovich-Hubbard trans- lent to the assumption of hyperscaling (Abrahams and formation for the interaction part with the replica trick Lee, 1986). We introduce here the scaling function for for the random potential, we can reduce the effective an external field h relevant to the order parameter action to the nonlinear sigma model,

Ϫ͑dϩz͒ 1/␯ yh 1 fs͑⌬,T,h͒ϭb sh͑b ⌬,b h͒ (2.304) S͓Q͔ϭϪ drtr ជ Q͑r͒ 2ϩ2H dr tr ⍀Q͑r͒ F 2G ͵ „“ … ͵ „ … with the exponent yh . Since the order parameter for a disordered transition is the single-particle density of ϩSint͑Q͒, (2.312) states ␳ at the Fermi level, we obtain the scaling of the where Q is an infinite matrix with matrix element Q␣␤ density of states as nm given by spin quarternions represented by 4ϫ4 complex f matrices (Finkelstein, 1983, 1984). The indices n,m areץ 1 ␳͑⌬͒ϭ ϭbϪ͑dϪyh͒␳͑b1/␯⌬,byhh͒ h ͯ for the Matsubara frequencies, while ␣,␤ denote replicaץ T hϭ0 indices. The c-number variables Q are obtained as the ϰ⌬␯͑dϪyh͒. (2.305) Stratonovich variables from the interaction and the ran-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1107

dom potential combined, and Sint(Q) represents the in- clearly distinguishes Mott insulators from Anderson- teraction part. The coupling constants Gϭ4/␲NFDc and localized insulators, whose compressibility is generally Hϭ␲NF/4 are given from the density of states NF and nonzero, because incoherent states localized around im- the diffusion constant Dc . The critical exponent ␯ for purities contribute to the compressibility. The Drude the transition by disorder is known to follow ␯у2/d weight and the charge compressibility are both zero in (Chayes et al., 1986). Experimentally and theoretically, the Mott insulating phase, whereas they are both non- the critical exponents are examined for various cases zero in metallic phases. Therefore these two quantities with and without magnetic field, magnetic impurities, must both be nonanalytic at the Mott transition point. and spin-orbit scattering. For example, under the spin- The existence of nonanalyticities in these two quantities flip mechanism, as in magnetic fields, ␯ϭ1 has been sug- at the transition point is also common in the transition gested experimentally [for example Dai, Zhang, and Sa- to a band insulator. However, as we shall see below, the rachik (1992) and Bishop, Spencer, and Dynes (1985)]. way of reaching the transition point, in other words, the When antiferromagnetic symmetry breaking exists on critical exponents and hence the universality class, can both sides of the MIT, it has been argued (Kirkpatrick be different for a Mott transition and for the usual band- and Belitz, 1995) that the effect of disorder can be mapped onto the random-field problem where the hy- insulator/metal transition. perscaling description is known to be modified even be- To understand the fundamental importance of the low the upper critical dimension. Drude weight and the charge compressibility, it is help- So far the effect of interaction has been considered in ful to represent them as stiffness constants to the twist in this nonlinear sigma model only by the saddle-point ap- boundary conditions in the path-integral formalism. The proximation and hence at the Hartree-Fock level. The stiffness constant to the twist is in general useful in iden- critical exponents have been calculated by perturbative tifying phase transitions. One simple example is found in renormalization-group analysis in the ␧ expansion. In the ferromagnetic transition of the Ising model at the the latter part of Sec. II.F, we shall see that the Mott critical temperature. Suppose one imposes two different transition must be treated by taking the interaction ef- types of fixed boundary conditions at the boundaries in fect beyond the Hartree-Fock level because the critical one spatial direction, x, in a system with a hypercubic behavior of the MIT itself is due to the critical fluctua- structure. In one of the boundary conditions, one fixes tion of interaction effects. Therefore qualitatively differ- all the up spins at the boundary rxϭ0 and rxϭL, where ent aspects of scaling behavior at the Mott transition rx denotes the x coordinate of the spatial point r.Inthe point appear, as we discuss later. other boundary condition, one fixes all the up spins at For more detailed discussions of the MIT caused by rxϭ0 and all the down spins at rxϭL. Let the averaged disorder in the case of minor contribution of electron energy under the former boundary condition be Eu and under the latter, E . Then ␦Eϭlim E ϪE must be correlation effects at the Hartree-Fock level, readers are t L ϱ t u → referred to the extensive review articles of Lee and Ra- zero in the paramagnetic phase because the spin corre- makrishnan (1985) and Belitz and Kirkpatrick (1994). lation length is finite, whereas ␦E remains nonzero as An important open question remains concerning the na- L ϱ in the ferromagnetically long-range-ordered ture of the MIT when both correlation effects and dis- phase→ because a domain wall in the plane perpendicular order are crucially important near the Mott insulator. A to the x axis has to be introduced only for the latter dynamic mean-field theory (infinite-dimensional ap- boundary condition due to the twist. The latter bound- proach) was recently formulated to discuss the MIT ary condition has higher energy coming from the forma- when both electron correlation and disorder are strong tion energy of a domain wall. (Dobrosavljevic´ and Kotliar, 1997). It shows the appear- When one wishes to establish similar criteria for the ance of a metallic non-Fermi-liquid phase near the MIT MIT, one has to work with the path-integral formalism and strong effects of correlation even away from half because this transition is intrinsically quantum mechani- filling. A recent experiment by Kravchenko et al. (1995) cal. Because of this quantum-mechanical nature, one has appears to show the existence of a true MIT in 2D in- two ways of imposing twists in the (dϩ1)-dimensional teracting systems, in contrast with the conventional ar- space of the path integral. One way is to impose twist in gument for noninteracting systems in which 2D systems the spatial direction while the other is to impose it in the become insulating even with weak disorder. The scaling temporal direction. Since the wave function in a metal is properties of this transition have been discussed theo- characterized by phase coherence, while that in an insu- retically, for example, by Dobrosavljevic´ et al. (1997). lator is characterized by incoherence of the phase, these two types of wave functions respond differently to a twist of the phase in boundary conditions in terms of 3. Drude weight and charge compressibility rigidity of the phase. In particular, in the Mott insulator, The Drude weight defined in Eq. (2.273) and the space-time localized wave functions result in the absence charge compressibility in Eq. (2.286) are the two impor- of phase rigidity in both the space and the time direc- tant and relevant quantities needed to describe the Mott tions. Below it is shown that spatial and temporal transition. A Mott insulator has two basic properties, rigidities are associated with nonzero Drude weight and one its insulating property and the other its incompress- nonzero charge compressibility, respectively (Continen- ibility. In particular, incompressibility is a property that tino, 1992 Imada, 1994c, 1995a, 1995b).

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1108 Imada, Fujimori, and Tokura: Metal-insulator transitions

˜ A An infinitesimally small twist of the phase, ␾L , im- Under the frequency-dependent vector potential x ϭA0ei␻t x posed between two boundaries rxϭ0 and rxϭL may al- x , the electric field of the component is given ternatively be expressed by a transformation of the by Grassmann variable introduced in Sec. II.C in the form 0 ExϭAxi␻/c. (2.321) When one uses the linear-response theory (Kubo, 1957), (␺␴͑r,␶͒ ␺␴͑r,␶͒exp͓irxٌ␾͔, (2.313 → the current response to this electric field yields the con- ˜ where ٌ␾ϭ␾L /L and rx denotes the x coordinate of r. ductivity as It has been shown by Kohn (1964), Thouless (1974), and 2e2 Shastry and Sutherland (1990) that this type of phase Im ␴͑␻͒ϭ D˜ . (2.322) twist measures the stiffness, which is proportional to the ␻ Drude weight. From the Kramers-Kronig relation In a tight-binding Hamiltonian, the transformation (2.313) may equivalently be replaced by transforming ϱ Im ␴͑␻͒ϭ d␻Ј P Re ␴͑␻Ј͒d␻Ј, the off-diagonal transfer term t(r,rЈ)¯␺(r)␺(rЈ) in the ͵Ϫϱ ␻Ϫ␻Ј Hamiltonian as the change in the energy density ␦fϭ␦E/Ld is given by using the Drude weight defined in Eq. (2.273) as ˜ t͑r,rЈ͒ t͑r,rЈ͒exp͓i͑rxϪrЈ͒␾L /L͔. (2.314) x 4 → 1 ␾˜ 2 ␾˜ This is nothing but the Peierls factor. In other words, L L ␦fϭ 2 2 DϩO . (2.323) Eq. (2.314) is the same as applying a vector potential 2␲e L ͩͩ L ͪ ͪ Equation (2.323) shows that the stiffness constant for ˜ Axϭc␾L /Le (2.315) the spatial phase twist is proportional to the Drude in the x direction. Because a small uniform phase twist weight. An important point to keep in mind is that the ˜ and hence a uniform vector potential is imposed, it Drude weight is obtained by taking the limit ␾L 0 first and L ϱ afterwards. If the opposite→ limit should be associated with the electromagnetic response → at kϭ␻ϭ0 in the second-order perturbation. This is in- lim␾ 0limL ϱ is taken, one has to take account of pos- L→ → ˜ deed shown as follows: Under the vector potential sible level crossings at finite ␾L in the limit L ϱ, which (2.315), the change in the Hamiltonian up to the second yields in principle a different quantity, namely,→ the su- ˜ order in ␾L reads perfluid density (Byers and Yang, 1961; Scalapino, White, and Zhang, 1993). jx Kx An infinitesimally small twist of the phase ␾˜ imposed ␦ ϭϪ ␾˜ Ϫ ␾˜ 2 ϩO͑␾˜ 3 ͒, (2.316) ␤ H L L 2L2 L L between two boundaries ␶ϭ0 and ␶ϭ␤ is equivalent to transforming the Grassmann variable as where jx and Kx are the current and the kinetic energy ˙ operator in the x direction, respectively. In the case of ␺␴͑r,␶͒ ␺␴͑r,␶͒exp͓i␶␾͔ (2.324) the nearest-neighbor Hubbard model (1.1a), these op- → ˙ ˜ erators in the coherent-state representation have the with ␾ϭ␾␤ /␤, where ␴ in general denotes quantum form numbers other than the spatial coordinate. In case of the Hubbard model, ␴ is the spin. The transformation (2.324) may be replaced with the transformation of the ¯ jxϭ2t sin kx␺k␴␺k␴ , (2.317) Hamiltonian ͚k,␴ ˙ ¯ ϩi␾͚ ␺␴͑r͒␺␴͑r͒. (2.325) ¯ H→H r,␴ KxϭϪ2t͚ cos kx␺k␴␺k␴ . (2.318) k,␴ This equivalence is proven as follows: To calculate the ˜ 2 partition function, one calculates the matrix element in a The change in the ground-state energy up to ␾L is ob- tained from the second-order perturbation as Trotter slice of the path integral: Ϫ⌬␶ [c†,c] ͑␶,␺, ͒ϭ͗␺͑␶ϩ⌬␶͉͒e •H ͉␺͑␶͒͘ ␾˜ 2 L H L 4 ␦EϭD˜ ϩO ␾˜ , (2.319) † 2Ϫd ͑ L͒ ¯ Ϫ⌬␶ [c ,c] L ϭ͗0͉ „1ϩ␺␣͑␶ϩ⌬␶͒c␣…e H ͟␣ 2 Ϫ1 1 ͉͗0͉jx͉␯͉͘ D˜ ϭ K ϩ , (2.320) † d ͗ ͘ ͚ ϫ 1Ϫ␺␣͑␶͒c␣ ͉0͘ L ͩ 2d ␯Þ0 E␯ϪE0 ͪ ͟ „ … ␥ for a d-dimensional hypercubic lattice. In Eq. (2.320), ͉␯͘ Ϫ⌬␶ [¯␺͑␶ϩ⌬␶͒,␺͑␶͒] ϭe H . (2.326) denotes an eigenstate with the energy E␯ and ␯ϭ0 de- notes the ground state. The total average kinetic energy If one takes the transformation (2.325), this matrix ele- is ͗K͘. ment is transformed to

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1109

i␶␾ ˙ ¯ through the parameter space of ⌬, temperature, and ͑␶,␺e , ͒ϭ ͑␶,␺, ͒Ϫi͑⌬␶͒␾␺␣͑␶ϩ⌬␶͒␺␣͑␶͒ L H L H staggered magnetic fields. On the metallic side, the scal- 2 is given from 2␤ץ/(ln Z)2ץϩO„͑⌬␶͒ … ing of the specific heat Cϭ␤2 Ϫ͑⌬␶͕͒ [c†,c]ϩi␾˙ c†c͖ Eq. (2.302): ϭ͗␺͑␶ϩ⌬␶͉͒e H ͉␺͑␶͒͘ 2ץ ϩO ͑⌬␶͒2 . (2.327) Cϭ⌬␯͑dϩz͒T ͑␰/Lϭ0,␰zT͒. (2.334) T2 Fץ … „ This proves equivalence to the transformation (2.325) in the functional integral limit ⌬␶ 0. At the critical point, the specific heat is given from the From Eq. (2.325) it turns out→ that Ϫi␾ is the same as ⌬-independent term as an additional change in the chemical potential ␮ ␮ CϰTd/z. (2.335) → Ϫi␾˙ . Therefore the free energy as a function of the im- If Cϭ␥T is satisfied at low temperatures in the metallic posed small twist should have the form phase, the coefficient ␥ follows: ˜ ˜ 2 i␾␤ ␾␤ ␯͑dϪz͒ f͑␾˙ ͒ϭf͑␾˙ ϭ0͒ϩ ␳ϩ ␬ϩO͑␾˜ 3 ͒, (2.328) ␥ϰ⌬ . (2.336) ␤ 2␤2 ␤ In general, if C is not linear in T but proportional to Tq ␮ and the charge as Cϰϳ¯␥Tq, ¯␥ is scaled asץ/fץwhere the electron density ␳ϭϪ -␮ are obtained as expansion coץ/␳ץcompressibility ␬ϭ efficients. Equation (2.328) shows that ␬ indeed mea- ¯␥ϰ⌬␯͑dϪqz͒. (2.337) sures the stiffness constant of the phase in the temporal direction. Another interesting quantity is the coherence tempera- ture TF below which the electron motion becomes quan- tum mechanical and degenerate. It is clear that TF has to 4. Scaling of physical quantities approach zero as ⌬ 0 in a continuous transition. In → Combining Eqs. (2.323) and (2.328) for the Drude case of the Fermi liquid, TF is nothing but the Fermi weight and the compressibility, respectively, with the temperature. The existence of a single characteristic en- finite-size scaling form derived from hyperscaling, one ergy scale with singularity at ⌬ϭ0 leads to the scaling of obtains useful scaling forms for physical quantities TF in the form ␯z (Imada, 1994c, 1995a, 1995b; Continentino, 1994). This TFϰ⌬ . (2.338) is because the finite-size scaling form (2.302) is also valid for ␦f, namely, the difference of the free energy be- 5. Filling-control transition tween the cases in the presence and the absence of the twists. From a comparison of Eqs. (2.323) and (2.302) in When the control parameter ⌬ is the chemical poten- Ϫ the order of L 2, the scaling of D is obtained: tial ␮ measured from the critical point ␮c , it represents the filling-control MIT (FC-MIT) and, in the terminol- ␨ Dϰ⌬ (2.329) ogy of critical phenomena, it is called a generic transi- with ␨ϭ␯(dϩzϪ2). Comparison of Eqs. (2.302) and tion. In this case, it is possible to derive a useful scaling (2.328) yields, in the order of ␤Ϫ1 and ␤Ϫ2, relation. Because the doping concentration is ␦ϭ it scales as ,⌬ץ/ fץ␮ϭϪץ/ fץϪ ␦ϰ⌬Ϫ␣ϩ1 (2.330) s s ␦ϳ⌬␯͑dϩz͒Ϫ1. (2.339) and Ϫ␣ From the comparison of Eqs. (2.339) and (2.330), we ␬ϰ⌬ (2.331) derive with ␣ϭ␯(zϪd), where ␦ denotes the doping concen- ␯zϭ1. (2.340) tration, that is, the density measured from the Mott in- sulator. The characteristic length scale is then When hyperscaling holds, scaling of other physical ␰ϳ␦Ϫ1/d, (2.341) quantities can also be derived from Eqs. (2.300) and (2.302). In the insulating phase, the charge excitation which is the length scale of the ‘‘mean hole distance.’’ This is a natural consequence because the mean hole gap Eg and the localization length ␰l defined in Eq. (2.296) should be determined from the single character- distance is indeed a characteristic length scale which di- istic energy and length scale ⍀ and ␰, respectively. Then, verges at ⌬ϭ0. From the scaling relation (2.340) the z␯ number of independent critical exponents is reduced to Egϰ͉⌬͉ (2.332) one in the above physical quantities. In the case of the and FC-MIT, because the doping concentration ␦ is easier to ␯ control than the chemical potential, various exponents ␰lϰ͉⌬͉ (2.333) in terms of ␦ are summarized for convenience as follows: immediately follows, where we have used the fact that ␯ ␰ϳ␦Ϫ1/d, (2.342a) and z are identical on both sides of the transition point. This is shown by connecting the two sides via a route ⌬ϳ␦z/d, (2.342b)

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1110 Imada, Fujimori, and Tokura: Metal-insulator transitions

␬ϳ␦1Ϫz/d, (2.342c) citations in the metallic phase. In the metallic phase, the low-energy spin and charge excitations can be character- 1ϩ͑zϪ2͒/d Dϳ␦ , (2.342d) ized either by free fermions or by some type of Tomonaga-Luttinger liquid. However, the critical expo- ␥ϳ␦1Ϫz/d, (2.342e) nents of the MIT do not reflect this difference. In the z/d TFϳ␦ . (2.342f) metallic phase, the characteristic length and energy scales for the MIT are finite and always determine a 6. Critical exponents of the filling-control metal-insulator crossover scale, while the character of a liquid is associ- transition in one dimension ated with behavior below this energy scale. The present scaling approach treats equally the zero-temperature In several 1D systems, we can derive rigorous esti- transition for different types of metals, while the univer- mates of various critical exponents for the cases of inte- sality class of the transition does not necessarily specify grable models. After comparison of various cases, it the nature of coherent metals. turns out that the hyperscaling scenario holds for the MIT in 1D with ␯ϭ1/2 and zϭ2 for all cases (Imada, 7. Critical exponents of the filling-control metal-insulator 1995b; see also Stafford and Millis, 1993). This means transition in two and three dimensions that 1D is between the upper and lower critical dimen- In two or three dimensions, we have no exact solution sions of MIT. We start with the case of the Hubbard for this problem. Therefore we have to rely on com- model. From the Bethe Ansatz solution, it is known that bined analyses of several numerical results. The most the compressibility (Usuki, Kawakami, and Okiji, 1989), important question to be answered is whether 2D and the Drude weight (Shastry and Sutherland, 1990), and 3D are lower than the upper critical dimension dc or the specific-heat coefficient (Takahashi, 1972; Usuki, not. If dcϽ2, some type of mean-field theory would be Kawakami, and Okiji, 1989) scale as justified even in 2D, whereas we have to go beyond the mean-field level if d Ͼ2. We first review the predictions ␹ ϰ␦Ϫ1, (2.343a) c c of various mean-field theories. In the Hartree-Fock ap- Dϰ␦, (2.343b) proximation, the 2D and 3D Hubbard model in general has a phase diagram with an extended region of antifer- Ϫ ␥ϰ␦ 1, (2.343c) romagnetic order in the metallic region around the Mott for the 1D Hubbard model. For the supersymmetric t-J insulator, as in Figs. 10 and 11. Therefore the MIT takes model, with tϭJ in 1D, the Bethe Ansatz solution shows place as a transition between insulating and metallic states, both with antiferromagnetic order. The solution the same scaling as in Eqs. (2.343a)–(2.343c) of the unrestricted Hartree-Fock approximation shows (Kawakami and Yang, 1991). This is actually the same that incommensurate periodicity at the wave-number k scaling form as we saw in Eq. (2.303) for noninteracting Þ(␲,␲) appears at finite doping. Irrespective of com- fermions near a band insulator. To understand the hy- mensurate or incommensurate cases, the metal-insulator perscaling in 1D more intuitively, it is useful to consider transition is in general achieved by the shrinking of hole the bosonized Hamiltonian derived after the asymptotic pockets. This is the case for the transition ͓I-1 M-1͔ spin-charge separation. As will be derived in Sec. II.G.1, discussed in Sec. II.B. It appears to be realized in↔ some the Hamiltonian for the charge part has the form 3D compounds with orbital degeneracy, such as NiS Se , as will be discussed in Sec. IV.A.2. ␲v␳K␳ v␳ 2Ϫx x ,␾ ͒2 , (2.344) Various experimental and numerical results, however ץ͑ H ϭ dx ⌸2ϩ ␳ ͵ 2 ␳ 2␲K x ␳ ͩ ␳ ͪ do not support this class in 2D. This is suggested by the where the charge-density phase field ␾␳ and the conju- absence of an antiferromagnetic metal in 2D. It means gate momentum ⌸␳ satisfy the Bose commutation rela- dmlϾ2 at least unless the orbital degeneracy is too large. tion and constitute a noninteracting Hamiltonian with In this case possible types of transition in 2D would be linear dispersion, ␻␳(k)ϭv␳͉k͉ (Emery, 1979; Haldane, ͓I-1 M-2͔ and ͓I-1 M-3͔. Quantum Monte Carlo cal- 1980; Schulz, 1990b). This free-boson system indeed de- culations↔ first performed↔ on the metallic side have Ϫ1 Ϫ1 scribes the case zϭ1/␯ϭ2 when v␳ϰ␰ . It should also shown ␹cϰ␦ and an antiferromagnetic correlation Ϫ1/2 be noted that, when the Hubbard model is mapped to length ␰mϰ␦ , which suggests a new universality class the massive Thirring model, the transition to a Mott in- (Furukawa and Imada, 1992, 1993). This implication on sulator with a Mott gap is mapped to the transition to a the metallic side has been examined in terms of the scal- band insulator with a hybridization gap (Hida, Imada, ing theory, leading to the conclusion that there does ex- and Ishikawa, 1983; Mori, Fukuyama, and Imada, 1994). ist a new universality class zϭ1/␯ϭ4 in this case (Imada, This also shows that in 1D, the universality class should 1994c, 1995b). Moreover, a localization length ␰l scaled Ϫ␯ be the same between the metal Mott-insulator and metal by ␰lϰ͉␮Ϫ␮c͉ is predicted on the insulating side. This band-insulator transitions. Even in bosonic systems, the prediction for the insulating side has been critically Mott transition in 1D is characterized by the same uni- tested by a quantum Monte Carlo calculation. The result Ϫ␯ versality class, zϭ1/␯ϭ2 (Imada, 1995b). is ␰lϰ͉␮Ϫ␮c͉ , ␯ϭ0.26Ϯ0.05. Therefore all the data As for the MIT in 1D, the universality class is always are consistently explained by the scaling theory with a the same irrespective of the character of low-energy ex- universality class of zϭ4 and ␯ϭ1/4 for the case

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1111

͓I-1 M-3͔. These two results can be explained neither turn is associated with strong incoherent scattering of by the↔ Hartree-Fock approximation nor by the dϭϱ re- charge by a large degeneracy of component excitations. sults. Indeed, neither the Hartree Fock approximation This degeneracy, coming from softened excitations of nor the dϭϱ results satisfies the assumption of hyper- component degrees of freedom, is due to critical fluctua- scaling [Eq. (2.300)]. For example, the Hartree-Fock ap- tion of component correlations. proximation predicts ␯ϭ1/2. Furthermore, scaling of the 9. Suppression of coherence localization length ␰l excludes the possibility that the MITs in 1D and 2D belong to the same universality In the new universality class discussed in the previous subsection (Sec. II.F.7), physical quantities show many class, because ␰l scales following ␯ϭ1/2 in 1D. Therefore numerical results on the 2D Hubbard model support the anomalous features in terms of the standard MIT, the following four points for single-band systems in 2D on a most important of which is the suppression of coher- square lattice: ence. In this subsection we discuss several unusual as- (1) The upper critical dimension d of the MIT satis- pects of this suppression for the new universality class. c Scaling theory predicts that the Drude weight is D fies d Ͼ2 so that none of the mean-field approaches, c ϰ␦1ϩ(zϪ2)/d. As is well known, Dϰ␦ is satisfied in the including the Hartree-Fock approximation and the usual MIT, which is consistent with scaling theory when infinite-dimensional approach, is justified. zϭ2. When the universality class is characterized by z (2) Hyperscaling is satisfied in 2D with a new univer- ϭ1/␯Ͼ2, D is suppressed more strongly than ␦-linear sality class of ␯ϭ1/4 and zϭ4. dependence at small ␦. For example, when zϭ4 is satis- (3) dmlϾ2Ͼdil , which means there are no antiferro- fied in 2D, as suggested in numerical studies, Dϰ␦2 is magnetic metals close to the Mott insulator in 2D. derived. This is one of the indications for the unusual (4) The universality class is different for 1D and 2D suppression of coherence in this universality class and because zϭ1/␯ϭ2 in 1D. indeed supported by recent results for the t-J model In 3D systems, the universality class of MIT for the case (Tsunetsugu and Imada, 1998). Relative suppression of of ͓I-1 M-3͔ is not well known. Experimental indica- D as compared to the value D0 for the metal band- ↔ (zϪ2)/d tions in (La,Sr)TiO3 appear to support the mass- insulator transition is given by D/D0 , ϰ␦ . diverging-type MIT, will be as discussed in Sec. IV.B.1, From the sum rule, the ␻-integrated conductivity and hence suggest that zϾ2 is also satisfied. gives the averaged kinetic energy ϱ Ϫ͗K͘ϭ ␴͑␻͒d␻. (2.345) 8. Two universality classes of the filling-control ͵0 metal-insulator transition In the strong-coupling limit, as in the t-J model, the † Two different types of filling-control Mott transition total kinetic energy ͗K͘ϭϪ(t/N)͚͗ij͗͘ci␴cj␴ϩH.c.͘ is have been shown to exist in the previous two subsec- expected to be proportional to ␦, that is, Ϫ͗K͘ϰ␦. tions. One is characterized by zϭ1/␯ϭ2, which is the Therefore, from the above sum rule, the Drude part D same as the band-insulator/metal transition. This is the ϰ␦1ϩ(zϪ2)/d becomes negligibly small at small ␦ in the case of all 1D systems as well as of several transitions at total weight Ϫ͗K͘ϰ␦ if zϾ2. This means most of the higher dimensions, where zϭ1/␯ϭ2 is expected for total weight Ϫ͗K͘ is exhausted in the incoherent part ͓I-1 M-1͔, ͓I-1 M-2͔, ͓I-2 M-2͔, and ͓I-2 M-4͔ in ␴ (␻) in Eq. (2.273). In scaling theory, the form of ↔ ↔ ↔ ↔ reg the classification scheme in Sec. II.B. In general, this ␴reg(␻) is not specified, and indeed it may depend on class of transition appears when the component degrees details of systems including the interband transition. of freedom do not contribute to low-energy excitations However, for the intraband contribution, ␴(␻)isex- except in the Goldstone mode near the critical point. pected to follow ␴(␻)ϳ(1ϪeϪ␤␻)/␻ at very high tem- One way such contributions are realized is when both peratures tϽT because the current correlation function sides of the critical point are in broken-symmetry phases C(␻) defined in Eq. (2.275) becomes temperature and such as the magnetic ordered and superconducting or- frequency insensitive for ␻Ͻt. In fact, ␴(␻)ϳ1/T for dered phases. The component degrees of freedom are small ␻ and ␴(␻)ϳ1/␻ for larger ␻ is a general feature also inert when an excitation gap such as a spin gap is at high temperatures (Ohata and Kubo, 1970; Rice and formed. To understand the case ͓I-2 M-2͔, it is helpful Zhang, 1989; Metzner et al., 1992; Kato and Imada, to consider the strong-coupling limit,↔ where the problem 1997). When the temperature is lowered, this incoherent is essentially reduced to single-component bosons be- part ϳ(1ϪeϪ␤␻)/␻ is in general transferred to the cause of the bound-state formation. Then it is equivalent Drude part ϳ1/(Ϫi␻ϩ␥) below the coherence tem- to the problem of a superfluid-insulator transition, to be perature. In the usual band-insulator metal transition, discussed in Sec. II.F.12. this transfer can take place for the dominant part of the The other case is characterized by values of z larger weight because Ϫ͗K͘ and D both may be proportional than two. This is seen in the case of ͓I-1 M-3͔. Two- to ␦. Therefore the incoherent part is totally recon- dimensional systems with a single orbital, as↔ in the Hub- structed to the Drude part with decreasing temperature. bard model on a square lattice, are shown in numerical However, for zϾ2, this transfer takes place only in a calculations to follow zϭ1/␯ϭ4. Such a large dynamic tiny part of the total weight because the relative weight exponent is associated with suppression of coherence, as of the Drude part to the whole weight is ␦(zϪ2)/d, which we discuss in the next subsection, and this suppression in vanishes as ␦ 0 for zϾ2. Even at Tϭ0, a large fraction →

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1112 Imada, Fujimori, and Tokura: Metal-insulator transitions of the conductivity weight must be exhausted in the in- tion length ␰ϰ1/␦1/d. This additional entropy ϰ␦ is re- coherent part. Since scaling theory gives us no reason to leased below the coherence temperature Tcoh . expect a dramatic change of the form for ␴reg(␻) at any Therefore we are led to a natural consequence that if temperature at the critical point, the optical conductivity T-linear specific heat characterizes the degenerate (co- more or less follows the form herent) temperature region, the coefficient ␥ should be given by ␥ϳ␦/T ϳ␦1Ϫz/d. This is indeed the scaling 1ϪeϪ␤␻ coh law we obtained in Eq. (2.336). From this heuristic ar- ␴reg͑␻͒ϳC (2.346) ␻ gument, it turns out that the metallic phase near the even for tϾ␻ and tϾT, where we have neglected Mott insulator is characterized by large residual entropy model-dependent features such as the interband transi- at low temperatures, which is also related to the sup- tion. The intraband part of the incoherent weight C may pression of the antiferromagnetic correlation. This large follow Cϰ␦. Such dominance of the incoherent part is residual entropy again has numerical support in the 2D consistent with the numerical results for 2D systems, as t-J model (Jaklicˇ and Prelovsˇek, 1996). It may also be is discussed in Sec. II.E.1. It is also consistent with the said that anomalous suppression of coherence at small ␦ is caused by short-ranged component correlations, which experimental indications of high-Tc cuprates as will be discussed in Secs. IV.C.1 and IV.C.3. The scaling of ␴(␻) scatter carriers incoherently. by Eq. (2.346) is another argument in support of zϾ2in In a Mott insulator, because the single-particle pro- 2D. We discuss in Sec. II.G.2 how the form of Eq. cess is suppressed due to the charge gap, we have to (2.346) is equivalent to the scaling of the marginal Fermi consider the two-particle process (electron-hole excita- liquid. In contrast, the numerical results in 1D by tions) expressed by the superexchange interaction. Simi- Stephan and Horsch (1990), discussed in Sec. II.E.1, sup- larly, even in the metallic region, suppression of coher- port a small incoherent part at low temperatures. This ence for the single-particle process leads us to consider may be due to zϭ2 in 1D because the majority of the the relevance of two-particle processes. In contrast with weight seems to be exhausted in the Drude weight. In the insulating phase, the two-particle processes in metals 3D perovskite systems, the dominance of the incoherent contain not only the particle-hole process (the exchange part is a common feature, as we see in systems such as term) but also particle-particle processes through an ef- fective pair hopping term. Such processes may be ge- (La,Sr)TiO3, (La,Sr)VO3, and (La,Sr)MnO3 (see Sec. IV for details). This seems to suggest zϾ2 in 3D, as nerically described by well. In the high-Tc cuprates, it appears that the coher- 2 † † ent part is almost absent, while in 3D systems, a small WϭϪtW͚ ͚ ͑ci␴ciϩ␦␴ϩciϩ␦␴ci␴͒ . (2.347) H i ␦␴ fraction of the coherent weight appears to exist which ͩ ͪ yields T2 resistivity. Coherence is not suppressed anomalously near the Mott Another aspect of the suppression of coherence is insulator in the case of a two-particle process because seen in the coherence temperature Tcoh . The scaling of the universality class of the two-particle transfer is given z/d Tcoh is given by Tcohϰ␦ . Because the standard MIT is by zϭ1/␯ϭ2, making it more relevant than single- 2/d characterized by Tcoh0ϰ␦ , the relative suppression particle transfer for small ␦. In fact, this term does not (zϪ2)/d Tcoh /Tcoh0 is again proportional to ␦ . When z suffer from the mass enhancement effect. This new term ϭ4, as suggested by numerical results in 2D, we obtain added to the Hubband model, namely the t-U-W model 2 Tcohϰ␦ in 2D in contrast with Tcoh0ϰ␦ for zϭ2. leads to the appearance of the d-wave superconducting In the Mott insulating phase, the charge degrees of phase (Assaad, Imada, and Scalapino, 1996, 1997; As- freedom are not vital and only the spin or orbital de- saad and Imada, 1997), although it is not clear whether grees of freedom (that is, the component degrees of the bare Hubband model implicitly contains sufficiently freedom) contribute to low-energy excitations. When large coupling of this form. The filling-control transition symmetry is broken by the component degrees of free- from superconductor to Mott insulator is indeed charac- dom, the entropy coming from the components is re- terized by zϭ1/␯ϭ2. This shows an instability of metals duced essentially to zero. The entropy is also reduced near the Mott insulator to superconducting pairing or, when the component excitation has a gap, as in the spin- more generally, an instability of the zϭ4 universality gap phase. Even without such a clear phase change, the phase to a broken symmetry state. growth of short-ranged component correlation progres- A related but different view was proposed for the role sively reduces the entropy with decreasing temperature. of incoherence in the mechanism of superconductivity For example, in 2D single-band systems, exponential (Wheatley, Hsu, and Anderson, 1988). According to this growth of the antiferromagnetic correlations leads to ex- proposal, lack of coherent transport in the interlayer di- ponentially small residual entropy at low temperatures, rection occurs when electrons are confined due to spin- except for the contribution from the Goldstone mode. charge separation in 2D. The coherence recovers When carriers are doped, an additional entropy due to through interlayer tunneling after pair formation, which the charge degrees of freedom is expected proportional leads to a kinetic-energy gain in the interlayer direction. to ␦. Of course, this additional entropy is assigned not In the analysis given in this section, the driving force for solely to the charge but also to the spin entropy, because recovering coherence by pair formation already exists in they are coupled, causing the destruction of antiferro- a single layer because of the novel universality class of magnetic long-range order and yielding a finite correla- the MIT. The enhancement of the pairing correlation for

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1113 systems with a flat band near the Fermi level was also coherent region of metals. This nontrivial definition of suggested in numerical studies (Yamaji and Shimoi, the correlation length ␰m is indeed known to exist in the 1994; Kuroki, Kimura, and Aoki, 1996; Noack et al., 1D Hubbard model (Imada, Furukawa, and Rice, 1992), 1997). where the antiferromagnetic correlation decays as 3/2 ϳ1/r at rϾ␰m while it decays as ϳ1/r for rϽ␰m ex- 3/2 10. Scaling of spin correlations cepting logarithmic corrections. The decay rate ϰ1/r is indeed the asymptotic correlation in the metallic region In the previous subsections, an unusually large z was for ␦ ϩ0, while the decay rate ϰ1/r is the long- obtained when the MIT was accompanied by a simulta- → distance behavior in the insulating phase, because ␰m neous component order transition, as in the 2D Hub- ϰ1/␦. In 2D, the insulating behavior at rϽ␰m is charac- bard model. In this case, a component correlation, such terized by a constant because of the presence of long- as the antiferromagnetic correlation or the orbital corre- range order in a Mott insulator. The exponent of the lation, follows a particular scaling behavior near the power-law decay beyond ␰m in 2D is not rigorously transition point. Here as an example we discuss the scal- known. In a Fermi liquid, it has to follow ͗S(0)•S(r)͘ ing of the antiferromagnetic correlation in a single-band Ϫ␥ ϰr with ␥ϭ3atrϾ␰m in 2D. model. The destruction of magnetic order in the metallic With this nontrivial definition of ␰m in mind, the scal- region is ascribed to charge dynamics. In this case, the ing form for the singular part of the free energy is gen- scaling of the magnetic correlation length ␰m follows eralized to include dependence on the staggered mag- ␰ ϳ␰, where ␰ is the length scale of the MIT itself, and m netic field hs with wave number Q as hence the mean distance of doped holes is given by Eq. ␯͑dϩz͒ ␸ (2.342a). In other words, the magnetic correlation length fs͑⌬,hs͒ϰ͉⌬͉ Y͑hs /͉⌬͉ ͒ (2.352) is controlled by ␰ because long-range order is destroyed at Tϭ0, where ␸ is the exponent for hs . From this ex- by the charge motion. We need to be careful about the pression, the staggered magnetization is derived as definition of ␰m in metals because they have gapless fsץ magnetic excitations, leading to a power-law decay at M ϭ ϭ͉⌬͉␤s. (2.353) ͯ hץ asymptotically long distances at Tϭ0. In general, at the s s h 0 critical point of the magnetic transition, the spin corre- s→ lation follows with the exponent Ϫ͑dϩzϪ2ϩ␩͒ iQ r ͗S͑0͒•S͑r͒͘ϳr e • (2.348) ␤sϭ␯͑dϩz͒Ϫ␸s . (2.354) with the ordering vector Q and the exponent ␩. Here, Therefore, from ␤sϭ0, we obtain ␸sϭ␯(dϩz). Then the magnetic correlation in the insulating phase has to the staggered susceptibility ␹(Q,␻ϭ0) is scaled as be kept unchanged up to the critical point as a function Msץ of the control parameter, the chemical potential. The ␹͑Q,␻ϭ0͒ϭ ϳ⌬Ϫ␥s (2.355) ͯ hץ reason is the following: The spectral representation of s h 0 s→ the spin correlation with ␥sϭ␯(dϩz). Re ␹(q,␻) and Im ␹(q,␻)/␻ both have 1 a peak around qϭQ at ␻ϭ0 with the same critical be- Ϫ␤͑EmϪ␮N͒ ͗S͑0͒•S͑r͒͘ϭ ͗m͉S͑0͒•S͑r͉͒m͘e Ϫ␥s Ϫz Z ͚m havior of the peak height ϰ⌬ and widths ϳ␰ in ␻ Ϫ1 (2.349) space and ϳ␰ in k space. The dynamical structure factor is given from the fluctuation-dissipation theorem does not contain an explicit dependence on the chemical Ϫ␤␻ potential ␮ at Tϭ0 because the number of particles is Im ␹͑q,␻͒ϭ͑1Ϫe ͒S͑q,␻͒. (2.356) ␤␮N kept constant in the Mott insulating phase and e in At Tϭ0, S(q,␻) also has a peak height ϰ⌬Ϫ␥s with the the numerator is canceled by the same factor in the par- widths ␰Ϫz and ␰Ϫ1 in ␻ and q spaces, respectively. The tition function Z. This means that the staggered magne- consistency of this scaling with the spin correlation tization defined by ͗S(r)S(0)͘ discussed above is confirmed from the rela- 1 tion z iQ•ri Msϭ Si e (2.350) N ͚i ϱ S͑Q,tϭ0͒ϭ S͑Q,␻͒d␻ is a constant in the insulating phase. Therefore the mag- ͵Ϫϱ netization exponent ␤ , defined by Ϫ␥ Ϫz 1Ϫ␥ d Ϫ1 s ϰ⌬ s␰ ϳ⌬ sϳ␰ ϳ␦ . (2.357) ␤s Msϰ͉⌬s͉ , (2.351) This is indeed the structure factor has to satisfy ␤sϭ0 and ␩ϭ2ϪdϪz. Note that this is a iQ r continuous transition because ␰m diverges continuously, S͑Q,tϭ0͒ϭ ͵ ͗S͑r͒S͑0͒͘e • dr (2.358) although Ms jumps at ⌬ϭ0. In the metallic region, this constant staggered magnetization is cut off at ␰m and the expected for spins with correlation length ␰mϭ␰, assum- antiferromagnetic correlation has to decay as a power ing the correlation is cut off at ␰m . law beyond ␰m . The exponent for the power law at r However, from the nontrivial definition of ␰m above, Ͼ␰m is determined from the Fermi-surface effect in the we should keep in mind that the power-law correlation

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1114 Imada, Fujimori, and Tokura: Metal-insulator transitions

beyond ␰m is neglected in this argument. When the spin correlation follows the exponential form eϪr/␰m, S(q,␻) satisfies simple Lorentzian form, indeed with the width discussed above. However, if the correlation is not ex- ponentially cut off but obeys a power-law decay beyond ␰m , this contribution of the long-ranged part has to be superimposed around qϭQ and ␻ϭ0inS(q,␻)asa non-Lorentzian contribution. We call this contribution the coherent part of the spin correlation, as discussed in Sec. II.E.4. Because of the presence of the coherent con- tribution, the width of the peak in q or ␻ space becomes FIG. 34. Scaling of Tscr and Tcoh in the plane of temperature ill defined, and the peak structure in general may have a and doping concentration in a typical situation of 2D systems. singular structure such as a cusp at incommensurate po- sitions, as schematically illustrated in Fig. 31. Such non- (2␲ͱ␦)zJ in ␻. Therefore the dominant ␻ dependence Lorentzian structure is due to the coherent contribution at small ␻ in Im ␹ is given by the factor 1ϪeϪ␤␻. Due to and grows below Tcoh . We discuss temperature depen- this simplification, the neutron scattering weight dence later in greater detail. Although the width is not ͐dq Im ␹(q,␻) follows well defined, the total weight in q or ␻ space satisfies the above scaling because the weight in the coherent part dq Im ␹͑q,␻͒ϳ1ϪeϪ␤␻ (2.362) does not exceed the incoherent contribution when the ͵ exponent of the power-law decay ␥ satisfies ␥Ͼd,asin and the NMR relaxation rate T1 is given as the Fermi-liquid case (in a Fermi liquid, ␥ϭdϩ1). Then 1 Im ␹͑q,␻͒ 1 the q-integrated and ␻-integrated intensities at Tϭ0 fol- ϭ lim ͵ f͑q͒dqϳ (2.363) low T1T ␻ 0 ␻ T → when the atomic form factor f(q) is a smooth function ␯Ϫ␥s Ϫ1Ϫ͑zϪ1͒/d ͵ dq Im ␹͑q,␻͒ϰ⌬ ϳ␦ , (2.359) with an appreciable weight around Q. Numerical results in accordance with the above scaling and also in agree- d␻ Im ␹͑q,␻͒ϰ⌬1Ϫ␥sϳ␦Ϫ1. (2.360) ment with experimental indications in the high-Tc cu- ͵ prates are discussed in Sec. II.E.4 and IV.C. When the Next we discuss the temperature dependence of the temperature is lowered to TϽTcoh , complexity arises because of the growth of the coherent part. In the co- scaling. The intrinsic correlation length ␰m is severely controlled by the hole concentration ␦. However, the herent temperature region, no microscopic calculation is z available for the explicit temperature dependence. How- incoherent part at distance rϽ␰m and time tϽ␰m has basically the same structure as for a Mott insulator ever, because the weight of the incoherent part is domi- where there is no effect of holes. At finite temperatures, nant over the coherent weight for small ␦, the scaling for ␦ dependence given above at Tϭ0 should be valid in the thermal antiferromagnetic correlation length ␰mT is a well-defined quantity because the correlation should this temperature range. decay exponentially at TÞ0aseϪr/␰mT in both metals A qualitatively different feature may appear in the and Mott insulators in 2D. If the temperature is lowered, pseudogap region of underdoped high-Tc cuprates due this thermal correlation length diverges as T 0. In the to an increase in the coherent contribution from pairing Mott insulating phase it leads to long-range→ order at T fluctuations. ϭ0. In metals, although ␰ diverges, the intrinsic cor- mT 11. Possible realization of scaling relation length ␰m cuts off the correlation even when ␰mT exceeds ␰m . It is rather trivial that the spin correla- A microscopic description of the Mott transition tion shows essentially the same behavior at high tem- which satisfies the scaling theory has not yet been fully peratures if ␰mTϽ␰m whether in metals or in insulators. worked out. Hyperscaling implies the existence of singu- The crossover temperature Tscr where ␰mT exceeds ␰m lar points at the MIT, around which charge excitations was introduced in Sec. II.E.4 and is schematically shown can be described. The large dynamic exponent zϭ4 im- in Fig. 34. At low doping concentration, Tscr is in gen- plies that the relevant charge excitation has a flat disper- 4 eral higher than Tcoh because of the suppression of sion ␧ϰk around those points near the MIT. This flat Tcoh , as can be seen in Fig. 34. As is discussed in Sec. dispersion has to be realized around some point in k II.E.4, the dominant weight Sinc(q,␻) has no appre- space near the large Fermi surface if the quasiparticle ciable temperature dependence below Tscr with ␰m un- description is to be justified. We consider the case in changed. This leads to the scaling form which the paramagnetic phase is kept near the antifer- Im ␹͑q,␻͒ϭ͑1ϪeϪ␤␻͒S , (2.361) romagnetic insulator (͓I-1 M-3͔ in the category of Sec. inc II.B). In this case, the Fermi↔ surface must be renormal- Ϫ1 where Sinc has the temperature-insensitive widths ϳ␰m ized to the completely nested one when the doping con- Ϫz in q and ϳ␰m J in ␻ space. These widths are rather wide centration decreases and the system approaches the an- because they are of the order of ϳ2␲ͱ␦ in q and tiferromagnetic insulator in a continuous fashion

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1115

(Miyake, 1996). This is because the shape of the Fermi in this subsection. The situation is simpler for this tran- surface changes continuously with doping while on the sition than for the fermionic Mott transition because it verge of the transition, it must coincide with the com- can be associated with the explicitly defined order pa- pletely nested shape of the Fermi level of the antiferro- rameter of the superfluid phase. This makes it possible magnetic insulator. Therefore at the Mott transition to write down an effective action explicitly using the su- point on the metallic side, a flat dispersion should be perfluid order parameter (Fisher et al., 1989). The observed around the completely nested surface. In nu- Hamiltonian for interacting bosons may be written as merical studies in the low-doping regime, as discussed in ϭ ϩ , (2.364) Sec. II.E.2, the spectral function A(k,␻) in 2D Hubbard H H0 H1 and t-J models indeed shows a flat dispersion around k ϭ(␲,0) (Bulut et al., 1994b; Dagotto et al., 1994; Preuss † 0ϭϪ͚ tij͑bi bjϩH.c.͒, (2.365) et al., 1997). In addition, there are several experimental H ͗ij͘ indications from angle-resolved photoemission which 2 suggest a flat dispersion around kϭ(␲,0) (Liu et al., 1ϭU͚ ni Ϫ␮͚ ni (2.366) 1992a; Dessau et al., 1993; Gofron et al., 1994; Marshall H i i † † et al., 1996), as is discussed in Secs. IV.C.3 and IV.C.4. with niϭb bi and the boson operators b and bi . When 4 i i This flat dispersion appears to be proportional to k , U is large, the Mott insulating phase appears at integer which means that the original Van Hove singularity fillings similarly to the fermionic case. The system un- present in the noninteracting case for the nearest- dergoes the Mott transition from Mott insulator to su- neighbor Hubbard model is not sufficient and the flat- perfluid when one changes filling or U/tij . The action is ness must be much more enhanced by electron correla- given as tion (Imada, Assaad, et al., 1998) The description ␤ ‘‘extended Van Hove singularity’’ inadequately de- † (␶biϩ ͔. (2.367ץ Sϭ d␶͓bi scribes this circumstance, which we call a ‘‘flat disper- ͵0 H sion.’’ A possible way to realize the zϭ4 universality class is, for example, for the dispersion to have the form After making the Stratonovich-Hubbard transformation 4 4 by introducing the c-number Stratonovich variable ␧ϳqxϪqy , around k0ϭ(␲,0) with qϭkϪk0 . In fact, this can occur if there is next-nearest-neighbor hopping ␺i(␶), we can expand the action in terms of ␺i as but no nearest-neighbor hopping. It may occur if antifer- romagnetic correlation is enhanced critically, because Sϭ ͵ dk d␻ r͑k,␻͉͒␺͑k,␻͉͒2 hopping to a different sublattice is suppressed by anti- ferromagnetic ordering. This flat dispersion may gener- 2 ate a strong channel of antiferromagnetic as well as Um- ϩg͵ dk d␻ i␻͉␺͑k,␻͉͒ klapp scattering with the momentum transfer kϭ(␲,␲) by scattering the quasiparticle around kϭ(␲,0) to that ϩu dx d␶͉␺͑x,␶͉͒4ϩS ͑␾,␺͒. (2.368) around kϭ(0,␲). Therefore spin degeneracies (correla- ͵ phase tions) are strongly coupled with the charge degeneracy The superfluid order is realized by ͉͗␺i(␶)͉͘Þ0, which is in momentum space, which further enhances the inco- linearly related to the superfluid order parameter herence of the charge excitations in a self-consistent ⌽(r,␶)ϭ͉⌽(r,␶)͉ei␾(r,␶) with the amplitude mode ͉⌽͉ fashion. The renormalization of the Fermi surface to a and the phase ␾. The mean-field action is obtained by perfectly nested one, as well as the appearance of the neglecting the phase fluctuation part Sphase(␾,␺). How- flat dispersion around (␲,0), implies that the self-energy ever, low-energy fluctuations are generated from the of the single particle has a singular k dependence phase mode after integrating out the amplitude mode as around (␲,0). The instability of single-particle states ␤ ␳ against the two-particle process for this flat dispersion ˙ ˙ 2 s 2 (Seffϭ dr d␶ ␳͑i␾͒ϩ␬␾ ϩ ٌ͑␾͒ , (2.369 may result in the formation of bound states around ͵ ͵0 ͩ m ͪ (␲,0), as is discussed in Sec. II.F.9 which may be associ- where ␳, ␬, m, and ␳ are the average density, the com- ated with the pseudogap phenomena in the high-Tc cu- s prates. Relationship between the scope here and experi- pressibility, the mass, and the superfluid density, respec- mental results is dicussed in Secs. IV.C.3 and IV.C.4. tively. To derive Eq. (2.369), we used the fact that from From both theoretical (numerical) and experimental the similar derivations to D and ␬ in the fermionic case, points of view, the wave-number dependence of charge the superfluid density ␳s and ␬ are related to the phase excitations with high-energy resolution is one of the stiffness constant when twists are imposed in the bound- hardest quantities to measure. Further studies are ary condition of the path integral. From a comparison of needed for a more accurate and quantitative picture. the hyperscaling form for the singular part of the free energy, f ϳ⌬␯͑dϩz͒ ͑␰/L,␰z/␤͒, (2.370) 12. Superfluid-insulator transition s F with the form for ␬ and ␳ , we obtain The scaling theory of the superfluid-insulator transi- s tion is a closely related subject, which we review briefly ␬ϰ⌬␯͑dϪz͒ (2.371)

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1116 Imada, Fujimori, and Tokura: Metal-insulator transitions and lic states (Ogata and Shiba, 1990). Recently, finite-size corrections to the Bethe ansatz solution have been com- ␳ ϰ⌬␨ (2.372) s bined with the conformal field theory to calculate the with ␨ϭ␯(dϩzϪ2) similarly to the fermionic case. For critical exponents of the Tomonaga-Luttinger liquid in a filling-control Mott transition (in other words, a ge- the strongly correlated regime (Frahm and Korepin, neric transition) in a pure system, it was deduced that 1990; Kawakami and Yang, 1990; Schulz, 1990b, 1991). the exponents zϭ1/␯ϭ2 and ␨ϭ1 hold in all dimensions Here we summarize the case of simple systems such as (Fisher et al., 1989). For a bandwidth-control Mott tran- the Hubbard model and the t-J model. In 1D interacting sition (that is, a multicritical transition) in a pure system, systems, the elementary excitation is not described by dϩ the effective action is mapped to the ( 1)- the electrons but by the spin density and charge-density dimensional XY model. Because of space-time isotropy, degrees separately. When the spin and charge degrees of zϭ1 holds. The upper critical dimension d is thus given c freedom both have gapless excitations, as in the case of by d ϭ3 and at dϾ3, zϭ1 and ␯ϭ1/2 hold. For a c the Hubbard model in the metallic phase, this quantum disorder-driven transition between a Bose glass and a superfluid, the dynamic exponent z is considered to be liquid is called a Tomonaga-Luttinger liquid, as com- zϭd as in the case of the disordered transition of fermi- pared to a Luther-Emery liquid, which we discuss later. ons because the compressibility is nonsingular. The The momentum distribution function n(k)ofa Bose-glass superfluid transition has been numerically Tomonaga-Luttinger liquid is given by studied by the Monte Carlo method. The conclusion z n͑k͒ ϭ n Ϫconstϫ kϪk ␪ sgn͑kϪk ͒, ͗ ͘ ͗ kF͘ ͉ F͉ F ϭ2 in 2D is consistent with scaling theory (Wallin et al., (2.373) 1994; Zhang, Kawashima, Carlson, and Gubernatis, 1995). where the exponent ␪ is related to the exponents for spin and charge correlation functions through K␳ as 2 G. Non-fermi-liquid description of metals ␪ϭ͑K␳Ϫ1͒ /4K␳ . (2.374) It is clear that the renormalization factor Z of the Fermi 1. Tomonaga-Luttinger liquid in one dimension liquid vanishes in a Tomonaga-Luttinger liquid because This article does not attempt to review recent ͗n(k)͘ does not have a jump at kϭkF . The charge and progress in theoretical studies of interacting 1D systems. spin correlation functions ͗␳(r)␳(0)͘ and ͗s(r)s(0)͘, We summarize the elementary properties of the respectively, are given at the asymptotically long dis- Tomonaga-Luttinger liquid in 1D theoretical systems tance r as Ϫ2 Ϫ͑1ϩK␳͒ only insofar as these are needed to discuss the present ͗␳͑r͒␳͑0͒͘ϳconstϩa0r ϩa2r experimental situation in d-electron systems. It is well Ϫ4K␳ known and can be easily shown that a perturbation ex- ϫcos 2kFrϩa4r cos 4kFr (2.375) pansion in terms of the interaction breaks down in 1D and systems. For instance, in a second-order perturbation ex- Ϫ2 Ϫ͑1ϩK␳͒ pansion of the Hubbard model in terms of U, the mo- ͗s͑r͒s͑0͒͘ϳb0r ϩb2r cos 2kFr. (2.376) † mentum distribution function n(k)ϵ͚␴͗ck␴ck␴͘ shows The Fermi-liquid state is reproduced by taking K␳ϭ1. logarithmic divergences at the Fermi level kF even in the However, in 1D interacting systems, K␳ is generally dif- second order of U, which is an indication of the break- ferent from one. In the Hubbard model, K␳ depends on 1 down of the Fermi-liquid state. After summing the so- U and filling in the range 2 рK␳р1. In the limit of half called parquet diagrams up to infinite order, we find that filling, ␦ 0, K 1 irrespective of UÞ0 while K ϭ 1 is → ␳→ 2 ␳ 2 n(k)atkF has a power-law singularity. Other ap- satisfied at any filling in the limit U ϱ. This power-law proaches such as the bosonization technique also sup- decay of spin and charge correlations→ is related to the port this singularity. The breakdown of the Fermi-liquid branch cut in the single-particle Green’s function given state in 1D was first suggested by Tomonaga (1950), who by carried out the bosonization procedure for spinless in- teracting fermions in the continuum limit. The power- 1 1 G͑r,t͒ϳ 1/2 1/2 law singularity of n(k) was analyzed in a similar model ͓xϪvstϩ͑i/⌳͒sgn t͔ ͓xϪvctϩ͑i/⌳͒sgn t͔ by Luttinger (1963). In the seventies, experimental stud- ϫ͓⌳2͑xϪv tϩi/⌳͒͑xϩv tϪi/⌳͔͒Ϫ␣c, (2.377) ies on quasi-1D conductors further stimulated theoreti- c c cal studies on 1D systems. Exact solutions by the Bethe where ⌳ is a cutoff on the order of vF /a, where a is the ansatz, perturbation expansions, renormalization-group lattice constant, while ␣cϭ␪/2. The spin and charge de- approaches, and bosonization techniques were devel- grees of freedom are decoupled in the asymptotic low- oped (see, for example, Emery, 1979 Solyom, 1979). The energy scale to cause the separation of spin and charge Tomonaga-Luttinger liquid as a universal feature of in- group velocities vs and vc . An electron propagates with teracting 1D systems was emphasized by Haldane (1980, the group velocity vF in a Fermi liquid at the Fermi 1981). The Bethe ansatz wave function for finite-size sys- level, while it separates into spin and charge degrees of tems of the Hubbard model in the limit of Uϭϱ was freedom with different group velocities vs and vc in a utilized to estimate numerically the singularity of metal- Tomonaga-Luttinger liquid. Because of this separation,

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1117

the single-particle spectral weight G(q,␻) does not have So far we have discussed the case of a quantum liquid a pole structure as in the Fermi liquid. After taking the for both spin and charge degrees of freedom. In some Fourier transform of Eq. (2.377), we find that it has the cases, such as the attractive Hubbard model with UϽ0, structure the spin degrees of freedom have an excitation gap due ␪ to the formation of a singlet or triplet bound state. ͉q/kF͉ ␻ vs G͑q,␻͒ϭ , (2.378) When the spin gap is formed, the charge correlation has qv v q v c Fͩ s c ͪ the asymptotic form with power-law singularities. ␳͑r͒␳͑0͒ ϳconstϩa rϪ2ϩa rϪK␳cos 2k r Thus the spectral function A(k,␻) of the Tomonaga- ͗ ͘ 0 2 F Luttinger model shows power-law singularities instead Ϫ4K␳ ϩa4r cos 4kFr. (2.384) of Lorentzian-like quasiparticle peaks. The thresholds of the power-law singularities for the Tomonaga-Luttinger This charge correlation function is compared with the singlet and triplet pair correlation functions given by model are at ␻ϭv␯(kϪkF)ϩ␮ and Ϫv␯(kϩkF)ϩ␮, where v ϭv and v (Meden and Scho¨ nhammer, 1993a, † Ϫ1/K␳ ␯ s c ͗O ͑r͒O͑0͒͘ϳc1r (2.385) 1993b; Voit, 1993). The density of states near the chemi- cal potential then behaves as (except for logarithmic corrections), where O is the Cooper pair order parameter. This class of quantum liq- ␪ ␳͑␻͒ϰ͉␻Ϫ␮͉ . (2.379) uid with spin gap is called a Luther-Emery-type liquid. Thus the photoemission spectra do not show a Fermi When K␳Ͼ1, the singlet pairing correlation dominates edge but instead have an intensity that vanishes at the over the charge-density correlation at kϭ2kF . This is Fermi level as a power law. the case for the attractive Hubbard model at ␦Þ0. Sev- The disappearance of the Fermi edge has been eral attempts at models for other interacting 1D systems claimed for many organic and inorganic quasi-1D metals such as the dimerized t-J model (Imada, 1991), a frus- (Dardel et al., 1992a, 1992b; Coluzza et al., 1993; Naka- trated t-J model (Ogata et al., 1991), and the ladder sys- mura et al., 1994a). The photoemission spectra of tem (Dagotto, Riera, and Scalapino, 1992; Rice et al., quasi-1D metals near the Fermi level indeed show the 1993; and for a review see Dagotto and Rice, 1996) also disappearance of the Fermi edge, as predicted by Eq. indicate that the singlet pairing correlation can be the (2.379). However, the exponent ␪ obtained from the ex- dominant correlation at asymptotically long distance. perimental data is typically 0.7–2, which is much larger An interesting but controversial problem is the evolu- than the maximum value of 0.125 predicted for the Hub- tion of three dimensionality with increasing interchain bard model (Kawakami and Yang, 1990). It has been transfer when we start from separate chains. This ques- suggested that this is related to long-ranged electron- tion was posed by Anderson (1991b) and prompted electron interactions in these compounds because arbi- many studies, which appear to have reached a consensus trarily large ␪ is possible for long-range interactions that infinitesimally small interchain transfer is relevant (Luttinger, 1963). Another puzzling aspect of the experi- and destroys the Tomonaga-Luttinger liquid in the sense mental results is that the energy range in which the of relevance in the renormalization group (Schulz, 1991; power-law fit is successful is as large as ϳ0.5 eV, in spite Castellani, di Castro, and Metzner, 1992; Fabrizio, Pa- of the fact that the power-law behavior of ␳(␻) is a result rola, and Tosatti, 1992; Haldane, 1994; Nersesyan, of spin-charge separation and therefore should be lim- Luther, and Kusmartsev, 1993). This problem was read- ited at most to the maximum spinon energy, which is of dressed by Clarke, Strong, and Anderson (1994), who the order of the superexchange interaction JՇ0.1 eV. claimed that the incoherence of the interchain hopping Thus, at the moment, it is not clear whether the experi- process still remains up to a finite critical value of inter- mental indications mentioned above constitute evidence chain transfer. A related controversy concerns the ques- for the Tomonaga-Luttinger liquid. tion whether the Tomonaga-Luttinger liquid can be re- The thermodynamic quantities of a Tomonaga- alized in 2D systems as hypothesized by Anderson Luttinger liquid are at a first glance similar to those of a (1990, 1997). We shall not discuss this controversy. Fermi liquid. For example, the specific heat, the uniform Below, we assume the relevance of interchain hopping magnetic susceptibility, and the charge susceptibility at and discuss dimensional crossover in quasi-1D systems low temperatures are given as with weak coupling between the chains. The problem here is how the physical properties change from those of Cϳ␥T, (2.380) low-dimensional systems to those of 3D ones as a func- tion of temperature and excitation energy. If we denote ␹sϳ␹s0ϳconst, (2.381) the renormalized transfer integral between the chains const, (2.382) ␹cϳ␹c0ϳ (planes) by tЌ* , the spectral function will behave as that respectively. At asymptotically low temperatures ␥ is ac- of a low-dimensional system at ͉␻Ϫ␮͉ӷtЌ* or at ␲kT tually given as the sum of spin and charge contributions, ӷtЌ* whereas it will show 3D character at ͉␻Ϫ␮͉ϽtЌ* and ␲kTϽtЌ*. The photoemission spectra of the ␥ϭ␥cϩ␥s , (2.383) quasi-1D organic conductor (DCNQI)2Cu salt, which is where ␥cϰ 1/vc and ␥sϰ 1/vs while ␹s0ϰ 1/vs and ␹c0 known to have considerable interchain coupling through ϰ 1/vc . intervening Cu atoms, exhibits such a behavior

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1118 Imada, Fujimori, and Tokura: Metal-insulator transitions

prates are unusual in the normal state in terms of the standard Fermi-liquid theory of metals. Aside from some unusual properties observed in the underdoped re- gion associated with the so-called pseudogap formation, typical examples of these properties include the follow- ing, to be discussed in greater detail in Sec. IV.C: (i) The resistivity has more or less linear dependence on T, ␳ϰT. (ii) The optical conductivity ␴(␻) has a long tail pro- portional to 1/␻ up to the order of several tenths or even 1 eV, while the true Lorentzian part of the Drude con- ductivity seems to have unusually low weight. (iii) The Raman scattering intensity has a flat back- FIG. 35. A schematic phase diagram describing a dimensional ground extending to several tenths of an eV. crossover between 1D and 3D. tЌ* is the (renormalized) inter- (iv) In photoemission experiments, the quasiparticle 3D chain transfer integral. TF ϵtЌ*/␲ is the 3D ‘‘Fermi tempera- peak is unusually broad and accompanied by an intense MF ture,’’ Tt is the mean-field CDW or SDW transition tem- incoherent background, leading to a very small spectral perature, Tt is the MIT temperature with long-range CDW or weight of the quasiparticle, if any exists. 1D SDW order, and TTL is the characteristic temperature of the (v) Nuclear magnetic resonance of Cu shows an un- 1D Tomonaga-Luttinger liquid. usually flat temperature dependence of the longitudinal relaxation time T1 above room temperature for a certain (Sekiyama et al., 1995). The spectra show that the Fermi range of doping, excepting the very low doping region. edge disappears at high temperatures, whereas a weak On phenomenological grounds, Varma et al. (1989) Fermi edge is observed at lower temperatures. On the proposed a form of single-particle self-energy ⌺ to ex- other hand, in the temperature dependence of the pho- plain all the above unusual features. This is substanti- toemission spectra of BaVS3, which is thought to be ated by more 1D like than the DCNQI-Cu salt, no sign of a y i␲ Fermi edge has been observed, but rather a suppression 2 2 ͚ ͑k,␻͒ϳg Nb͑␮͒ ͑␻Ϫ␮͒ln Ϫ y , (2.386) of the intensity at the Fermi level even more pro- ͫ ␻c 2 ͬ nounced than the power-law behavior at low tempera- where y stands for yϳmax(͉␻Ϫ␮͉,T). Here g is a cou- tures just above the MIT (at ϳ70 K), as discussed in Sec. pling constant and ␻c is a high-energy cutoff energy. IV.D.2. The schematic TϪtЌ* phase diagram of Fig. 35 Clearly, this form leads to breakdown of the Fermi- shows these experimental results in the context of di- liquid theory in a marginal way. The imaginary part of mensional crossover as a function of temperature: At the self-energy is proportional to ͉␻Ϫ␮͉ or T so that the low temperatures, TӶtЌ*/␲k, the charge transport be- inverse lifetime of the quasiparticle is comparable to its tween chains is coherent and the coupled chains behave energy. Because the quasiparticle can be defined as an as a 3D system (Kopietz, Meden, and Scho¨ nhammer, elementary excitation only when its energy ␻ is much 1995). Therefore we denote the characteristic tempera- larger than the inverse lifetime ␶Ϫ1, this form makes the 3D Ϫ1 ture of the 3D Fermi liquid tЌ*/␲k as TF . On the other quasiparticle only marginally defined because ␻ and ␶ hand, Fermi-surface nesting is more easily realized for are comparable. The form of the real part of the self- smaller tЌ* , leading to increase of the charge-density- energy in Eq. (2.386) is a consequence of causality wave or spin-density-wave transition temperature in the through the Kramers-Kronig relation. This self-energy is MF mean-field sense, Tt . Since true long-range order is singular at ͉␻Ϫ␮͉ϭ0 but retains Fermi-liquid properties possible in 3D but not in 1D, the MIT with CDW or in a marginal way. That is, it conserves the ‘‘Fermi sur- MF 3D SDW formation occurs at the lower of Tt and TF . face’’ (as in the 1D Tomonaga-Luttinger liquids), but 3D Z For small tЌ* , there is a temperature region TF ϽT the quasiparticle weight k vanishes logarithmically as ϽTMF one approaches the Fermi surface, ͉␻Ϫ␮͉ 0. Because t , where there is short-range order due to dynami- → cal CDW or SDW fluctuations. The characteristic en- ͉Im ͚(k,␻)͉ rises steeply with energy, i.e., rises linearly 1D 3D with ͉␻Ϫ␮͉ rather than quadratically as in a Fermi liq- ergy of the Tomonaga-Luttinger liquid TTL (ӷTF )is uid, the ‘‘quasiparticle’’ width dramatically increases as determined by intrachain coupling (t*ʈ , U, etc.) and is therefore independent of t* .Ift*is small, T3DϽTMF it disperses away from the chemical potential, in agree- Ќ Ќ F t et al. and no Fermi-liquid region exists, while (DCNQI) Cu ment with experimental results (Olson , 1989; Des- 2 sau et al., 1993). may belong to the large-t* region, where the Fermi- Ќ This phenomenological description of excitation spec- liquid region exists. For small tЌ* , the short-range order 3D MF tra does not assume explicit wave-number dependence. at TF ϽTϽTt may create a pseudogap in the spectral To understand the origin of this marginal behavior from function even if the system is metallic. more microscopic points of view, Virosztek and Ruvalds 2. Marginal Fermi liquid (1990) examined the effect of Fermi-surface nesting, ex- pected near the antiferromagnetic instability, as a candi- Since the discovery of the high-Tc cuprates, it has date for incoherent character. A nearly nested Fermi turned out that many electronic properties of the cu- surface may have a strong scattering mechanism at the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1119 nesting wave vector. However, this proposal does not we see from Eq. (2.274). The true Drude contribution necessarily reproduce marginal behavior in a wide re- has stronger ␦ dependence. This is closely related to the gion of doping because the nesting condition is easily large dynamic exponent of the Mott transition. We fur- destroyed by doping and, in addition, in the YBCO com- ther discuss this problem in Sec. II.F.9. In the case of pounds (YBa2Cu3O7Ϫy), the nesting condition, mean- spin correlations, S(q,␻) also has a strong incoherent ingfully defined only from the weak-correlation ap- contribution, as discussed in Sec. II.E.4. A point to be proach, is clearly not satisfied even near the Mott noted here is that the dominant incoherent contribution insulator, because of these compounds’ rather round gives again marginal Fermi-liquid-like behavior because Fermi surface (see Sec. IV.C.3). S(q,␻) has broad structure but the Bose factor 1 As is discussed in Secs. II.E.1 and II.E.3, this marginal ϪeϪ␤␻ yields marginal character. For example, the Ϫ1 Ϫ1 behavior has recently received support from numerical NMR relaxation rate T1 follows (T1T) ϰ1/T as in analyses of the t-J and Hubbard models. The frequency- Eq. (2.363) for the contribution from the incoherent dependent spin susceptibility and conductivity are given part. Another point to be stressed here is that although by Eqs. (2.274) and (2.293). According to the above nu- S(q,␻) has a broad peak structure at ␦у0.1 as C/͓i␻ merical analyses, the spin-dynamical structure factor ϩD(q2ϩ␬2)͔, as discussed in Sec. II.E.4, its widths in q Re S(q,␻) and the current-current correlation function and ␻ space become critically sharpened as ␦ 0. Origi- C(␻), discussed in Secs. II.E.1 and II.E.3, are more or nally the marginal Fermi-liquid hypothesis was→ proposed less temperature independent for the dominant part, without considering appreciable q dependence. This ␻- with a broad peak around ␻ϭ0orqϭQ,␻ϭ0 where dependent but q-independent structure is naturally un- their widths in ␻ space are on the order of eV for C(␻) derstood from the form of Eqs. (2.274) and (2.293) when and comparable or less than a tenth of an eV for S(q,␻) we ascribe this behavior to incoherent dynamics. The at ␦ϳ0.1ϳ0.2. This implies that the spin and charge dy- origin of this incoherence is discussed in Sec. II.F. namics are strongly incoherent with a short time scale of There are two different views concerning the micro- relaxation whose inverse is of the order of 1 eV for scopic origin of marginal behavior, or incoherent dy- C(␻). In the conductivity this means that C(␻) namics. Marginal behavior or more generally non- 2 ϭ͚␯␮͉͗␯͉J͉␮͉͘ ␦(␧␯Ϫ␧␮Ϫ␻)is␻insensitive at small ␻. Fermi-liquid behavior can appear when charge is When we take the Kramers-Kronig transformation of coupled to other degenerate degrees of freedom in ad- Im ␹(q,␻)orRe␴(␻), it turns out that it necessarily dition to the Fermi degeneracy. This indeed happens at leads to a form of ‘‘marginal Fermi liquid.’’ For ex- a critical point such as the magnetic instability point, ample, in the case of conductivity, one can show from where there is an arbitrarily small energy scale due to Eq. (2.274) the degeneracy. In one view, the critical point exists at 0 while in the other the critical point is at ϭ0, the b␤␻ ␦Þ ␦ Im ␴͑␻͒ϳ C͑␻͒, (2.387) Mott transition point. ␤␻ϩa In the first view, degeneracy, arises at the critical point where a and b are constants. When we wish to interpret between the phases of normal metal and the supercon- this behavior by extending the Drude analysis and intro- ducting state at Tϭ0. For example, Varma (1996b) ducing an ␻-dependent mass m*(␻) and a relaxation claimed that the critical point is at the optimal doping time ␶(␻) in the form concentration for the highest superconducting Tc and that degeneracy is due to a double-well-type structure of ␴0͑␻͒ ␴͑␻͒ϭ , potentials for local charge which comes from valence Ϫi␻ϩ␶Ϫ1͑␻͒ fluctuation. Another possible origin of marginal behavior with in- ne2 coherent charge dynamics is the Mott transition itself ␴0͑␻͒ϭ , m*͑␻͒ (Imada, 1995b). The Mott transition has a unique char- we obtain ␶Ϫ1(␻)ϵ␻ ͓Re ␴(␻)/Im ␴(␻)͔ Ӎ (1/b) ␻ acter with a large dynamic exponent zϭ4, as explained ϩ (a/b) T. This leads to ␶Ϫ1 that is linear in T at small in Sec. II.F, which causes quantum critical behavior in a ␻, just as in the phenomenological description of a mar- wide region of the (␦,T) plane. In consequence there is 2 ginal Fermi liquid, in overall agreement with experi- suppression of the coherence temperature Tcohϰ␦ , ments in cuprates, as is discussed in Sec. IV.C (Thomas which leads to incoherent charge dynamics, consistent et al., 1988; Schlesinger et al., 1990; Uchida et al., 1991; with the observation described in the first part of this Cooper et al., 1992; for a review see Tajima, 1997). This subsection. This also naturally explains why the wave- interpretation is based on the observation that the number dependence is absent in the ‘‘marginal-Fermi- charge dynamics is totally incoherent because C(␻) has liquid’’ behavior. very broad structure with more or less ␻-independent value C0 up to the order of eV. The constant C0 has a H. Orbital degeneracy and other complexities doping concentration dependence C ϰ␦, so that the ␻- 0 1. Jahn-Teller distortion, spin, and orbital fluctuations integrated conductivity up to the order of 1 eV shows and orderings the scaling ͐␴(␻)d␻ϰ␦. However, it should be noted that this contribution comes from totally incoherent In transition-metal compounds, various degrees of charge dynamics without any true Drude response, as freedom combined together can play roles in physical

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1120 Imada, Fujimori, and Tokura: Metal-insulator transitions properties. Roles of spin degrees of freedom are dis- cussed in previous subsections in detail. However, in ϭ ͑JsSi SjϩJ␶␶ជ i ␶ជ jϩJs␶Si Sj␶ជ i ␶ជ j͒, (2.388) H ͚ • • • • transition-metal compounds, as mentioned in Sec. II.A, ͗i,j͘ the orbital degrees of freedom are degenerate in various where the usual spin-orbit interaction S␶ is neglected ways depending on the crystal field coming from the ba- (Inagaki and Kubo, 1973, 1975; Kugel and Khomskii, sic lattice structure and Jahn-Teller distortions. This de- 1982). Here, the parameters are given by generacy produces a rich structure of low-energy excita- 1 1 3 2 tions through orbital fluctuations and orbital orderings. Jsϭt Ϫ ϩ ϩ , (2.389a) In this subsection the role of orbital degeneracies, com- ͩ UЈϪJH UЈ Uͪ bined sometimes with the Jahn-Teller effect, is dis- 3 1 1 2 cussed. The degenerate Hubbard model defined in Eqs. J␶ϭt ϩ Ϫ , (2.389b) (2.6a)–(2.6e) may be a good starting point, if we may ͩ UЈϪJH UЈ Uͪ neglect the coupling to the lattice. As is mentioned in 1 1 1 Sec. II.A.2, the first term of is the origin of the J ϭ4t2 Ϫ ϩ . (2.389c) DUJ s␶ U ϪJ U U Hund’s-rule coupling, which isH the strong driving force ͩ Ј H Ј ͪ that aligns spins of d electrons on the same atom. If Js␶ is positive and J␶ϾJs , this leads to stabilization of We discuss here the basic structure of degeneracy in a the ferromagnetic state with the ‘‘antiferro-orbital’’ or- cubic crystal field. When three t2g orbitals are degener- dering (Roth, 1966, 1967; Inagaki, 1973). In more realis- 2 3 ate, d and d states usually have spin quantum number tic cases, the transfer is anisotropic with finite off- Sϭ1 and Sϭ3/2 where two or three t2g orbitals are oc- diagonal hoppings, and hence J and J become highly 3ϩ 3ϩ ␶ s␶ cupied with parallel spins, as in V and Cr . If the anisotropic depending on the direction. crystal-field splitting between t2g and eg is smaller than The orbital degeneracy can be lifted by several on-site 4 5 the Hund’s-rule coupling, d , and d states are realized mechanisms, including the spin-orbit interaction and the by occupying three spin-aligned t2g orbitals and one or Jahn-Teller distortion. In the case of spin-orbit interac- two electrons with the same spin in the eg orbitals, as in tion, splitting of orbital levels is usually accompanied by Mn3ϩ and Fe3ϩ, respectively. As in Co3ϩ, there may be 6 simultaneous spin symmetry breaking. For example, if subtlety in d due to competing states of fully occupied the spins order along the z axis, then the LS interaction t orbitals (a low-spin state), five electrons in the t 2g 2g ␭L–S immediately stabilizes the orbitals with the z com- orbitals in addition to one eg electron (an Sϭ1 state), ponent of the angular momentum LzϭϮ1orLzϭϮ2 and four electrons in the t2g orbital plus two spin-aligned L ϭ 7 3ϩ instead of z 0. Depending on the sign of ␭, the orbit- eg electrons (Sϭ2 state). For the case of d as in Ni , als are chosen so that Lz and the spin align in a parallel the low-spin state is realized with fully occupied t2g or- or antiparallel way. (Strictly speaking, the spin and or- bital and singly occupied eg orbital with Sϭ1/2, as in bital quantum numbers are not good quantum numbers 2ϩ 8 RNiO3.AsinNi ,d is given by a state in which the t2g separately, but the total angular momentum is con- orbitals are completely filled and two e orbitals are oc- g served.) When a finite Lz state is chosen, it may also cupied by spin-aligned electrons (Sϭ1 state) if the induce lattice distortion. For example, in the octahedron crystal-field splitting of dx2Ϫy2 and d3z2Ϫr2 is weak. In surroundings of the ligand atoms, a Jahn-Teller distor- such ways, depending on these electron fillings, different tion occurs in such a way that the octahedron elongates physics appears. Dynamic coupling to the lattice is an- in the z direction because the electrons occupy nonzero other source of complexity. Lz states, such as yz and zx orbitals. We start with the simplest degenerate case, with two Coupling between orbital occupancy and the Jahn- orbitals with one electron (or one hole) occupied at each Teller distortion can play a major role as a driving force site on average. This is indeed the case, for example, in of symmetry breaking, because the orbital occupation 9 KCuF3, in which the Cu d state leads to one hole per may strongly couple to the lattice in some cases. The site in the twofold degenerate eg orbitals, under the cu- simplest example of this coupling is seen in Fig. 36, bic crystal field of the perovskite structure. RNiO (R 3 where an electron in a d3z2Ϫr2 or dx2Ϫy2 orbital induces ϭLa, Nd, Y, Sm) is another example, in which eg orbit- distortion of anions in the octahedron structure around als are occupied with one electron per site although t2g the transition-metal elements. The static Jahn-Teller dis- orbitals are fully occupied. If the intrasite Coulomb in- tortion of course lifts the original degeneracy of t2g or eg teraction U␯␯Ј in Eq. (2.6c) is large enough, as in orbitals. The Jahn-Teller distortion occurs because the KCuF3, this integer filling state should be the Mott insu- energy gain due to this splitting of the degeneracy, lator. Here, we discuss the effective Hamiltonian for the which is linear in the distortion, always wins in compari- strong-coupling limit to derive the orbital and spin ex- son to the loss of energy due to the lattice elastic energy, change as in Eq. (2.8). To understand the basic physics, which is quadratic in the distortion. When the crystal we make a drastic simplification to Eq. (2.8). We first lattice of these transition-metal elements and anions is ␯␯Ј assume only the diagonal transfer tij ϭt␦␯␯Ј and ␯- formed, the static Jahn-Teller distortion frequently fa- dependent interaction U␯␯ЈϭUЈ for ␯Þ␯Ј and U␯␯ vors a specific type of distortion pattern. For example, in ϭU. Here JH is the Hund’s-rule coupling. Through the the configuration of Fig. 36, the Jahn-Teller distortion second-order perturbation, the exchange interaction is stabilizes the ‘‘antiferro-orbital’’ ordering of the pseu- given by the simplified Hamiltonian dospin ␶ for dx2Ϫy2 and d3z2Ϫr2 orbitals. This induces the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1121

This is because the octahedron has corner-sharing struc- ture for adjacent transition-metal elements, so that the displacement of an anion at the corner induces occupa- tions of opposite orbitals for the transition-metal ion in the neighborhood of the distorted anion, as in Fig. 36. In contrast, an edge-sharing system such as the spinel struc- ture has a tendency for ‘‘ferro-orbital’’ ordering. In addition to the static Jahn-Teller distortion, the dy- namic process of virtual distortion also mediates the ex- change interaction of pseudospins ␶ on adjacent sites ␣ ␣ ␣ through a coupling of the form gk qk␶k where gk (␣ ϭx,y,z) is the coupling constant, qk is the anion distor- ␣ tion (phonon), and ␶k is the Fourier transform of the ␣ pseudospin operator ␶i . By eliminating phonon degrees of freedom in the second-order perturbation, we obtain the pseudospin exchange in the form

z z z xy ϩ Ϫ Ϫ ϩ ϭ͚ ͓Jij␶i ␶j ϩJij ͑␶i ␶j ϩ␶i ␶j ͔͒. (2.390) H ͗ij͘ Direct quadrupole-quadrupole interaction between d electrons on adjacent sites also generates a similar effec- tive Hamiltonian for ␶. These exchange interactions be- tween orbitals can be the origin of orbital ordering even without the static Jahn-Teller distortion. In realistic situ- ations, exchange coupling of pseudospins through the dynamic Jahn-Teller interaction or the quadrupole- quadrupole interaction, and static Jahn-Teller distor- tions as well as super-exchange-type coupling as in Eq. FIG. 36. Examples of the Jahn-Teller distortion by antiferro- (2.388) or (2.8) due to the strong-correlation effect may orbital occupation of (upper panel) dx2Ϫy2 and dy2Ϫz2 orbitals be entangled and compete or cooperate in rich ways. In and (lower panel) dx2Ϫy2 and d3z2Ϫr2 orbitals. this article, we do not go into a comprehensive descrip- tion of such interplay. Readers are referred to the re- view article by Kugel and Khomskii (1982) for this prob- cooperative Jahn-Teller effect, accompanied by symme- lem. Below, a few typical examples of recent try breaking of both lattice and orbitals (Kanamori, developments are described. 1959). A typical example of such ‘‘antiferro-orbital’’ Several mean-field-level calculations have been done symmetry breaking can be seen in KCuF , as shown in 3 on models with orbital degeneracy. They include Fig. 37. A general tendency of the cooperative Jahn- Hartree-Fock calculations (Roth, 1966, 1967; Inagaki Teller effect is that some type of ‘‘antiferro-orbital’’ or- and Kubo, 1973, 1975), the Gutzwiller approximation dering takes place, usually in the perovskite structure. (Chao and Gutzwiller, 1971; Lu, 1994), slave-boson ap- proximations (Fre´sard and Kotliar, 1997; Hasegawa, 1997a, 1997b), and the infinite-dimensional approach (Rozenberg, 1996a, 1996b; Kotliar and Kajueter, 1996; Kajueter and Kotliar, 1996). Quantum Monte Carlo cal- culations were also carried out to study MIT, and spin and orbital fluctuations (Motome and Imada, 1997, 1998). In V2O3 compounds, as is discussed in Sec. IV.A.1, the t2g levels may split into a single a1g orbital and twofold- degenerate eg orbitals due to the corundum structure. In particular, the strong orbital fluctuation combined with spin fluctuation can be the origin of the first-order tran- sition between a paramagnetic insulator and an antifer- romagnetic insulator (Rice, 1995). See Sec. IV.A.1 for further discussions. In terms of effects of orbital degen- eracy, mechanisms of gapped spin excitations in the Mott insulating phase of another vanadium compound, CaV4O9, were examined (Katoh and Imada, 1997). Recently, the interplay of electron correlation with spin orderings, orbital orderings, and the Jahn-Teller

FIG. 37. Antiferro-orbital order in a phase of KCuF3. distortion has been examined in the end materials of 3D

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1122 Imada, Fujimori, and Tokura: Metal-insulator transitions

IV.F.1. Various types of Jahn-Teller distortions and an- tiferromagnetic order realized in 3D perovskite struc- ture are shown in Figs. 38(b) and 38(c), respectively. Although the spin degrees of freedom have been stud- ied rather thoroughly on the level of their fluctuations in the paramagnetic phase, fluctuations of the orbital de- grees of freedom have been neither experimentally nor theoretically well studied. These fluctuation effects may be important when the system is metallized away from the end material. Orbital ordering phenomena are also widely observed in f-electron systems, where they are called quadrupole orderings. An example is seen in CeB6. The interplay of spin and orbital fluctuations may be very similar in d- and f-electron systems, though it is not fully understood so far.

2. Ferromagnetic metal near a Mott insulator

When the Jahn-Teller distortion is absent in LaMnO3, the antiferro-orbitally ordered 3x2Ϫr2 and 3y2Ϫr2 states strongly hybridize with the 3z2Ϫr2 orbital be- cause of high symmetry and hence enhance the ferro- magnetic coupling, basically the same as in the Hund FIG. 38. Types of lattice distortion and antiferromagnetic or- mechanism. It has been argued that this is also the origin der. (a) GdFeO3-type tilting distortion of MO6 octahedron; (b) of ferromagnetism in metallic doped systems like types of Jahn-Teller distortion in 3D perovskite structure; (c) La1ϪxSrxMnO3 because the static Jahn-Teller distortion types of antiferromagnetic order in 3D perovskite structure. seems to disappear in the metallic phase. The basic physics of this mechanism was pointed out by Kanamori perovskite compounds. In the 3D perovskite com- (1959). pounds, RMO3, in addition to the Jahn-Teller distor- Although the underlying physics and the mechanism tion, tilting of the MO6 octahedron, illustrated in Fig. are similar, a slightly different way of explaining the fer- 38(a), plays an important role in determining the hybrid- romagnetism in metallic La1ϪxSrxMnO3 is the so-called ization of different orbitals on adjacent sites and conse- double-exchange mechanism (Zener, 1951; de Gennes, quently in controlling the bandwidth. Considering this 1960; Anderson and Hasegawa, 1955). Because the lo- tilting as well as the Jahn-Teller distortion and spin and calized 3t2g electrons are perfectly aligned due to the orbital orderings, Mizokawa and Fujimori (1996a, Hund’s-rule coupling, the double-exchange Hamiltonian 1996b) performed unrestricted Hartree-Fock calcula- defined in Eqs. (2.9a)–(2.9c) may be a good starting tions on RMO3. In general, the spin and orbital order- point if degeneracy of the eg orbitals can be neglected. ings obtained were in good agreement with those of the In the completely ferromagnetic phase, eg electrons can experimentally observed results when the observed hop coherently without magnetic scattering by t2g spins, Jahn-Teller and tilting distortions were assumed. Elec- while they become strongly incoherent due to scattering tronic structure calculations with the local spin-density through the JH term if the localized t2g spins are disor- approximation (Hamada, Sawada and Terakura, 1995; dered. In other words, if t2g spins on adjacent sites have Sarma et al., 1995), with GGA (Sawada, Hamada, Tera- different directions, the eg electron feels a higher energy kura, and Asada, 1996) and with the LDAϩU method on one of the two sites because of Hund’s-rule coupling. (Solovyev, Hamada, and Terakura, 1996b) also suc- The gain in kinetic energy due to coherence favors the ceeded in improving and reproducing general tendency ferromagnetic order in the metallic phase. When the t2g of spin and orbital orderings by assuming the proper orbital electrons are treated as a classical spin coupled Jahn-Teller and tilting distortion. However, these two with eg electrons by very large Hund’s-rule coupling, the approaches show different tendencies for the band gap. spin axis of the quantization may be taken in the direc- The Hartree-Fock and LDAϩU calculations in general tion of the t2g spins as in the discussion below Eq. overestimate the band gap, while the LSDA and GGA (2.156). Then the transfer between the sites i and j in calculations tend to underestimate the gap, as is ex- Eq. (2.9b) is modified through the relative angle of the plained in Sec. II.D.3. At the moment, self-consistent t2g spins at the i and j sites as determination of lattice distortion, orbital occupation, i͑␾iϪ␾j͒ and spin ordering have not been attempted successfully tijϭt͓cos͑␪i/2͒cos͑␪j/2͒ϩsin͑␪i/2͒sin͑␪j/2͒e ͔, (2.391) except in the case of KCuF3 by Liechtenstein, Anisimov, and Zaanen (1995). Detailed discussions of each com- where ␾ is the phase of the wave function. Neglecting pound are given in Secs. IV.A.3, IV.B.1, IV.B.2, and the phase fluctuations, this leads to

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1123

tijϭt cos͑␪ij/2͒, (2.392) charge ordering with some commensurate periodicity to the lattice may occur. This commensurate charge order where ␪ij is the relative angle of the t2g spins. In the may indeed be stabilized in the presence of intersite Hartree-Fock treatment, the effective averaged transfer Coulomb interactions, as in the extended Hubbard is given by model [Eqs. (2.5b)–(2.5c) with VÞ0]. The charge order ˜ at a simple commensurate filling may also be stabilized t ϭt͗cos͑␪ij/2͒͘. (2.393) by the underlying mass enhancement due to the proxim- Recently, giant or colossal negative magnetoresis- ity of the Mott transition. In this sense, the charge or- tance has attracted renewed interest in ferromagnetic dering effect is enhanced in strongly correlated systems metals in Mn and Co perovskite compounds. This is dis- as a secondary effect of the Mott transition. Charge or- cussed in detail in Sec. IV.F. From the theoretical point dering at a commensurate filling leads to an incompress- of view, d electrons in Mn and Co compounds have ible state, which may have similarities to the most funda- strong coupling to Jahn-Teller-type distortions and fluc- mental Mott insulator itself. Experimentally, charge tuations. These couplings as well as orbital degeneracy orderings are observed for some simple commensurate in eg orbitals make magnetic and transport properties periodicities, as will be discussed in Sec. IV.E. This is complicated, especially near the Mott insulating phase. also related to the valence fluctuation problem in heavy- The hardest and most challenging subject is to under- fermion compounds. We discuss theoretical aspects of stand the interplay of spin, orbital, and lattice fluctua- charge ordering in each compound in connection with tions when the critical fluctuations of the Mott transi- the experimental results in more detail in Sec. IV. Pos- tion, exemplified by incoherent charge dynamics, sible pairing mechanisms due to dynamic fluctuation of appear. We review the present status of this theoretical charge ordering or phase separation will be discussed in effort in Sec. IV.F.1 and compare theory with experi- Sec. IV.C.3. mental indications. III. MATERIALS SYSTEMATICS 3. Charge ordering Periodic structures of spin and charge orderings are A. Local electronic structure of transition-metal oxides discussed within the framework of the Hartree-Fock ap- 1. Configuration-interaction cluster model proximation in Sec. II.D.2, where the periodicity of the spin or charge orderings is assumed to be commensurate Prior to the early 1980s, the most widely accepted pic- to the lattice periodicity. This commensurate order, as- ture of band-gap formation in a Mott insulator was that sumed in the above results, has been reexamined in 2D the d band was split into lower and upper Hubbard within the unrestricted Hartree-Fock approximation to bands, separated by the on-site d-d Coulomb interac- allow an incommensurate ordering vector (Machida, tion U, as discussed in Sec. II and shown in Fig. 5(a). 1989; Poilblanc and Rice, 1989; Schulz, 1990a; Inui and Charge excitations across that band gap were thus real- n n Littlewood, 1991; Verge´s et al., 1991). These analyses in- ized by a charge transfer of the type (d )iϩ(d )j nϪ1 nϩ1 dicate that, within the Hartree-Fock approximation, an- (d )iϩ(d )j , where i and j denote transition- → tiferromagnetic ordering at the commensurate wave vec- metal sites, and cost energy U. This classical picture was tor (␲,␲) gives way to incommensurate ordering if the first questioned by resonance photoemission studies of filling deviates away from half filling. Incommensurate NiO (Oh et al., 1982; Thuler et al., 1983) and of a series ordering may also be realized by introducing a stripe of transition-metal halides (Kakizaki et al., 1983). Until structure in real space, such as the periodic stripe of hole then the photoemission spectra of NiO had been inter- arrays (Zaanen and Gunarson, 1989). In this case, preted within ligand-field theory (Sugano et al., 1970) as Hartree-Fock solutions give an insulating phase up to due to transitions from the d8 ground state of the Ni2ϩ some finite doping concentration of holes even away ion to the d7 final-state multiplet. As shown in Fig. 39, from half filling. This problem was further explored by these transitions constitute the topmost part of the the variational Monte Carlo method by Giamarchi and valence-band photoemission spectra of NiO, below Lhuiller (1990) and by a Monte Carlo method with which is located the oxygen 2p band. In addition to fixed-node approximation by An and van Leeuwen these ‘‘main’’ features, there is another structure below (1991). The possibility of phase separation into hole-rich theO2pband called a ‘‘shake-up satellite’’ or ‘‘multi- and hole-free phases was also examined, although calcu- electron satellite.’’ It was thought that the satellite arose lational results are limited to small clusters (Emery, Kiv- from a transition from the filled oxygen p orbitals to the elson, and Lin, 1990). As we discuss later, the problem unoccupied metal d orbitals subsequent to the emission with the above arguments is that the effects of quantum of the d electron from the d8 ground state. If this assign- fluctuations, which may destroy the stripe phase and ment were correct, the satellite would be due to the d8Lគ phase separation, have not been seriously considered. In final states as opposed to the d7 final states of the main Sec. II.E.6, we show that phase separation is clearly seen band, where Lគ denotes a hole in the ligand O 2p band. by numerical methods in some other models, as it should Then photoemission spectra recorded using photons in be, whereas it is not observed numerically in the 2D the Ni 3p 3d absorption region should have shown Hubbard model. Although the stability of incommensu- enhancement→ of the main band derived from the 3d8 rate phases against quantum fluctuations is questionable, 3d7 spectral weight through the resonance photoemis- →

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1124 Imada, Fujimori, and Tokura: Metal-insulator transitions

action (CI) within a simplified metal-ligand cluster model, in which one ligand Lគ orbital and one d orbital are considered. He showed that the satellite should be due to d7 final states rather than d8Lគ final states if it showed resonance enhancement at the 3p 3d thresh- old. Subsequently, a more realistic CI cluster-model→ analysis of the photoemission spectra of NiO was made by Fujimori, Minami, and Sugano (1984), who took into account the multiplet structure of the Ni ion and symmetry-adopted molecular orbitals consisting of the ligand p orbitals. According to their analysis, shown in Fig. 39, it was confirmed that the satellite corresponds to the lower Hubbard band (d8 d7 spectral weight) and is → located below the O 2p band (d8 d8Lគ spectral weight), corresponding to the situation schematically→ de- picted in Fig. 5(b). In NiO, the O 2p-to-Ni 3d charge- transfer energy ⌬ (ϳ4 eV) is smaller than the on-site d-d Coulomb energy U (ϳ 7 eV), which makes the n n n p-to-d charge transfer, (d )iϩ(d )j (d Lគ )i nϩ1 → ϩ(d )j , the lowest-energy charge fluctuation and hence the band gap of the system (see Fig. 39). Here, the charge-transfer energy ⌬ is defined as the energy re- quired to transfer an electron from the filled ligand p FIG. 39. Photoemission and inverse-photoemission (BIS) orbitals to the empty metal d orbitals, dn dnϩ1Lគ . spectra of NiO (Sawatzky and Allen, 1984) and their analysis → using ligand-field (LF) theory with the Ni2ϩ ion and that using Because the on-site Coulomb energy U is often larger than the d bandwidths in 3d transition-metal com- CI theory for the NiO6 cluster model (Fujimori, Minami, and Sugano, 1984). pounds (U being typically as large as 4–8 eV), it is often more convenient to consider the total energies of a metal-ligand cluster such as NiO6 within CI theory sion process, 3p63d8ϩh␯ 3p53d9 3p63d7ϩ␧l, rather than as the energies of one-electron states, at where ␧l is a photoelectron. However,→ the→ experimental least for insulating compounds. Figure 40 shows the results showed the enhancement of the satellite rather total-energy diagram of the NiO6 cluster in the neutral, than the main band, which implied that the satellite was positively ionized, and negatively ionized states, which in fact due to the d7 final states, that is, the lower Hub- we refer to as N-electron, (NϪ1)-electron, and (Nϩ1)- bard band is located below the O 2p band. Davis (1982) electron states, respectively (Fujimori and Minami, made an analysis of the resonance photoemission results 1984). Here, it should be noted that the ‘‘neutral’’ NiO6 12Ϫ of the Ni compounds by treating the configuration inter- cluster has negative charges of 12- ͓(NiO6) ͔ in order

FIG. 40. Total-energy diagram of the NiO6 cluster in the ‘‘neutral’’ (N-electron), positively ionized [(NϪ1)-electron], and negatively ionized [(Nϩ1)-electron] states. Eh and Ee are the minimum energies required to create an electron and hole, respectively. eF denotes an electron at the Fermi level.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1125 to ensure the complete filling of the O 2p band and the therefore, one has to remove an electron from the d-band filling of eight as in bulk NiO. The ground state N-electron ground state of one cluster and to add that of the N-electron state primarily consists of the d8 con- electron to the N-electron ground state of another clus- figuration (corresponding to the ionic Ni2ϩ and O2Ϫ con- ter and calculate the total energy difference: figurations) with admixture of charge-transfer configura- tions d9Lគ and to a lesser extent d10Lគ 2: EgapϭEg͑Nϩ1͒ϩEg͑NϪ1͒Ϫ2Eg͑N͒, (3.5) where E (NϪ1) and E (Nϩ1) denote the lowest en- 8 9 10 2 g g ⌽g͑N͒ϭ␣1͉d ͘ϩ␣2͉d Lគ ͘ϩ␣3͉d Lគ ͘. (3.1) ergies of the (NϪ1)- and (Nϩ1)-electron states, re- The ground state of the N-electron state is split into spectively. In Fig. 40 are plotted the total energies of the multiplet terms due to intra-atomic Coulomb and ex- (NϪ1)- and (Nϩ1)-electron states as well as those of 3 the N-electron state of the NiO6 cluster. In the change interactions. The lowest multiplet term A2g has 7 6 2 (NϪ1)-electron state, the d states are located above a high-spin (Sϭ1) t e configuration, as predicted by 2g g the charge-transfer d8L states because ⌬ϽU, and the ligand-field theory. Hybridization with the charge- គ energy separation between the two configurations is U transfer states d9Lគ , which are located by ϳ⌬ higher, Ϫ⌬. In analogy with Eq. (3.1), the (NϪ1)-electron causes a ‘‘crystal-field splitting’’ of the d8 multiplet as state of a given spin and spatial symmetry is given by well as a lowering of the ground-state energy Eg(N). This energy lowering signifies the covalency contribu- ⌽ NϪ1͒ϭ␤ d7 ϩ␤ d8L ϩ . (3.6) tion to the cohesive energy of the ground state. i͑ i1͉ ͘ i2͉ គ ͘ ¯ If there were no multiplet splitting and p-d hybridiza- In the (Nϩ1)-electron state, the d10Lគ configuration is tion, the total energy of the cluster with the central ion located well above d9 because of the large energy sepa- in the dn configuration would be given by (Bocquet ration Uϩ⌬ӷ0. From Fig. 40, one can also recognize et al., 1992a; Fujimori et al., 1993) that the magnitude of the band gap (EgapϭEeϩEh , where Ee and Eh are the lowest energies of the extra 1 electron and the extra hole added to the N-electron E͑dn͒ϭE ϩn␧0 ϩ n͑nϪ1͒U, (3.2) 0 d 2 ground state, respectively, as shown in the figure) is ϳ⌬ 0 rather than ϳU. where E0 is a reference energy and ␧d is the energy of a bare d electron (including the Madelung potential at the cation site). The charge-transfer energy ⌬ is then given by 2. Cluster-model analyses of photoemission spectra nϩ1 n 0 To obtain the electronic structure parameters ⌬, U, ⌬ϭE͑d ͒Ϫ␧pϪE͑d ͒ϭ␧dϪ␧pϩnUϵ␧dϪ␧p , (3.3) and T, the CI cluster model has been widely used to analyze photoemission spectra. The wave functions of where ␧ is the energy of the ligand p orbital neglecting p the ground state, photoemission final states, and inverse- the finite p bandwidth. Here, the Coulomb interaction photoemission final states of the MO metal-oxygen among the p electrons and that between neighboring p m cluster are given by and d electrons are ignored. One can obtain the ground state Eq. (3.1) by solving a secular equation for the ⌽ ͑N͒ϭ␣ ͉dn͘ϩ␣ ͉dnϩ1Lគ ͘ϩ␣ ͉dnϩ2Lគ 2͘ϩ , Hamiltonian, g 1 2 3 ¯ nϪ1 n nϩ1 2 8 ⌽i͑NϪ1͒ϭ␤i1͉d ͘ϩ␤i2͉d Lគ ͘ϩ␤i3͉d Lគ ͘ϩ¯ , E͑d ͒Ϫ⌬E8 Ϫ&T 0 8 ϭ Ϫ&TE͑d͒ϩ⌬ Ϫ&T , Nϩ ϭ dnϩ1 ϩ dnϩ2L ϩ dnϩ3L2 ϩ H ⌽i͑ 1͒ ␥i1͉ ͘ ␥i2͉ គ ͘ ␥i3͉ គ ͘ ¯ . ͩ 0 Ϫ&TE͑d8͒ϩ2⌬ϩUͪ (3.7) (3.4) Here, each wave function is an eigenstate of the total for NiO. Here, TϵϪ)(pd␴)ϵϪ2tpd [with (pd␴) be- spin S and Sz and has the proper point-group symmetry ing a Slater-Koster parameter (Harrison, 1989)] is a p-d of the cluster. The d-derived spectral function 3 transfer integral and Ϫ⌬E8 is the shift of the A2g term (d-derived density of states) ␳d(␻) is thus calculated us- 8 † of the d multiplet relative to the center of gravity ing Eq. (2.54), with ck and ck replaced by the annihila- 8 8 † E(d ) of the d multiplet. In the present CI treatment, tion and creation operators of the d electron, cd and cd : the crystal-field splitting between the t2g and eg orbitals is included not through an input parameter 10Dq as in * * * 2 ␳d͑␻͒ϭ͚ ͉␤i1␣1ϩ␤i2␣2ϩ␤i3␣3͉ ligand-field theory but as a result of the different p-d i hybridization strengths for the t and e orbitals. (One 2g g ϫ␦ ␻ϩE ͑NϪ1͒ϪE ͑N͒ has also to consider a small nonorthogonality contribu- „ i g … tion to 10Dq, typically ϳ0.5 eV.) 2 E ϩ ͉␥i*1␣1ϩ␥2*␣i2͉ The band gap gap of the system is given by the lowest ͚i energy of an electron plus that of a hole, which are cre- ated in the crystal independently. In the cluster model, ϫ␦„␻ϪEi͑Nϩ1͒ϩEg͑N͒…. (3.8)

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1126 Imada, Fujimori, and Tokura: Metal-insulator transitions

Since the energy levels of the cluster are discrete, the diagonal matrix elements of ϳT, whereas in the CI delta functions in Eq. (3.8) have to be appropriately treatment the off-diagonal matrix elements of the many- broadened in order to be compared with experiment. electron Hamiltonian are ͱ10ϪnT, ͱ11ϪnT,or Let us consider the simplest case in which only two ͱ9ϪnT in the N-electron, (NϪ1)-electron, or configurations are used as basis functions for each of the (Nϩ1)-electron states [Eqs. (3.10) and (3.11)], respec- N- and NϪ1-electron states: tively. Such enhanced off-diagonal matrix elements en- n nϩ1 hance spectral weight transfer from high-energy to low- ⌽g͑N͒ϭcos ␪͉d ͘ϩsin ␪͉d Lគ ͘, energy states (satellites to main features) as well as nϪ1 n enhance the energy splitting between them. This can be ⌽Ϫ͑NϪ1͒ϭcos ␾͉d ͘ϩsin ␾͉d Lគ ͘, seen by comparing the Hartree-Fock band density of nϪ1 n ⌽ϩ͑NϪ1͒ϭϪsin ␾͉d ͘ϩcos ␾͉d Lគ ͘, (3.9) states (DOS) of NiO and the actual photoemission spec- where Ϫ␲/2Ͻ␪, ␾Ͻ␲/2. The Hamiltonian for the trum [or the Hartree-Fock band DOS corrected for the N-electron state is self-energy (Mizokawa and Fujimori, 1996a; Takahashi and Igarashi, 1996)], where the lower Hubbard band, n E͑d ͒ Ϫͱ10ϪnT well (ϳ7 eV) below EF in the Hartree-Fock approxima- ϭ . (3.10) H ͩ Ϫͱ10ϪnT E͑dn͒ϩ⌬ ͪ tion, is shifted further away from EF and spread in en- ergy and the spectral weight of the lower Hubbard band Here, the multiplet structure and the different transfer is transferred to the main band within ϳ5eVofEF. integrals for the eg and t2g orbitals have been ignored. For the (NϪ1)-electron final states 3. Zaanen-Sawatzky-Allen classification scheme E͑dnϪ1͒ Ϫͱ11ϪnT In the (NϪ1)-electron state of NiO, where ⌬ϽU, the ϭ nϪ1 , (3.11) H ͩ Ϫͱ11ϪnT E͑d ͒ϩ⌬ϪUͪ dnLគ configuration is located below the dnϪ1 configura- whose lower and higher energy eigenvalues are denoted tion, resulting in a p-to-d band gap. If ⌬ϾU, on the nϪ1 by E (NϪ1) and E (NϪ1), respectively. These final other hand, the d configuration would lie below the Ϫ ϩ n states correspond to the main feature and the satellite of d Lគ configuration, leading to a d-to-d band gap. the photoemission spectrum, respectively. Then the Zaanen, Sawatzky, and Allen (1985) and Hu¨ fner (1985) d-derived photoemission spectrum (the ␻Ͻ␮ part of the extended this view of electronic structure to a wide d spectral function) is given by range of 3d transition-metal compounds, including light transition-metal (Ti,V,...) compounds. For ⌬ϽU, the 2 ␳d͑␻͒ϭ͉cos ␪ cos ␾ϩsin ␪ sin ␾͉ band gap is of the p-d type; the p band is located be- ϫ␦ ␻ϩE ͑NϪ1͒ϪE ͑N͒ tween the upper and lower Hubbard bands in the spec- „ Ϫ g … tral function, as shown in Fig. 5(b). The gap is thus ϩ͉Ϫcos ␪ sin ␾ϩsin ␪ cos ␾͉2 called a charge-transfer gap and the compound is called a charge-transfer insulator. It has a band gap of magni- ϫ␦ ␻ϩE Nϩ1 ϪE N , (3.12) „ ϩ͑ ͒ g͑ ͒… tude ϳ⌬.If⌬ϾU, on the other hand, the band gap is of where the first term gives a peak at the lower binding the d-d type as in the original Hubbard picture [Fig. energy ͓␮Ϫ␻ϭEϪ(NϪ1)ϪEg(N)ϩ␮͔ and the second 5(a)]. The gap is then called a Mott-Hubbard gap and term at the higher binding energy ͓␮Ϫ␻ϭEϩ(NϪ1) the compound is called a Mott-Hubbard insulator. It has ϪEg(N)ϩ␮͔. The ratio between the spectral weight a band gap of magnitude ϳU. Figure 41 shows a two- (Iϩ) of the higher final state ⌽ϩ(NϪ1) and that (IϪ)of dimensional plot, the so-called Zaanen-Sawatzky-Allen the lower final state ⌽Ϫ(NϪ1) is given by Iϩ /IϪ plot, of the transition-metal compounds according to the ϳtan2(␪Ϫ␾) according to Eq. (3.12). Because the differ- relative magnitudes of U and ⌬. The straight line U ence between ␪ and ␾ increases with decreasing ͉T͉ or ϭ⌬ separates the U-⌬ space into Mott-Hubbard and with increasing U, spectral weight transfers from the charge-transfer regimes. It should be noted, however, lower to the higher binding energy peak with increasing that because of hybridization between the d and p U/͉T͉. states, the boundary Uϭ⌬ is not a sharp phase bound- The above spectral weight transfer may be viewed in a ary but is a crossover line at which the character of the different way if we refer to the basis functions of fixed band gap changes continuously. local configurations and regard the p-d transfer integrals Whether doped carriers in a ‘‘filling-control’’ system as a perturbation. In this picture, T causes spectral have oxygen p character or transition-metal d character weight transfer in the opposite direction, namely, from is an important question in considering the transport higher to lower energies. By increasing ͉T͉, the differ- properties of the transition-metal oxides. Since the ence between ␪ and ␾ becomes small, resulting in an ground state of the N-electron state is usually domi- 2 n increase in Iϩ /IϪϳtan (␪Ϫ␾). Here, we point out that nated by the d configuration, the doped holes will have the p-d transfer integrals effectively increase due to mainly d character if the lowest (NϪ1)-electron state is electron correlation, i.e., in going from the one-electron dominated by the dnϪ1 configuration and will have Hartree-Fock picture to the many-body CI picture. If mainly p character if it is dominated by the dnLគ configu- one treats the cluster model in the Hartree-Fock ap- ration. That is, doped holes have mainly d character in a proximation, the one-electron Hamiltonian has off- Mott-Hubbard insulator (⌬ϾU) and p character in a

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1127

added to the above Egap , where ␦(N) is the decrease in energy of the lowest N-electron state due to p-d hybrid- ization. The effect of the finite p bandwidth Wp can be incor- porated by replacing the ligand molecular orbitals by a continuum of the p band (Zaanen et al., 1985). The problem should then be formulated using the Anderson impurity model instead of the cluster model. The Hamil- tonian of this model is given by 1 ϭ ␧ n ϩ ␯␯ ␮␮ d† d†d d ͚ d␯ d␯ ͚ ͗ Ј͉͉ Ј͘ ␯Ј ␯ ␮ ␮Ј H ␯ 2 ␯␯Ј␮␮Ј

† ϩ ␧pknpkϪ ͑T␯kd␯pkϩH.c.͒, (3.13) ͚k ͚␯k where ␯, ␮, etc., denote the spin and orbital of the d * 2 FIG. 41. Zaanen-Sawatzky-Allen plot for classifying 3d states and ͗␯␯Ј͉͉␮␮Ј͘ϵ͐dr͐drЈ␾␯*(r)␾␯ (rЈ)e /(͉r † † Ј transition-metal compounds. The straight line Uϭ⌬ separates ϪrЈ͉) ␾␮(r)␾␮Ј(rЈ). d␯ and pk are the creation opera- the Mott-Hubbard regime (A) and the charge-transfer regime tors of localized d and bandlike p electrons, respec- (B). Between the two regimes is located an intermediate re- tively. This Hamiltonian can be rewritten as gime (AB). The shaded areas near Uϭ0 and ⌬ϭ0 are metal- lic regions where a Mott-Hubbard-type or charge-transfer-type 1 † † ϭ ␧ n ϩ ͗␯␯Ј͉͉␮␮Ј͘d d d d band gap is closed. From Zaanen, Sawatzky, and Allen, 1985. ͚ d␯ d␯ ͚ ␯Ј ␯ ␮ ␮Ј H ␯ 2 ␯␯Ј␮␮Ј charge-transfer insulator (⌬ϽU). In either case, the ␧pϪ Wp/2 ϩ ␧n ͑␧͒d␧ doped electrons have mainly d character because the ͵ p ␧pϪ Wp/2 lowest unoccupied states are dominated by the dnϩ1 configuration, since the dnϩ1 dnϩ2Lគ charge-transfer ␧pϪ Wp/2 Ϫ d␧͓T ͑␧͒d†p͑␧͒ϩH.c.͔, (3.14) energy ⌬ϩU is always a large→ positive number. ͚ ͵ ␯ ␯ ␯ ␧pϪ Wp/2 In going from lighter to heavier transition-metal ele- ments, the d level is lowered and therefore ⌬ is de- where creased, whereas U is increased because the spatial ex- 2 * tent of the d orbital shrinks with atomic number. This ͉T␯͑␧͉͒ ␦␯␯ ϵ T␯kT␯ k␦͑␧pkϪ␧͒. (3.15) Ј ͚k Ј means that early transition-metal compounds tend to be- come Mott-Hubbard type and late transition-metal com- The ground state of the formally d8 impurity embedded pounds tend to become charge-transfer type. When the in the filled p-band continuum is given by ligand atom becomes less electronegative, the p level is ␧pϩ Wp/2 raised and hence ⌬ is decreased. For instance, transition- 8 9 ⌽g͑N͒ϭA ͉d ͘ϩ ͵ a͑␧͉͒d Lគ ͑␧͒͘d␧ϩ¯ , metal chalcogenides are more likely to become charge- ͫ ␧pϪ Wp/2 ͬ transfer type than transition-metal oxides. Compounds (3.16) to be plotted in Fig. 41 indeed follow this chemical where ␧ is the energy of an electron in the ligand p trend, where Ti and V oxides are classified as typical band. By inserting Eq. (3.16) into the Anderson impu- Mott-Hubbard-type compounds and Cu oxides as typi- rity Hamiltonian (3.14) [neglecting the ͉d10Lគ (␧)2͘ cal charge-transfer compounds. Systematic variation of term], we obtain an equation the electronic structure is described below in more de- ␧ ϩ W /2 2 p p 2͉T␯͑␧͉͒ d␧ tail. ␦ϭ , (3.17) ͵ Ϫ ϩ ϩ ϩ The diagram of Fig. 41 contains a metallic region ␧pϪ Wp/2 ␦ ␧ ␧p ⌬ ⌬E8 (shaded area) near the ⌬- and U-axes, where the mag- where ␯ is one of the eg orbitals and &T␯(␧) nitude of the band gap, ϳ⌬ or ϳU, becomes small. ⌬ 8 9 2 ϵ͗d ͉ ͉d Lគ (␧)͘, which satisfies Tpd and U are defined with respect to the center of the p and ␧pϩHWp/2 2 ϭ͐ ͉T␯(␧)͉ d␧. The lowest-energy solution of d bands and therefore the actual band gap is smaller ␧pϪWp/2 than ⌬ or U by the half widths of the p and d bands: If Eq. (3.16) yields ␦ϭ␦(N). Figure 42(a) shows that p-d for simplicity, one ignores hybridization between the hybridization lowers the energy of the d8 multiplet com- 3 metal d and ligand p orbitals and multiplet effects, the ponent (the A2g term in the case of NiO). In the (N 1 8 magnitude of the band gap is given by Egapϳ⌬Ϫ 2 (Wp Ϫ1)-electron state, although the d Lគ continuum lies be- 7 ϩWd) for a charge-transfer insulator and Egapϳ⌬ low the discrete d state, strong p-d hybridization cre- 8 ϪWd for a Mott-Hubbard insulator, where Wd and Wp ates a split-off state of predominantly d Lគ character be- are the bandwidths of d and p, respectively (see Fig. 5). low the d8Lគ continuum, as shown in Fig. 42(b). The If the energy shifts due to p-d hybridization are taken diagram of Fig. 41 was indeed calculated using such an into account, 2␦(N)Ϫ␦(NϪ1)Ϫ␦(Nϩ1) should be Anderson impurity model for the N-, (NϪ1)-, and (N

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1128 Imada, Fujimori, and Tokura: Metal-insulator transitions

TABLE II. Relation between Kanamori parameters, u, uЈ, and j, Racah parameters, A, B, and C, and Slater integrals, F0 , F2 , and F4 . Here rϾ and rϽ denote, respectively, the larger and smaller ones of r1 and r2 .

AϭF0Ϫ49F4 ,BϭF2Ϫ5F4 ,Cϭ35F4 7 14 7 F0ϭAϩ 5 C,UϭAϪ 9 Bϩ 9 C uϭAϩ4Bϩ3CϭF0ϩ4F2ϩ36F4 uЈϭAϪBϩCϭF0ϪF2Ϫ4F4 5 5 45 jϭ 2 BϩCϭ 2 F2ϩ 2 F4 0 1 2 1 4 F0ϭF ,F2ϭ 49 F ,F4ϭ 441 F k ϱ 2 ϱ 2 k kϩ1 2 2 F ϭ͐0 r1dr1͐0 r2dr2(rϽ/rϾ )R3d(r1) R3d(r2)

The energy E(dn) given by Eq. (3.2), where multiplet effect is neglected, yields the center of gravity of the multiplet terms of the dn configuration. If we on the multiplet splitting in the above energy-level scheme, as shown in Fig. 43, each multiplet term 2Sϩ1⌫ is shifted from the center of gravity E(dn) [Eq. (3.2)] by a ‘‘multiplet correction’’ ⌬E(dn: 2Sϩ1⌫). Then the energy of each multiplet term is given by E(dn: 2Sϩ1⌫)ϭE(dn)ϩ⌬E(dn: 2Sϩ1⌫). For the lowest term of the dn multiplet, we denote the multiplet correc- n 2Sϩ1 tion by ⌬E(d : ⌫)ϵϪ⌬En(Ͻ0). The charge- transfer energy ⌬ and the d-d Coulomb energy U can alternatively be defined using the lowest multiplet of each configuration instead of the center of gravity as described in Fig. 43. We denote the parameters thus re- defined by ⌬eff and Ueff . Most of the physical properties

FIG. 42. Hybridization between discrete and continuum states in the impurity mode for a charge-transfer insulator: (a) Hy- bridized d8 and d9Lគ configurations in the N-electron state for NiO; (b) d7 and d8Lគ configurations in the (NϪ1)-electron state.

ϩ1)-electron states (and assuming Wpϭ3T). Because the effect of the finite d bandwidth cannot be included in the impurity model, a finite d bandwidth Wdϭ0.5T was assumed to produce the metal-insulator phase boundary shown in Fig. 41.

4. Multiplet effects Multiplet splitting of the energy levels of a transition- metal ion is of a magnitude comparable to ⌬, U, and T and is therefore important in characterizing electronic structure. Intra-atomic exchange and anisotropic Cou- lomb interactions are the cause of multiplet splitting and are represented by Racah parameters (B and C)or Slater integrals (F2 and F4). [The Racah parameter A or the Slater integral F0 represents the isotropic part of the Coulomb interaction and therefore is close to U, which is a multiplet averaged value defined in Eq. (3.2).] Defi- nitions of these parameters and the interrelation be- FIG. 43. Definition of ⌬, U, ⌬eff , Ueff , and ⌬En . From Fuji- tween them are given in Table II. mori et al., 1993.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1129

i.e., calculated in the Hartree-Fock approximation (Brandow, 1977). Here, for simplicity, Kanamori’s pa- rameters u, uЈ, and j are used instead of the more ac- curate Racah parameters or Slater integrals: u ϵ͗␯␯͉H͉␯␯͘ϵU␯␯ , uЈϵ͗␯␯Ј͉H͉␯␯Ј͘ϵU␯␯Ј [Eq. (2.6d)] and jϵ͗␯␯Ј͉H͉␯Ј␯͘ϵJ0␯␯Ј [Eq. (2.6e)] (␯Þ␯Ј), where ␯ and ␯Ј denote different d orbitals. u and uЈ are Cou- lomb integrals and j is an exchange integral between d electrons. The relationship between Kanamori’s param- eters, Racah parameters, and Slater integrals is summa- rized in Table III. The multiplet corrections to ⌬ and U modify the mag- nitudes of the band gaps as follows. From the behavior of ⌬effϪ⌬ and UeffϪU shown in Fig. 44: (i) for a charge- transfer insulator, the band gap is reduced for nр4 whereas it increases for nу5, with the most pronounced effect occurring at nϭ4 and 5; (ii) for a Mott-Hubbard insulator, the band gap greatly increases at nϭ5 and is FIG. 44. Multiplet corrections ⌬ Ϫ⌬, U ϪU, and (U slightly reduced otherwise. A dramatic manifestation of eff eff eff the multiplet effect is seen, for example, in Fe and Mn Ϫ⌬eff)Ϫ(UϪ⌬) for the high-spin states as functions of the d-electron number n. Racah parameters of divalent ions are oxides, in which n is changed around 4 and 5 when the used. From Fujimori et al., 1993. transition-metal atomic number and valence are varied: 3ϩ 5 LaFeO3 (Fe : d ) is a wide-gap insulator because n 4ϩ 4 ϭ5. For nϭ4, SrFeO3 (Fe : d ) is a metal (Bocquet of 3d transition-metal compounds are determined by 4ϩ 4 et al., 1992b) and CaFeO3 (Fe : d ) has a small gap ⌬eff and Ueff rather than by ⌬ and U, since the lowest that is closed at high temperature or under high pressure multiplet terms contribute most significantly to the (Takano et al., 1991). As for Mn compounds, the band ground state and the low-energy excited states of the 3ϩ 4 gap of LaMnO3 (Mn : d ) is smaller than that of system, including the magnitude of the band gap. Using 4ϩ 3 SrMnO3 (Mn : d ) (van Santen and Jonker, 1950); the multiplet correction ⌬En , we obtain MnO (Mn2ϩ: d5) is a well-known wide-gap insulator. ⌬effϭ⌬ϩ⌬EnϪ⌬Enϩ1 , (3.18) Multiplet splitting also affects the character of the band gap and hence the character of doped carriers be- UeffϭUϩ2⌬EnϪ⌬EnϪ1Ϫ⌬Enϩ1 . (3.19) cause now the relative magnitudes of ⌬eff and Ueff , in- In a free ion, the lowest multiplet term of each configu- stead of ⌬ and U, determines the character of doped nϪ1 ration is always a Hund’s-rule high-spin state. This is not holes and electrons. UeffϪ⌬eff ͓ϵϪ⌬eff(d )͔ may be always the case in solids because p-d hybridization may taken as a measure of the p character of the top of the lower the energy of the low-spin state below the high- valence band, i.e., a measure of the p-hole character of spin state. We restrict the following argument to those doped holes. Multiplet corrections to UϪ⌬ [i.e., Ueff cases in which the high-spin state is the lowest term of Ϫ⌬effϪ(UϪ⌬)] are also shown in Fig. 44. From the sign each multiplet. of the correction, one can see that the multiplet effect As we shall show below, ⌬ and U are smooth func- tends to increase the p-hole character for nр5 whereas tions of the atomic number Z and valence v of the it tends to increase the d-hole character for nу6; the transition-metal ion, whereas ⌬eff and Ueff are appar- maximum of the p-hole character occurs at nϭ5. In- ently irregular functions of these variables (Bocquet deed, it was found from the O 1s x-ray absorption spec- et al., 1992c). The irregularity arises from the fact that troscopy (XAS) study of La1ϪxSrxFeO3 that the p-hole the multiplet correction ⌬En is a strongly nonmonotonic character of holes doped into LaFeO3 is particularly function of the nominal d-electron number n, showing a pronounced (Abbate et al., 1992). Another measure of maximum at nϭ5, i.e., at half filling. Figure 44 shows the the character of doped holes is the difference between multiplet corrections for ⌬ and U as functions of n: the undoped ground state and the hole-doped state, i.e, ⌬effϪ⌬ (ϵ⌬EnϪ⌬Enϩ1) and UeffϪU (ϵ2⌬En the difference between the weight of the ligand hole Lគ Ϫ⌬EnϪ1Ϫ⌬Enϩ1). ⌬effϪ⌬ increases with n except for a character in the undoped state and that in the hole- sharp drop from nϭ4tonϭ5, resulting in a minimum at doped state. Since the p-hole character in the ground nϭ4 and a maximum at nϭ5. UeffϪU exhibits a sharp state decreases with ⌬eff , i.e., increases with Ϫ⌬eff , and maximum at nϭ5 and otherwise is a small negative that in the doped state increases with UeffϪ⌬eff , their number. In Fig. 44, the solid circles yield energies of full difference UeffϪ⌬effϪ(Ϫ⌬eff)ϭUeff may serve as a crude multiplet calculations using Racah parameters whereas measure of the p-hole character of doped holes. From the open circles yield energies calculated assuming a this viewpoint, the maximum p-hole character should single Slater determinant of the type occur at nϭ5 (see Fig. 44). In the case of electron dop- ing, the difference between the p-hole (dnϩ1Lគ ) charac- mЈ mЉ nЈ nЉ ⌽͑N͒ϭ͉t2g t2g eg eg ͉, (3.20) ter in the ground state (measured by Ϫ⌬ ) and that in ↑ ↓ ↑ ↓ eff

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1130 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 45. Modified Zaanen-Sawatzky-Allen diagram including the effect of strong p-d hybridization in the region ⌬ϳ0 (Mi- zokawa et al., 1994). In this region, the gap increases with in- creasing p-d transfer or decreasing p bandwidth wp .

nϩ2 the doped state (negligibly small d Lគ character) FIG. 46. Schematic picture of the one-electron density of yields the p-electron character of the doped electrons. states of the negative-⌬ insulator (Mizokawa et al., 1994). Therefore one can conclude from Fig. 44 that nϭ5 com- pounds have maximum d character for doped electrons and that nϭ4 compounds have maximum p character pendent on the p-d hybridization strength T. Therefore, as the ligand bandwidth Wp decreases or the transfer for doped electrons. From this we expect that SrFeO3 (Abbate et al., 1992; Bocquet et al., 1992b) would ac- integral T increases, the band gap tends to increase. In commodate doped electrons predominantly in the ligand Fig. 45 are plotted sets of U and ⌬ which give Egap p orbitals to form the stable d5 configuration. ϭ0.4 eV as a measure of the metal-insulator phase boundary. The effect of lattice periodicity is much more compli- 5. Small or negative charge-transfer energies cated in the case of negative ⌬ than in the case of posi- In a charge-transfer insulator, if the charge-transfer tive ⌬. In the periodic lattice, d orbitals at neighboring energy ⌬ or ⌬eff is small compared to T, hybridization transition-metal atoms share intervening anion p orbit- between the dn and dnϩ1Lគ configurations becomes als, which leads to a complicated many-body problem, strong and the contribution of the dnϩ1Lគ configuration especially for a small or negative ⌬. This will effectively to the ground state becomes substantial. If ⌬ or ⌬eff decrease the 3d-ligand hybridization strength T. Also, Ͻ0, the weight of the dnϩ1Lគ configuration becomes the dispersional widths of the (NϪ1)- and dominant in the ground state. This situation is expected (Nϩ1)-electron states resulting from intercluster hy- to occur when the valence of the transition-metal ion is bridization would effectively increase the ligand band- unusually high because ⌬ is expected to decrease with width Wp . Therefore intercluster hybridization effec- nϩ1 metal valence. The d Lគ configuration is a continuum. tively decreases T or increases Wp , thereby shrinking However, within the Anderson impurity model, it can be the insulating region in the U-⌬ phase diagram. The shown that sufficiently strong p-d hybridization causes a strength of the intercluster hybridization is dependent discrete state of dnϩ1Lគ character to be split off below on the metal-oxygen-metal bond angle (see Sec. III B): the continuum in the same way as in the formation of a It is maximized at 180° and minimized at 90°. This may split-off state in the (NϪ1)-electron state of a charge- explain why LaCuO3, in which octahedral CuO6 clusters transfer insulator described above [see Fig. 42(b)]. The are interconnected with ϳ180° Cu-O-Cu bond angles, is insulating region was restricted to positive ⌬ values in metallic and NaCuO2, in which square-planar CuO4 the original Zaanen-Sawatzky-Allen diagram (Fig. 41) clusters are interconnected with ϳ 90° Cu-O-Cu angles, because it was assumed that the p-d hybridization is insulating, although the Cu is formally trivalent in 1 strength was small Tϭ 3 Wp (Zaanen et al., 1985). In or- both compounds (Mizokawa et al., 1991). For a negative der to extend the Zaanen-Sawatzky-Allen framework to ⌬, the lowest (NϪ1)-electron state is dnϩ1Lគ 2 [because small and negative ⌬ values, the magnitude of the band E(dnϩ1Lគ 2)ϭE(dnLគ )ϩ⌬ϽE(dnLគ ); see Eq. (3.2)] and gaps has been calculated using the Anderson impurity the lowest (Nϩ1)-electron state is dnϩ1. Since the main model for various sets of ⌬, U, T ͓ϵϪ)(pd␴)͔, and constituent of the ground state is the dnϩ1Lគ configura- nϩ1 nϩ1 Wp as shown in Fig. 45 (Mizokawa et al., 1994), where tion, the charge transfer (d Lគ )iϩ(d Lគ )j nϩ1 2 nϩ1 one can see that the insulating region is extended to (d Lគ )iϩ(d )j gives the lowest-energy charge negative ⌬ values for certain parameters. In the small or fluctuation→ and therefore the band gap has predomi- negative ⌬ regime, the magnitude of the band gap is nantly p-to-p character. This situation is illustrated in proportional neither to ⌬ nor to U but is sensitively de- Fig. 46.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1131

vϩ The fact that NaCuO2 is a nonmagnetic insulator (S Using the ionization potential I(M ) and the elec- ϭ0) was explained by the low-spin (Sϭ0) solution of tron affinity A(Mvϩ) of the transition-metal Mvϩ ion, the CI cluster model, whereas band-structure calcula- Ueff is given by tions using the local-density approximation (LDA; e2 Karlsson et al., 1992; Singh, 1994) have also given a non- vϩ vϩ UeffϭI͑M ͒ϪA͑M ͒Ϫ , (3.21) magnetic insulating ground state. Therefore it is not ob- dMϪM vious whether NaCuO2 is a correlated insulator or a 2 where e /(dMϪM) is the attractive Coulomb energy be- band insulator. According to the configuration- tween nearest-neighbor metal ions that are separated by interaction picture, the nonmagnetic ground state domi- d . Here, the inclusion of the e2/(d ) term cor- nated by the d9Lគ configuration may be viewed as a MϪM MϪM responds to the optical excitation, where an electron and valence-bond state in which the local spin of the d hole a hole are created on nearest-neighbor cation sites. (d9) and that of the ligand hole (Lគ ) form a local singlet. Therefore the negative ⌬ insulator with low-spin con- Without the latter term, Ueff would correspond to pho- figuration may be called a ‘‘local-singlet lattice’’ or toemission and inverse photoemission, where the elec- ‘‘valence-bond (Heitler-London) insulator.’’ In this pic- tron and hole are well separated from each other and do ture, the wave function of the ground state is quite dif- not interact. Correspondingly, ⌬eff is given by ferent from a single Slater determinant without electron e2 correlation. Such a lattice consisting of the local singlets ⌬ ϭe⌬V ϩI͑O2Ϫ͒ϪA͑Mvϩ͒Ϫ , (3.22) eff Mad d is, however, analytically connected to the nonmagnetic MϪO band insulator. Because the origin of the band gaps in where ⌬VMad is the difference in the Madelung potential insulators with small or negative ⌬ is neither U nor ⌬ between the metal site and the oxygen site and dMϪO is but rather the strong p-d hybridization. Sarma and co- the nearest-neighbor metal-oxygen distance. The defini- workers referred to them as ‘‘covalent insulators’’ tion of ⌬eff also contains the electron-hole interaction (Nimkar et al., 1993). According to their Hartree-Fock term relevant to optical transitions. band-structure calculations on the p-d model, the char- In Fig. 47 are plotted values for ⌬eff and Ueff esti- acter of the band gap is of strongly hybridized pd-p type mated using Eqs. (3.21) and (3.22). These ⌬eff and Ueff rather than pure p-p type. On the other hand, it is closer values have been calculated using the bare electrostatic to the p-p type according to the LDA calculation of potential without screening and are therefore expected Singh (1994). to overestimate the actual parameter values. Indeed, they range from Ϫ10 eV to ϩ20 eV, extending over a much wider energy range than parameter values esti- B. Systematics in model parameters and charge gaps mated using first-principles methods or the spectro- To understand the diverse physical properties of 3d scopic method. One can see from Fig. 47 that, although transition-metal compounds from a unified point of the absolute magnitudes of ⌬eff and Ueff are overesti- view, it is important to clarify the systematics underlying mated, the qualitative trends are the same as those esti- changes in the electronic structure and hence in the mated using spectroscopic methods. The charge-transfer model parameters, ⌬, U, and T as functions of chemical type (⌬effϽUeff) versus Mott-Hubbard type (⌬effϾUeff) environment, namely, as functions of the atomic number seems almost correctly predicted by this method: Late and valence of the transition-metal ion, the nature (elec- transition-metal compounds fall into the charge-transfer tronegativity, ionic radius, etc.) of the ligand atom, and regime and early transition-metal compounds into the the crystal structure. In the following, we present three Mott-Hubbard regime with some V compounds with different approaches to this task, namely, (1) methods high valences in the charge-transfer regime. According based on the ionic crystal model, (2) semiempirical to the Zaanen-Sawatzky-Allen picture, a charge-transfer insulator becomes metallic when the charge-transfer gap methods based on the analysis of photoemission spectra, 1 and (3) theoretical methods using first-principles elec- of magnitude ⌬effϪ 2(WpϩWd) becomes negative, while tronic structure calculations. a Mott-Hubbard-type insulator becomes metallic when the Mott-Hubbard-type gap of magnitude UeffϪWd be- comes negative. The vertical (⌬ ϭ⌬ ) and horizontal 1. Ionic crystal model eff B (UeffϭUB) lines in Fig. 47 are drawn in an attempt to The simplest method of estimating values for the pa- separate metals from insulators. ⌬Bϳ10 eV and UB rameters ⌬ and U is the ionic crystal model, which was ϳ11 eV would mean that WpϳWdϳ10 eV. The latter applied to 3d transition-metal oxides by Torrance et al. values are, however, unrealistically large and would con- (1991). Assuming that each atom is a point charge hav- tain other contributions (such as incomplete screening of ing its formal ionic charge, they calculated the bare the bare electrostatic potentials in the ionic model). Madelung potential energy at each atomic site and esti- The band gaps predicted by the ionic model were mated ⌬eff and Ueff using the ionization potentials and compared with the optical band gaps of perovskite-type electron affinities of the constituent ions. Here, the mul- RMO3 compounds, where R is La or Y, by Arima, tiplet corrections are included in the parameters thus Tokura, and Torrance (1993). The experimentally deter- estimated since the experimental ionization potentials mined optical gaps are shown in Fig. 48. The figure are used. shows that the lowest-energy gap is of the Mott-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1132 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 49. Experimental optical gaps vs charge-transfer (open symbols, ⌬) or Mott-Hubbard (solid symbols, U) gaps calcu- lated using the ionic model. From Arima, Tokura, and Tor- rance, 1993.

Hubbard type for MϭTi and V and is of the charge- transfer type for MϭMn to Cu. The magnitude of the gap shows a maximum at nϭ5(MϭFe), due to the maximum stabilization of the completely filled t2g and ↑ eg levels as shown in Fig. 44. The peak at nϭ3(M ↑ ϭCr) is due to the complete filling of the t2g level. The dip at nϭ4(MϭMn) seen in Fig. 48 can also↑ be under- stood from the multiplet correction (Fig. 44). A com- parison of the optical gap values with those calculated using the ionic model is made in Fig. 49, where one can see excellent correlation between the experimental and FIG. 47. Charge-transfer energies ⌬eff and on-site d-d Cou- the calculated values except for the ϳ10 eV overesti- lomb energies U of transition-metal oxides calculated using eff mate of the calculated values. The ϳ10 eV difference the ionic model by Torrance et al. (1991). The vertical (⌬ eff should have the same origin as the metal-insulator ϭ⌬B) and horizontal (UeffϭUB) lines are drawn to separate metals (solid symbols) and insulators (open symbols). The boundary ⌬effϭ⌬B and UeffϭUB noticed in Fig. 47. number next to each element gives its oxidation state, e.g., There are slightly different methods of estimating the parameters U and ⌬ using the ionic crystal model. Ohta Fe3: ␣-Fe2O3 and LaFeO3. et al. (1991a, 1991b) assumed that the Madelung term e⌬VMad is screened by the optical dielectric constant ␧ϱ : e⌬VMad /␧ϱ . Zaanen and Sawatzky (1990) argued that the dielectric constant, which describes screening between charges separated by many lattice spacings, is not appropriate for describing the screening process considered here. Instead they added the polarization en- ergies of surrounding atoms Ϫ2Epol and Ϫ4Epol to Eqs. (3.21) and (3.22), respectively. Here, Epol is the polariza- tion energy for a unit point charge Ϯe in the crystal and 1 2 is given by Epolϭ 2 ͚j␣jFj , where Fj is the electric field produced by the point charge at ion j and ␣ is the po- larizability of ion j. The Racah parameter A and charge- transfer energy ⌬ estimated by this method for various 3d transition-metal monoxides are listed in Table III. Here, in order to estimate Epol , parameters for NiO have been determined so as to fit the photoemission 4 spectra, and the empirical 1/d dependence of Epol on the interatomic distance d has been used. FIG. 48. Measured optical gaps for RMO3 (RϭLa or Y) plot- ted against the transition-metal atom (Arima, Tokura, and 2. Spectroscopic methods Torrance, 1993). The open symbols give charge-transfer (p-to- d) gaps, while the solid symbols give Mott-Hubbard (d-to-d) The valence-band photoemission spectra of a number gaps. of 3d transition-metal compounds have so far been ana-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1133

TABLE III. On-site Coulomb (U) and exchange (J) energies estimated using the constrained LDA method (Anisimov et al., 1991), compared with the empirical estimates for the Racah parameter A and charge-transfer energy ⌬ (Zaanen and Sa- nϪ1 nϩ1 watzky, 1990). Ueff is the splitting between the d and d high-spin states. Energies are in eV.

UJUeff A ⌬

CuO/CaCuO2 7.5 0.98 6.5 4.0 4.0 NiO 8.0 0.95 7.1 (6.0) (4.92) CoO 7.8 0.92 6.9 5.02 5.53 FeO 6.8 0.89 5.9 4.46 6.21 MnO 6.9 0.86 10.3 5.43 8.23 VO 6.7 0.81 5.9 3.54 10.47 TiO 6.6 0.78 5.8 3.02 10.01 lyzed using the CI cluster model and the Anderson im- purity model (Fujimori et al., 1986, 1990; Eskes et al., 1990; van Elp et al., 1991) and the model parameters ⌬, U, and T have been estimated. Having obtained N- and (NϪ1)-electron eigenstates, one calculates the photo- emission spectrum by

2 ␳͑␻͒ϭ ͉͗⌽i͑NϪ1͉͒d͉⌽g͑N͉͒͘ ͚i

ϫ␦„␻ϩEi͑NϪ1͒ϪEg͑N͒…, (3.23) where ⌽g(N) and ⌽i(NϪ1) are the ground state and the final states of photoemission, Eg(N) and Ei(NϪ1) are their energies, d is the annihilation operator of a d electron and ␻ϵ␧kinϪh␯. Therefore ␮Ϫ␻, where ␮ is the electron chemical potential, i.e., the Fermi level, is FIG. 50. Valence-band photoemission spectra of CuO (Eskes the electron binding energy measured from the Fermi et al., 1990), NiO (Fujimori and Minami, 1984), MnO (Fuji- mori et al., 1990) and VO (Werfel et al., 1981) and their level and ␳(␻) with ␻Ͻ␮ gives a photoemission spec- cluster-model analyses. ⌬ϳ2.75 eV and Uϳ6.5 eV for CuO, trum. The summation in Eq. (3.23) runs over all final ⌬ϳ4.0 eV and Uϳ7.5 eV for NiO, and ⌬ϳ7.0 eV and U states that can be reached by the annihilation of one d ϳ7.5 eV for MnO. VO is metallic and no cluster-model analy- electron. By inserting Eqs. (3.1) and (3.6) into Eq. 2 sis could be made. (3.23), one obtains ␳(␻)ϭ͚i͉a1bi1ϩa2bi2ϩ¯͉ ␦„␻ ϩEi(NϪ1)ϪEg(N)…. When one compares the calcu- transfer-type CuO and NiO, and UϽ⌬ for the Mott- lated spectra with experiment, the delta function in Eq. Hubbard-type VO (although this compound is metallic). (3.23) is replaced by a Gaussian function (broadened by MnO with Uϳ⌬ is intermediate between the charge- a Lorentzian function) in order to represent the finite transfer-type and Mott-Hubbard-type compounds. bandwidths, lifetime widths, and other broadening ef- The photoemission spectra of the core levels of the 3d fects. In addition, the oxygen p band is superposed on transition-metal atom also yield nearly the same infor- the d-electron spectrum. mation as the valence-band spectra (van der Laan et al., Figure 50 shows the photoemission spectra and the 1981). The core-level spectra have an advantage over best-fit results for a series of transition-metal monox- the valence-band spectra in that there are no overlap- ides. Since CuO and NiO are charge-transfer insulators, n ping features, such as the O 2p band, that obscure the the main band closer to the Fermi level is due to d Lគ cluster-model analysis, while they have the disadvantage final states and the satellite at higher binding energies is nϪ1 that the unknown strength of the core-hole potential in- due to d final states. The separation between the troduces an additional parameter. The final state of main and satellite features is given by ϳUϪ⌬ (plus core-level photoemission is given by shifts due to p-d hybridization). In going from heavier n nϩ1 to lighter transition-metal atoms, the separation be- ⌽͑cគ N͒ϭe1͉cគ d ͘ϩe2͉cគ d Lគ ͘ϩ¯ (3.24) tween the main band and the satellite decreases, mean- and the photoemission spectrum by ing that UϪ⌬ decreases. The cluster-model analysis shown in Fig. 50 yields a systematic increase of ⌬ with 2 ␳͑␻͒ϭ ͉a1ei1ϩa2ei2ϩ¯͉ ␦„␻ϩEi͑cគ N͒ϪEg͑N͒… decreasing atomic number, while the systematic de- ͚i crease of U is not so significant. UϾ⌬ for the charge- (3.25)

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1134 Imada, Fujimori, and Tokura: Metal-insulator transitions

A systematic analysis of the metal 2p core-level spec- in the Zaanen-Sawatzky-Allen diagram (Fig. 41) but tra for a wide range of 3d transition-metal compounds with U/Teff and ⌬/Teff shown in the plot instead of U/T was attempted by Bocquet et al. (1992a; 1996) using a and ⌬/T. Most of the V and Ti oxides are found near simplified version of the CI cluster model, where wave the origin (U/Teffϳ0 and ⌬/Teffϳ0) because of the large functions of single Slater determinants [Eq. (3.20)] and Teff values. The parameter values shown in Fig. 54 indi- Kanamori’s parameters were employed for both the cate that the ⌬ of compounds with high metal valencies, ground state and the final states. The parameter values such as Ti4ϩ,V4ϩ, and V5ϩ, is comparable to or lower thus obtained are not exactly the same as, but are rea- than U. sonably close to, those obtained from analysis of the If there were no p-d hybridization and the band- valence-band photoemission spectra. More elaborate widths (Wp and Wd) were zero, the band gap would be full multiplet cluster-model calculations have been made given by ⌬eff for a charge-transfer insulator and by Ueff for the transition-metal dihalides MX2 and trivalent for a Mott-Hubbard insulator. Values for ⌬eff and Ueff transition-metal oxides M2O3 by Kotani and co-workers deduced from the cluster-model analysis of the metal 2p (Okada et al., 1992; Uozumi et al., 1997). The ⌬ values core-level spectra are plotted for various transition- thus obtained for M2O3 are generally smaller than those metal compounds in Fig. 55 as functions of n. Figure for the monoxides MO but show a tendency to increase 55(a) demonstrates that ⌬eff decreases monotonically with decreasing transition-metal atomic number, as in with n except for the large discontinuity between nϭ4 the case of the monoxides. Thus ⌬ seems to decrease and 5. This variation reflects both the smooth variation systematically with atomic number Z and valence v of of ⌬ with n and its multiplet correction, ⌬effϪ⌬ (Fig. 44), the transition-metal ion. This tendency is clearly seen in which is a large negative value for nϭ4 and a large posi- Fig. 51, where ⌬ values deduced from the 2p core-level tive value for nϭ5. Figure 55 also shows that, for a fixed spectra are plotted as a function of Z and v. In Fig. 52 n, ⌬eff (and hence ⌬) decreases as the valence of the are plotted U values, which weakly increase with the metal ion increases, reflecting the lowering of the 3d atomic number and valence of the transition-metal due energy level due to the increasing positive charges at the to shrinkage of the spatial extent of the 3d orbitals. Fig- transition-metal ion. Figure 55(b) shows that Ueff slowly ure 52 also shows that the decrease in the electronega- and monotonically increases with n except for the sharp tivity of the ligand atom (in going from oxygen to sulfur) peak at nϭ5. Calculated band gaps are shown in Fig. 56. decreases the ⌬ value. The systematic variation of ⌬ and Due to p-d hybridization, the d-d and p-d characters of U with Z and v demonstrated above may be very the band gaps are mixed with each other and the varia- crudely written as tion of the band gap with n reflects both that of ⌬eff and ⌬ϳ⌬0Ϫ0.6ZϪ2.5v, (3.26) that of Ueff . The p-d hybridization weakens the degree of variation of the band gap compared to that of ⌬eff or UϳU ϩ0.3Zϩ0.5v, (3.27) 0 Ueff . Moreover, p-d hybridization lowers the ground- E N E where ⌬0ϳ26 eV and U0ϳϪ2.5 eV for oxides and ⌬0 state energy g( ), thereby increasing gap [see, Eq. E nϭ ϳ23.5 eV and U0ϳϪ4.5 eV for sulfides. (3.5)]. In addition to the gap peak at 5, which is Figure 52 shows that light transition-metal compounds derived from the peak in ⌬eff and/or Ueff , there is a still have sizable U values, which in some cases exceed small peak at nϭ3. This comes from the stabilization of 3 ⌬. This means that many light transition-metal oxides the d configuration due to intra-atomic exchange cou- are not ideal Mott-Hubbard-type compounds, in which pling (Hund’s rule coupling), since the t2g orbitals are ⌬ is larger than U and the magnitude of the band gap is half-filled by three electrons. The calculated Egap for ox- ϳU. Indeed, recent cluster-model analyses of the metal ides of composition RMO3 compare favorably with the 2p core levels of Ti and V oxides have shown that U is experimental values (Fig. 48). comparable to or larger than ⌬ (Uozumi et al., 1993; Using the cluster-model calculation, one can study Okada et al., 1994; Bocquet et al., 1996). The prefactor whether doped carriers occupy the metal d or ligand p orbitals. We may define the p character of doped holes ͱ10Ϫn in the off-diagonal matrix elements Teff ϵͱ10ϪnT increases with decreasing atomic number Z, and the d character of doped electrons from the nd val- not only because of the increase in T itself (Fig. 53), but ues of the N-, (NϪ1)-, and (Nϩ1)-electron states, p also because of the increase in (10Ϫn). Therefore, in nd(N)(ϵnd), nd(NϪ1), and nd(Nϩ1), as Choleϭ1 d the light transition-metal oxides, Teff is often larger than Ϫ͓nd(N)Ϫnd(NϪ1)͔ and Celϭnd(Nϩ1)Ϫnd(N). p U and ⌬ and sets the largest energy scale. Then Teff The Chole shown in Fig. 57(a) generally increases with n, plays the most important role in determining the magni- reflecting the decrease in ⌬. The curves have a peak at tude of the band gap, unlike in the original Zaanen- nϭ5 and a dip at nϭ6 due to the multiplet effect. In d Sawatzky-Allen picture. This situation may be described Fig. 57(b), Cel is plotted against n. The exchange stabi- 1 5 as follows: For example, for the d system Ti2O3, the lization and hence the reduced covalency for the d con- occupied d band seen by photoemission is actually a figurations enhances the d character of doped electrons split-off state of predominantly d1Lគ character rather and the p character of doped holes. One can also see than d0 in the photoemission final state (Uozumi et al., that the d character of doped electrons decreases with 1996) in a manner similar to that shown in Fig. 42(b). the valence of the transition-metal ion, reflecting Figure 54 shows various transition-metal oxides plotted the decrease in ⌬ with metal valence. In the above

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1135

FIG. 51. Charge-transfer energies ⌬ of various 3d transition-metal compounds displayed with respect to (a) the transition-metal atomic number (b) its valence, and (c) the ligand electronegativity (Bocquet et al., 1992a, 1992c, 1996).

p 3. First-principles methods definition those compounds which have Chole smaller d than 0.5 and Cel larger than 0.5 can be said to have d-d There has also been considerable progress in first- p gaps, while those compounds which have Chole larger principles methods for estimating the parameter values d than 0.5 and Cel larger than 0.5 can be said to have p-d using the local(-spin)-density approximation [L(S)DA]. gaps. Although the L(S)DA often fails to predict the correct

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1136 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 52. On-site d-d Coulomb energies U of various 3d transition-metal compounds displayed with respect to (a) the transition- metal atomic number, (b) its valence and (c) the ligand electronegativity (Bocquet et al., 1992a, 1992c, 1996).

ground-state properties of Mott insulators, it gives basi- ͓(pp␴), (pp␲)͔, the transition-metal d level ␧d , and cally correct model parameters. the anion p level ␧p . Systematic changes of these pa- The p-d transfer integral T [or (pd␴), (pd␲)] can be rameters with transition-metal atomic number were estimated by fitting the LDA energy band structure us- found in an early study of 3d transition-metal monox- ing a tight-binding Hamiltonian, which contains the p-d ides (Mattheiss, 1972): Tight-binding fits to non-self- transfer integrals as well as the p-p transfer integrals consistent band-structure calculations showed an in-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1137

FIG. 53. p-d transfer integral (pd␴) plotted against the nomi- nal d-electron number n (Bocquet et al., 1996).

crease in ␧dϪ␧p by ϳ0.7 eV per unit increase in Z,in good agreement with the spectroscopic estimates de- scribed above. Using the tight-binding fit approach to band structures calculated within the LDA, Mahadevan, Shanti, and Sarma (1996) systematically studied a series of LaMO3 compounds. Figure 58 shows deduced values for various transfer integrals including (pd␴) and (pd␲) between the metal and other oxygen orbitals. The decreasing

FIG. 55. Effective charge-transfer energy ⌬eff (a) and effective averaged d-d Coulomb energy Ueff (b) of 3d transition-metal oxides plotted against the nominal d-electron number n. The dotted line for MO is deduced from the extrapolated values of ⌬ and U for MO. Dashed lines represent behaviors expected from the systematic variation of the parameters. From Saitoh, Bocquet, et al., 1995a.

trend of their magnitudes with atomic number is in ac- cordance with the spectroscopic estimates shown in Fig. 53. Values for ␧dϪ␧p deduced simultaneously are plot- ted in Fig. 59, which again shows the same decreasing trend with metal atomic number as the ⌬ values deduced from electron spectroscopy (Fig. 51). One can see from the figures that ⌬ is generally larger than ␧dϪ␧p by a few eV. This can be understood from the definition of 0 1 the LDA eigenvalue ␧LDA,dϵ␧dϩ(nϪ 2 )U [see Eq.

FIG. 54. Zaanen-Sawatzky-Allen U/Teff-⌬/Teff plot for 3d (3.29) below] and that of the d-electron affinity level 0 transition-metal oxides (Bocquet et al., 1996). ␧dϵ␧dϩnU [see Eqs. (3.2) and (3.3)].

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1138 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 56. Band gaps Egap of 3d transition-metal compounds calculated using the cluster model. The dotted line for MO represents the calculated results using the extrapolated values of ⌬ and U for MO (Saitoh, Bocquet, et al., 1995a).

The Coulomb interaction parameter U in solids can be calculated using constrained local-density functional techniques (McMahan et al., 1989, 1990; Gunnarsson et al., 1989; Hybertsen et al., 1990). In the L(S)DA, the one-electron potential at a given spatial point r is ap- FIG. 57. Character of doped electron and hole in 3d proximated by a function of the charge density at that transition-metal compounds: (a) p character of the doped hole point r, ␳(r) [and the spin density ␴(r)ϵ␳ (r)Ϫ␳ (r)]. p d ↑ ↓ Chole ; (b) d character of the doped electron Cel . From Saitoh, If we use atomic orbitals as the basis set, the total energy Bocquet, et al., 1995a. of a transition-metal atom is given by the d-electron number n at the atomic site (which is here regarded as a the change in n in order to include solid-state screening continuous variable) as effects in self-consistent calculations for the U value. In order to deduce the charge-transfer energy ⌬, the 0 1 nϩ1 n E ͑n͒ϭE ϩn␧ ϩ n͑nϪ1͒U, (3.28) affinity level, E(d )ϪE(d )ϭ␧d [Eq. (3.3)], or the LDA 0 d 2 n nϪ1 ionization level E(d )ϪE(d )ϭ␧dϪU of the d or- when the orbitals are not spin polarized. The one- bital has to be evaluated. For this purpose, Slater’s electron eigenvalue of the atomic d orbital is given by transition-state rule (Slater, 1974) is utilized: the derivative of the total-energy functional with respect 1 ␧ ϵE ͑dnϩ1͒ϪE ͑dn͒ϭ␧ ͑nϩ ͒, (3.31) to n: d LDA LDA LDA,d 2 which also follows from Eqs. (3.28) and (3.29). Slater’s ELDA͑n͒ 1ץ ␧ ͑n͒ϭ ϭ␧0 ϩ nϪ U. (3.29) transition-state rule can also be applied to evaluate U: n d ͩ 2ͪץ LDA,d Uϵ͓E͑dnϩ1͒ϪE͑dn͔͒Ϫ͓E͑dn͒ϪE͑dnϪ1͔͒ Therefore U is given by the n derivative of the eigen- 1 1 value ␧LDA : ϭ␧LDA,d͑nϩ 2 ͒Ϫ␧LDA,d͑nϪ 2 ͒. (3.32)

␧LDA,d The p-level energy ␧ is assumed to be correctly givenץ Uϭ . (3.30) p .n by the LDA eigenvalueץ When the system is spin polarized, the LSDA func- In order to evaluate this quantity, one calculates ␧ LDA,d tional becomes a function not only of nϵn ϩn but also for various fixed n values. To do so, one employs the ↑ ↓ of the d-electron number of each spin, n and n : constrained LDA method in the impurity model or the ↑ ↓ supercell model, in which the transfer integrals connect- 0 1 ELSDA͑n ,n ͒ϭE0ϩn␧dϩ n͑nϪ1͒U ing the d atomic orbitals of the central transition-metal ↑ ↓ 2 ion with surrounding orbitals (including anion p) are set to zero, making n a good quantum number. The sur- 1 1 Ϫ n ͑n Ϫ1͒JϪ n ͑n Ϫ1͒J, (3.33) rounding orbitals are allowed to relax as a function of 2 ↑ ↑ 2 ↓ ↓

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1139

large, as shown in Table III, and hence these compounds are incorrectly predicted to be wide gap insulators. This is related to the problem of how well the constrained LDA works in describing the screening process in the solid state and also to the problem of how one can de- fine localized orbitals in a strongly hybridized system.

C. Control of model parameters in materials

1. Bandwidth control The electron correlation strength can be controlled by modifying the lattice parameters or the chemical compo- sition while essentially maintaining the original lattice structure. The on-site (U) or intersite (V) Coulomb in- teraction is kept almost unchanged during the above procedure and hence control of electron correlation strength is usually achieved by control of the transfer interaction (t) or the one-electron bandwidth (W). One method of controlling W is the application of pressure. In general, an application of hydrostatic pressure de- creases the interatomic distance and hence increases the transfer interaction. Pressure-induced Mott-insulator-to- metal transitions are observed typically for V2O3 (see Sec. IV.A.1) and RNiO3 (RϭPr and Nd; see Sec. FIG. 58. Various p-d and p-p transfer integrals in the 3d IV.A.3). Hydrostatic pressure is an ideal perturbation transition-metal oxide series LaMO3 derived from a tight- that modifies W and that is suitable for investigating the binding fit to the LDA band structures (Mahadevan et al., critical behavior of the phase transition. To derive a 1996). The inset shows the metal-oxygen and oxygen-oxygen more quantitative picture of the transition, however, we distances in Å. need to know the lattice structural change under a pres- sure that is usually not so easy to determine. In particu- lar, in the anisotropic crystals, such as a layered perov- where J(ϵjϵJH/2) is the exchange integral between the skite structure, anisotropic compressibility affects W in a d electrons. The one-electron eigenvalues of the spin-up complex manner. and spin-down d levels are given by (Anisimov et al., Another method of W control is modification of the 1991) chemical composition using the solid solution or mixed- crystal effect. In the case of transition-metal compounds, ͒ ELSDA͑n ,nץ ␧LSDA,d␴͑n ,n ͒ϭ ↑ ↓ the electron correlation arises from a narrow d band. n␴ץ ↓ ↑

0 1 1 ϭ␧dϩ͑nϪ 2 ͒UϪ͑n␴Ϫ 2 ͒J, (3.34) where ␴ϭ or . Using this expression, one can estimate the parameters↑ ↓U and J by n 1 n n 1 n Uϭ␧LSDA,d ϩ , Ϫ␧LSDA,d ϩ , Ϫ1 , ↑ͩ 2 2 2 ͪ ↑ͩ 2 2 2 ͪ (3.35) n 1 n 1 n 1 n 1 Jϭ␧LSDA,d ϩ , Ϫ Ϫ␧LSDA,d ϩ , Ϫ . ↓ͩ 2 2 2 2ͪ ↑ͩ 2 2 2 2ͪ (3.36) 1 Here, it should be noted that setting n ϭn ϭ 2 n in Eq. (3.35) does not reduce to Eq. (3.32), meaning↑ ↓ that the definition of U is different between the two expressions. J in Eq. (3.36) is equal to j, and U in Eq. (3.35) is a certain average of u and uЈ (see Table II). Parameters thus estimated are listed in Table III for some 3d transition-metal monoxides (Anisimov et al., 1991). A serious shortcoming of this method is that the Coulomb energy tends to be overestimated for light 3d transition- FIG. 59. Variation of ␧dϪ␧p for the LaMO3 series with the metal oxides: U’s for TiO and VO were calculated too metal M element. From Mahadevan et al., 1996.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1140 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 61. Orthorhombically distorted perovskite (GdFeO3- FIG. 60. Phase diagram of NiS2ϪxSex in the plane of tempera- ture and x (Czjzek et al., 1976; Wilson and Pitt, 1971). type) structure.

Therefore alloying of the transition-metal cations should 1970; MacLean et al., 1979), as shown in Fig. 62, nearly be avoided for W control, and the solid solution in other irrespective of the species of A and B. The bond angle chemical sites is usually attempted. One of the best- distortion decreases the one-electron bandwidth W, known examples of this W control is the case of since the effective d electron transfer interaction be- tween the neighboring B sites is governed by the super- NiS2ϪxSex crystals (Wilson, 1985). The schematic metal- transfer process via the O 2p state. For example, let us insulator phase diagram for NiS2ϪxSex is shown in Fig. 60 (see Sec. IV.A.2 for the details and references). The consider the hybridization between the 3deg state and the 2p ␴ state in a GdFeO -type lattice composed of the parent compound NiS2 is a charge-transfer insulator 3 with the charge gap between the Ni 3d-like state (upper quasi-right BO6 octahedral tilting alternatively (see Fig. Hubbard band) and the S 2p band. The partial substi- 61). In the strong-ligand-field approximation, the p-d 2 0 tution of Se on the S site enlarges the amalgamated 2p transfer interaction tpd is scaled as tpd cos(␲Ϫ␪), where band and increases d-p hybridization. Around xϭ0.6 the mixed crystal NiS2ϪxSex undergoes the insulator- metal transition at room temperature. At low tempera- tures, the antiferromagnetic metal (AFM) state exists in between the Pauli paramagnetic metal (PM) and antifer- romagnetic insulator phase. It has been confirmed that an application of hydrostatic pressure near the AFM-PM phase boundary reproduces a change in the transport properties like that observed in the case of x control, indicating a nearly identical role of pressure and compositional control. Another useful method of W control for a perovskite- type compound (ABO3, see Fig. 61) is modification of the ionic radius of the A site. The lattice distortion of the perovskite ABO3 is governed by the so-called toler- ance factor f, which is defined as

fϭ͑rAϩrO͒/&͑rBϩrO͒. (3.37)

Here, ri (iϭA, B, or O) represents the ionic size of each element. When f is close to 1, the cubic perovskite struc- ture is realized, as shown in Fig. 62. As rA or equiva- lently f decreases, the lattice structure transforms to rhombohedral and then to orthorhombic (GdFeO3-type) structure (Fig. 61), in which the B-O-B bond is bent and the angle is deviated from 180°. In the case of the ortho- FIG. 62. The metal-oxygen-metal bond angles in an ortho- rhombic lattice (the so-called GdFeO3-type lattice), the rhombically distorted perovskite (GdFeO3-type) as a function bond angle (␪) varies continuously with f (Marezio et al., of tolerance factor.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1141

compounds such as R1ϪxAxMO3 (MϭTi, V, Mn, and Co), where the averaged ionic size of the rare-earth (R) and alkaline-earth (A) ions on the perovskite A sites can be systematically modified.

2. Filling control The importance of filling control in correlated metals has been widely recognized since the discovery of high- temperature superconductivity as a function of filling in the layered cuprate compounds. The standard method of filling control is to utilize ternary or multinary com- pounds in which ionic sites other than the 3d (4d)or2p electron related sites can be occupied by different- valence ions. For example, the band filling (n)of La2ϪxSrxCuO4 is controlled by substitution of divalent Sr on the trivalent La sites and is given by the relation; nϭ1Ϫx. Similarly, a number of filling-controlled com- FIG. 63. Electronic phase diagram of RNiO (Torrance et al., 3 pounds can be made by forming the A-site mixed crys- 1992). tals of perovskites, such as La1ϪxSrxMO3, M being the 3d transition-metal element. Taking the correlated insu- 0 tpd is the value for the cubic perovskite. Thus W is ap- lator (Mott or CT insulator with nϭinteger filling) as proximately proportional to cos2␪. A similar relation will the parent compound, we customarily use the nomencla- hold approximately for the one-electron bandwidth of ture ‘‘hole doping’’ when the band filling is decreased the t2g electron state, although a more complicated cal- and ‘‘electron-doping’’ when it is increased, although the culation for distortion-induced mixing between the t2g carriers are not necessarily hole-like or electron-like. and 2p ␴ orbitals is necessary (Okimoto et al., 1995b). Among a number of ternary and multinary com- An advantage of using a distorted perovskite for W pounds, perovskite and related compounds are quite control is that the A site is not directly relevant to the suitable for filling control since their structure is very electronic properties inherent in the B-O network and tough in withstanding chemical modification on the also that the W value can be varied to a considerable A-site. We show in Fig. 64 a guide map for the synthesis extent (by 30–40 %) by choice of different rare-earth or of quasicubic and single-layered (K2NiF4-type) perov- alkaline-earth cations as the A-site element. The best skite oxides of 3d transition-metals (Tokura, 1994), e.g., example of successful W control is the MIT for RNiO3, (La,Sr)MO3 and (La,Sr)2MO4, with a variable number with R being the trivalent rare-earth ions (La to Lu). of 3d electrons (or with a variable filling of the The MI phase diagram for RNiO3 derived by Torrance 3d-electron-related band). Black bars indicate the range et al. (1992) is shown in Fig. 63. The tolerance factor f as of the solid solution (mixed-crystal) compounds that the abscissa represents the variation of the one-electron have so far been successfully synthesized. We can imme- bandwidth or the degree of p-d hybridization tpd . The diately see that quite a wide range of the band fillings effective valence of Ni in RNiO3 is 3ϩ and hence the d can be achieved by A-site substitution of perovskite- 6 1 electron configuration is t2geg with Sϭ1/2. LaNiO3 with related structures. the maximal tpd is metallic, while the other RNiO3 com- For the pseudocubic perovskites of 3d transition- pounds with smaller f or tpd are charge-transfer insula- metal oxides, we show in Fig. 65 a schematic metal- tors with an antiferromagnetic ground state. The MI insulator phase diagram (Fujimori, 1992) with the rela- phase boundary is slanting in the T-f plane, as shown in tive electron correlation strength represented by U/W as Fig. 63. For example, PrNiO3 shows a first-order I-M an ordinate and the band filling of the 3d band as an transition around 200 K associated with an abrupt abscissa. As described in the previous section, YMO3 change in the lattice parameters and the sudden onset of shows stronger electron correlation than LaMO3 be- antiferromagnetic spin ordering. For the smaller-f re- cause of reduced W due to M-O-M bending in the dis- gion, however, the paramagnetic insulating phase is torted perovskite structure. The nϭinteger filled 3d present above the antiferromagnetic insulating phase, as transition-metal oxides with perovskite-type structure in most of the correlated insulators. Again, an applica- are mostly correlated insulators, as can be seen in Fig. tion of pressure increases tpd and in fact drives the IM 65, apart from LaCuO3 and LaNiO3. A fractional va- transition for PrNiO3 and NdNiO3 because these two lence or filling drives the system metallic, yet in some compounds are located in the vicinity of the MI phase cases the compounds remain insulating for large U/W. boundary. (See Sec. IV.A.3 for the details.) Nonstoichiometry or altered stoichiometry for the el- Such a strategy for W control in the perovskite-type ement composition sometimes plays the role of filling compounds has been widely used. In Sec. IV, we shall control. A well-known example of this is the case of see ample examples of W control not only for the parent YBa2Cu3O6ϩy with variable oxygen content. The yϭ1 (integer-n) compounds but also for the carrier-doped compound (YBa2Cu3O7) is composed of CuO2 sheets

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1142 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 64. A guide map for the synthesis of filling-controlled (FC) 3d transition-metal oxides with perovskite and layered perovskite (K2NiF4-type) structures. and CuO chains. The nominal valence of Cu on the Filling control by use of nonstoichiometry (offstoichi- sheet and chain is estimated to be approximately ϩ2.25 ometry) has also been carried out for other systems, for and 2.50, respectively (Tokura et al., 1988). In other example V2ϪyO3 (see Sec. IV.A.1) and LaTiO3ϩy (Sec. words, the holes are almost optimally doped into the IV.B.1), which both show the Mott-insulator-to-metal sheet to produce superconductivity in this stoichiometric transition with such slight offstoichiometry as yр0.03. compound. The oxygen content on the chain site can be The advantage of utilizing oxygen nonstoichiometry is reduced, either partially (0ϽyϽ1) or totally (yϭ0), that one can accurately vary the filling on the same which results in a decrease in the nominal valence of Cu specimen by a post-annealing procedure under oxidizing on the sheet, lowering the superconducting transition or reducing atmosphere. Since vacancies or interstitials temperature Tc and finally (yϽ0.4) driving the system may cause an additional random potential, the above to a Mott insulator (or more rigorously, a CT insulator). method is not appropriate for covering a broad range of The nominal valence of the chain-site Cu in the yϭ0 fillings. compound is ϩ1 due to the twofold coordination, while that for the yϭ1 compounds is Ϸϩ2.5. Thus the nomi- 3. Dimensionality control nal hole concentration or Cu valence in the sheet can apparently be controlled by oxygen nonstoichiometry on Anisotropic electronic structure and the resultant an- the chain site, yet it bears a complicated relation to the isotropy in the electrical and magnetic properties of d oxygen content (y) and furthermore depends on the de- electron systems arises in general from anisotropic net- tailed ordering pattern of the oxygen on the chain sites. work patterns of covalent bondings in the compounds.

FIG. 65. A schematic metal-insulator diagram for the filling-control (FC) and bandwidth-control (BC) 3d transition-metal oxides with perovskite structure. From Fujimori, 1992.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1143

octahedron (e.g., Jahn-Teller distortion) have to be taken into account when considering the anisotropic na- ture of the electronic states near the Fermi level. A prototypical series of layered perovskites are the 4ϩ titanates, Srnϩ1TinO3nϩ1 . They have Ti valence (no 3d electron) irrespective of n, showing the fairly wide 2p – 3d gap exceeding 3 eV. The best known example of a layered perovskite is La2ϪxAxCuO4 (AϭBa, Sr, or Ca) with nϭ1 structure or the so-called K2NiF4 struc- ture. This was the first compound of the high- temperature superconducting cuprates discovered by Bednorz and Mu¨ ller (1986). The nϭ2 analogs of the cuprates are La2ϪxCa1ϩxCu2O6 or (La2ϪxSrx)CaCuO6, which are also superconducting for the appropriate dop- ing range, e.g., at xϳ0.3 (Cava et al., 1990). The com- pounds are regularly oxygen deficient (yϭ1) at the api- cal positions in between a pair of CuO2 sheets and hence form a nearly isolated Cu-O sheet composed of CuO5 pyramidal, not octahedral, units. Similar nϭ1,2,3,... ana- logs of the layered perovskite structure are often seen as essential building blocks of the high-temperature super- conducting cuprates, e.g., in the compounds containing BiO, TlO, or HgO layers. In those cases, however, each FIG. 66. Ruddlesden-Popper series (layered perovskite struc- CuO2 sheet is always nearly isolated by the oxygen- tures) (R,A)nϩ1MnO3nϩ1 (nϭ1, 2, and 3). deficient layer or by the absence of the apical oxygens and retains its essentially single-CuO2-layer structure as The lowering of the electronic dimensionality causes a well. variety of essential changes in the electronic properties In contrast, a literally dimensional crossover with n as well as changes in the interactions among the elec- (the number of MO2 sheets) has been observed in some tron, spin, and lattice dynamics. Systematic control of 3d transition-metal oxides with layered perovskite struc- the electronic dimensionality while keeping the funda- ture. One such example is Srnϩ1VnO3nϩ1 (nϭ1, 2, 3, mental electronic parameters is usually not so easy. In and infinite), in which the nominal electronic configura- the ternary or multinary transition-metal oxide com- tion of the V ion is always 3d1 (Sϭ1/2). The single- pounds, however, we may sometimes find a homologous layered (nϭ1) compound shows an insulating behavior series in which dimensionality control is possible to a over the whole temperature region, whereas the nу2 considerable extent, as we show in the following ex- compound is metallic down to near-zero temperature amples. (Nozaki et al., 1991). Another example of a layered per- Typical examples of such a homologous series are the ovskite family showing the insulator-metal transition transition-metal oxide compounds with layered perov- with n is the case of (La,Sr)nϩ1MnnO3nϩ1 . As described skite structure (called the Ruddlesden-Popper phase), in detail in Sec. IV.F, the perovskite manganites un- which are represented by the chemical formula dergo a phase change from antiferromagnetic insulator (R,A)nϩ1MnO3nϩ1Ϫy , where R and A are the trivalent (Mott insulator) to ferromagnetic metal with hole dop- rare-earth and divalent alkaline-earth ions, respectively, ing, e.g., by partial substitution of Sr on the La sites. and y is the oxygen vacancy. The layered perovskite Such a ferromagnetic metallic state is stabilized to gain structure, (R,A)nϩ1MnO3nϩ1 , includes in the unit cell the kinetic energy of doped carriers under the so-called two sets of n-MO2 sheets, which are connected via api- double-exchange interaction. The single-layered (nϭ1) cal oxygens and separated by an intergrowth (R,A)O compounds La1ϪxSr1ϩxMnO4 are insulators with anti- layer of rock salt structure, as depicted for nϭ1, 2, and 3 ferromagnetic or spin-glass-like ground states over the in Fig. 66. The infinite-n analog of a layered perovskite whole nominal hole concentration (x). (The xϭ0 corre- is nothing but the conventional cubic perovskite. Filling sponds to a parent Mott insulator with a nominal va- control can be executed by alloying the (R,A)-sites, lence of Mn3ϩ). On the other hand, the ferromagnetic such as Rnϩ1ϪxAxMnO3nϩ1Ϫy . The nominal valence of metallic state appears around xϭ0.3– 0.4 in the double- the transition metal M(v) is given by the relation v layer (nϭ2) compounds, La2ϪxSr1ϩxMn2O7, perhaps ϭ(3nϩxϪ2yϪ1)/n. Thus varying n in the layered per- due to an increased double-exchange interaction with ovskite structures corresponds to dimensionality control effective increase of dimensionality, in which high an- from two to three dimensions as far as the M-O network isotropy in magnetic and electronic properties is ob- shape is concerned. However, the lifting of the served (see Sec. IV.F.2). A doped hole in the mangan- d-electron orbital degeneracy in the (pseudo)tetragonal ites has an orbital degree of freedom in the eg state, and structure and/or the anisotropic distortion of the MO6 hence the lifting of orbital degeneracy as well as the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1144 Imada, Fujimori, and Tokura: Metal-insulator transitions

A. Bandwidth-control metal-insulator transition systems

1. V2O3

Vanadium sesquioxide V2O3 and its derivatives have been extensively studied over the past three decades as the canonical Mott-Hubbard system. The discovery of high-temperature superconductivity in the doped cu- prates has revived interest in the vanadium and titanium oxides with many similar ingredients for correlated me- tallic states in close proximity to insulating antiferro- magnetic states. As for undoped and doped V2O3, the literature is vast (Ͼ500 papers), yet a review by Mott ¨ FIG. 67. Schematic structures of spin-ladder chain systems, (1990) summarizes the basic features (see also Bruckner et al., 1983). Here, we should like to focus on recent SrnϪ1Cunϩ1O2n (nϭ3 and 5). From Azuma et al., 1994. advances in the study of V2O3 achieved in the 1990s. V2O3 shows the corundum structure (Fig. 68), in effective decrease of W should be taken into account which the V ions are arranged in V-V pairs along the c when considering the variation of electronic properties axis and form a honeycomb lattice in the ab plane. The with n. oxidation state of V ions is V3ϩ with 3d2 configuration. A new homologous series of the ladder chain systems Each V ion is surrounded by an octahedron of O atoms. of the copper oxides, SrnϪ1Cunϩ1O2n , may be viewed as The t2g orbitals are subject to lifting of degeneracy due a successful example of dimensionality control between to trigonal distortion of the V O lattice and splitting 1D and 2D. The structures are depicted for nϭ3 and 5 2 3 into nondegenerate a g and doubly degenerate eg levels. in Fig. 67. The nϭϱ compound corresponds to the so- 1 This set of t2g levels forms a group of subbands which called infinite-layer compound, SrCuO2 (higher-pressure are isolated in energy but do not decouple further within phase), which is composed of alternating sheets of CuO2 the subbands. A recent LDA band-structure calculation and Sr. A strong antiferromagnetic exchange interaction by Mattheiss (1994) confirms this feature. On the other (of the order of 0.1 eV) works between the Sϭ1/2 spins hand, the c-axis V-V pairs have a large overlap of orbit- on the Cu2ϩ sites connected via the corner oxygens (i.e., als in the a1g levels and hence it has been assumed in on the corner-sharing Cu sites) but not on the edge- many theoretical studies that one electron per V atom sharing Cu sites. Thus the Cu-O networks shown in Fig. resides in a singlet bond of a g levels and the remaining 67 can be viewed as representing the two-leg and three- 1 electron is accommodated in the doublet eg levels. How- leg antiferromagnetic Heisenberg spin-ladder systems. It ever, as we discuss below, the electronic state of this was theoretically predicted (Dagotto, Riera, and Scala- material is still controversial. pino 1992; Rice et al., 1993) that the even-leg chains Direct information about the electronic structure of would show the singlet ground-state with a spin gap, V2O3 has been obtained by photoemission and x-ray ab- while the odd-leg chains would have no spin gap. sorption spectroscopic studies since the 1970s. As shown Takano and co-workers (Hiroi et al., 1991; Azuma in Fig. 69, the valence-band photoemission spectra of et al., 1994; Hiroi and Takano, 1995) have succeeded in V2O3 show the O 2p band located from ϳϪ4to preparing a series of these spin-ladder compounds, n ϳϪ10 eV and the V 3d band within ϳ3 eV of the Fermi ϭ3, 5, and infinite, with use of an ultrahigh-pressure fur- level (EF), apparently indicating a typical Mott- nace. In fact, they have observed a spin-gap behavior in Hubbard-type electronic structure, i.e., ⌬ϾU (Shin the temperature dependence of the magnetic suscepti- et al., 1990). The peak of the V 3d band (at 1–1.5 eV bility and the NMR relaxation rate (T1) for the nϭ3 below EF) shows a weak dispersion of ϳ0.5 eV accord- compound, but not for the nϭ5 compound. Dagotto, ing to angle-resolved photoemission spectroscopy Riera, and Scalapino (1992) and Rice et al. (1993) pre- (Smith and Henrich, 1988). This assignment was con- dicted theoretically that hole-doping in a spin-ladder firmed by a resonant photoemission study, in which a system with a spin gap might produce the metallic state strong enhancement of the near-EF feature was ob- with a pseudo spin gap in the underdoped region and served for photon energies above the V 3p 3d core give rise to novel superconductivity. Unfortunately, car- absorption threshold (h␯տ45 eV), as shown→ in the fig- rier doping has not been successful up to now for the ure. However, hybridization between the O 2p and V spin-ladder systems shown in Fig. 67. However, it is pos- 3d states is substantial, as indicated by their having sible for somewhat modified spin-ladder systems, and in nearly the same ⌬ and U values (ϳ4 eV), as described fact superconductivity was discovered (Uehara et al., in Sec. III.B (Bocquet et al., 1996). If one applies the 1996), as detailed in Sec. IV.D.1. local-cluster configuration-interaction picture to V2O3, theV3dband just below EF in the photoemission spec- IV. SOME ANOMALOUS METALS tra is described as made up of bonding states which are formed between the d1 and d2Lគ configurations in the In Table IV we summarize the basic properties of (NϪ1)-electron final state and are pushed out of the compounds discussed in this section. p-band continuum (i.e., d2Lគ ) due to strong p-d

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 e.Md hs,Vl 0 o ,Otbr1998 October 4, No. 70, Vol. Phys., Mod. Rev. TABLE IV. Basic physical properties near metal-insulator transitions for d-electron systems discussed in this article. The compounds are listed in the same order as their appearance in Sec. IV. The representative cases of the properties summarized here can be found in figures in the corresponding subsections of Sec. IV. The following abbreviations are used: Mott-Hubbard type (MH), charge-transfer type (CT), antiferromagnetic order (AF), ferrimagnetic order (Ferri), paramagnetic (P), superconducting (SC), spin-density wave (SDW), filling control (FC), bandwidth control (BC), temperature control (T), continuous transition (C), crossover (CR), accompanied by a change of structural symmetry (S), enhanced Pauli paramagnetic (ePp), enhanced specific heat ␥(e). The following notations are also used: TCO (charge ordering transition temperature), Tc (superconducting transitions temperature), TC (ferromagnetic Curie temperature), TN (Ne´el temperature), xc (critical concentration for MIT). The asterisks indicate that data are not available, while ‘‘text’’ indicates that details are given in the text (Sec. IV). The resistivity denoted as ‘‘diffuse’’ indicates that the temperature dependence is rather flat with weakly metallic dependence at high temperatures. The description of the magnetic susceptibility in metals is for the paramagnetic states outside the magnetically ordered state near the MIT. The unit of temperature T, the uniform magnetic susceptibility ␹, and the specific-heat coefficient ␥ are K, 10Ϫ4 emu/(Avogadro’s number of transition-metal ions), and (mJ/K2)/(Avogadro’s number of transition-metal ions), respectively. Mott insulator MIT Metals near MIT Spatial Spin Control dimension Compounds Type order TN(K) parameter order ␳␹ ␥Spin order of anisotropy Remark md,Fjmr,adTkr:Mtlisltrtransitions Metal-insulator Tokura: and Fujimori, Imada, FC T2 for ePpϳ10 e SDW for FC strong mass enhancement 2 V2O3Ϫy d MH AF ϳ180 BC 1 BC for BC р40 P for BC 3 ϳ30– 40 for FC for BC T (yϭ0.013) and FC corundum 8 2 1.5 NiS2ϪxSex d CT AF 40–80 BC weakly T ePp e AF 3 ␳ϳT at AF-P transition Sϭ1 T 1 ϳ5 [1] ϳ20 pyrite RH enhanced 7 2 RNiO3 d CT AF 130Ϫ BC 1 T ePp e P3 Sϭ1/2 240 T 5 14 perovskite 8 2 NiS1ϪxSex d CT AF 260 (FC)BC 1 T ePp e P3 T(260 at NiAs Sϭ1 xϭ0) 1.6Ϫ2.4 ϳ6 – 7 2 Ca1ϪxSrxVO3 ——— — — — TePp e P3 ϳ2ϳ9 perovskite 1 2 La1ϪxSrxTiO3 d MH AF 140 FC C T ePp e P 3 typical mass enhancement by FC Sϭ1/2 xcϭ0.05 р5 р17 perovskite 2 1.5 La1ϪxSrxVO3 d MH AF ϳ150 FC C ϳT ePp e P3 Sϭ1 xcϭ0.2 ϳ9.5 [2] * perovskite 9 La2ϪxSrxCuO4 d CT AF 300 FC C ϳT text P 2 high-Tc superconductor Sϭ1/2 xcϳ0.05Ϫ K2NiF4 Tcр40 K 0.06 р14 SC 9 2 Nd2ϪxCexCuO4 d CT AF ϳ240 FC C ϳT P2high-Tc superconductor Sϭ1/2 xcϳ0.14 **SC Tcр24 K text 9 YBa2Cu3O7Ϫy d CT AF ϳ400 FC C ϳT (pseudo р18 P2high-Tc superconductor Sϭ1/2 ycϳ0.6 gap) SC perovskite-like Tcр90 K 9 Bi2Sr2Ca1ϪxRxCu2O8ϩ␦ d CT ** FC C ϳT * P2high-Tc superconductor sϭ1/2 р0.2 [3] SC Tcр90 K 9 2 La1ϪxSrxCuO2.5 d CT AF 110 FC C ϳT ϳ1 р4 [4] P 1 (or 3) near critical point between AF and Sϭ1/2 xcϳ0.15 gapped phase, spin-ladder system 9 2 Sr14ϪxCaxCu24O41 d CT gap — FC C ϳT ϳ2 * P 1 (or 2) superconductor under pressure at

sϭ1/2 xcϳ13.5 SC xϳ13.5 (Tcϳ10 K), spin-ladder 1145 system 1146 e.Md hs,Vl 0 o ,Otbr1998 October 4, No. 70, Vol. Phys., Mod. Rev. TABLE IV. (Continued). Mott insulator MIT Metals near MIT Spatial Spin Control dimension Compounds Type order TN(K) parameter order ␳␹␥Spin order of anisotropy Remark 1 BaVS3 d MH AF ϳ35 T C diffuse CW * P 1 TL liquid? Sϭ1/2 (74 K) hexagonal ferrimagnetic Tcϭ858 K 6 5 Fe3O4 d ,d Ferri T 1 diffuse — — P 3 charge order Tcoϭ121 K Sϭ2,5/2 (121 K) S Ferri spinel resistivity is metallic (d␳/dTϾ0) only at TϾ350 K 5 La1ϪxSrxFeO3 d CT AF 740(xϭ0) FC C diffuse CW * P(xϽ0.7) or 3 charge order Sϭ5/2 134(xϭ1) T (at Tco) AF(xϭ1) perovskite xϭ0.66 Tcoϭ210 K, d␳/dTϽ0 even at TϾTco ϳ500 at charge order transitions Metal-insulator Tokura: and Fujimori, Imada, 8 La2ϪxSrxNiO4ϩy d CT AF xϭyϭ0 [5] FC C diffuse ϳ4[6] * P2 yϭ0.125 TNϭ110 K xcϳ0.7Ϫ K2NiF4 xϭ0.33 Tcoϭ240 K, TNϭ180 K xϭ0.5 0.8 Tcoϭ340 K, d␳/dTϽ0 even at TуTco 4 La1ϪxSr1ϩxMnO4 d CT AF 120 — — — — — — 2 charge order Sϭ2 K2NiF4 xϭ0.5 Tcoϭ217 K, TNϭ110 K 4 2 La1ϪxSrxMnO3 d CT AF 140 FC C T — ϳ3 – 5 F 3 double-exchange system Sϭ2(xcϳ0.17) perovskite ferromagnetic Tcϳ370 K (xϳ0.3) 4 La2Ϫ2xSr1ϩ2xMn2O7 d CT ** FC C * — * F 2 double-exchange system Sϭ2(xcϳ0.3) perovskite FeSi gap — T CR diffuse CW — P 3 (300 K) distorted NaCl 1 VO2 d (MH) gap — T 1 T? ePp — P 1or3 (340 K) S ϳ6 rutile 1 Ti2O3 d MH gap — (FC) T CR diffuse * —P 3 400 K–600 K corundum 6 LaCoO3 d CT gap — T CR diffuse CW — P 3 low-spin, intermediate-spin Sϭ0,1, ϳ500 K perovskite (high-spin) transition 2 2 La1.17ϪxAxVS3.17 d (MH) gap — FC(xcϳ0.35) CR diffuse ePp e P2 T(280 K misfit at xϭ0.17) 4.4 Ͻ20 2 Sr2RuO4 ———— — —ϳTePp 9.7 39 P 2 superconducting Tcϳ1K SC K2NiF4 2 Ca1ϪxSrxRuO3 ———— — —T CW 30 P(xϽ0.4) 3 (xϽ0.4) xϭ1F(xϾ0.4) perovskite Imada, Fujimori, and Tokura: Metal-insulator transitions 1147

FIG. 68. Corundum structure of V2O3.

(d1-d2Lគ ) hybridization (Uozumi et al., 1993). Cluster- model analysis has revealed considerable weight of charge-transfer configurations, d3Lគ ,d4Lគ 2,..., mixed into the ionic d2 configuration, resulting in a net d-electron number of ndӍ3.1 (Bocquet et al., 1996). This value is considerably larger than the d-band filling or the formal d-electron number nϭ2 and is in good agreement with the value (ndϭ3.0) deduced from an analysis of core- FIG. 69. Photoemission spectra of V2O3 in the metallic phase core-valence Auger spectra (Sawatzky and Post, 1979). taken using photon energies in the 3p-3d core excitation re- If the above local-cluster CI picture is relevant to experi- gion. From Shin et al., 1990. ment, the antibonding counterpart of the split-off bond- ing state is predicted to be observed as a satellite on the high-binding-energy side of the O 2p band, although its spectral weight may be much smaller than the bonding state [due to interference between the d2 d1ϩe and d3Lគ d2Lគ ϩe photoemission channels; see→ Eq. (3.12)]. Such→ a spectral feature was indeed observed in an ultra- violet photoemission study by Smith and Henrich (1988) and in a resonant photoemission study by Park and Allen (1997). In spite of the strong p-d hybridization and the resulting charge-transfer satellite mechanism de- scribed above, it is not only convenient but also realistic to regard the d1-d2Lគ bonding band as an effective V 3d band (lower Hubbard band). The 3d wave function is thus considerably hybridized with oxygen p orbitals and hence has a relatively small effective U of 1–2 eV (Sa- watzky and Post, 1979) instead of the bare value U ϳ4 eV. Therefore the effective d bandwidth W becomes comparable to the effective U: WϳU. With these facts in mind, one can regard V2O3 as a model Mott-Hubbard system and the (degenerate) Hubbard model as a rel- evant model for analyzing the physical properties of V2O3. The time-honored phase diagram for doped V2O3 sys- tems, (V1ϪxCrx)2O3 and (V1ϪxTix)2O3 , is reproduced in Fig. 70. The phase boundary represented by the solid line is of first order, accompanied by thermal hysteresis

(Kuwamoto, Honig, and Appel, 1980). In a Cr-doped FIG. 70. Phase diagram for doped V2O3 systems, system (V1ϪxCrx)2O3 , a gradual crossover is observed (V1ϪxCrx)2O3 and (V1ϪxTix)2O3. From McWhan et al., 1971, from the high-temperature paramagnetic metal (PM) to 1973.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1148 Imada, Fujimori, and Tokura: Metal-insulator transitions the paramagnetic insulator (PI) with decreasing tem- perature from above the 500-K region. In a (V1ϪxCrx)2O3 (xϭ0.03) crystal, for example, one en- counters the three transitions from the high-temperature side: PM to PI, PI to PM, and PM to antiferromagnetic insulator, the last two of which are of first order. Such a reentrant MIT disappears for x(Cr)Ͻ0.05, and with x(Ti)Ͼ0.05 the metallic phase dominates over the whole temperature region. As first theoretically demonstrated by Castellani et al. (1978a, 1978b, 1978c), orbital spin coupling of the type (2.8) is particularly important for our understanding of the ordered spin structure in the antiferromagnetic insu- lating (AFI) phase, since the sign and magnitude of the spin-spin superexchange couplings are dependent on the orbital occupancy. More recently, Rice (1995) stressed that the interaction between orbital and spin degrees of freedom is strong and mutually frustrating, so that the paramagnetic phases may have strong spin and orbital fluctuations which are unrelated to the ordered phase. This may explain the first-order nature of the magnetic order transition because, in general, two competing or- FIG. 71. Phase diagram for V2ϪyO3 system as a function of der parameters easily trigger a first-order transition, as both pressure (P) and y (Carter et al., 1993, 1994). From Bao can be seen in the Landau free-energy expansion. Rice’s et al., 1993. conjecture has been confirmed in a recent inelastic neu- tron scattering measurement by Bao et al. (1997). The diagram for a V2ϪyO3 system as a function of both pres- momentum dependence of the spin fluctuation in the sure (P) and y is displayed in Fig. 71. Notably, the me- PM state is not related at all to that of the spin ordering tallic V2ϪyO3 develops a spiral spin-density wave in the AFI state, where the orbital ordering effect is (SDW) below TNϭ9 K (in the AFM phase), with incom- supposed to be important. mensurate wave vector qϷ1.7c*. The incommensurate The pioneering and extensive work by McWhan and spin structure is depicted in Fig. 72: The ordered spin for coworkers (McWhan et al., 1973) demonstrated that yϭ0.027 is in the basal plane with a magnitude of 3ϩ there is an empirical scaling relating the addition of Ti 0.15␮B , forming a perfect planar honeycomb antiferro- (Cr3ϩ) and external pressures, 0.4 GPa per 0.01 addition magnet, and spins in the nearest-neighbor plane are al- of Ti3ϩ and 0.4 Gpa per 0.01 removal of Cr3ϩ. It had most perpendicular. More recently, a nearly identical been anticipated that the doping of Cr or Ti would not alter the electron counting of the V sites but rather would affect the one-electron bandwidth via a slight modification of the lattice. In the case of Cr doping, the three 3d electrons on the impurity Cr-site may be tightly bound, so the above hypothesis holds fairly well. By con- trast, the doped Ti atom can be in a mixed valence be- tween 3ϩ and 4ϩ, indicating a possible change in the band filling. Thus the Ti-doping-induced MIT is not purely a BC-MIT but has the character of an FC-MIT. A difference between Ti doping and application of pressure is seen in the ground state of the metallic state: The Ti-doped metallic crystal (0.06ϽxϽ0.30) undergoes an antiferromagnetic transition at 10–50 K (Ueda et al., 1979), though this is not shown in Fig. 70. Ueda, Kosuge, and Kachi (1980) also found that an electronic phase diagram similar to the case of Ti doping occurs for O- rich (or equivalently V-deficient) samples (V2ϪyO3): The AFI abruptly disappears with yϾ0.02 and instead the antiferromagnetic metal (AFM) phase is present be- low 8–10Kuptoyϭ0.05, beyond which the sample shows a phase separation. The spin structure in the AFM phase was recently determined by neutron scattering for an O-rich crystal FIG. 72. Incommensurate spin structure of V2ϪyO3. From Bao (Bao et al., 1993; Bao, Broholm, et al., 1996). The phase et al., 1993.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1149 spin structure and moment were confirmed to be present in the AFM phase for a Ti-doped (xϭ0.05) crystal. In contrast with the robust helical SDW phase against y in V2ϪyO3, pressure above 2 GPa transforms a neat V2O3 crystal to the low-temperature metallic state with- out magnetic order (McWhan et al., 1973; Carter et al., 1994). Importantly, there is no sign of superconductivity down to 350 mK in this BC-MIT (Carter et al., 1994). The magnetic properties of such a pressurized metallic state were investigated by measurements of the 51V NMR Knight shift and the spin-lattice relaxation rate (1/T1 ; Takigawa et al., 1996). [At 1.83 GPa, the highest pressure attainable in that measurement, a major frac- tion of the sample showed the MIT around 70 K, while a finite fraction (about 4%) of the PM phase remained at the lowest temperature for unidentified reasons but not due to the presence of impurities or nonstoichiometry.] FIG. 73. Temperature dependence of Hall coefficient for (a) a The spin susceptibility deduced from the Knight-shift series of V-deficient crystals in the metallic phase and (b) an data has a maximum at 40 K. Also, a peak was observed insulating crystal driven to metallic with pressure. From Carter et al., 1993, 1994. in the 1/T1T versus T curve at 8 K. These phenomena resemble the spin-gap behavior observed in lightly doped (underdoped) cuprate superconductors (see Sec. tive Hall coefficient versus temperature curves show IV.C.3), although the temperature scale differs by nearly maxima at around the SDW transition temperature one order of magnitude. At the moment, it is not clear (ϳ10 K). As the paramagnetic high-temperature state whether this ‘‘gap’’ behavior arises from the spin contri- approaches the SDW transition with decrease of tem- bution or the orbital one due to orbital degeneracy (Cas- perature, the Hall coefficient is critically enhanced from tellani et al., 1978a, 1978b, 1978c; Rice, 1995). Such an the high-temperature value (3 – 4ϫ10Ϫ4 cm3/C), which anomalous metallic state in pressurized V2O3 is particu- is comparable with the value observed for a pressure- larly interesting in light of the possibility of a spin- driven metallic V2O3 (20 GPa, 4.2 K; McWhan et al., gapped metal or an orbital contribution and warrants 1973). Such an anomalous enhancement of the Hall co- further study. efficient is reminiscent of the high-temperature super- The Ti-doping-induced or V-deficiency-induced MIT conducting cuprates (see Sec. IV.C), although the en- is easier to investigate experimentally than the undoped hanced amplitude is substantially smaller than that in case, which needs relatively high pressure. However, the the high-Tc cuprates. The enhancement ceases as the robust presence of the SDW in the AFM phase compli- temperature is decreased below TN . cates the interpretation of Tϭ0 MIT in terms of the Another notable feature in Fig. 73 is that the V defi- canonical Hubbard model. Nearly a quarter of a century ciency y, or equivalently the change in the band filling, ago, McWhan and co-workers (1971) found that the remarkably suppresses such an enhancement of the Hall T-linear term in the heat capacity in the Ti-doped (x coefficient, which is nearly linear with y around TN . The ϭ0.08) metallic state was very large (␥ϭ80 mJ KϪ2 per observed behavior bears a qualitative resemblance to mol. of V2O3) and that the Wilson ratio ␹/␥ϭ1.8. On the case of the hole-doped cuprates, e.g., La2ϪxSrxCuO4 the other hand, the Hall coefficient for the metallic V2O3 (see Sec. IV.D), in which the Hall coefficient shows a both under pressure (2 GPa, Tϭ4.2 K) and at ambient strong temperature dependence in the underdoped re- pressure (Tϭ150 K), is ϩ(3.5Ϯ0.4)ϫ10Ϫ4 cm3/C, indi- gion and the enhanced coefficient nearly scales with 1/x cating nearly one hole-type carrier per V site. Such a rather than the inverse band filling (1Ϫx)Ϫ1. One might large mass enhancement effect has been interpreted as a speculate that the two systems share a common under- clear manifestation of Brinkman-Rice-type mass renor- lying physics caused by strong antiferromagnetic spin malization on the verge of the BC-MIT. One may no- fluctuations. However, we need careful and further tice, however, that the low-temperature metallic state is analyses, because the enhanced amplitude is consider- different for the different cases, namely, those under ably different (typically by an order of magnitude). pressures of more than 2 GPa and those induced by Ti The critical behavior of the electronic specific-heat co- doping (or V deficiency). The critical behavior of ␥ as efficient ␥ is shown in Fig. 74 as a function of y (at well as of the Hall coefficient was reinvestigated recently ambient pressure) as well as of pressure P for the spe- by Carter and her co-workers (1993, 1994). Figure 73 cific (yϭ0.013) crystal (Carter et al., 1993). In the light shows the Hall coefficient versus temperature for (a) a of the phase diagram depicted in Fig. 71, the metallic series of barely metallic V-deficient crystals and (b) an crystal whose ␥ is shown in Fig. 74 is in the SDW state insulating crystal driven to metallic under pressure. and not in the paramagnetic phase. The aforementioned [Note that both cases show the SDW phase below about large ␥ for the Ti-doped (xϭ0.95) crystal that was re- 10 K in the measured composition (y) and pressure re- ported by McWhan et al. (1971) must also be for the gion. See also the phase diagram in Fig. 71.] The respec- SDW state. As seen in Fig. 74, the ␥ for the pressurized

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1150 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 74. T-linear electronic specific-heat coefficient ␥ vs the oxygen deficiency y at ambient pressure as well as vs the pres- sure P at yϭ0.013. From Carter et al., 1993, 1994. system is steeply increased as the compound approaches FIG. 75. Optical conductivity spectra of V2ϪyO3 in the metallic phase (full lines) at Tϭ170 K (upper) and Tϭ300 K (lower). the metal-insulator (AFI-AFM) phase boundary. This The inset contains the difference of the two spectra ⌬␴(␻) was considered as a hallmark of Brinkman-Rice-type ϭ␴170 K(␻)Ϫ␴300 K(␻). Diamonds indicate the measured dc mass enhancement, although the Brinkman-Rice model conductivity. Dotted lines indicate ␴(␻) of insulating phase does not consider the SDW-like spin order. The critical yϭ yϭ Ϫ1 with 0.013 at 10 K (upper) and 0 at 70 K (lower). From behavior is given by ␥ϭ␥0͓1Ϫ(U/Uc)͔ , where ␥0 is Rozemberg et al., 1995. the ␥ expected for the noninteracting V 3d in a band and U is the intra-atomic Coulomb interaction with the value Uc at the MIT (Brinkman and Rice, 1970). The titatively and strongly enhanced simply due to correla- data presented in Fig. 74 are consistent with the assump- tion effects. However, we know of no reason for diver- tion that the ratio U/Uc varies linearly with P/Pc . gence of ␥ at the AFM-AFI boundary. The experimental In the case of an FC-MIT, on the other hand, the ␥ results shown in Fig. 74 appear to be consistent with decreases as the V deficiency y approaches the MI phase these arguments. boundary. Carter et al. (1993) compared the different The effect of electron correlation in V2O3 and the re- cases, variations of y and P, by sorting the ␥ into con- sultant temperature- and filling-dependent change in the tributions from the spin-fluctuation term (Moriya, 1979) electronic state are also manifested by the optical con- and the Brinkman-Rice term. Then the result could be ductivity, as reported by Thomas et al. (1994) and Ro- explained by assuming that the spin-fluctuation term in- zenberg et al. (1995). Figure 75 shows the optical con- creases in proportion to the doping y, while the ductivity spectra (essentially of the Tϭ0 limit) for the Brinkman-Rice term is nearly unchanged. However, it is insulating V2ϪyO3 phase with MIT critical temperatures not conceptionally clear whether the ␥ can be consid- Tcϭ154 K (yϭ0) and Tcϭ50 K (perhaps yϭ0.13, ered as a sum of the two terms, which must have the though not specified in the original report). Although same physical origin, i.e., electron correlation. If further the experimental data were analyzed based on the cal- doping were possible, the ␥ would eventually show a culated result for a BC-MIT of the infinite-dimensional peak at the AFM(spin-density wave)-PM phase bound- Hubbard model, the experimental conditions imply that ary and then steeply decrease, as observed in another the dominant character may not be BC-MIT but FC- doped Mott-Hubbard system, La1ϪxSrxTiO3 (see Sec. MIT. IV.B.1). In a real V2ϪyO3 system, however, the spin or- The temperature-dependent MIT in pure V2O3 be- dering in the AFM phase is robust against y and no tween the low-temperature antiferromagnetic insulating paramagnetic metallic state at Tϭ0 K emerges up to the phase and the high-temperature paramagnetic metallic chemical phase separation (yϷ0.06; Ueda, Kosuge, and phase was studied using a high-resolution photoemission Kachi, 1980; Carter et al., 1991). When the filling is con- technique as shown in Fig. 76 (Shin et al., 1995). In the trolled for the AFM phase to approach the Mott insula- insulating phase below Tϭ155 K, the total width of the tor by fixing U/t, ␥ should rather decrease because lower Hubbard band was ϳ3 eV, considerably larger the electron and hole pockets with small Fermi surfaces than that (ϳ0.5– 1 eV) deduced from the optical study shrink further. This is in contrast to the mass divergence by Thomas et al. (1994). The low-intensity region near expected for the MIT between the PM and AFI phases. EF in the insulating phase has a width of ϳ0.2 eV, much As discussed in Secs. II.C, II.G.1, and II.G.8, metal- smaller than the band gap (ϳ0.6 eV) deduced from an insulator transitions between the AFI and AFM phases optical study of a sample with the same transition tem- are categorized as ͓I-1 M-1͔ where the carrier number perature but twice the transport activation energy (ϳ0.1 vanishes at the MIT,↔ and ␥ may decrease toward the eV; McWhan and Remeika, 1970). Since the insulating MIT in three dimensions due to a vanishing density of phase of V2O3 is a p-type due to a small states at the band edge. When U/t is increased with fixed number of V vacancies, the Fermi level should be lo- filling, on the other hand, the mass and ␥ may be quan- cated closer to the top of the occupied valence band

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1151

than to the bottom of the unoccupied conduction band tensity at EF to be reduced again above the transition, and should thus be separated from the top of the va- but the high-temperature insulating phase had a larger lence band by a p-type transport activation energy. The bandwidth than that of the low-temperature insulating origin of the small discrepancy between the photoemis- phase, which would partly be due to thermal broadening sion gap (ϳ0.2 eV) and the activation energy ϳ0.1 eV and partly due to the absence of magnetic ordering. is not clear but might be due to a polaronic effect, which As mentioned above, most of the theoretical studies generally increases the photoemission gap compared to on V2O3 (Ashkenazi and Weger, 1973; Castellani, Na- the transport gap for p-type semiconductors. In the in- toli, and Ranninger, 1978a, 1978b, 1978c; Hertel and Ap- sulating phase, the spectrum shows a shoulder on the pel, 1986) employed a model in which one electron oc- low-binding-energy side of the lower Hubbard band, i.e., cupied the eg level and the other electron occupied the at ϳ0.5 eV below EF , which may correspond to the co- a1g level. Here, the a1g electrons were assumed to form herent part of the spectral function as predicted by a Heitler-London-type singlet bond between the two ad- Hubbard-model calculations for the Mott insulating jacent V atoms along the c axis (or equivalently, two phase (Shin et al., 1995). In the metallic phase (T electrons occupied the bonding state formed between Ͼ155 K), the gap is closed and finite spectral weight the a1g orbitals of the two V ions). Thus the remaining appears at EF . However, the lower Hubbard band per- eg electron contributed to the magnetic and transport sists in the metallic phase without significant change in properties and the Hubbard model, with one electron its line shape from that in the insulating phase, except per site used as an appropriate starting point for V2O3. for a broadening by ϳ0.5 eV. This indicates that elec- Recently, based on their LDAϩU band-structure calcu- tron correlation remains substantial even in the metallic lations, Czyzyk and Sawatzky (1997) proposed a new phase, in accordance with the strong mass renormaliza- model in which the two d electrons occupy the eg orbit- tion of conduction electrons. If we attribute the spectral als, forming a high-spin (Sϭ1) state. Then the experi- weight near EF , which appeared above the MIT tem- mental value of the ordered magnetic moment, 1.2␮B ,is perature, to that of quasiparticles, the quasiparticle spec- naturally explained by the reduction of the high-spin tral weight Zϵmb /m␻ [Eq. (2.73c)] is as small as ϳ0.05. moment (2␮B) due to a covalency effect. Indeed, the On the other hand, the spectral intensity at EF nearest-neighbor V-V distance is elongated compared to ͓ϭ(mk /mb)Nb(␮)͔ is only 10–20 % of Nb(␮) deduced that of the ideal corundum structure, implying that the from the band-structure calculation (Ashkenazi and neighboring V ions do not form a chemical bond. This Weger, 1973). Thus the effective mass m*/mb model was supported by recent V 2p x-ray absorption ϭ(mk /mb)(m␻ /mb)ϳ2Ϫ4, which is consistent with spectroscopy (Park et al., 1997b). If this model is correct, m*/mbӍ5 deduced from the electronic specific heat of the whole scenario of V2O3 should be fundamentally the metallic phase of V2O3 (stabilized at low tempera- modified. Further experimental and theoretical studies tures by pressure or V vacancies). The small quasiparti- are necessary to confirm this point in the future. cle weight Zϳ0.05 indicates that the metallic carriers are largely incoherent in V2O3, consistent with the largely incoherent optical conductivity spectra (Thomas et al., 1994). 2. NiS2ϪxSex The high-temperature MIT in Cr-doped samples be- The pyrite system NiS2ϪxSex has long been one of the tween the high-temperature paramagnetic insulating typical compounds for BC-MIT systems. The progress (PI) phase and the low-temperature paramagnetic me- up to the mid 1980s was thoroughly reviewed by Wilson tallic (PM) phase was studied by photoemission by (1985). A more concise review on the MIT in NiS2ϪxSex Smith and Henrich (1994). They found the spectral in- was also given by Ogawa (1979), who with his co- workers had done pioneering work on this system. Here, let us confine ourselves to the nature of the metallic state on the verge of the BC-MIT in NiS2ϪxSex . The end member NiS2 shows a pyrite structure, in which Ni and S2 form a NaCl-type lattice. Due to strong intradimer bonding, the S2 pair accepts only two elec- trons in the bonding state and hence serves as a single 2Ϫ anion S2 . Ni in the pyrite is hence formally divalent with two d␥ electrons. As we shall discuss below, ac- cording to the renewed view on the correlated electronic structure, NiS2 is a charge-transfer insulator (Fujimori, Mamiya, et al., 1996). The other end compound NiSe2, which is isostructural with NiS2, is a paramagnetic metal over the whole temperature range. This is because the p 2Ϫ orbital of the Se2 ion has larger spatial extent than 2Ϫ FIG. 76. High-resolution photoemission spectra of V2O3 above that of S2 and hence larger p-d and p-p transfer in- and below the MIT temperature (Shin et al., 1995). The Fermi teractions. Thus NiSe2 may be viewed as derived from a edge disappears below the transition temperature. charge-transfer insulator by closing of the CT gap.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1152 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 78. Photoemission and inverse-photoemission (BIS) FIG. 77. The x dependence of ␹, A, and residual resistivity spectra of FeS (Mamiya et al., 1997b) compared with the (␳ ) for NiS Se . From Miyasaka et al., 1997. 2 0 2Ϫx x band-structure calculation (Folkerts et al., 1987). Good agree- ment is found between theory and experiment.

The solid solution NiS2ϪxSex with amalgamated p bands undergoes the BC-MIT around xϭ0.8 at room originating from the corresponding molecular orbitals of 2Ϫ temperature and around xϭ0.45 at low temperatures the S2 molecule. The narrow t2g band of Fe 3d origin below about 100 K (Wilson and Pitt, 1971; Bouchard is completely filled and located within the p␲-p␴* gap; et al., 1973; Jarrett et al., 1973; Czjzek et al., 1976), as the relatively wide eg band is empty and overlaps the shown in Fig. 60 and also reproduced in the top panel of empty p␴* band. In going from FeS2 to CoS2 to NiS2, Fig. 77 (Miyasaka et al., 1998). The antiferromagnetic the t2g band is progressively broadened and an addi- metal phase is present in between the antiferromagnetic tional feature due to the occupation of the eg band ap- insulator and paramagnetic metal phases, which bears pears near EF . The transition-metal ions in the pyrite- some analogy to the case of V2O3 (see Sec. IV.B.1). The type compounds tend to take low-spin configurations: AFI-AFM transition is weakly first order, while the Sϭ0 for Fe2ϩ(d6), Sϭ1/2 for the Co2ϩ(d7) ion, and S AFM-PM transition is second order. The AFI state of ϭ1 for Ni2ϩ(d8). [No low-spin (Sϭ0) ground state is 8 NiS2 shows a complicated spin structure (Miyadai et al., possible for the d ion in a cubic crystal field.] Earlier 1978; Panissod et al., 1979), which is a superposition of photoemission studies utilized ligand field theory (van two kinds of sublattice structures, M1 and M2. In the der Heide et al., 1980), molecular-orbital calculations on AFM phase, only the M1-type (MnTe2-type) magnetic MS6 clusters (E. K. Li et al., 1974) and band-structure structure appears to be preserved through the AFI- calculations (Krill and Amamou, 1980; Folkerts et al., AFM transition and the averaged staggered moment 1987) and the structures within ϳ4 eV from EF , were shows a gradual decrease with x, for example, from attributed to metal 3d states. A recent resonant photo- Ϸ1.0␮B for xϭ0 (AFI) to Ϸ0.5␮B for xϭ0.65 (AFM) emission study of NiS2 (Fujimori, Mamiya, et al., 1996), (Miyadai et al., 1983). however, showed that 3p-to-3d resonance enhancement The electronic structure of NiS2 and NiS2ϪySey as well occurs at 5–10 eV below EF rather than in the ‘‘d-band’’ as of the pyrite-type chalcogenides of Fe and Co has region near EF , indicating that NiS2 is a charge-transfer been studied by photoemission and inverse- insulator. In going from NiS2 through CoS2 to FeS2, the photoemission spectroscopy for decades. Figure 78 near-EF region becomes more strongly enhanced, indi- shows combined photoemission and BIS spectra of the cating that the system gradually changes from charge- nonmagnetic semiconductor FeS2 (Mamiya et al., transfer-like to Mott-Hubbard-like because of the in- 1997b). There is a large gap between the occupied p␲ creasing ⌬ (and decreasing U). The photoemission band and the unoccupied antibonding p␴* band, both spectra of NiS2 were analyzed using the CI cluster model

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1153

near the AFM-PM phase boundary, A is nearly propor- tional to ␥2, obeying the Kadowaki-Woods relation for a strongly correlated Fermi liquid (Kadowaki and Woods, 1986). The T2 dependence of the resistivity in the PM phase is seen up to relatively high temperatures (say, 200 K), yet the deviation of the T2 law above some tem- perature is conspicuous near the PM-AFM phase boundary (xϭ1.0) and the resistivity follows the T1.5 dependence rather than T2 dependence down to low temperatures. In the immediate vicinity of the AFM-PM phase boundary, T1.5 dependence persists down to liquid-helium temperatures and hence the T2 coefficient tends to diverge (Ogawa, 1979), as seen in the third panel of Fig. 77. This is in line with the SCR theory (Ueda, 1977; Moriya, 1995) and can be ascribed to strong spin fluctuations, as discussed in Secs. II.D.1 and II.D.8. The T1.5 dependence in this framework is derived in Eq. (2.118). All these observations are consistent with the canonical view of a Fermi liquid with strong electron correlation. Because A diverges while ␥ remains finite, FIG. 79. Temperature dependence of Hall coefficient in the the Kadowaki-Woods relation discussed in Sec. II.D.1 is PM phase for xϾ1 of NiS2ϪxSex . From Miyasaka et al., 1998. clearly not satisfied near the AFM-PM boundary. Let us now turn to the feature in the AFM phase near (Fujimori, Mamiya, et al., 1996). The electronic structure the MI phase boundary. Hall coefficients in the AFM parameters were obtained as ⌬ϭ1.8 eV, Uϭ3.3 eV, phase (Fig. 79, Miyasaka et al., 1998) show a remarkable and the charge transfer energy (pd␴)ϭ1.5 eV, confirm- increase with lowering of temperature. The low- ing the p d charge-transfer nature of the band gap. We temperature value increases with decrease of x, imply- note that→ the global spectral line shape does not change ing that in the AFM phase the carrier number decreases very much across the Se-substitution-induced MIT ex- as the AFI phase is approached. It is worth noting that cept for a slight narrowing of the photoemission main the temperature-dependent enhancement of the Hall co- (d8Lគ final-state) peak in the metallic phase and overall efficient is discernible in the high-temperature PM phase shifts of the empty d and p␴* bands in the inverse- above TN , perhaps due to antiferromagnetic spin fluc- photoemission spectra toward EF in going from the me- tuations. Such a shrinkage of the Fermi surface in the tallic to the insulating phase by 0.2–0.3 eV (Mamiya AFM phase is also reflected in a change in the residual et al., 1997b) due to closure of the band gap. resistivity, as shown in the bottom panel of Fig. 77. The Viewed from the paramagnetic metallic side, the MIT residual resistivity, which is free from renormalization is accompanied by strong enhancement of the electronic effects of electron correlation, can be expressed with a specific heat(␥), the paramagnetic susceptibility (␹), and Fermi surface area SF and mean impurity distance Li as the coefficient of the T2 term (A) in the temperature- (Rice and Brinkman, 1972) dependent resistivity, as noted by Ogawa (1979) and re- ␳ ϭ͑e2S L /12␲3͒Ϫ1. (4.1) cently reinvestigated by Miyasaka et al. (1998). The x 0 F i dependence of ␹, A, and the residual resistivity (␳0)is In the high-pressure measurements, of which results are summarized in Fig. 77 (Miyasaka et al., 1998). In the PM plotted with open circles in the figure, Li was kept con- phase (1.0ϽxϽ2.0), ␥ increases critically with decreas- stant and hence ␳0 directly probed the Fermi surface ing x down to the PM-AFM phase boundary (xϭ1.0) size SF . In this system, the application of pressure was 2 2 from 10 mJ/mol K for xϭ2.0 (NiSe2) to 28 mJ/mol K . equivalent to Se substitution for bandwidth control, as On the other hand, the Hall coefficient in the PM phase the scaling relation indicates on the upper scale of Fig. is almost temperature independent and approximately 77 for the xϭ0.7 crystal. The residual resistivity under 10Ϫ4 cm3/C with positive sign, as shown in the xϾ1 data pressure remained constant in the paramagnetic metal of Fig. 79, which is consistent with a large Fermi surface phase (PϾ2 GPa) irrespective of the x-dependent mass containing Ϸ2 carriers per atom of Ni. These results renormalization, indicating that the carrier density was clearly indicate mass renormalization as the insulator essentially unchanged in the PM phase. By contrast, the phase is approached, while the Fermi surface is pre- residual resistivity in the AFM phase (PϽ2 GPa) in- served in volume. In accord with this, the nearly creased rapidly towards the metal-insulator phase temperature-independent ␹ is also enhanced (Ogawa, boundary, indicating shrinkage of the Fermi surface or 1979) and the Wilson ratio ␹/␥ is around 2, nearly con- reduction in carrier density, in accordance with the Hall stant against x. coefficient. The T2 coefficient A of the resistivity also reflects As the AFI-AFM phase boundary (xϭ0.45) is ap- such an x-dependent mass-renormalization effect, as proached from the metallic side, the T2 coefficient of the shown in the third panel of Fig. 77. Except for the region resistivity (A) is again enhanced by the Brinkman-Rice

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1154 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 80. Photoemission spectra of NiS1.55Se0.45 in the antifer- romagnetic metallic state (15 K) and the paramagnetic insulat- ing state (300 K) (Mamiya et al., 1997a). The solid curve is the 15-K spectrum broadened with a Gaussian to ‘‘300 K.’’ mechanism (Husmann et al., 1996). In the low- temperature region below 1 K near the MI phase bound- ary, however, the conductivity ␴ shows the T1/2 depen- dence that is characteristic of a MIT by disorder due to FIG. 81. Angle-resolved photoemission spectra of NiS1.5Se0.5 Anderson localization under the electron-electron inter- in the antiferromagnetic phase (Matsuura et al., 1996). The actions, as discussed in Sec. II.G.2 (Lee and Ramakrish- spectra label the analyzer angle. The shaded feature due to a narrow Ni 3d band shows little dispersion. nan, 1985; Belitz and Kirkpatrick, 1994). However, in the pressure-tuned AFM phase on the verge of the MI phase boundary (at a distance with ⌬PϭPϪPc band narrowing due to electron correlation near EF Ͻ0.2 Kbar), the conductivity shows a T0.22 dependence. were strong enough. The 50-meV peak was observed to Husmann et al. (1996) argued that this supports viola- be even sharper and only very weakly dispersing in the tion of hyperscaling for the AFM-AFI transition. Such a angle-resolved photoemission study of NiS1.5Se0.5 by violation had indeed been discussed in connection with Matsuura et al. (1996; Fig. 81), indicating that the band the mapping to the random-field problem (Kirkpatrick narrowing is indeed very strong. The spectrum of the PI and Belitz, 1995). For theoretical aspects of this issue, phase in Fig. 80 shows no gap at EF in spite of the fact see Sec. II.G.2. that the electrical conductivity is of the activated p type The temperature-induced MIT in NiS2ϪySey with x with an activation energy of ϳ75 meV. This means that ϳ0.5 has been the subject of recent high-resolution pho- the activated transport comes from the mobility, consis- toemission studies. The angle-integrated spectra of tent with the observed low mobility Ӷ1cm2VϪ1secϪ1, NiS1.55Se0.45 near EF shown in Fig. 80 exhibit spectral which is in the range of small hopping (Kwizera weight transfer from the near-EF region to 0.2–0.4 eV et al., 1980; Thio and Bennett, 1994). Indeed, the Hall below it as the temperature increases from below Ttϳ60 coefficient is small and temperature independent in the K (AFM) to above it (PI) (Mamiya et al., 1997a). (This high-temperature paramagnetic insulator phase, mean- can be more clearly seen by comparing the solid curve, ing that the carrier number is large (ϳ2 hole/Ni; Mi- which represents the 15-K spectrum temperature- yasaka et al., 1997). For pure NiS2 also, the photoemis- broadened to 300 K, with the actual 300-K spectrum in sion spectrum shows almost no gap at EF , implying that Fig. 80). The peak at ϳ50 meV below EF in the AFM the transport activation energy (ϳ300 meV at 300 K) is state may correspond to the quasiparticle spectral mostly due to the mobility. weight, as predicted by the calculations for the Hubbard model in three and higher dimensions discussed in 3. RNiO Sec. II.E.6 for the infinite-dimensional approach 3 (Georges et al., 1996; Ulmke et al., 1996). However, it The interest in RNiO3 has recently been revived by a should be noted that the calculations were done for the systematic study on MITs by Torrance et al. (1992). The paramagnetic metal state and not for the AFM. A simi- MI phase diagram as a function of the perovskite toler- lar change in the spectrum near EF can be identified in a ance factor (or equivalently the p-d transfer interaction) composition-dependent MIT, between the AFM (x was shown (Fig. 63) in Sec. III.C as a prototypical ex- Ͼ0.5) and AFI (xϽ0.5). In the one-electron band pic- ample of a BC-MIT. The metal-insulator phase diagram ture, energy bands of the paramagnetic phase are folded qualitatively resembles that of V2O3 (Fig. 68) or the the- into the antiferromagnetic Brillouin zone, and accord- oretical prediction for the infinite-dimensional Hubbard ingly a gap or a pseudogap tends to open at EF . Al- model, although the insulating state of RNiO3 should be though this picture contrasts with the observation of the classified as a charge-transfer (CT) insulator rather than 50-meV peak, the pseudogap at EF might collapse, if a Mott-Hubbard insulator. Among charge-transfer-type

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1155

FIG. 82. The magnetic and orbital order structure of RNiO3 FIG. 83. Temperature dependence of the magnetic moment determined by a neutron-scattering study by Garcı´a-Mun˜oz for RNiO3. From Gracı´a-Mun˜oz et al., 1992. et al. (1992).

footing with the spin (Sϭ1/2) degree of freedom, as dis- compounds, the high valency of the Ni3ϩ ions in RNiO 3 cussed in Sec. II.I.1. Hartree-Fock band-structure calcu- makes its charge-transfer energy ⌬ as small as ϳ1eV lations were made to investigate the possibility of orbital and stabilizes its low-spin configuration t3 t3 e ,as 2g 2g g ordering in these materials using the parameters ⌬, U, shown by CI cluster-model analysis of photoemission↑ ↓ ↑ and (pd␴) deduced from a cluster-model analysis of the spectra (Barman et al., 1994; Mizokawa et al., 1995). Be- photoemission spectra (Mizokawa et al., 1995; Mi- cause of the small ⌬, the ground state is a strong mixture zokawa and Fujimori, 1995). The ground states with of d7, d8L, and d9L2 configurations, resulting in a net គ គ various spin- and orbital-ordered structures (ferromag- d-electron number as large as n ϭ7.8. On the other d netic and G-, A-, and C-type antiferromagnetic order- hand, the Ni magnetic moment is calculated to be ing) were found to fall within a narrow energy range of ϳ0.91␮ , in good agreement with experiment, and is B ϳ70 meV/Ni, indicating that the complicated orbital or- not strongly reduced from the ionic value, 1␮B . This is 3ϩ dering proposed by the neutron study, which can be because for the low-spin Ni ion charge transfer occurs viewed as a combination of these spin-orbital ordered p e e from O 2 orbitals to g and g orbitals. structures, is quite likely. However, the Hartree-Fock ↑ ↓ e Since there is nominally one electron in the g orbital, calculations predicted that an orbital superlattice of the as in LaMnO3, one may expect a Jahn-Teller distortion x2-y2 – 3z2-r2-type would have an energy lower than the and associated orbital ordering in the insulating RNiO3 above spin-orbital-ordered states. A possible cause for compounds. Indeed, the unusual spin ordering struc- the absence of Jahn-Teller distortion in RNiO3 is hy- tures in PrNiO3 and NdNiO3 was attributed to the for- bridization of the eg and t2g orbitals as a result of the mation of an orbital superlattice (Garcı´a-Mun˜oz et al., † † pair-transfer term (ct ct ce ce and its Hermitian 1992), but so far no evidence for Jahn-Teller distortion 2g↑ 2g↓ g↑ g↓ has been found. The magnetic ground state, described conjugate) in the degenerate Hubbard-model Hamil- kϭ tonian [Eq. (2.6e)]: The Hartree-Fock approximation by a commensurate (1/2,0,1/2) spin-density wave, † shown schematically in Fig. 82, was determined from a gives terms of the type ct ce , which cause a one- 2g↑ g↑ neutron-scattering study by Garcı´a-Mun˜oz et al. (1992). electron hybridization between the t2g and eg orbitals on The 3D magnetic ordering consists of alternating in- the mean-field level. There is no pair-transfer contribu- verted bilayers: AA(AA)AA(AA)... along the [001] di- tion in the high-spin Mn3ϩ ion because the spin is fully rection, with A representing the magnetic arrangement polarized. of Ni moments in one (001) plane (Fig. 82). The tem- Torrance et al. (1992) discussed the metal-insulator perature dependence of the magnetic moment for both transition in terms of a closing of the charge-transfer gap compounds is shown in Fig. 83. Reflecting the first-order with increase of p-d hybridization. To substantiate this, nature of the transition, the abrupt appearance of the Arima and Tokura (1995) investigated optical conduc- ordered moment (Ϸ0.9␮B) is observed at the metal- tivity spectra at room temperature for a series of RNiO3 insulator transition. The coexistence of F and AF inter- (metallic for RϭLa and Nd and insulating for RϭSm actions, which manifest themselves in the ordered spin and Y) as shown in Fig. 84. In accord with the metal- structure, suggests the existence of an orbital superlat- insulator phenomena in dc conductivity, the RϭSm and tice, such as regular occupation of either d3z2Ϫr2 or Y compounds showed decreasing optical conductivity as dx2Ϫy2 type orbitals, as shown in Fig. 82. The nominal ␻ approached zero, that is, the gap feature. On the other 3ϩ Ni in RNiO3 indicates one electron accommodated in hand, the optical conductivity appeared to remain at a the doubly degenerate eg orbital. Thus one should con- finite value for the metallic (RϭLa and Nd) compounds. sider the orbital degree of freedom (or orbital pseu- The spectral feature was not a simple Drude form, but dospin; Kugel and Khomskii, 1973, 1982) on an equal the 2p – 3d interband transitions seemed to contribute

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1156 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 84. Optical conductivity spectra at room temperature for a series of RNiO3. From Arima and Tokura, 1995. down to the low-energy part of the conductivity spectra. A point to be noted is the conspicuous change in the conductivity spectra over a wide energy region up to several eV with variation of R site: This change in the conductivity spectra is apparently accompanied by the isosbetic (namely, equal-absorption) point at 3.5–4.0 eV, across which the spectral weight is transferred from high- to low-energy regions upon closing of the CT gap. Such a large energy scale arises from the large-energy quantities that govern the opening and closing of the CT FIG. 85. Pressure dependence of ␳(T) for PrNiO3. From gap, namely, 3d electron correlation and the p-d trans- Obradors et al., 1993. fer interaction. Thus collapse of the CT gap with in- crease of the tolerance factor (e.g., from RϭSm to R ϭLa) is obviously not a simple closing of the gap like accompanied by a subtle increase in the cell volume by that expected for a semiconductor-to- transi- Ϸ0.02% (Torrance et al., 1992). A more detailed tion, but accompanies a reshuffle of the whole electronic neutron-diffraction study (Garcı´a-Mun˜oz et al., 1994) re- structure relating to the 3d – 2p states. vealed that the lattice transition is accompanied by Among the RNiO3 compounds, LaNiO3 remains me- coupled tilts of NiO6 octahedra, which implies changes tallic down to low temperature, and shows features of a in the Ni-O-Ni angles (by Ϫ0.5°) governing the transfer correlated metallic system (Mohan Ram et al., 1984; interaction between the Ni eg andO2porbitals. How- Vasanthacharya et al., 1984; Sreedhar et al., 1992). The ever, the structural change with the transition is a subtle 2 2 resistivity shows T dependence, ␳ϭ␳0ϩAT , below one and the MIT may be viewed as electronically driven. 50–70 K, although the (␳0 ,A) values are scattered in The control of the p-d transfer interaction by changes reports (Sreedhar et al., 1992; Xu et al., 1993), perhaps in the tolerance factor can be mimicked by application because of difficulty in the characterization of the of external pressure (Canfield et al., 1993; Obradors sample quality (particularly the oxygen content). The et al., 1993; Takagi et al., 1995). We show an example of heat capacity data below 10 K can be fitted to the rela- the pressure dependence of the ␳(T) curves for PrNiO3 tion (Sreedhar et al., 1992) (Obrandors et al., 1993) in Fig. 85. The metal-insulator phase boundary under pressure for PrNiO and NdNiO Cϭ␥Tϩ␤T3ϩ␦T3 lnT. (4.2) 3 3 can be related to the phase diagram shown in Fig. 63 by The last term arises as a consequence of spin fluctua- assuming that the pressure effectively increases the Ni- tions (Pethick and Carneiro, 1973). The electronic O-Ni angle (or tolerance factor). It is also noted in Fig. specific-heat coefficient ␥ϭ14 mJ/mol K2 and the Pauli 85 that the ␳(T) curve shows an apparent reentrant fea- magnetic susceptibility ␹ϭ5.1ϫ10Ϫ4 emu/mol are en- ture like the M-I-M transition in the vicinity of the hanced well above their free-electron gas values. The pressure-induced disappearance of the AFI phase. This Wilson ratio ␥/␹ for LaNiO3 is 2.4, which is typical of a was ascribed to the nature of the first-order phase tran- correlated system. sition (Obradors et al., 1993): The first-order phase As shown in the MI phase diagram (Fig. 63), RNiO3 transformation occurring at TMI is incomplete due to a metals (RϭPr, Nd, Sm, and Eu) undergo a MIT with slowdown of the growth rate of the insulating phase, and decrease of temperature. Among them, PrNiO3 and the metallic phase remains dominant, producing reduced NdNiO3 exhibit a transition from the paramagnetic resistivity at low temperatures. metal (PM) to the antiferromagnetic insulator (AFI) at Another important aspect of the MIT to be kept in around 135 K and 200 K, respectively. The MIT is first mind is the complicated ordered spin structure for the order with thermal hysteresis (Granados et al., 1992) and AFI phase of PrNiO3 and NdNiO3 and possible orbital

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1157

FIG. 86. Temperature dependence of optical conductivity spectra of NdNiO3. From Katsufuji et al., 1995. ordering. It is essential to consider the orbital contribu- tion to the total free energy not only for bandwidth- controlled MITs in RNiO but also possibly for electron 3 FIG. 87. Pressure-temperature phase diagram of NiS samples: properties in the metallic phase. circles, Ttϭ230 K; squares, Ttϭ210 K. Open symbols repre- We show in Fig. 86 the temperature dependence of sent increasing pressures and closed symbols decreasing pres- the optical conductivity spectrum of NdNiO3 (Katsufuji, sures. From McWhan et al., 1972. Okimoto, Arima, Tokura, and Torrance, 1995), which undergoes the PM-to-AFI transition at TMIϭ200 K. Drude-like increase of the optical conductivity as ␻ 0 der pressure (Sparks and Komoto, 1967; Anzai and is suppressed, and the opening of the charge gap is→ ob- Ozawa, 1968). Figure 87 shows the pressure- temperature phase diagram (McWhan et al., 1972). The served below TMI . The missing spectral weight in the low-energy part is redistributed over the energy region discovery of the MIT in NiS has attracted considerable above 0.3 eV, but such a change in spectral features in attention because the transition is not accompanied by a the course of a thermally induced MIT is obviously dif- change in the symmetry of the crystal structure and is ferent from the case of an MIT with variation of the R therefore a candidate for a Mott transition driven by site (Fig. 84) affecting the energy scale for the spectral electron-electron interaction. The ␤-NiS, which is a metastable phase at room temperature and can only be weight transfer. The evolution of the gap below TMI is gradual with temperature, which appears to be well fit- obtained by quenching from high temperatures above ted with the BCS function (Katsufuji, Okimoto, Arima, ϳ600 K, crystallizes in the NiAs structure as shown in Tokura, and Torrance, 1995). The feature is reminiscent Fig. 88. Here the NiS6 octahedra shares edges within the of the SDW gap, whose ground-state magnitude should ab plane and share faces along the c direction. The Ni-Ni distance along the c direction is therefore short, be 2⌬(0)Ϸ3.5kB (Overhauser, 1962). However, what dNi-Niϭ2.65– 2.7 A, making the Ni-Ni interaction along happens in NdNiO3 is beyond this picture. In the case of Cr metal, which is a prototypical SDW system, the ex- the c axis important. At the MIT, the hexagonal lattice perimentally observed peak energy relevant to the gap structure in the conductivity spectrum is 5.2kBTc . This was interpreted in terms of a slight modification of the mean-field theory by the phonon-scattering effect (Barker et al., 1968). By contrast, the peak conductivity energy for NdNiO3 at9KisϷ20kBTc as can be seen in Fig. 86. This is much larger than the value expected for a simple SDW transition and implies that electron corre- lation plays an important role in producing the AFI phase. The effect of the simultaneous formation of an orbital superlattice may also be relevant to the large-gap feature in the low-temperature phase of NdNiO3.

4. NiS The hexagonal form of NiS (␤-NiS) undergoes a first- order phase transition from an antiferromagnetic non- FIG. 88. Crystal structure of NiS (Matoba et al., 1991). Small metal to a paramagnetic metal with temperature or un- circles, Ni atoms; large circles, S atoms.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1158 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 89. Electrical resistivity of NiS1ϪySey : Open symbols are for heating runs, closed symbols for cooling runs. From Anzai et al., 1986.

FIG. 90. Magnetic susceptibility of NiS1ϪySey . Open circles are for heating runs, filled circles for cooling runs. From Anzai parameters a (Ӎ3.45 A) and c (Ӎ5.3– 5.4 A) decrease et al., 1986. by 0.3% and 1%, respectively, in going from the nonme- tallic phase to the metallic phase, resulting in a volume moelectric power is positive and relatively large below collapse as large as ϳ2%. Tt and becomes negative and small above Tt , consistent Electrical resistivity as a function of temperature with the sign of the Hall coefficient (Ohtani, 1974; Ma- shows a jump by a factor of ϳ20 at the MIT temperature toba, Anzai, and Fujimori, 1986). Below Tt , the Seebeck Tt , as shown in Fig. 89 (Anzai and Ozawa, 1968; Sparks coefficient is roughly proportional to the temperature, a and Komoto, 1968; Anzai et al., 1986). Above Tt , the characteristic of a degenerate semiconductor, with some Ϫ5 Ϫ4 resistivity is as low as 10 – 10 ⍀ cm and increases phonon-drag contribution. Above Tt , the Seebeck coef- with temperature. Below Tt , the resistivity is nearly ficient is negative and small, but again is proportional to temperature independent in the 10Ϫ3 ⍀ cm range down the temperature, showing more typical metallic behav- to Tϳ0, in contrast to typical semiconductors or insula- ior. If there is one type of carrier in each of the metallic tors, indicating that the low-temperature phase is a semi- and nonmetallic phases, the jump of the resistivity at the metal (White and Mott, 1971; Koehler and White, 1973) MIT by a factor of ϳ20 and that of the Hall coefficient or a degenerate semiconductor (Ohtani, Kosuge, and by a factor of ϳ40 means that the mobility in the metal- Kachi, 1970; Barthelemy, Gorochov, and McKinzie, lic phase is similar to, or somewhat smaller than, that in 1973). Accordingly, we refer to the low-temperature the nonmetallic phase. However, according to band- phase as the ‘‘nonmetallic’’ phase rather than the insu- structure calculations, the metallic phase has several lating or semiconducting phase. The nonmetallic phase Fermi surfaces (Mattheiss, 1974; Nakamura et al., is always doped with holes due to a small concentration 1994b), and the simple relation nϭ1/eRH may not hold of naturally present Ni vacancies (typically xϭ0.002 in in the metallic phase. Ni1ϪxS). Ttϭ260 K for nearly stoichiometric samples According to neutron-diffraction studies (Sparks and and decreases with Ni vacancies. Komoto, 1968; Coey et al., 1974), the magnetic moment The Hall coefficient RH of Ni1ϪxS with small x of Ni in NiS is 1.5– 1.7␮B at Tϭ4.2 K and it is ferromag- (0.004ϽxϽ0.017) is positive in the nonmetallic phase netically coupled within the ab plane and antiferromag- and nϭ1/eRH is nearly equal to 2x per unit formula, netically coupled along the c axis. The magnetic suscep- which indicates that one Ni vacancy produces two hole tibility of NiS is flat below Tt and abruptly decreases at carriers (Ohtani, Kosuge, and Kachi, 1970; Ohtani, TϭTt , as shown in Fig. 90 (Anzai et al., 1986). The sus- 2 Ϫ1 Ϫ1 1974). The mobility is small (ϭ3cm V sec ) and ceptibility above Tt is not typical Pauli nearly temperature independent, indicating transport in and has a positive slope (d␹/dTϾ0). (Unfortunately, a narrow band. The Hall coefficient in the metallic phase measurements above ϳ300 K are prohibited by the is smaller than in the nonmetallic phase by a factor of phase instability). The magnetic susceptibility of single ϳ40 and becomes negative (Barthelemy, Gorochov, and crystals has shown that the susceptibility ␹Ќ perpendicu- McKinzie, 1973), meaning that in going from the non- lar to the magnetization axis (c axis) is very small below metallic to the metallic phase, the number of carriers Tt (Koehler and White, 1973); therefore the exchange dramatically increases. Presumably the character of the coupling constant J between two Ni atoms along the c 2 Fermi surfaces changes from ‘‘small’’ hole pockets to direction (ϭN␮B/2z␹Ќ , where zϭ2 is the number of ‘‘large’’ electronlike (Luttinger) Fermi surfaces. Ther- nearest-neighbor Ni ions) is as large as 0.14 eV. An in-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1159

FIG. 91. Composition dependence of the transition tempera- ture Tt in Ni1ϪxS1ϪySey (Matoba, Anzai, and Fujimori, 1991). Solid symbols are cooling runs and open symbols heating runs. The inset shows a schematic T-x-y phase diagram. elastic neutron-scattering study indeed revealed a high spin-wave velocity along the c axis and Jϭ0.1 eV was obtained (Hutchings, Parisot, and Tocchetti, 1978). From this J value, the Ne´el temperature was predicted to be TNϳ1000 K, which is much higher than Tt ϭ260 K. This indicates that the antiferromagnetic-to- paramagnetic transition is not driven by a spin disorder- ing along the c axis but by a spin disorder within the ab plane or by the appearance of the metallic phase itself. The strong exchange coupling J suggests the existence of antiferromagnetic correlation in the paramagnetic me- tallic state and is possibly the origin of the positive slope in the magnetic susceptibility of the metallic phase. This has some analogy with the metallic phase of high-Tc cu- prates, although in NiS the electronic transport is three FIG. 92. Electrical resistivity of Ni0.98S1ϪySey . (a) Tempera- dimensional and the magnetic correlation is rather one ture vs resistivity curves; (b) its low-temperature part is fitted to ␳ϭ␳ ϩAT2 and A and ␳ are plotted against Se concentra- dimensional. 0 0 tion y. From Matoba and Anzai, 1987. A pressure-induced MIT is an ideal BC-MIT, for which interpretation of experimental data is most straightforward. On the other hand, it is rather difficult only a weak composition (y) dependence and does not to obtain a set of accurate experimental data under high show an increase towards the MIT (yϭ0.1 in pressure. Substitution of Se for S causes an effect almost Ni0.98S1ϪySey). This is contrasted with the A of identical to pressure as far as the electronic phase dia- NiS2ϪySey in the PM phase, which exhibits divergent grams are concerned (see Fig. 87 and the xϭ0 curve in behavior towards the PM-AFM transition, as predicted Fig. 91). However, it should be noted that Se substitu- by spin-fluctuation theory (Moriya, 1985). Presumably, tion increases the crystal volume rather than decreases it the first-order composition-dependent MIT in Se- (Barthelemy et al., 1976). NiS1ϪySey is a charge-transfer- substituted NiS has prevented our approaching the criti- type compound (see below) whose band gap is formed cal region. Indeed, the discontinuity of the lattice pa- between the occupied chalcogen p and empty Ni 3d rameters at the y-dependent MIT is substantial at Tϳ0. states. Since the Se 4p band is higher than the S 3p Figure 90 shows that the magnetic susceptibility as well band, Se substitution reduces the band gap and drives as the positive slope d␹/dT in the metallic phase de- the system toward the metallic phase. The transition creases with Se concentration (Matoba, Anzai, and Fuji- temperature Tt and the magnitude of the resistivity mori, 1991). This may indicate that the antiferromag- jump at Tt decrease with Se substitution in NiS1ϪySey . netic spin fluctuations are weakened as the system is In fact, Tt is completely suppressed above yӍ0.13, driven away from the MIT boundary into the metallic where the electrical resistivity of the metallic phase can region. (The y dependence of ␳0 was explained as due to be studied down to Tϳ0. The resistivity of Ni0.98S1ϪySey disorder scattering.) 2 shows a T behavior at low temperatures and is ana- An FC-MIT in NiS can be studied in Ni1ϪxS (Sparks 2 lyzed using the formula ␳ϭ␳0ϩAT (Fig. 92) (Matoba and Komoto, 1968). The MIT temperature Tt decreases and Anzai, 1987). The coefficient A, however, shows with x in a manner analogous to that for pressure and Se

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1160 Imada, Fujimori, and Tokura: Metal-insulator transitions

substitution, and Tt is completely suppressed at x Ӎ0.033. This can be seen from the phase diagram of the Ni1ϪxS1ϪySey system shown in Fig. 91: 4x and y show almost the same effect on Tt . The changes in the ther- moelectric power of NiS1ϪySey with Se content (Ma- toba, Anzai, and Fujimori, 1991) are also almost identi- cal to those of Ni1ϪxS (Ohtani, 1974) if ‘‘4x-y scaling’’ is applied. It should be noted that the Ni vacancies not only introduce holes but also decrease the volume, so that the Ni1ϪxS system is influenced by a combination of filling and bandwidth control. The volume contraction of Ni1ϪxS in going from xϭ0toxϭ0.033 corresponds to that of NiS under pressure of ϳ10 kbar, which is half of the pressure (23 kbar) necessary to induce the MIT in NiS. There have been studies of another type of substitu- tion carried out on NiS, namely, 3d transition-metal sub- stitution for Ni. It is not clear a priori whether this kind of substitution belongs to the BC type or the FC type; various experimental results and chemical consider- ations have been combined to estimate the number and sign of carriers in the substituted materials. Substituted Fe and Co are considered to form Fe2ϩ and Co2ϩ ions from the chemical point of view. This is consistent with the increase of Tt for Fe and the decrease of Tt for Co because FeS is more insulating and CoS more metallic FIG. 93. Magnetic susceptibility ␹ at 300 K, the electronic (Coey and Roux-Buisson, 1979; Futami and Anzai, specific-heat coefficient ␥, and the Wilson ratio for 1984). Substitution of Ti, V, or Cr for Ni is more subtle, Ni0.98S1ϪxSex . From Wada et al., 1997. since these elements can take various valence states in sulfides, although Ti and V tend to take valences higher can see that the ␥ is constant and that the magnetic sus- than two (Anzai, Futami, and Sawa, 1981). Chemical ceptibility is weakly enhanced near xϭ0, resulting in a shifts of core-level x-ray photoemission spectroscopy weak enhancement of the Wilson ratio near xϭ0. If we peaks indicated that Ti, V, and Cr are in the trivalent compare the ␥ and the DOS at EF deduced from an state and act as donors. Indeed, thermoelectric power in LDA band-structure calculation in the nonmagnetic the nonmetallic phase changes the sign from positive to state (Terakura, 1987), we obtain a mass enhancement negative at around 2–5% V or Cr substitutions, indicat- factor m*/mbϭ1.7. If A for Ni0.98S1ϪySey with yϳ0.1 is ing n-type doping to the NiS host (Matoba, Anzai, and employed, we obtain a Kadowaki-Woods ratio of A/␥2 Fujimori, 1994). Thermoelectric power in the metallic ϳ5ϫ10Ϫ5 ␮⍀ cm(mol K/mJ)2, which is closer to phase, on the other hand, remains negative, probably (though significantly larger than) the strong-correlation indicating that the transport properties of the metallic limit 1.0ϫ10Ϫ5 rather than the weak correlation limit state with large Fermi surfaces is not sensitive to a small 0.4ϫ10Ϫ6. change in the band filling. Direct information about low-energy excitations was Now, we characterize more quantitatively the proper- obtained in an infrared optical reflection study by ties of conduction electrons in the metallic phase of NiS. Barker and Remeika (1974). Figure 94(a) shows the op- The electronic contribution to the specific heat of NiS in tical conductivity spectra obtained from the Kramers- the metallic phase is difficult to estimate because of the Kronig analysis of the reflectivity data. The spectrum of high MIT temperature. Brusetti et al. (1980) estimated the nonmetallic phase shows a threshold at 0.14 eV and the ␥T term by subtracting calculated lattice contribu- a peak at 0.4 eV due to gap excitations. If the minimum tions from the measured entropy change at the MIT. band gap is an indirect one, the transport gap (indirect The deduced value was ␥Ӎ6mJmolϪ1KϪ2, which was gap) can be much smaller than the optical gap. A non- incidentally close to the ␥ value of Ni1ϪxS with xϭ0.04 self-consistent band-structure calculation by Mattheiss Ϫ0.05, in which Tt is suppressed to zero. On the other (1974) predicts a direct gap of 0.15 eV, whereas the hand, the magnetic susceptibility decreases with increas- minimum gap of indirect type is vanishingly small [Fig. ing Se concentration, as shown in Fig. 90. Thus the Wil- 94(b)]. The ratio between the magnitude of the gap and 2 2 son ratio (␲kB/3␮B)(␹/␥) decreases from ϳ2.9 for NiS that of the transition temperature, 2⌬(0)/kBTt ,isϳ6if (␹ϭ2.36ϫ10Ϫ4 emu molϪ1)toϳ2.1 for the 18% Se- we take the threshold energy 2⌬(0)ϭ0.14 eV or ϳ18 if substituted sample (␹ϭ1.75ϫ10Ϫ4 emu molϪ1) if the we take the peak energy 2⌬(0)ϭ0.4 eV. This indicates same ␥ is assumed. The ␥ and ␹ of Ni0.98S1ϪxSex were that the gap cannot be an ordinary SDW gap and that its recently studied over the wide range of 0.1Ͻxр1by opening involves electron correlation. Figure 94(a) also Wada et al. (1997) and are plotted in Fig. 93, where one shows that the spectrum changes over a wide energy

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1161

able to predict the insulating behavior of the antiferro- magnetic phase, whereas the calculated DOS well ex- plains the overall line shape of the photoemission spec- tra (Hu¨ fner and Wertheim, 1973). Mattheiss (1974) simulated the antiferromagnetic state within the non- self-consistent augmented plane-wave (APW) method by introducing a potential difference (ϮVs) between the muffin-tin spheres of the spin-up and spin-down Ni sites and succeeded in opening a gap as described in Fig. 94(b). More recently, Terakura (1987) performed LSDA calculations for the antiferromagnetic as well as for the paramagnetic state, using the self-consistent linearized APW (LAPW) method. However, the self-consistent so- lution converged to a nonmagnetic state for the experi- mental lattice constants. In order to obtain the antifer- romagnetic solution with a finite band gap, an external staggered magnetic field was applied, but the induced magnetic moment was still as small as ϳ1␮B , compared to the experimental value of 1.5– 1.7␮B (Fujimori, Tera- kura, et al., 1988). Figure 95(a) shows that although the calculated DOS of the paramagnetic state is in rather good agreement with experiment, the DOS for the anti- ferromagnetic state is quite different from the experi- mental line shape. The discrepancy was traced back to the exchange splitting as large as ϳ2 eV of the Ni 3d band in the calculated antiferromagnetic DOS, which vanishes in the nonmagnetic state. As many difficulties are encountered in the band- theoretical description of the electronic structure of NiS, a different approach has been taken based on the NiS6 cluster model, in which electron correlation is fully FIG. 94. Optical conductivity and band structure of NiS: (a) taken into account within a small cluster (Fujimori, Optical conductivity. Solid curve, metallic phase; dashed curve, Terakura, et al., 1988), in analogy with the local cluster nonmetallic phase (Barker and Remeika, 1974). (b) Band picture of NiO. Figure 95(b) shows that the theoretical structure of NiS in the antiferromagnetic state calculated by photoemission spectrum calculated using the CI cluster Mattheiss (1974). The arrows show direct gaps while the indi- model and the measured photoemission spectrum are in rect gap is vanishingly small. good agreement with each other. The result shows that the peak located from 0 to ϳ3 eV below EF , which was range (even above h␯ϳ1 eV) between the two phases, previously assigned to the Ni 3d band, is now assigned indicating that the MIT reorganizes the electronic struc- to the d8Lគ charge-transfer final states, and the d7 final ture up to high energies. The Drude peak in the metallic states appear as satellites 5–10 eV below EF (Hu¨ fner, phase is broad and the scattering rate of charge carriers Riesterer, and Hulliger, 1985). Although the cluster is estimated to be as high as 1/␶ϳ0.3 eV (if one ignores model is too crude to describe the electronic structure the frequency dependence of 1/␶). The width of the near EF , such as the semiconducting band gap of order Drude peak in the nonmetallic phase is an order of mag- 0.1 eV and the Fermi surfaces, it well describes the spec- nitude smaller than that in the metallic phase: 1/␶ tral weight distribution on the coarse energy scale of 1 ϳ0.03 eV. From the mobility ratio between the metallic eV to several eV. The calculated magnetic moment at (M) and nonmetallic (NM) phases, ␮M /␮NMϳ1/2 (if we the Ni site is found to be 1.7␮B , in good agreement with assume nϭ1/eRH in both phases) and from the scatter- experiment. This means that the reduction of the mag- 2ϩ ing rate ratio ␶M /␶NMϳ1/10, we may deduce the effec- netic moment from the ionic value of Ni ion is largely tive mass ratio mM* /mNM* ϳ1/5. due to p-d covalency and we do not need to invoke In order to explain the unusual transport and mag- itinerant-electron antiferromagnetism as previously sug- netic properties of NiS and its MIT mechanism, various gested (Coey et al., 1974). models of the electronic structure have been proposed Since the CI cluster model can describe localized elec- (White and Mott, 1971; Koehler and White, 1973). tronic states with a finite local magnetic moment, the These models consider the eg band of Ni near EF in the good agreement between the experimental and theoret- nonmagnetic and magnetic states, with the t2g andS3p ical photoemission spectra and the magnetic moment in- bands completely occupied. Earlier band-structure cal- dicates that the electronic structure of the nonmetallic culations (Tyler and Fry, 1971; Kasowski, 1973) did not phase of NiS is essentially that of a Mott insulator. On take into account the magnetic ordering and were not the other hand, the loss of local magnetic moment in the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1162 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 96. Photoemission spectra of NiS (Ttϭ260 K) and Fe- substituted NiS (Ttϳ350 K) near the Fermi level compared with LDA (paramagnetic metal), LSDA [AF(1): metallic; AF(2): insulating], and LDAϩU [AF(0)] band-structure cal- culations. From Nakamura et al., 1994b.

metallic phase would naturally cast doubt upon the va- lidity of the model in that phase, even though the calcu- lated and measured photoemission spectra are in good agreement with each other. In order to reconcile the spectroscopic data and the nonmagnetic behavior, one could argue that the local magnetic moment exists on the time scale of photoemission (10Ϫ15 sec), but its fast quantum fluctuations lead to the Pauli paramagnetism. The unusual paramagnetic metallic state has been cor- roborated by an exact diagonalization study of a Ni4X4 cluster, where X is a sulfur atom in the present case and the degeneracy of the eg orbitals is taken into account (Takahashi and Kanamori, 1991). Takahashi and Kan- amori found that between the insulating phase (for large ⌬), where the Ni ions are in the ‘‘high-spin’’ state 2 (͗SNi͘ϳ2) and are antiferromagnetically coupled, and the metallic phase (for small ⌬), where they are in the 2 ‘‘low-spin’’ state (͗SNi͘ϳ0.4), there is an intermediate phase where the local moment remains substantial 2 (͗SNi͘ϳ1) but the antiferromagnetic correlation be- tween nearest neighbors is reduced. Calculated photo- emission spectra do not significantly change between the insulating and intermediate phases, which may explain the result of the photoemission experiment. The low-energy electronic structure and its tempera- FIG. 95. Photoemission spectra of NiS in the metallic (T ture dependence were investigated using high-resolution ϭ300 K) and nonmetallic (Tϭ80 K) phases compared with (a) photoemission spectroscopy by Nakamura et al. (1994b). the band-structure DOS of the nonmagnetic and antiferromag- Figure 96 shows the photoemission spectra near EF netic states and (b) the CI cluster-model calculation (Fujimori taken with an energy resolution of ϳ25 meV. Starting et al., 1988). Best-fit parameters are ⌬ϭ2, Uϭ4, and (pd␴) from the thermally broadened Fermi edge of the metal- ϭϪ1.5 eV. lic phase, a gap of ϳ10 meV is opened below Tt .A

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1163

FIG. 98. Magnetic susceptibility of Ca1ϪxSrxVO3. From Inoue et al., 1997b.

oue et al., 1997b). Stoichiometric SrVO crystallizes in a FIG. 97. Electrical resistivity of CaVO3 and SrVO3. From 3 Onoda, Ohata, and Nagasawa, 1991. cubic perovskite structure, while CaVO3, because of the small ionic radius of Ca2ϩ compared to Sr2ϩ, forms a GdFeO -type orthorhombic structure. Although they do remarkable feature in the spectra of the nonmetallic 3 not show a metal-insulator transition, magnetic order- state is that the band edge is an almost resolution- ing, or superconductivity, these compounds (and their limited step function, whereas the band-structure calcu- alloy system Ca Sr VO ) allow us to deduce impor- lation (the LSDA calculation with the applied staggered 1Ϫx x 3 tant information about the nature of electron correlation magnetic field) predicts a much broader band edge. This in the normal metallic states when the d-band width is means that actual energy bands near the top of the va- continuously varied. Both compounds are thought to be lence band are strongly narrowed compared to those in the Mott-Hubbard regime (for details see below), and predicted by band-structure calculations, most probably theV3dbands are occupied by one electron per V due to strong electron correlation. Comparison between atom. The on-site d-d Coulomb repulsion U is thought the band DOS and the photoemission spectra of the to be nearly the same for each of the two compounds nonmetallic phase shown in Fig. 96 suggests a band- because of the same valencies of the V ions. The differ- narrowing factor (mass enhancement factor) of m*/m b ence between SrVO and CaVO lies in the different տ10. This value is larger than the mass enhancement 3 3 V-O-V bond angles, ␪ϭ180° in SrVO and 155° – 160° in factor in the metallic phase (m*/m ϭ1.7) estimated 3 b CaVO , which should lead to a decrease in the one- from the electronic specific heat, and it is consistent with 3 electron d-band width Wϰcos2␪ by ϳ15% in going from the relationship m* /m* ϳ1/5 deduced from the trans- M NM SrVO to CaVO , or equivalently an increase in the port and optical data above. It is tempting to view the 3 3 bare band mass m by ϳ15%. very narrow quasiparticle bands near the band edge of b For CaVO , from the T-independent spin susceptibil- the nonmetallic phase as reminiscent of a band structure 3 ity ␹ϭ2.3ϫ10Ϫ4 emu molϪ1 (corrected for the core with a nearly k-independent SDW gap, which is opened and the Van Vleck susceptibility) on the entire Fermi surfaces of the metallic states as in a and the electronic specific-heat coefficient ␥ BCS superconductor, even though the nonmetallic ϭ9.25 mJ molϪ1 KϪ2 (Inoue et al., 1997a), we obtain a phase of NiS is not near a (hypothetical) second-order Wilson ratio of (␲k2 /3␮2 )(␹/␥)ϭ1.8, a value similar to MIT. Recently, Miyake (1996) showed, using Luttinger’s B B sum rule and the Ward identity, that in any spatial di- those found for La1ϪxSrxTiO3 and Y1ϪxCaxTiO3 et al. et al. mension, if a continuous phase transition occurs be- (Tokura, Taguchi, , 1993; Taguchi , 1993). As ϭ ϫ Ϫ4 Ϫ1 tween a paramagnetic metal and an antiferromagnetic for SrVO3, ␹ 1.65 10 emu mol and ␥ ϭ8.18 mJ molϪ1 KϪ2, Inoue et al. (1997a, 1997b) give insulator, the Fermi surface should exhibit perfect nest- 2 2 ing at the MIT on the metallic side. (␲kB/3␮B)(␹/␥)ϭ1.5, which is smaller than the ratio of CaVO3 by ϳ20%. Electrical resistivity shows a nearly T2 behavior (Fig. 97), indicating electron-electron scat- 5. Ca Sr VO 1Ϫx x 3 tering, but surprisingly the T2 behavior persists up to 2 Both SrVO3 and CaVO3 exhibit metallic conductivity, room temperature. The Kadowaki-Woods ratio is A/␥ as shown in Fig. 97, and nearly T-independent paramag- ϭ0.9ϫ10Ϫ5 and 1.2ϫ10Ϫ5 ␮⍀ cm(mol K/mJ)2 for netism (apart from the increase at low temperatures CaVO3 and SrVO3, respectively, in the same range as probably due to impurities), as shown in Fig. 98 (Onoda, heavy fermions in the case of perovskite-type Ti oxides Ohta, and Nagasawa, 1991; Fukushima et al., 1994; In- (Tokura, Taguchi, et al., 1993). In going from SrVO3 to

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1164 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 100. Photoemission spectra of Ca1ϪxSrxVO3 in the V 3d band region (Inoue et al., 1995). The upper panel shows the DOS derived from LDA band-structure calculation.

band’’ is even more substantial. A comparison between FIG. 99. Photoemission and inverse-photoemission spectra of the photoemission spectra and the band-structure DOS of the t2g subband of V 3d is made in Fig. 99. Spectral SrVO3 and CaVO3 in the V 3d band region compared with a LDA band-structure calculation of Takegahara (1994). From features from ϳϪ0.7 to ϩ1.5 eV correspond to the V Morikawa et al., 1995. 3d band in the LDA band structure and are therefore attributed to the coherent contribution or the quasipar- ticle part of the V 3d-derived spectral function. The CaVO3, since the electronic specific heat indicates that peak at ϳϪ1.6 eV has no corresponding feature in the the quasiparticle mass m* increases by ϳ15% and the band-structure DOS and is therefore attributed to the bare band mass mb increases by ϳ15%, the mass en- incoherent part of the spectral function, reminiscent of hancement factor m*/mb does not increase. This is con- the lower Hubbard band. The photoemission spectra of trary to the general belief that the effective mass is en- Ca1ϪxSrxVO3 shown in Fig. 100 indicate a systematic hanced (or even diverges) as one approaches the BC- spectral weight transfer with increasing U/W.TheV3d MIT point with increasing U/W. As further evidence for spectral weight is transferred from the coherent part to normal Fermi-liquid behavior, the Hall coefficient of the incoherent part of the spectral function (Inoue et al., SrVO3 is negative and has a nearly T-independent 1995), in good agreement with the dynamic mean-field value, which corresponds to one electron per unit cell, as calculation of the Hubbard model (Zhang, Rozenberg, in La1ϪxSrxTiO3 with xտ1 (Eisaki, 1991). An NMR and Kotliar, 1993). However, there is an important dif- study of SrVO3 (Onoda, Ohta, and Nagasawa, 1991) ference between the dynamic mean-field calculation and showed a nearly T-independent Knight shift and a con- the experimental results. That is, with increasing U/W, stant T1T. the quasiparticle band is narrowed in the calculation Further insight into the mass renormalization in whereas the overall intensity of the quasiparticle band SrVO3 and CaVO3 was obtained from a series of photo- decreases without significant band narrowing in the emission studies (Fujimori, Hase, et al., 1992a; Inoue measured spectra. A high-resolution photoemission et al., 1995; Morikawa et al., 1995). Figure 99 shows com- study has confirmed the absence of a sharp quasiparticle bined photoemission and BIS spectra of SrVO3 and peak at EF (Morikawa et al., 1995). CaVO3. They show a typical Mott-Hubbard-type elec- The deviation of the spectral function ␳(␻) from the tronic structure as in V2O3:TheV3dband is located LDA band structure can be expressed in terms of a self- around EF and the filled O 2p band well (3–10 eV) energy correction ⌺(k,␻) to the LDA eigenvalues ␧0(k) below EF . It should be noted, however, that the ‘‘V 3d as described in Sec. II.D.1. The spectral intensity at EF , band’’ is not a simple V 3d band in the original Mott- ␳(␮), differs from that of the LDA calculation by a fac- Hubbard sense but is made up of states split off from the tor mk /mb , where mk is the k mass defined by the mo- O2pcontinuum due to strong p-d hybridization as in mentum derivative of the self-energy [Eq. (2.73e)]. The 4ϩ the case of V2O3. Since ⌬’s for V compounds are quasiparticle mass m*, which is inversely proportional 3ϩ smaller than those for V compounds, SrVO3 and to the coherent bandwidth and is equal to the thermal CaVO3 fall into the ⌬ϽU region, meaning that the ad- (and transport) masses, is given by m*/mbϭ(m␻ / mixture of states with O 2p character into the ‘‘V 3d mb)(mk /mb), where m␻ is the ␻ mass defined by the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1165

FIG. 102. Optical conductivity in YTiO3. From Okimoto et al., 1995b.

FIG. 101. Electronic and magnetic phase diagram for the R Sr TiO . the transfer energy (t) of the 3d electron or the one- 1Ϫx x 3 electron bandwidth (W), depends critically on the ionic radii of the (R,A) ions or the tolerance factor. For ex- energy derivative of the self-energy [Eq. (2.73d)]. Here, ample, the Ti-O-Ti bond angle is 157° for LaTiO3 but we have assumed that the quasiparticle residue Zk(␻) decreases to 144° (ab plane) and 140° (c axis) for [Eq. (2.70)] does not vary significantly within the quasi- YTiO3 (MacLean et al., 1979), which causes a reduction particle band (Z (␻)ϵm /m ). Then the coherent-to- k b ␻ in W of the t2g state by as much as 20%, according to an incoherent spectral weight ratio is given by mb /m␻ :(1 estimate by a simple tight-binding approximation Ϫmb /m␻). The remarkable spectral weight transfer (Okimoto et al., 1995b). The W value can be finely con- from the coherent part to the incoherent part in going trolled by use of the solid solution La1ϪyYyTiO3 or in from SrVO to CaVO means that m increases as the 3 3 ␻ RTiO3 by varying the R ions. The abscissa in Fig. 101 system approaches the Mott transition from the metallic thus represents the effective correlation strength U/W. side. The observed simultaneous decrease in spectral in- On the end (xϭ0) plane, the variation of the antifer- tensity at EF indicates that mk decreases towards the romagnetic (AF) and ferromagnetic (F) transition tem- Mott transition: for SrVO , m /m Ӎ0.25, m /m Ӎ6, 3 k b ␻ b peratures is plotted for La1ϪyYyTiO3 (Goral, Greedan, and m*/mbӍ1.5; for CaVO3, mk /mbӍ0.07, m␻ /mb and MacLean, 1982; Okimoto et al., 1995b). Due to or- Ӎ20, and m*/mbӍ1.4. Almost the same m* for the two bital degeneracy and orbital ordering, the xϭ0 insula- compounds is consistent with the specific heat ␥ and the tors with large U/W value or a Y-rich (yϽ0.2) region effective electron numbers (ϰ1/m*) deduced from an show a ferromagnetic ground state. In fact, a recent evaluation of the optical conductivity at about 1.5 eV study using polarized neutrons showed the presence of (Dougier, Fan, and Goodenough, 1975). t2g orbital ordering associated with a minimal Jahn- Teller distortion of TiO6 octahedron (Akimitsu et al., B. Filling-control metal-insulator transition systems 1998). Figure 102 shows the optical conductivity spectrum of the gap excitations in YTiO (Okimoto et al., 1995b): 1. R A TiO 3 1Ϫx x 3 The onset of the Mott-Hubbard gap transition can be The perovskitelike RTiO3 (where R is a trivalent clearly seen around 1 eV, while the rise in optical con- rare-earth ion) and its ‘‘hole-doped’’ analog ductivity around 4 eV is ascribed to the charge-transfer R1ϪxAxTiO3 (where A is a divalent alkaline-earth ion) gap between the O 2p filled state and the Ti 3d upper are among the most appropriate systems for experimen- Hubbard band. The observation of the two kinds of gap tal investigations of the FC-MIT. The end member transition and their relative positions indicate that RTiO3 is a typical Mott-Hubbard (MH) insulator with RTiO3 is a Mott-Hubbard insulator rather than a CT Ti3ϩ 3d1 configuration, and the A (Sr or Ca) content (x) insulator in the scheme of Zaanen, Sawatzky, and Allen. represents a nominal ‘‘hole’’ concentration per Ti site, The U/W dependence of the Mott-Hubbard gap magni- or equivalently the 3d band filing (n) is given by n tude (Eg) normalized by W, which was obtained by op- ϭ1Ϫx. tical measurements of La1ϪyYyTiO3 (Okimoto et al., The electronic and magnetic phase diagram for 1995b), is plotted in the xϭ0 end plane of Fig. 101. The R1ϪxSrxTiO3 is depicted in Fig. 101. The crystalline lat- Mott-Hubbard gap magnitude changes critically with a tice is an orthorhombically distorted perovskite (of the rather gentle change (ϭ20%) of W. A similar W de- GdFeO3 type; see Fig. 61). As described in detail in Sec. pendence of the Mott-Hubbard gap was also observed III.C.1, the Ti-O-Ti bond angle distortion, which affects for a series of RTiO3 (Crandles et al., 1994; Katsufuji,

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1166 Imada, Fujimori, and Tokura: Metal-insulator transitions

Okimoto, and Tokura, 1995; Katsufuji, Okimoto, Arima, Tokura, and Torrance, 1995). Before considering the electronic structure of R1ϪxAxTiO3, it is necessary to understand the elec- tronic structure of the parent Mott insulators RTiO3. Neither the local spin-density approximation nor the generalized gradient approximation is sufficient to ex- plain the insulating nature of these compounds within one-electron band theory; calculations on the level of the Hartree-Fock or LDAϩU approximation, in which the spin and orbital degrees of freedom are taken into account on an equal footing, are found to be necessary (Mizokawa and Fujimori, 1995, 1996b; Solovyev, Ha- mada, and Terakura, 1996b). With decreasing size of the R ion, RTiO3 changes from an antiferromagnetic insula- tor (LaTiO3) to a ferromagnetic insulator (YTiO3). Ac- cording to the Hartree-Fock calculation (which includes 1 the Ti 3d spin-orbit interaction), the t2g configuration in the nearly cubic LaTiO3 is in the quadruply degenerate spin-orbit ground state, out of which two states with an- tiparallel orbital and spin moments [both directed in the (111) direction] are alternating between nearest neigh- bors, resulting in a G-type antiferromagnetic state (see Fig. 38). Larger GdFeO3-type distortion [see Figs. 38(a) and 61] induces greater hybridization between neighbor- ing xy, yz, and zx orbitals, which makes the antiferro- orbital exchange as well as ferromagnetic exchange larger through the Hund’s-rule coupling. The ordering stabilizes the Jahn-Teller distortion of type d (see Fig. 38) and the t2g level is split into two lower-lying levels and one higher-lying level (for example, for an octahe- dron elongated along the x direction, the energies of the xy and zx orbitals are lowered relative to the yz or- bital). At each site, one of the two low-lying orbitals is occupied so that the neighboring orbitals are approxi- mately orthogonal to each other with the spins ferro- FIG. 103. Temperature dependence of the resistivity for magnetically aligned. This is why the G-type antiferro- LaTiO3ϩ␦/2 or La1ϪxSrxTiO3 and for Y1ϪxCaxTiO3. From Kat- magnetic order and the small Jahn-Teller distortion in sufuji and Tokura, 1994. LaTiO3 is changed to ferromagnetic order with d-type Jahn-Teller distortion in YTiO3. It has always been con- troversial whether orbital ordering is driven by electron- to xϭ0.4, though ferromagnetic spin ordering disap- electron interaction or by electron-lattice interaction pears around xϭ0.2 (see also Fig. 101; Taguchi et al. (the Jahn-Teller effect). Considering the small Jahn- 1993; Tokura, Taguchi, et al., 1993). In the immediate Teller distortion in the present Ti perovskites, the or- vicinity of the MIT, the resistivity curve of Y1ϪxCaxTiO3 bital ordering would be largely of electronic origin. shows a maximum around 150–200 K. The resistivity be- Filling control in the 3d band can be achieved by par- havior below the temperature for the maximum is ac- tially replacing the trivalent ions with divalent Ca or Sr companied by thermal hysteresis, indicating the occur- ions. The solid solution can be formed for an arbitrary rence of a first-order insulator-metal transition, although R/A ratio and hence the filling (n) can be varied from 1 this feature is blurred, perhaps due to the inevitable ran- to 0. Figure 103 shows the temperature dependence of domness of a mixed-crystal system. A similar tempera- the resistivity for LaTiO3ϩ␦/2 or La1ϪxSrxTiO3 and for ture dependence or reentrant MIT with change of tem- Y1ϪxCaxTiO3 near the FC-MIT. For LaTiO3, which has perature was observed in Sm1ϪxCaxTiO3 near the FC- relatively weak electron correlation, several % of hole- MIT (Katsufuji, Taguchi, and Tokura, 1997). However, doping is sufficient to destroy the antiferromagnetic or- no structural transition has been detected so far near the dering and cause the MIT. To control the filling on a fine hysteresis. scale, nonstoichiometry (␦/2) of oxygen is also utilized, Metallic compounds, La1ϪxSrxTiO3 for xϾ0.05 and as shown in the figure: nϭ1Ϫ␦. The upturn or inflection Y1ϪxCaxTiO3 for xϾ0.4, show strongly filling- point in the resistivity curve corresponds to the antifer- dependent properties near the metal-insulator phase romagnetic transition temperature. For YTiO3 with boundary (Kumagai et al., 1993; Taguchi et al., 1993; larger U/W, by contrast, the insulating phase persists up Tokura, Taguchi, et al., 1993). The upper panel of Fig.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1167

FIG. 105. Temperature dependence of the resistivity in metal- lic La Sr TiO . From Tokura et al., 1993. FIG. 104. Upper panel, x-dependence of the electronic 1Ϫx x 3 specific-heat coefficient (␥) and the Pauli paramagnetic suscep- The temperature dependence of the resistivity in me- tibility (␹) in the metallic region of La1ϪxSrxTiO3 and Y1ϪxCaxTiO3; Lower panel, the effective carrier number per tallic La1ϪxSrxTiO3 can be well expressed up to 300 K 2 Ti-site (nc), deduced from the Hall coefficient (RH) as a func- by the relation ␳ϭ␳0ϩAT , as can be seen in Fig. 105. tion of xϭ(1Ϫn) (Tokura et al., 1993; Taguchi et al., 1993). The T2 dependence of the resistivity is reminiscent of the dominant electron-electron scattering process as de- 104 shows the x dependence of the electronic specific- rived in Eq. (2.83) (see Sec. II.E.1). The coefficient A is heat coefficient (␥) and the Pauli paramagnetic suscep- also enhanced as the Mott-Hubbard insulator is ap- tibility (␹) in the metallic region of La1ϪxSrxTiO3 and proached from the metallic side. In the inset of the fig- Y1ϪxCaxTiO3. In the lower panel of the figure, the ef- ure, the relation between A and ␥ is shown for the fective carrier number per Ti site (nc), deduced from samples near the metal-insulator boundary. The well- 2 the Hall coefficient (RH) via the relation RHϭ1/ncec,is known relation for heavy-fermion systems, AϭC␥ shown as a function of xϭ(1Ϫn). Thus the estimated (Kadowaki and Woods, 1986), also seems to hold ap- value of nc approximately equals the filling n, as indi- proximately in this system, with C͓ϭ1.0 Ϫ cated by a solid line in the lower panel of the figure. ϫ10 11 ⍀ cm(mol K/mJ)2͔, which is nearly the same as Such a relation is consistent with a simple band picture. the universal constant (Tokura, Taguchi, et al., 1993). On the other hand, ␥ shows a critical increase as the This proportional constant seems to be in agreement system approaches the Mott-Hubbard insulator phase with the infinite-dimensional result (Moeller et al., (xӍ0). The x dependence of ␹ is almost parallel with 1995). that of ␥, showing a nearly constant Wilson ratio of Such a critical enhancement of the carrier effective about 2. Provided that the band is parabolic with a con- mass near the MIT phase boundary can also be probed stant effective mass m*, ␥ or the density of states at the by Raman spectroscopy. According to the theory of 1/3 Fermi level (EF) should be proportional to n ,as Mills, Maradudin, and Burstein (1970) for phonon Ra- shown by a solid line in the upper panel of Fig. 104. The man scattering in metals, the scattering cross section of critical enhancement of ␥ and ␹ as compared with this an even-parity phonon is given by reference line is ascribed to the increase of m* as the d2I system approaches the Mott-Hubbard insulator phase. ϰ͕␥ /͓␥2 ϩ͑␻Ϫ␻ ͒2͔͖͉␦␹͉2, (4.3) d d ph ph ph Such a filling-dependent mass renormalization is a ge- ⍀ ␻ neric feature of the FC-MIT, as predicted theoretically where ␻ph is the phonon frequency and ␥ph is the width (Furukawa and Imada, 1992; Imada, 1993a; see Sec. of the phonon spectrum. ␦␹ is the modulation of the II.G). The mass enhancement observed in the experi- charge susceptibility caused by the atomic displace- ments also is consistent with the result in infinite dimen- ments. It was shown by a crude approximation (Katsu- sions (Rozenberg, Kotliar, and Zhang, 1994; see for a fuji and Tokura, 1994) that ␦␹ is proportional to the review, Georges et al., 1996). The observed Wilson ratio Drude weight, nc /m*. The Raman process in metals is is clearly not inconsistent with the dϭϱ result, as dis- mediated by particle-hole excitations across the Fermi cussed in Sec. II.D.6. In the case of Y1ϪxCaxTiO3,a level, while that in insulators is mediated by interband large mass enhancement in metals near xϳ0.4 may be transitions. Thus, roughly speaking, the spectral inten- 2 associated with orbital fluctuations instead of spin fluc- sity of the phonon is proportional to (nc /m*) or the tuations. square of the Drude weight in metals. [However, we

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1168 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 106. Variation of phonon Raman intensities for the Ti-O bending A1g mode as a function of x in La1ϪxSrxTiO3 and Y1ϪxCaxTiO3. From Katsufuji and Tokura, 1994.

FIG. 107. Optical conductivity spectra in R1ϪxSrxTiO3ϩy or need more careful analysis on the frequency depen- R1ϪxCaxTiO3ϩy (RϭLa, Nd, Sm, and Y). From Katsufuji, dence. As we shall discuss later, the true Drude weight is Okimoto, and Tokura, 1995. very small, with a presumably smaller frequency width than the phonon frequency ␻ph in the Raman process. weight transfer from the correlation gap (Mott-Hubbard In the frequency range out of the Drude-dominant fre- or CT gap) excitations to the inner gap excitations with quency, ␴(␻) is governed by the incoherent charge dy- carrier doping is a common feature of FC-MIT systems. namics discussed in Secs. II.F.1 and II.G.9. Therefore, The doping-induced changes in the optical conductivity strictly speaking, the intensity picks up the incoherent spectra in R1ϪxSrxTiO3ϩy or R1ϪxCaxTiO3ϩy (where weight of intraband excitations.] We show in Fig. 106 the RϭLa, Nd, Sm, and Y and y is a small oxygen non- variation of phonon Raman intensities for the Ti-O stoichiometry, mostly Ͻ0.02; Katsufuji, Okimoto, and A x bending 1g mode as a function of in La1ϪxSrxTiO3 Tokura, 1995) are shown in Fig. 107. The spectra of the and Y1ϪxCaxTiO3. This mode becomes Raman active as respective samples are labeled with the nominal hole a result of the orthorhombic distortion of the crystalline concentration ␦, where ␦ϭxϩ2yϭ1Ϫn, and with the lattice, which is present for xϽ0.4 in La1ϪxSrxTiO3 and value of the one-electron bandwidth W˜ normalized by for the whole compositional range in Y1ϪxCaxTiO3.As can be seen in the figure, the spectral intensity in the that of LaTiO3. The open circles show the magnitude of metallic regions steeply decreases as the metal-insulator the dc conductivity, ␴(␻ϭ0), at room temperature. In phase boundary (vertical dotted lines) is approached, the case of RϭNd (Nd1ϪxCaxTiO3), for example, ␴(␻) and almost disappears in the insulating regions. This in- for ␦ϭ0(nϭ1) shows negligible spectral weight below dicates that this phonon Raman mode is activated in the 0.8 eV but steeply increases above 0.8 eV due to the metallic state via coupling with the intraband excitation onset of the Mott-Hubbard gap excitations. The spectral across the Fermi level, as described above, but not in the weight of the Mott-Hubbard gap excitations is reduced insulating state with a charge gap larger than the phonon with increase of ␦, while that inside the gap increases. energy. The phonon intensity can be corrected by the The spectral weight appears to be transferred gradually x-dependent orthorhombic distortion. It decreases to one from the above-gap region to the inner-gap region zero with decreasing x, indicating that the intraband re- across an isosbetic (equal-absorption) point around 1.2 eV. sponse at ␻ph vanishes as the MIT is approached. As is noticed by Katsufuji and Tokura (1994), this x depen- It is remarkable in Fig. 107 that, though ␴(␻) changes dence of the phonon intensity is roughly scaled by ͓(1 with ␦ in a similar way irrespective of the element R, the Ϫx)/␥͔2 where ␥ is the specific-heat coefficient. Al- evolution rate of the inner-gap excitation (quasi-Drude though the simple interpretation of ␦␹ as the Drude part) with ␦ is critically dependent on the magnitude of weight needs caution as discussed above, it is interesting W. That is, spectral weight transfer into the quasi-Drude to note that ␦␹ appears to follow the scaling (1Ϫx)/␥, part decreases with a decrease of W. In quantitative implying that the low-energy incoherent weight given by terms, the effective number of electrons is ␻ the ␻-independent prefactor for the incoherent part of 2m0 xϭ Neff͑␻͒ϭ ␴͑␻Ј͒d␻ (4.4) ␴(␻) follows the same scaling as 1/␥ near 0. ␲e2N ͵ The FC-MIT causes a change in electronic structure 0 over a fairly large energy scale. In particular, spectral where m0 is the free-electron mass and N the number of

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1169

FIG. 109. Lower panel, the WϪ1 (strength of electron correla- Ϫ1 FIG. 108. ND to ND0 as a function of ␦ (Katsufuji, Okimoto, tion) dependence of C : upper panel, the value of 2⌬/W (the and Tokura, 1995) for La1ϪxSrxTiO3. optical energy gap normalized to the bandwidth) as well as the activation energy (2⌬act) of the resistivity for R1ϪxCaxTiO3. From Katsufuji, Okimoto, and Tokura, 1995. Ti atoms per unit volume. Neff(␻) is nearly independent of ␦, with ␻ϭ2.5 eV, which includes the spectral weight of both the inner-gap and the Mott-Hubbard-gap excita- ergy of the intraband part containing both the true tions. The sum of these two parts is nearly conserved Drude and the incoherent responses) seems to be scaled under a variation of ␦, but the spectral weight is trans- by the doping level and is consistent with both the scal- ferred from the Mott-Hubbard part to the inner-gap ing theory in Sec. II.G.9 and the dϭϱ result (Kajueter part across the isosbetic point ប␻c (Ϸ1.1 eV for and Kotliar, 1997). Ϫ1 Nd1ϪxCaxTiO3). Thus the value defined as ND Figure 109 (lower panel) shows the W (strength of Ϫ1 ϭNeff(␻c)␦ϪNeff(␻c)␦ϭ0 is the low-energy spectral electron correlation) dependence of C , while the up- weight transferred from the higher-energy region with per panel shows the value of 2⌬/W (the optical energy doping. In Fig. 108, ND , which is normalized to ND0 , gap normalized to the bandwidth) as well as the normal- the value of ND for Uϭ0 and ␦ϭ0(nϭ1), is plotted as ized activation energy (2⌬act /W) of the resistivity. The a function of ␦. The observed linear increase of ND with solid line, which is a linear fit to the data points of ␦, 2⌬act /W, crosses the 2⌬ϭ0 line (abscissa) at a finite value of W(Ϸ0.97) corresponding to the critical value ND /ND0ϭC␦, (4.5) for the BC-MIT ͓(U/W)c͔ at which the charge gap is consistent with theoretical studies of the Hubbard closes and the Mott transition occurs. The relation ⌬act ␣ model in dϭϱ (Jarrell, Fredericks, and Pruschke, 1995), ϰ͓U/WϪ(U/W)c͔ with the exponent ␣ϳ1 implies provided that ND is interpreted as the Drude weight at that the critical exponents of a BC-MIT here are char- ␻ϭ0. However, because ND also includes incoherent acterized by ␯zϳ1 where ␯ and z are the correlation excitations up to ␻c , this ␦ dependence is dominated by length and dynamic exponents, respectively, as is dis- that of the intraband kinetic energy and hence domi- cussed in Sec. II.F.4 [see Eq. (2.332)]. nated by the incoherent charge response discussed in On the other hand, the lower panel of Fig. 109 indi- Sec. II.F.1. In fact, the optical conductivity shown in Fig. cates that CϪ1 decreases linearly with a decrease in 107 shows a rather similar feature to the incoherent re- WϪ1. It is to be noted that the value of WϪ1 at which sponse ␴(␻)Ӎ(1ϪeϪ␤␻)/␻ discussed in Secs. II.F.1, CϪ1 becomes zero nearly coincides with the value at II.G.9, and II.M.2 because both have a long tail at large which 2⌬ becomes zero. This means that the rate of evo- ␻ roughly scaled as 1/␻, although ␴(0)ϳ1/T2 seen in lution of the low-energy incoherent part with ␦ is criti- Fig. 105 does not fit this form. This implies that the true cally (divergently) enhanced as the electron correlation 2 Drude weight associated with T resistivity is very small, (U/W) approaches the critical value (U/W)c for the only prominent around ␻ϳ0, while the dominant weight BC-MIT, following the relation in the observed ␴(␻) is exhausted in the incoherent part Cϰ͓͑U/W͒Ϫ͑U/W͒ ͔Ϫ1. (4.6) with the long tail up to large ␻ϳ1 eV. This is in contrast c with the theoretical results in dϭϱ and similar to the Strictly speaking, the MIT is defined as the point where incoherent response discussed in Secs. II.F.1 and II.G.9. the true Drude weight vanishes. However, it is interest- The integrated intensity itself (namely, the kinetic en- ing to note that the weight of the broad incoherent part

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1170 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 111. Evolution of the single-particle density of states ␳(␻) as a function of U/W and band filling n, derived from the photoemission and inverse-photoemission studies of Ti and V perovskite oxides. From Morikawa et al., 1996.

FIG. 110. Photoemission spectra of La1ϪxSrxTiO3. From Fuji- mori et al., 1992b. studied in that work. Recent measurements on stoichio- with the 1/␻ tail also more or less vanishes at the BC- metric samples yielded a vanishing intensity at EF .) With increasing x, i.e., with hole doping, the ϳ1.5 eV MIT point because Neff(␻c)␦ϭ0 is small. Since Neff mea- sures the kinetic energy when ␻ is taken to be large, C peak loses its spectral weight without changing energy measures approximately the singular part of the kinetic position or line shape down to xϳ1, at which composi- energy near the MIT. The exponent in Eq. (4.6) contra- tion the peak almost disappears. The lost spectral weight dicts that in the Gutzwiller approximation given in Eq. is transferred to above EF , as observed by O 1s x-ray (2.198), which implies the breakdown of the mean-field absorption and inverse photoemission spectroscopy. In treatment in 3D. the metallic region (0ՇxՇ1 for La1ϪxSrxTiO3), a weak A proper microscopic description of mass enhance- but finite Fermi-edge intensity is found, indicating that ment near the FC-MIT is an important step in elucidat- upon hole doping a small portion of the spectral weight ing the nature of the Mott transition, as reflected by the is transferred from the lower Hubbard band to the qua- specific heat and the magnetic susceptibility mentioned siparticle band crossing EF . Qualitatively the same above. Photoemission spectroscopy gives such informa- spectral changes were observed for the other tion in the same way as was described for the R1ϪxAxTiO3 systems, but the quasiparticle spectral weight decreases with the decrease in ionic radius of the bandwidth-control system Ca1ϪxSrxVO3 (Sec. IV.A.5). For filling-control systems, in addition to the spectral A-site ion and hence with increasing U/W (Morikawa weight transfer between the coherent quasiparticle part et al., 1996). The changes in spectral weight distribution and the incoherent part, spectral weight should be trans- as a function of both band filling n and interaction ferred between the occupied (␻Ͻ␮) and unoccupied strength U/W are schematically depicted in Fig. 111. (␻Ͼ␮) portions of the spectral function because of the If we assume that the self-energy correction ͚(kជ ,␻)is ␮ Ϫ1 sum rule that ͐Ϫϱd␻␳(␻) is equal to the number of kជ independent and therefore that m*/nbϳZ (where electrons [see Eq. (2.58)]. Z is the quasiparticle spectral weight), the experimen- Spectral weight transfer in the photoemission spectra tally observed low quasiparticle weight in every has been studied for La1ϪxSrxTiO3 (Fujimori, Hase, R1ϪxAxTiO3 system is apparently incompatible with the et al., 1992b; Aiura et al., 1996), R1ϪxBaxTiO3 (RϭY, electronic specific heat, which implies m*ϳmb except La, and Nd) (Robey et al., 1993), Nd1ϪxSrxTiO3 (Robey for the vicinity of the Mott transition, where m* be- et al., 1995) and Y1ϪxCaxTiO3 (Morikawa et al., 1996). comes as large as ϳ5mb (Tokura, Taguchi, et al., 1993). These systems differ from each other in the magnitude This indicates the important role of nonlocal effects, i.e., of W and hence of U/W. Figure 110 shows the photo- of the momentum dependence of the self-energy correc- emission spectra of La1ϪxSrxTiO3 (Fujimori, Hase, et al., tion, in the mass renormalization and enables us to 1 1992b). The nϭ1(xϭ0) end member is a d Mott in- evaluate not only m*ϭmkm␻ /mb but also mk and m␻ sulator and shows a broad peak centered at ϳ1.5 eV separately, as in the case of Ca1ϪxSrxVO3 (see Sec. below EF , which is assigned to the lower Hubbard band. IV.A.5). In the present case, because such a decomposi- (The weak intensity at the Fermi edge is due to the me- tion of the spectra into the coherent and incoherent tallicity of the LaTiO3ϩ␦ sample with excess oxygen parts is difficult owing to the weakness of the coherent

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1171

part, the quasiparticle DOS at EF , N*(␮), obtained metallic conductivity may remain finite up to a consid- from the electronic specific-heat coefficient ␥, is utilized erable degree of potential disorder. In the other model, to estimate Z, using the relationship Zϭ␳(␮)/N*(␮) one considers the interaction of doped electrons with [Eq. (2.126)]. Because of the very low spectral intensity the lattice distortion through both short-range and long- at EF , one obtains ZՇ0.01 and ϳ0.05, respectively, for range interactions, which are usually described by the Holstein and Fro¨ hlich Hamiltonians, respectively (Fuji- metallic samples of Y1ϪxCaxTiO3 and La1ϪxSrxTiO3 mori, Bocquet, et al., 1996). In this model, a doped elec- (Morikawa et al., 1996). Since mk /mbϭ␳(␮)/Nb(␮) [Eq. (2.125)], the reduction of the spectral weight ␳(␮) tron is self-trapped due to the polaronic effect with a small binding energy (and hence a small transport acti- in going from La1ϪxSrxTiO3 to Y1ϪxCaxTiO3 means that vation energy) but, when the electron is emitted in the Z and mk decrease with U/W not only for nϭ1 photoemission process, the lattice is left with a large ex- (Ca1ϪxSrxVO3) but also for 0ϽnϽ1. The extremely low quasiparticle spectral weight ob- cess energy of ϳ1 eV, giving rise to the deep d-derived served in the photoemission study seems consistent with photoemission peak. This approach also has a problem, the overwhelmingly incoherent response in ␴(␻). Al- namely, that the observed metallic transport remains to though the results of finite-size calculations using the be explained using models with the tendency for self- Hubbard model indicate a relatively sharp peak of large trapping. quasiparticle spectral weight (of order 0.1–1) for a real- istic U/W (or U/t) values (Dagotto et al., 1991; Bulut, Scalapino, and White, 1994a; Eskes et al., 1994; Ulmke 2. R A VO et al., 1996), the real coherent weight of the quasiparticle 1Ϫx x 3 near the Mott insulating phase could be only a tiny por- La1ϪxSrxVO3 has long been known as a FC-MIT sys- tion of this as discussed in Sec. II.F.2. This view seems to tem among the 3d transition-metal oxide perovskites be consistent with the above observation, at least for a and was discussed in the earlier version of Mott’s text metal near the Mott insulator. However, this does not book (Mott, 1990). The end compound RVO3, where R explain the low weight at small electron fillings. In order is a trivalent rare-earth ion, possesses V3ϩ ions with 3d2 to explain the strong reduction of the spectral weight at configuration (Sϭ1). The crystalline lattice of RVO3 is 1 EF in the bandwidth control d Mott-Hubbard system tetragonal for RϭLa and Ce, while for the others it is of Ca1ϪxSrxVO3, Morikawa et al. (1995) proposed that it GdFeO3 type with orthorhombic distortion (Sakai, Ada- was the effect of the long-range Coulomb interaction, chi, and Shiokawa, 1977). Therefore the one-electron which is neglected in the Hubbard model. The effect of bandwidth (W) must decrease with decreasing radius of the long-range Coulomb force could be one origin of the R ion, as in the case of RTiO3 (Sec. IV.B.1). RVO3 small Z even far away from the integer filling because is typically a Mott-Hubbard insulator and undergoes an the screening becomes progressively ineffective at low antiferromagnetic transition at 120–150 K. The antifer- electron filling. Another origin of low Z could be disor- romagnetic phase is accompanied by a spin canting, giv- der. A recent dynamic mean-field approach has shown ing rise to weak ferromagnetism. that the typical quasiparticle weight can be reduced even The insulating nature of the parent RVO3 compounds far away from half filling if disorder and strong correla- is again insufficiently explained by LSDA band- tion coexist (Dobrosavljevic´ and Kotliar, 1997). structure calculations. LaVO3 and YVO3 show C-type Some remarks should be made on the experimental and G-type antiferromagnetic ordering, respectively. observation that the ϳ1.5 eV peak of the photoemission They are also accompanied by a Jahn-Teller distortion spectra persists in xϳ1(nϳ0) samples, where the d as in RTiO3, but of the a type for LaVO3 and d type for band is nearly empty. This is an unexpected experimen- YVO3. These phenomena were explained by Hartree- tal result by itself, since in the limit of small band filling, Fock (Mizokawa and Fujimori, 1996b) band-structure nӶ1, there should be only a small probability of calculations. For the ideal cubic lattice, the lowest- electron-electron interaction. A similar peak was ob- energy Hartree-Fock solution is the C-type antiferro- served (but at somewhat lower binding energies of magnetic state in which two adjacent V atoms have 1 1 1 1 ϳ1.2 eV) for oxygen-deficient SrTiO3Ϫy (Henrich and (xy) (yz) and (xy) (zx) configurations, which favors Kurtz, 1981; Barman et al., 1996). In order to explain the a-type Jahn-Teller distortion. Here, the occupation this behavior, two different models were proposed. One of the xy orbital at every site induces antiferromagnetic was based on the potential disorder caused by the sub- coupling within the ab plane, while the alternating oc- stituted A-ion sites (Sarma, Barman, et al., 1996). The cupation of the yz and zx orbitals induces ferromag- substitution of La for Sr lowers the potential at neigh- netic coupling along the c axis. (One can ignore the boring Ti sites and, if the potential lowering is large spin-orbit interaction, unlike the case of LaTiO3, be- enough, produces a peak ϳ1 eV below EF even for a cause the Jahn-Teller distortion is substantial in the small band filling (nӶ1). However, such a strong poten- RVO3 compounds.) For the d-type Jahn-Teller distor- tial disorder may cause a localization of electrons and tion, either the yz or the zx orbitals are occupied along hence seems difficult to explain the metallic conductivity the c axis, resulting in antiferromagnetic coupling along observed for the low n (especially for La1ϪxSrxTiO3). the c direction and hence G-type antiferromagnetic or- On the other hand, Dobrosavljevic´ and Kotliar (1997) dering. Generalized density gradient band-structure cal- have shown, using dynamical mean-field theory, that culations also explained the C-type antiferromagnetism

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1172 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 112. Phase diagram of La1ϪxSrxVO3 (Mahajan et al., FIG. 114. Optical conductivity spectra at room temperature 1992; Inaba et al., 1995). for La1ϪxSrxVO3. From Inaba et al., 1995.

of LaVO but failed to explain the G-type antiferromag- 3 Figure 113 presents the data for the temperature de- netism in YVO (Sawada et al., 1996). 3 pendence of the resistivity in arc-melted polycrystals With hole doping by partial substitution of divalent ions (Sr or Ca) for the rare-earth ions, the insulating (which suffer less from grain boundary effects than ce- antiferromagnetic phase eventually disappears, as shown ramics samples) of La1ϪxSrxVO3 (Inaba et al., 1995). The conductivity in the insulating region (xϽ0.15) in Fig. 112 for the case of La1ϪxSrxVO3 (Mahajan et al., 1992; Inaba et al., 1995). As compared with the case of shows a thermal activation type at relatively higher tem- peratures above the antiferromagnetic transition tem- V2O3 (Sec. IV.A.1), doping with more nominal holes is necessary to produce the metallic state. In the case of perature TN (Fig. 112), and the thermal activation en- La1ϪxSrxVO3 the FC-MIT takes place around xcϷ0.2 ergy, Ϸ0.15 eV for LaVO3, decreases steeply but (Dougier and Casalot, 1970; Dougier and Hagenmuller, continuously with x. Variation of the activation energy 1.8 1975; Inaba et al., 1995), while in Y1ϪxCaxVO3 it takes of La1ϪxSrxVO3 is shown to follow (xcϪx) (Dougier place around xcϷ0.5 with smaller W (Kasuya et al., and Hagenmuller, 1975). Mott (1990) argued that the ␨ 1993). critical variation in the form of (xcϪx) with ␨ϭ1.8 is in contradiction to the simple scenario, i.e., ␨ϭ1, of the Anderson localization and that some kind of mass en- hancement, such as a consequence of the formation of spin , must be taken into account. At lower temperatures below TN in the region of 0.05ϽxϽ0.15, the compound shows a variable-range hopping type of conduction. Again, Sayer et al. (1975) reported that the critical exponent for the localization length is ␨ϭ0.6 in- stead of the expected value ␨ϭ1. The FC-MIT and resultant change of the electronic structure show up clearly in the optical spectra. Figure 114 shows the low-energy part of the optical conductiv- ity spectra at room temperature for La1ϪxSrxVO3 to- gether with the inset for a broader energy region (Inaba et al., 1995). In LaVO3 a Mott-Hubbard gap between the lower and upper Hubbard bands and a charge-transfer (CT) gap between the occupied O 2p band and the up- per Hubbard band are observed at 1.1 eV and 3.6 eV, respectively; these can be discerned as rises in the opti- cal conductivity in the inset of Fig. 114. With hole dop-

FIG. 113. Resistivity of La1ϪxSrxVO3. From Inaba et al., 1995. ing by substitution of Sr (x), the CT gap shifts to lower

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1173

dependence of the resistivity can be interpreted in terms of the strong Curie-like enhancement of the spin suscep- tibility ␹(Q) as derived in Eq. (2.118). This T2ϪT1.5 crossover has clearly been observed near the antiferro- magnetic metal-normal metal phase boundary, for ex- ample in the case of the BC-MIT system NiS2ϪxSex ,as described in Sec. IV.B.2. In La1ϪxSxVO3, however, it is not yet clear whether the antiferromagnetic metallic state is present around xϷ0.2. Furthermore, the mea- Ϫ1 surement of the NMR relaxation rate T1 by Mahajan et al. (1992) shows a conventional Korringa behavior and no indication of antiferromagnetic fluctuation, at least for xу0.4. Detailed neutron-scattering and NMR measurements using single-crystalline La1ϪxSrxVO3 samples in the immediate vicinity (xϷ0.2) of the MIT are obviously needed to detect and characterize possibly anomalous magnetic fluctuations near the FC-MIT. The evolution of electronic states in the R1ϪxAxVO3 FIG. 115. Resistivity of La1ϪxSrxVO3 near the MI phase system with x should be different from that in boundary as a function of T1.5. From Inaba et al., 1995. R1ϪxAxTiO3 in that the parent insulator RVO3 is a Mott insulator with d2 configuration, where there is a Hund’s energy mainly due to the shift in the 2p band, and the rule coupling between the d electrons, and the x 1 → Mott-Hubbard gap is closed. As for the doping-induced (100% hole doping) limit still has one d electron per V change in the low-energy region, the spectral weight is site. The photoemission and oxygen 1s x-ray absorption first increased in the gap region and eventually forms the spectra of RVO3 (RϭLa,Y) revealed lower and upper inner-gap absorption band. In barely metallic samples ( Hubbard bands, respectively, corresponding to the d2 xϭ0.20 and 0.22) at room temperature (see Fig. 113), d1 and d2 d3 spectral weight (Edgell et al., 1984; Ei- → → the conductivity spectra show a non-Drude-like shape, saki, 1991; Pen et al., 1998). Upon hole doping, new indicating a dominant contribution from the incoherent spectral weight grows rapidly just above EF , which cor- motion of the charge carriers. However, the spectra for responds to the d1 d2 spectral weight induced by hole → xϾ0.3 are typical of a metal. In contrast with the case of doping, as in the high-Tc cuprates. At the same time the R1ϪxAxTiO3 (Sec. IV.B.1), the spectral weight does not lower Hubbard band loses its spectral weight. As the appear to be transferred from the gap excitation but system becomes metallic (xϾ0.5 for RϭY and xϾ0.2 from a higher-lying p-d interband transition, since no for RϭLa), a Fermi edge is developed in the photoemis- isosbetic point can be observed near the Mott-Hubbard sion spectra. If one starts from the xϭ1(d1) limit of the gap energy. A similar feature of the evolution of the system La1ϪxSrxVO3 and increases the band filling, the inner-gap excitation with doping is observed for remnant of the lower Hubbard band increases its inten- Y1ϪxCaxVO3 (Kasuya et al., 1993) but at a higher dop- sity while the quasiparticle band in the photoemission ing level, probably reflecting stronger electron correla- and inverse photoemission spectra loses its intensity tion. (Tsujioka et al., 1997). The decrease of the quasiparticle Figure 115 shows the resistivity of the compounds weight as the system deviates from integer filling is op- near the metal-insulator phase boundary as a function of posite to what has been theoretically predicted for a T1.5 (Inaba et al., 1995). Apart from the upturns at low fixed U/W (Kotliar and Kajueter, 1996; Rozenberg temperatures (probably below TN), the temperature de- et al., 1996), and would be attributed to the increase in pendence of the resistivity (␳) can be well expressed by U/W with decreasing x. Here it should be noted that in 1.5 the relation ␳ϭ␳0ϩAT in the metallic region. The in- R1ϪxAxVO3 not only does the band filling n (ϵ2Ϫx) set of Fig. 115 shows a magnification of the low- decrease but also U/W decreases with x, since U/W temperature region for the xϭ0.28 sample, indicating Ͻ1 in the metallic AVO3 while U/WϾ1 in the insulat- that the above relation holds good at least down to 2.5 ing RVO3. K. The coefficient A critically increases as the metal- insulator phase boundary is approached from the C. High-Tc cuprates higher-x (metallic) region. Inaba et al. (1995) inter- 1. La Sr CuO preted this to mean that the enhanced T1.5 dependence 2Ϫx x 4 was due to antiferromagnetic spin fluctuations. The self- The discovery of high-Tc cuprate superconductors consistent renormalization (SCR) theory (Moriya, 1995; was brought about by the observation of a supercon- Ueda, 1977; see also Secs. II.E.1 and II.E.8) indicates ducting transition for the compound La2ϪxMxCuO4 with that the low-temperature resistivity near the critical MϭBa by Bednorz and Mu¨ ller (1986). The lattice struc- boundary between the antiferromagnetic metal and nor- ture of this compound was soon identified as the K2NiF4 mal metal deviates from the normal T2 law and shows type, namely, the layered perovskite structure shown in T1.5 dependence, as discussed in Sec. II.E.8. The T1.5 Fig. 116 (Takagi, Uchida, Kitazawa, and Tanaka, 1987).

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1174 Imada, Fujimori, and Tokura: Metal-insulator transitions

for a review Scalapino, 1995; Annett et al., 1996). How- ever, in this review article, we focus on the unusual normal-state properties of high-Tc compounds; proper- ties in the superconducting state are beyond the scope of this article. Before going into a detailed discussion of experimen- tal data, we first outline the basic electronic structure and the phase diagram in the plane of doping concentra- tion x and temperature T. Because of the two- dimensional anisotropy of this lattice structure, CuO6 octahedra are slightly elongated along the direction per- pendicular to the CuO2 layer. This means that the dis- tance from Cu to the apex oxygen is slightly longer than the distance to the in-plane oxygen, which may be viewed as a Jahn-Teller-type distortion. This distortion lifts the degeneracy of the eg orbitals of Cu 3d electrons to the lower-lying dz2Ϫr2 and higher-lying dx2Ϫy2 orbit- als. Because the parent compound La2CuO4 has the va- lence of Cu2ϩ, and hence as the nominal d9 state the above crystal-field splitting leads to fully occupied t2g orbitals and d3z2Ϫr2 orbital, while the dx2Ϫy2 orbital re- mains half-filled, the Fermi level lies in a band con- structed mainly from the dx2Ϫy2 orbital, relatively close to the level of the 2p␴ oxygen orbital, compared with other transition-metal oxides. In La2CuO4, since the 2p␴ band lies within the Mott-Hubbard gap, it is a charge- transfer (CT) insulator, as discussed in Secs. II.A and FIG. 116. The layered perovskite structure of La2ϪxMxCuO4; large open circles, La(M); small open circle, Cu; large filled III.A. La2CuO4 is a typical insulator with antiferromag- circle, O. netic long-range order. Up to the present time, all the cuprate superconductors with high critical temperatures are known to be located near the antiferromagnetic Although a large number of compounds have been dis- Mott insulating phase. La2CuO4 is, in fact, a Mott insu- covered as high-Tc cuprate superconductors since then, lator with the Ne´el temperature TNϳ300 K. Because they all have a common structure of stacking of CuO2 the dx2Ϫy2 band is half-filled, it is clear that the insulat- layers sandwiched by block layers, as shown schemati- ing behavior experimentally observed even far above TN cally in Fig. 117. The superconductivity itself has been is due to strong correlation effects. an attractive and extensively studied subject. The sym- La2ϪxMxCuO4 with MϭSr and Ba provides a typical metry of the pairing order parameter of the high-Tc cu- example, with a wide range of controllability of carrier prates now appears to be settled as the dx2Ϫy2 symmetry concentration, from the antiferromagnetic Mott insulat- after enormous combined effort involving magnetic ing phase to overdoped good metals, by the doping of resonance experiments, Josephson junctions, etc. (see Sr, Ba, or Ca. Therefore La2ϪxMxCuO4 is a good ex- ample of an FC-MIT system with 2D anisotropy. The basic electronic structure of the cuprate superconductors in all regions, from insulator to metal, is believed to be described by the d-p model defined in Eqs. (2.11a)– (2.11d). The parameters deduced from photoemission (Shen et al., 1987; Fujimori, Takekawa, et al., 1989), LDA calculations (Park et al., 1988; Stechel and Jenni- son, 1988; Hybersten, Schlu¨ ter, and Christensen, 1989; McMahan, Martin, and Satpathy, 1989), and ␴(␻) give overall consistency with UddӍ6–10 eV, tpdӍ1–1.5 eV, ⌬Ӎ1 – 3 eV, UppӍ1 eV, and Vpdϳ1–1.5 eV. The de- tails of methods for determining the model parameters are given in Sec. III.B. With increasing x,La2ϪxSrxCuO4 very quickly under- goes a transition from an antiferromagnetic insulator to a paramagnetic metal with a superconducting phase at low temperatures. The phase diagram of La2ϪxSrCuO4 FIG. 117. Conceptual illustration of stacking of CuO2 layers is shown in Fig. 118. At sufficiently low temperatures, it sandwiched by block layers. is believed that the MIT with x is actually a

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1175

FIG. 119. The structure of the antiferromagnetic order in La2ϪxSrxCuO4. From Yamada et al., 1989.

microscopic level was discussed by Fukuzumi et al., (1996). Below we discuss physical properties in the plane of x and T in more detail. The existence of antiferromagnetic long-range order below TNϳ300 K at xϳ0 is clearly ob- FIG. 118. The phase diagram of La2ϪxSrxCuO4 in the plane of served in neutron-scattering experiments as well as in the doping concentration x and temperature T. The Ne´el tem- magnetic susceptibility measurements. Other quantities perature TN , superconducting transition temperature Tc , such as NMR and Raman scattering also show anoma- transition temperature between high-temperature tetragonal lies. The structure of the antiferromagnetic order was (HTT) and low-temperature orthorhombic (LTO) structures, determined by neutron-scattering experiments, as shown TTO , and spin-glass transition temperature Ts-glass are plotted in Fig. 119. The driving force of the antiferromagnetic from various experimental data (by courtesy of T. Sasagawa order is of course the superexchange interaction of and K. Kishio). neighboring Cu spins mediated by intermediate states of the bridging 2p␴ orbital. The localized spins at the Cu superconductor-insulator transition. Above the super- site have a moment ϳ0.4␮B , which is a consequence of conducting transition temperature, the normal metallic the reduction due to a strong quantum effect and strong phase shows various unusual properties which are far covalency. The ordered spins are primarily aligned in from the standard Fermi-liquid behavior, especially for the basal CuO2 plane (Vaknin et al., 1987; Yamada small x. These anomalous aspects will be described in et al., 1987). To understand the detailed structure of the detail below. The superconducting transition tempera- ordering vector, however, we need to consider small tilt- ture Tc has the maximum Tcϳ40 K around xϳ0.15 for ing of the CuO6 octahedron as well as the spin-orbit La2ϪxSrxCuO4 and decreases with further doping. The interaction to generate a small Dzyaloshinski-Moriya in- doping concentration xϭxm which shows the maximum teraction, which leads to an anisotropic interaction for Tc is frequently called the optimum concentration while the ordering vector. This leads to a slight canting of the the region xϽxm is called the underdoped region and spin vector out of the CuO2 plane with a small ferromag- xϾxm the overdoped region. Anomalous metallic be- netic moment. (Fukuda et al., 1988; Kastner et al., 1988; havior observed in the underdoped region gradually Thio et al., 1988; Yamada et al., 1989; Koshibae, Ohta, crosses over to more or less standard Fermi-liquid be- and Maekawa, 1993, 1994). havior in the overdoped region. Another point to be Aside from this slight anisotropy, the basic magnetic discussed in the global phase diagram is the existence of interaction in the Mott insulating phase is well described structural phase transitions between the tetragonal within the framework of the spin-1/2 Heisenberg model. structure at high temperatures and the orthorhombic The antiferromagnetic exchange interaction J deter- one at low temperatures, as illustrated in Fig. 118. At mined from two- Raman scattering is J xϭ0, the transition occurs at around 560 K and it ϳ1100 cmϪ1ϭ1600 K (Lyons, Fleury, Remeika, Coo- quickly decreases with increasing x. In the orthorhombic per, and Negran, 1988; Cooper et al., 1990b; Tokura, Ko- phase, the unit of octahedron CuO6 is slightly tilted in a shihara, et al., 1990), while Jϭ135 meV, equivalent to staggered way. In the overdoped region, xϾxm , Tc de- 1570 K, is identified from the dispersion in the neutron- creases continuously at least up to xϭ0.20 (see the scattering data (Aeppli et al., 1989), which has one of the phase diagram). On the other hand, the possibility of largest antiferromagnetic exchanges among d-electron phase separation and a discontinuity between under- systems. There are two reasons for the large J. One is doped and heavily doped regions around xϭ0.20 on a that the transfer between the neighboring Cu dx2Ϫy2 or-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1176 Imada, Fujimori, and Tokura: Metal-insulator transitions

bital and O2p␴ orbital tpd is large, ϳ1 eV because of the large overlap of these two orbitals. The second is that the level splitting ⌬ϭ͉␧dϪ␧p͉ is relatively small, ϳ1 – 3 eV, as shown in Eqs. (2.11a)–(2.11d) and the paragraph below these equations. Then in the fourth- order perturbation, J may be large because it is propor- 4 3 tional to tpd/͉␧dϪ␧p͉ when the Coulomb interaction at the oxygen site, Upp , and the intersite Coulomb inter- action Vpd are neglected while Udd is taken large enough. The Ne´el temperature TN is much smaller than J because of the strong two-dimensionality and resulting quantum fluctuation effects. In fact, in single-layer 2D systems, TN is theoretically expected to be zero, while small interlayer exchange coupling drives TN to a finite temperature. Short-ranged antiferromagnetic correla- tions are developed even well above TN , as observed in neutron-scattering experiments and NMR measure- ments. The growth of antiferromagnetic correlations with decreasing temperature is qualitatively described both by a renormalization-group study of the nonlinear sigma model (Chakravarty, Halperin, and Nelson, 1989) and by a modified spin-wave theory (Takahashi, 1989) as discussed in the neutron-scattering results (Aeppli et al., 1989; Yamada et al., 1989; Keimer et al., 1992). (The prefactor of the exponential dependence is not repro- duced by the spin-wave theory). At low temperatures above TN , as discussed in Sec. II.E.4, the antiferromag- netic correlation length is expected to grow as FIG. 121. Short-ranged incommensurate spin correlation ob- served by neutron-scattering experiments for Im ␹(q,␻)in 0.27បc 2␲␳s T La Sr CuO above T at ␻ϭ3 meV taken from the scan A ␰ϳ exp 1Ϫ ϩ (4.7) 2Ϫx x 4 c 2␲␳ ͩ T ͪͩ 4␲␳ ¯ͪ illustrated in the top panel. The closed circles in the top panel s s denote the peak positions of the incommensurate structure. J with the spin stiffness constant ␳sӍ1.15 (Chakraverty, The incommensurate peak observed above Tc (top panel) be- Halperin, and Nelson, 1989; Hasenfratz and Nieder- comes invisible below Tc (lower panel) at this energy ␻ mayer, 1991, 1993). The experimentally observed tem- ϭ3 meV. From Yamada et al., 1995. perature dependence of ␰ appears to agree with the above prediction of the nonlinear sigma model in the renormalized classical regime (Keimer et al., 1992; Bir- spin-wave theory of the Heisenberg model (Hayden geneau et al., 1995). The energy dispersion of spin exci- et al., 1991b; S. Itoh et al., 1994). tations determined by high-energy inelastic neutron- When Sr, Ba, or Ca are substituted for La, as in scattering experiments shows overall agreement with the La2ϪxAxCuO4 with AϭSr or Ba, holes are doped into CuO2 planes. A remarkable feature of the hole doping is that it causes a quick collapse of the antiferromagnetic order and a MIT. When the doping concentration is low enough, the system is still insulating for xр0.05, as illus- trated in the phase diagram Fig. 118, while antiferro- magnetic long-range order is destroyed around x ϭ0.02, where the localization effect due to disorder is serious. For 0.02ϽxϽ0.05, the spin-glass state is realized at low temperatures (see, for example, Chou et al., 1995). In any case, antiferromagnetic long-range order is absent in the metallic region, xϾ0.05. The antiferromag- netic correlation length ␰ becomes progressively shorter and shorter when the doping concentration is increased. In the early stage of neutron studies on La2ϪxSrxCuO4, it was shown that ␰ is approximately given by ␰ϳa/ͱx, where a is the lattice constant, and hence ␰ is nothing FIG. 120. Doping concentration dependence of antiferromag- but the mean hole distance, as shown in Fig. 120 (Birge- netic correlation length ␰ suggested by neutron scattering of neau et al., 1988). La2ϪxSrxCuO4 (Birgeneau et al., 1988). ␰ appears to follow the In the metallic region xϾ0.05, the periodicity of mean hole distance 3.8 Å/ͱx. short-ranged antiferromagnetic correlations is known to

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1177

FIG. 122. The doping concentration (x) dependence of the incommensurability ␦ [deviation of the peak position from (␲,␲)] in spin correlations observed in neutron-scattering ex- periments for La2ϪxSrxCuO4 (Yamada et al., 1998). Incom- mensurability appears in the metallic phase. The inset shows FIG. 123. Single-parameter scaling of ͐ Im ␹(q,␻)dq for vari- the half width at half-maximum ⌫. ous choices of T and ␻ for La1.96Sr0.04CuO4. From Keimer et al., 1992. shift from (␲,␲) to fourfold-incommensurate wave vec- tors ͓␲(1Ϯ␦),␲͔ and ͓␲,␲(1Ϯ␦)͔, as can be seen in An overall feature of Im␹(q,␻) for La2CuO4 and Fig. 121 (Yoshizawa et al., 1988; Thurston et al., 1989, La1.85Sr0.15CuO4 up to high energies was measured by 1992; Cheong et al., 1991; Yamada, Endoh, et al., 1995; high-energy pulsed neutron inelastic scattering (Ya- Yamada, Wakimoto, et al., 1995). The hole doping mada, Endoh, et al., 1995; Hayden et al., 1996). The quickly increases ␦ at xϾ0.05 upon metallization; it magnetic scattering intensities around (␲,␲) due to anti- seems that ␦ approaches the value ␦ϭx at 0.05рx ferromagnetic fluctuations were clearly observed up to р0.1, as can be seen in Fig. 122 (Yamada et al., 1998). ␻ϭ120 meV. In addition, when the nonmagnetic com- At xϭ0.15 and 0.075, inelastic neutron-scattering ex- ponent was subtracted, the magnetic signals remained periments show that the width ⌬, the peak height up to 300 meV. S(Q,␻), and the intensity ͐dq S(q,␻) around the in- Here, it should be noted that a careful analysis of the commensurate peak depend on temperature for the en- neutron and NMR data is required for interpretation of ergy ␻Ͻ4 meV, that is, the peak grows and is sharpened spin fluctuations. As clarified in Secs. II.E.4 and II.F.10, with lowering temperature when measured at T the dynamic spin susceptibility is phenomenologically Ͻ100 K, while the incommensurate peak structure is given as the sum of coherent and incoherent contribu- more or less independent of temperature for ␻ tions. Although the coherent part may depend on tem- у4 meV (Cheong et al., 1991; Thurston et al., 1992; Ya- perature due to the Fermi-surface effect, the antiferro- mada, Wakimoto, et al., 1995) above Tc . This suggests magnetic correlation length ␰ must be determined from that rather sharp incommensurate peaks grow only in the incoherent part, and it appears to be temperature the low-energy range ␻Ͻ4 meV at xϳ0.15 due presum- insensitive, as discussed in Sec. II.E.4. The incoherent ably to a Fermi-surface effect combined with imperfect contribution becomes more and more dominant over the but finite nesting effects of the Fermi surface. This coherent contribution as the doping concentration is growth of incommensurate peaks may be ascribed to the lowered. growth of the coherent part of the magnetic response, as At low doping concentration, ͐Im␹(q,␻)dq follows a discussed in Secs. II.E.4 and II.F.10. Recent studies us- single-parameter scaling ing high-quality samples show that the incommensurate peak decreases and disappears with decreasing tempera- ͵ Im ␹͑q,␻͒dqϭg͑␤␻͒, (4.8) tures below Tc at low energies (Mason et al., 1992, 1993; Yamada, Wakimoto, et al., 1995). This is consistent with with g(x)ϳ1ϪeϪx or tanh x as in Fig. 123 (Keimer the opening of a superconducting gap at the incommen- et al., 1992; Hayden et al., 1991a). The same scaling was surate wave number. The wave-number dependence ap- also observed in a YBa2Cu3O7Ϫy compound, as dis- pears to be consistent with dx2Ϫy2 symmetry of the su- cussed in Sec. IV.C.3. As discussed in Secs. II.E.4 and perconducting gap. Another remarkable feature below II.G.2, this scaling is consistent with both a marginal Tc is that the incommensurate peak grows and becomes Fermi liquid (see Sec. II.G.2; Varma et al., 1989) and a more prominent and sharpened on the high-energy side view in terms of a totally incoherent response as in Eqs. (տ7 meV) with very long correlation length (Ͼ50 Å; (2.294) and (2.362) (see Secs. II.F.10 and II.E.4; Imada, Mason et al., 1996). This may be related to the compat- 1994a; Jaklicˇ and Prelovsˇek, 1995a). It should be noted ibility of antiferromagnetic correlation, especially in the that this single-parameter scaling is experimentally not d-wave superconductor (see, for example, Assaad, satisfied at higher doping at low temperatures (Matsuda Imada, and Scalapino, 1997; Assaad and Imada, 1997). et al., 1994). A general tendency appears to be that the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1178 Imada, Fujimori, and Tokura: Metal-insulator transitions single-parameter scaling by ␤␻ is satisfied in the region ization of the 4s orbital and the Wannier orbital of of incoherent charge dynamics at low doping or at high dx2Ϫy2 extended to the neighboring site, as proposed by temperatures, namely at TϾTcoh (see Fig. 34), which is Mila and Rice (1989). The q dependence of the nuclear consistent with the region of incoherent charge and spin form factor Aq has a large weight around qϭ(␲,␲) for dynamics discussed in the theoretical argument in Secs. the Cu site, while the weight around qϭ(␲,␲) is vanish- II.E.4 and II.F.10. In addition, it is clear that this single- ing in the case of Aq at the in-plane O site. At the Cu parameter scaling is not satisfied when the supercon- site, the nuclear form factor is such that the antiferro- ducting gap (or pseudogap with nodes) is open, as is magnetic fluctuation around (␲,␲) is selectively picked observed by (Yamada, Wakimoto, et al., 1995). The scal- up in 1/T1 . ing appears not to be satisfied in the pseudogap region Comprehensive measurements of the NMR nuclear above T in the Y compounds, as discussed in Sec. spin-lattice relaxation time T1 at the Cu site were per- c formed in the range 0 xϽ0.15 and T 900 K, as shown IV.C.3. However, the boundary of the region of single- р р in Fig. 124 (Imai et al., 1993). At the moment its inter- parameter scaling by ␤␻ in the plane of x and T has not pretation is not unique. However, at least several points yet been completely clarified experimentally, although it should be noted concerning on the T results. For the is important to establish the correct interpretation and 1 mother material La CuO ,1/T shows striking enhance- theory. 2 4 1 ment with decreasing temperature, in agreement with When the doping concentration x is around 1/8, the exponential growth of the antiferromagnetic corre- La Ba CuO and La Sr CuO are known to show 2Ϫx x 4 2Ϫx x 4 lation length ␰ in the renormalized classical regime of anomalies in their transport properties. In particular, the nonlinear sigma model. As doping proceeds, the en- La Ba CuO with xϳ0.125 shows a marked decrease 2Ϫx x 4 hancement of 1/T is rapidly suppressed, giving way to in T as compared with the other region, and it becomes 1 c more or less temperature independent 1/T for T even insulating in the region close to xϭ0.125. This may 1 Ͼ100– 200 K with gradual crossover to 1/(T T)ϳconst be due to the charge ordering of doped holes accompa- 1 at lower temperatures for xϳ0.1– 0.15. This temperature nied by spin ordering, as observed by muon (Kumagai dependence has been interpreted from several different et al., 1994), nuclear quadrupole resonance, and viewpoints. One interpretation is that 1/(T T) follows neutron-scattering measurements. The periodicity of the 1 the Curie-Weiss law 1/(T T)ϳC/(Tϩ⌰) [see Eq. charge ordering is commensurate to the CuO lattice 1 2 (2.115)] which comes from a temperature-dependent an- structure. In particular, recent neutron measurements tiferromagnetic correlation length ␰ϳ 1/ͱTϩ⌰, as dis- for La Nd Sr CuO at around xϭ0.12 showed that 1.6Ϫx 0.4 x 4 cussed in Secs. II.D.1 and II.D.8 and stressed by spin- a commensurate charge- and spin-ordered stripe phase fluctuation theories (Millis, Monien, and Pines, 1990; is stabilized (Tranquada, Sternlieb, et al., 1995). Moriya, Takahashi, and Ueda, 1990). Another is the in- Nuclear magnetic resonance provides us with another terpretation from marginal Fermi-liquid theory (see Sec. powerful tool for investigating spin correlations in II.G.2) where Im ␹ϳtanh ␤␻ is consistent with 1/T La M CuO . The longitudinal relaxation time of 1 2Ϫx x 4 ϳconst for TϾ200 K at ␦ϳ0.1– 0.15. The NMR data are nuclear magnetization, T , is inversely proportional to 1 also consistent with the discussion of the neutron data the dynamic spin susceptibility as above when we take the limit ␻0 0. A completely dif- 1 ␥2 k T Im ␹ q,␻ → n B ␣ 2 ␣␣͑ 0͒ ϭ 2 ͚ ͚ ͉Aq͉ , (4.9) T1 2␮B q ␣ϭx,y ␻0 where ␻0 is the frequency of the NMR field applied in the z direction. The frequency ␻0 is usually very small and it may be viewed as the limit ␻ 0 for the magnetic response in which we are interested.→ The shift of the NMR frequency due to the hyperfine interaction be- tween nuclei and electrons, called the Knight shift K,is given as the sum of the orbital part Korb and the spin part Kspin . The spin part is given by 1 Kspinϭ Aqϭ0Re ␹͑qϭ0,␻ϭ0͒, (4.10) ␮B where ␹(qϭ0,␻ϭ0) is the uniform magnetic suscepti- bility. Korb is usually assumed to be temperature inde- pendent. For the isolated Cu2ϩ ion, anisotropy of the dx2Ϫy2 wave function causes very anisotropic Korb , FIG. 124. Temperature dependence of inverse of NMR which well accounts for the experimental data (Taki- nuclear spin-lattice relaxation time T1 at the Cu site for gawa et al., 1989). The nuclear form factor A is deter- q La2ϪxSrxCuO4: filled circle, xϭ0; open circle, xϭ0.04; open mined from the hyperfine interaction. The anisotropy square, xϭ0.075; filled square, xϭ0.15. The inset shows the c ab between Aqϭ0 and Aqϭ0 is accounted for by considering scaling in the insulating phase. From Imai, Slichter, a transferred hyperfine interaction through the hybrid- Yoshimura, and Kosuge, 1993.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1179 ferent interpretation related to the marginal Fermi- liquid hypothesis is that the dynamic structure factor S(q,␻) is basically temperature independent and the temperature dependence of Im ␹(q,␻)ϭ(1ϪeϪ␤␻)S(q,␻) comes solely from the Bose factor 1ϪeϪ␤␻, as suggested from the temperature-independent ␰ and the incoherent character of spin correlations in numerical studies (Imada, 1994a; Jaklicˇ and Prelovsˇek, 1995b). This view- point is discussed in Secs. II.E.4 and II.F.10. The cross- over from 1/T1ϳconst to 1/(T1T)ϳconst has been at- tributed to the crossover from the incoherent charge and spin dynamics to coherent dynamics, which are sepa- rated by an anomalously suppressed coherence tempera- ture Tcoh as discussed in Sec. II.F.10. 1/T1 at site O does not show appreciable influence from antiferromagnetic fluctuations because of the vanishing contribution from the part around (␲,␲), as is evident from the nuclear FIG. 125. Temperature dependence of the uniform magnetic form factor of the oxygen site. This is more or less well susceptibility of La2ϪxSrxCuO4 for various x. From Nakano et al., 1994. described by Korringa-type behavior 1/(T1T)ϳconst over a wide range of temperatures. However, a puzzle remains in the temperature dependence when the peak 1989; Takagi, Uchida, and Tokura, 1989; Nakano et al., is at deeply incommensurate wave numbers, because the 1994). The peak structure and decrease at low tempera- NMR at Cu and O sites both show a substantial contri- tures appear to be more pronounced in the low-doping bution from those incommensurate wave numbers, while region ␦ϳ0.1 than in the undoped case. The doping de- the temperature dependences of those sites are differ- pendence of the peak temperature T is similar to the ent. ␹ crossover temperature TH where the Hall coefficient The Gaussian decay rate 1/TG associated with the shows a sharp increase with decreasing temperature, as nuclear spin-spin relaxation time measures the real part discussed below. As a possible interpretation, this of the static spin susceptibility ␹(q,␻ϭ0). In particular, anomaly was analyzed as the effect of preformed singlet 1/TG at the Cu site was measured in Tl2Ba2CuO6ϩ␦ , pairing fluctuations or antiferromagnetic fluctuations YBa2Cu3O7Ϫy , and YBa2Cu4O8, which selectively picks (Imada, 1993a, 1994b) which start below T␹ in the me- up ␹(Q) with the staggered wave vector Qϭ(␲,␲). Ex- tallic region as is discussed in Sec. IV.C.3. cept for the pseudogap region discussed below, the scal- 2 This ‘‘pseudogap’’ behavior is seen in other quantities ing 1/(T T)ϰ␹(Q)ϳ(1/TG) was suggested in a limited 1 such as the specific heat ␥. From far above Tc , ␥ and temperature range around 200 K (Y. Itoh et al., 1994b; hence the electronic entropy itself are anomalously sup- Y. Itoh, 1994). This behavior again can be interpreted in pressed in the low-doping region as in Fig. 126. (Loram two ways. From the viewpoint of spin-fluctuation theo- et al., 1997). The crossover temperature T␥ is above 300 ries, this scaling is due to Eqs. (2.114) and (2.115), where Ϫ1 K at low doping. This suppression of ␥ at low tempera- ␹(Q)ϰ(T1T) is satisfied in 2D. The other interpreta- tures TӶT␥ in the underdoped region was also pointed tion is that both ␹(Q,␻ϭ0) and Im␹(q,␻)/␻ are basi- cally scaled by the above-mentioned Bose factor Ϫ␤␻ lim␻ 0͓(1Ϫe )/␻͔, which leads to similar behavior between→ ␹(Q) and Im␹(Q,␻). This is because the factor ␤ also remains for Re␹ after the Kramers-Kronig trans- formation for T not much smaller than the frequency cutoff for the main contribution of Im ␹. Since TG and T1 do not seem to follow a universal relation over a wide temperature range, the scaling does not uniquely determine the mechanism. The uniform magnetic susceptibility ␹(qϭ0,␻ϭ0) can be measured by bulk magnetization measurements as well as by the Knight shift. In an undoped compound, it has a pronounced peak at high temperatures у1000 K and decreases with decreasing temperatures, in agree- ment with the behavior of the 2D Heisenberg model (Okabe and Kikuchi, 1988). When carriers are doped, the peak temperature T␹ decreases (T␹ϳ400– 500 K at FIG. 126. Temperature dependence of specific heat ␥ for ␦ϳ0.15). However the peak structure with a rather re- La2ϪxSrxCuO4. The numbers indicate the doping concentra- markable decrease at lower temperatures is retained, as tion in % (Loram et al., 1997). The entropy is released at shown in Fig. 125 (Johnston, 1989; Takagi, Ido, et al., higher temperatures at lower dopings.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1180 Imada, Fujimori, and Tokura: Metal-insulator transitions

dependence, ␳ϳ␳0ϩAЈT, in a wide range from Tc to 800 K, as shown in Fig. 127. The residual resistivity ␳0 appears to vanish for samples of higher quality. This T-linear behavior is widely observed in other high-Tc cuprates around optimum doping. In the overdoped re- gion, the resistivity roughly follows the form p ␳ϭ␳0ϩAЉT , (4.11) with 2уpϾ1, where p increases with increasing doping (Takagi, Batlogg, et al., 1992a). In the underdoped re- gion, ␳ has a feature around T␹ , in which the linear law ␳ϳ␳0ϩAЈT changes slope: AЈ increases and ␳0 de- creases in going from the high-temperature region T ϾT␳ӍT␹ to below T␳ , as shown in Fig. 127. Recent measurements of ␳ at low temperatures in the under- doped region revealed that ␳ eventually shows insulating behavior (logarithimic T dependence) at sufficiently low temperatures if the superconductivity is suppressed (Ando et al., 1995). Various mechanisms for this T-linear behavior have been proposed. We discuss here only some of these sug- gestions. Although the mechanism is not yet fully under- stood, it is useful to list some of the proposed ap- proaches to make connections with the recent extensive theoretical efforts discussed in Sec. II. From the Fermi-liquid point of view, the spin fluctua- tion theory discussed in Secs. II.D.1 and II.D.8 ascribes this behavior to control of the carrier relaxation time ␶ by the scattering of carriers (quasiparticles) by paramag- netic excitations. The T-linear resistivity in this ap- proach is obtained in Eq. (2.121) and, as discussed in the FIG. 127. Temperature dependence of the in-plane resistivity derivation of (2.121), is ascribed to the contribution of for various doping concentrations of La2ϪxSrxCuO4 for single antiferromagnetic excitations which follows the Curie- and poly crystals (Takagi et al., 1992a) T-linear resistivity is Weiss form (2.114) (Millis, Monien, and Pines, 1990; seen in a wide region. Moriya, Takahashi, and Ueda, 1990). Near the critical point, ⌰ϳ0, Im ⌺ and hence the inverse relaxation time out by Wada, Nakamura and Kumagai (1991) and by of the quasiparticles becomes proportional to T follow- Nishikawa, Takeda, and Sato (1994). It is in striking con- ing the form of Eqs. (2.110), (2.111), and (2.114). trast with the critical enhancement of ␥ for (La,Sr)TiO3 In the scenario of spin-charge separation, discussed in discussed in Sec. IV.B.1. From the specific-heat data, a Sec. II.D.7, the slave-boson approximation with cou- large entropy remaining above T␥ՇT␹ is suggested from pling between spinons and holons treated by the gauge the picture of a mass-diverging Mott transition at higher field leads to the T-linear resistivity due to the incoher- temperatures with crossover to vanishing carrier number ent character of the holons. (Ioffe and Wiegmann, 1990; below T␥ due to the formation of a pseudogap, as dis- Ioffe and Larkin, 1989; Lee and Nagaosa, 1992). The cussed in Sec. II.F (Imada, 1993a, 1994b). This means a T-linear resistivity is also obtained from marginal crossover of the scaling for the MIT between two differ- Fermi-liquid theory, where Im ⌺ is indeed proportional ent universality classes. We discuss the pseudogap be- to T, as discussed in Sec. II.G.2. havior and the significance of T␹ and T␥ in greater detail As discussed in Sec. II.E.1, recent numerical results of for the case of Y compounds in Sec. IV.C.3. an exact diagonalization for the 2D t-J model by Jaklicˇ The transport properties of this compound are also and Prelovsˇek (1995a) succeeded in reproducing this unusual in terms of standard Fermi-liquid theory (see, T-linear behavior; the prefactor AЈ was in quantitative for other reviews, Iye, 1990; Ong, 1990). One of the most agreement with the experimental data shown in Fig. 127, unusual properties is the temperature dependence of the although the temperature achieved by the numerical resistivity. As we have discussed in Sec. II.D.1, standard method was at least a few hundred K. It was suggested Fermi-liquid theory predicts that the temperature de- that this T-linear resistivity could be due solely to inco- 2 pendence of the resistivity ␳ is given as ␳ϭ␳0ϩAT [see herent charge dynamics, which should be observed Eq. (2.83)] with correlation effects contained in the T2 above the coherence temperature of the carriers. The term. In contrast to this, this compound especially conductivity is in fact consistent with incoherent charge around optimum doping xϳ0.15 shows robust T-linear dynamics of the form (2.274) with a broad C(␻), of the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1181

FIG. 129. Temperature dependence of the Hall coefficient for FIG. 128. Doping concentration dependence of the Hall coef- La2ϪxSrxCuO4 (Nishikawa, Takeda, and Sato, 1994). The thin line represents the nominal value of 1/nec at nϭ1 per Cu. ficient at low temperatures for La2ϪxSrxCuO4 and Strong temperature dependence is seen in the underdoped re- Nd2ϪxCexCuO4. Shaded windows show the superconducting gion. phase at low temperatures (Uchida et al., 1989). RH roughly follows 1/x in the underdoped region. ever, the large amplitude in the underdoped region is order of an eV, which implies that the current-current hard to explain from this approach. It has turned out correlation decays very quickly over the time scale that the temperature dependence of RH is even more Ϫ1 eV . Such anomalous suppressions of coherence are peculiar because a large positive RH quickly decreases consistent with the anomalous criticality of the Mott above a certain crossover temperature THӍT␹ ,as transition with the large dynamic exponent zϭ4, as dis- shown in Fig. 129 (Nishikawa, Takeda, and Sato, 1993, cussed in Sec. II.E.9 (Imada, 1995b; see Sec. II.F.9). 1994; Hwang et al., 1994). The crossover temperature The anisotropy of the resistivity shows another un- TH is close to T␹ for the uniform susceptibility and both usual property. The resistivity in the CuO2 plane, ␳ab ,is show similar x dependence. Above THӍT␹ , RH decays smaller by a factor of 200 than that along the c axis to a small amplitude consistent with the large Luttinger (out-of-plane), ␳c , at optimum doping in the room- volume given by 1Ϫx. As discussed above, this cross- temperature region (Ito et al., 1991). The anisotropy in- over at TH was predicted to occur due to the gradual creases for smaller x as well as for smaller T. Because formation and fluctuations of a preformed singlet pair the LDA calculation predicts a mass anisotropy of only below TH , as discussed in Sec. II.E.1 (Imada, 1993a, 28 (Allen, Pickett, and Krauer, 1987), this large anisot- 1994b, 1995b), although the interpretation of the ob- ropy has to be ascribed to strong electron correlations. served behavior has not been uniquely established. In Even more remarkable is that, in the low-doping and other words, the crossover at TH reflects a change in the low-temperature region, there exists a region where ␳ab character and the universality class of the Mott transi- shows metallic while ␳c shows insulating temperature tion at small x, as discussed above for the anomalies of dependence. Roughly speaking this region is TϽT␳ .We the specific heat T␥ and the susceptibility T␹ . Another discuss the anisotropy of ␳ further in Sec. IV.C.3. interpretation of the existence of TH is the anomalous The Hall coefficient RH also shows anomalous char- due to skew scattering arising from antiferro- acter in terms of standard Fermi-liquid theory. In the magnetic fluctuations. The effect of nesting has been dis- standard theory of weakly correlated systems, RH is in- cussed as the source of an anomalous contribution to RH dependent of temperature, and the amplitude is given by through a process similar to the Aslamazov-Larkin pro- RHϳ1/nec with carrier density n. However, at low tem- cess in superconducting fluctuations (Miyake and peratures, the observed RH in this compound is hole- Narikiyo, 1994a). like, that is, positive with large amplitude roughly scaled The optical conductivity ␴(␻) shows a spectacular by 1/x for xϽ0.3, as shown in Fig. 128 (Ong et al., 1987; change with increasing doping concentration, as illus- Takagi, Ido, et al., 1989). This contradicts a naive expec- trated in Fig. 130 (Uchida, et al., 1991). A charge- tation from the Fermi-liquid picture because the carrier transfer gap of the order of 1.5–2 eV is clearly observed number is given by 1Ϫx when the Luttinger theorem is in La2CuO4, although it is accompanied by a tail up to satisfied and the carrier should be electronlike. The sign ϳ1 eV whose origin is not uniquely identified. Upon of RH is determined from the average of the curvature doping, the spectral weight is transferred from the re- over the Fermi surface in the weak-correlation picture gion above the CT gap to the lower-frequency region. In and therefore the sign itself can be determined from a the low-doping region, it appears to develop a broad subtle balance of the electronic band structure. How- structure around or below 0.5 eV, which is replaced by

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1182 Imada, Fujimori, and Tokura: Metal-insulator transitions

ing gap. Since the c-axis reflectivity is lower, it is easier to distinguish from the gap region with 100% reflectiv- ity. Because the cuprate superconductors can be viewed as layered superconductors of CuO2 sheets coupled weakly by Josephson tunneling through the interlayer coupling, the frequency ␻p of this Josephson- coupled system may have a smaller value than the su- perconducting gap ⌬ if the Josephson coupling is weak. This situation, ␻pϽ⌬, is in sharp contrast with the usual case where ␻pϾ0.5 eVӷ⌬ is satisfied. Because ␻p is in the gap region, the Josephson plasma oscillation hardly decays. Various exotic electromagnetic phenomena due to this low ␻p were predicted by Tachiki, Koyama, and Takahashi (1994). Below we shall discuss the electronic structure studied by first-principles calculations together with results of spectroscopic measurements. This compound is one of FIG. 130. The in-plane optical conductivity ␴(␻) at room tem- the hardest cases for reproducing the correct electronic perature for various choices of x in La2ϪxSrxCuO4. From states by band-structure calculation because of strong- Uchida et al., 1991. correlation effects and strong p-d covalency. In order to explain the antiferromagnetic insulating state of the par- an overall 1/␻ dependence upon further doping. To un- ent compound La2CuO4, standard LDA calculations are derstand the ␻ dependence, two phenomenological ap- insufficient. The insulating state was correctly predicted proaches have been attempted using extended Drude by LDAϩU calculations (Anisimov et al., 1992), self- analyses (Thomas et al., 1988; Collins et al., 1989; see for interaction-corrected (SIC)-LDA calculations (Svane a review Tajima, 1997). In the first approach, ␴(␻)isfit and Gunnarsson, 1988a; Temmerman et al., 1993), and with a single-oscillator model, while in the other ap- parametrized Hartree-Fock calculations (Grant and Mc- proach a two-oscillator model has been employed. The Mahan, 1991). The effect of hole doping on an antifer- advantage of the single-oscillator model is that the romagnetic insulator, however, is usually beyond the parameters—namely, the effective plasma frequency scope of these band-structure calculations. The elec- 2 ␻p*ϭ4␲ne /m*(␻) and the relaxation time ␶(␻)—are tronic structure of La2ϪxSrxCuO4 with static (self- uniquely determined by assuming an ␻-dependent effec- trapped) holes was studied by Anisimov et al. (1992) us- tive mass m* and ␶, while the interpretation of the ␻ ing the LDAϩU method applied to supercell dependence is not straightforward for doped samples calculations. They showed that, even without lattice dis- because of the mid-IR structure. On the other hand, the tortion, a hole is self-trapped within a small ferromag- two-oscillator model fits the data better, while the pa- netic region where antiferromagnetic order is destroyed rameters are not uniquely determined. Because the by the doped hole. Then a localized state is split off from mid-IR peak is absent in an untwinned Y compound, the the top of the valence band into the band gap. In spite of two-oscillator model probably needs to take account of the structural simplicity and the presence of a well- the impurity localization effect and polaron effects defined parent insulator, La2ϪxSrxCuO4 is complicated (Thomas et al., 1991). In any case, the overall 1/␻ depen- in that doped holes enter not only the px,y-dx2Ϫy2 anti- dence and the 1/T dependence at ␻ϭ0 are consistent bonding orbitals of the CuO2 plane but also the pz or- with the form (2.274) and (2.275) with weak ␻ and T bitals of the apical oxygen hybridized with the Cu 3dz2 dependence for C(␻). The consistency of this form with orbital, as suggested theoretically (Kamimura and Suwa, numerical results is discussed in Sec. II.E.1. The connec- 1993) and observed experimentally (Chen et al., 1992). tion to the scaling theory of the Mott transition and the Experimentally, hole doping in La2CuO4 induces a marginal Fermi-liquid theory are given in Secs. II.F.9 new spectral weight (the so-called ‘‘in-gap spectral and II.G.2, respectively. An important point is that ex- weight’’) within the charge-transfer gap, as first revealed perimental data that are well described by the form by oxygen 1s core absorption spectroscopy (Romberg (2.274) with weakly ␻-dependent C(␻) should be the et al., 1990; Chen, Sette, et al., 1991). The in-gap spectral consequence of incoherent charge dynamics with the weight is distributed mostly above EF , coming largely complete lack of the true Drude weight. We discuss this from the doped holes themselves (as d9Lគ d9 spectral incoherence further in the case of Y compounds in Sec. weight) but is also transferred from the upper→ Hubbard 9 10 IV.C.3. The out-of-plane conductivity ␴c(␻) shows dif- band (d d spectral weight). A theoretical interpre- ferent charge dynamics from the in-plane one. A charge tation of→ the doping-induced spectral weight transfer gap is developed from T above Tc (Tamasaku, Naka- above EF is beyond the framework of the t-J model, mura, and Uchida, 1992). Below Tc , it is rather difficult since the model has no upper Hubbard band. Spectral to observe the superconducting gap in the in-plane weight transfer has been studied using a strong-coupling ␴a,b(␻), because the reflectivity is almost 100% around treatment (i.e., perturbation with respect to t/U)ofthe the gap energy, even in the absence of the superconduct- Hubbard model (Eskes and Oles´, 1994; Eskes et al.,

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1183

FIG. 132. Chemical potential shift as a function of hole con- centration in La2ϪxSrxCuO4 (Ino et al., 1997a). Shifts expected from band-structure calculation and the ␥T term in specific heat (which is proportional to the quasiparticle DOS at EF) are also plotted in (b).

energy Hamiltonian. The decrease in the spectral weight of the Zhang-Rice singlet with hole doping is not easy to detect because of the overlapping intense oxygen 2p band. Figure 131(a) shows combined photoemission and inverse-photoemission spectra of La2ϪxSrxCuO4 (Ino et al., 1997c). The doping-induced spectral weight is not concentrated just above EF , as suggested by previous x-ray absorption studies, but is distributed over the whole band-gap region with the DOS peak ϳ1.5 eV above EF . Thus spectral weight transfer occurs over a relatively wide energy range, within a few eV of EF , and the Fermi level is located at a broad minimum of the DOS, as schematically shown in Fig. 131(b). The spec- tral DOS at EF is much weaker than that given by LDA band-structure calculations even for optimally doped and overdoped samples, indicating that the quasiparticle weight is very small: ZӶ1. In underdoped samples, FIG. 131. Density of states of La2ϪxSrxCuO4 . (a) Photoemis- there is even a pseudogap feature at EF , which evolves sion and inverse-photoemission spectra near the Fermi level into the insulating gap at xϭ0. (Ino et al., 1997c); (b) Schematic representation of the evolu- The E or the electron chemical potential in tion of spectral weight with hole doping. F La1ϪxSrxCuO4 deduced from the shifts of various core levels exhibits an interesting behavior as a function of x 1994) and using numerical simulations of Hubbard clus- (Ino et al., 1997a). The chemical potential is shifted ters (Dagotto et al., 1991; Li et al., 1991; Bulut et al., downward with hole doping as expected, but the shift is 1994a). Decrease of the spectral weight below EF occurs fast in the overdoped regime (xՇ0.15) and slow in the 9 9 1 in the Zhang-Rice singlet band ͓d d Lគ ( A1g) spec- underdoped regime (xտ0.15), as shown in Fig. 132. It -x becomes vanishingly small within exץ/␮ץ tral weight], which was regarded as→ the effective lower appears that Hubbard band when the high-energy part of the original perimental errors for small x, implying a divergence of p-d Hamiltonian is projected out and the Hubbard charge compressibility/susceptibility towards xϳ0as model or the t-J model is employed as the effective low- predicted by Monte Carlo simulation studies of the 2D

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1184 Imada, Fujimori, and Tokura: Metal-insulator transitions

Hubbard model (Otsuka, 1990; Furukawa and Imada, 1993). This is also consistent with the scaling theory with the dynamic exponent zϭ4 (see Sec. II.F). On the other hand, the electronic specific-heat coefficient ␥ is de- creased towards xϳ0 (Momono et al., 1994), meaning that the effective mass of conduction electrons decreases as discussed above. In the overdoped region, the results of the chemical potential shift and the electronic specific heat are consistently interpreted within Fermi-liquid theory if the quasiparticle–quasiparticle repulsion is as- 0 0 sumed to be Fs ϳ7. In the underdoped region, Fs has to 0 xץ/␮ץbe assumed small or even negative: Fs Ϫ1if 0, invalidating the original Fermi-liquid→ assumption. Because→ the measurements were made in the normal state, the results indicate that the normal state of the underdoped cuprates is not a Fermi liquid but has a (charge) pseudogap at EF . This observation would be closely related to other aspects of pseudogap behavior in the underdoped region discussed above as well as in Secs. IV.C.3 and IV.C.4. In accord with other measure- ments of the pseudogap discussed in other part of Sec. IV.C, possible microscopic mechanisms for the observed FIG. 133. Crystal structure of TЈ phase. behavior may be (i) preformed Cooper pair fluctuations above Tc , (ii) spinon pairing in the resonating valence because high-quality filling-controlled single crystals (in bond scenario, (iii) antiferromagnetic spin fluctuations the sense of oxygen stoichiometry and distribution) have strong enough to destroy Fermi-liquid behavior, and (iv) been difficult to prepare. In particular, only minute oxy- fluctuations of spin-charge stripes. Concerning of (iv), gen nonstoichiometry is determinant not only for the Zaanen and Oles´ (1996) argued that spin-charge stripes appearance of superconductivity but for normal-state pin the chemical potential for low hole concentrations if transport and magnetic properties. For example, the 24– the stripes become static and may continue to do so 25-K superconductivity is known to show up around x even if the stripes are dynamic. ϭ0.15 for Nd2ϪxCexCuO4 and Pr2ϪxCexCuO4, only when the compound is subject to a thermal annealing procedure under reducing conditions (Takagi, Uchida,

2. Nd2ϪxCexCuO4 and Tokura, 1989). In these compounds, the oxygen sto- ichiometry corresponding to increase of ⌬yϷ0.01 is suf- Since the discovery of electron-doping-induced super- ficient to quench superconductivity completely (Idemoto conductivity in R2ϪxCexCuO4, where R may be Pr, Nd, et al., 1990; Suzuki et al., 1990). This implies that minute Sm, or Eu (Tokura, Takagi, and Uchida, 1989), the ex- oxygen interstitials positioned on other than the in-plane istence of high-Tc superconductivity for both hole and O(2) sites (Fig. 133) may serve as very effective carrier electron doping has imposed a strong constraint on its scatterers or localization centers, in contrast with the mechanism (Anderson, 1992). R2ϪxCexCuO4 has the so- other cases where slight oxygen offstoichiometry plays called TЈ structure (Fig. 133) in which the single CuO2 the role of fine filling control (see Sec. IV.B). According sheet without apical oxygen is sandwiched by fluorite- to single-crystal neutron-diffraction structural studies type (R,Ce)2O2 block layers. The Ce ion in the fluorite- (Radaelli et al., 1994; Schultz et al., 1996), it is probable type block is tetravalent (4ϩ), and hence Ce doping (x) that in as-grown crystals without a reducing procedure changes the band filling in the opposite direction to that interstitial oxygens may sit on the apical O(3) site imme- in the so-called hole-doped superconductors such as diately above the Cu site. La2ϪxSrxCuO4. In other words, high-Tc superconductiv- Keeping in mind the remarkable but still puzzling role ity emerges irrespective of the increase or decrease in of minute oxygen interstitials, let us survey the MIT- band filling from the CuO2-based half-filled Mott insula- related properties in R2ϪxCexCuO4 (where RϭNd and tor. Electron doping can only be performed for this type Pr). Figure 134 (Matsuda et al., 1992) shows the evolu- of square CuO2 sheet without apical oxygens, while the tion of 3D antiferromagnetic (AF) spin order (TN pyramidal or octahedral CuO2 sheet with apical oxygens ϭ160 K) for an as-grown Nd2ϪxCexCuO4 (xϭ0.15) or halogens are mostly hole-dopable (Tokura and crystal (nonsuperconducting) and its near disappearance Arima, 1990; Tokura 1992a, 1992b), probably due to for the reduced crystal (24-K superconducting). The electrostatic reasons (i.e., Madelung-type energy origins (1,0,1) Bragg-peak intensity is contributed to by both Cu from the valence change of Cu or O). and Nd moments and its decrease below 50 K is due to In spite of extensive studies on normal and supercon- the onset of the Nd spontaneous moment. By the de- ducting properties in these electron-doped supercon- composition procedure with the respective form factors, ductors, much less physics has been clarified than in the the saturation Cu moment at the lowest temperature cases of their hole-doped counterparts. This is simply was estimated as around 0.2␮B or less (Matsuda et al.,

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1185

FIG. 135. Electrical resistivity vs temperature for the Pr1.85Ce0.15CuO4ϩ␦ crystals A (drawn lines) and AЈ (dashed lines) annealed in different oxygen atmospheres as given in the figure. The inset shows the resistive superconducting transi- tions for samples A and AЈ.

interstitial oxygen may trigger long-range antiferromag- netic order, which is incompatible with superconductiv- ity, as observed in the spin-gapped Cu-O ladder com- pounds (Sec. IV.D.1). Figure 135 shows the temperature dependence of the FIG. 134. Upper panel: M/H vs temperature both before and after reduction and annealing. The samples were cooled in in-plane resistivity on single crystals of Pr2ϪxCexCuO4ϩy zero magnetic field. Lower panel: (101) Bragg-peak intensity (xϭ0.15) after several annealing steps at low-oxygen before and after reduction and annealing. partial pressure (Brinkmann et al., 1996). With a reduc- tion procedure, the superconducting transition tempera- ture T steadily increases, accompanying a gradual de- 1992). The as-grown xϭ0.15 crystal shows metallic be- c crease in residual resistivity (defined by the resistivity havior with resistivity ␳ less than 10Ϫ3 ⍀ cm, but the ␳ shows an upturn around the temperature corresponding immediately above Tc). In fact, the change of residual resistivity scales linearly with that of Tc , implying that to TN (usually ranging from 120 to 160 K for the as- grown xϭ0.15 crystal). As shown in Fig. 134 and also the oxygen interstitials in the less reduced crystals play demonstrated by earlier ␮SR (Luke et al., 1989) and the role of strong carrier scatterers as well as pair- NMR (Kambe et al., 1991) measurements, antiferromag- breaking impurities. netic order is extinguished by the reducing procedure, The resistivity in the normal state of 24-K supercon- which gives rise to superconductivity below TcϷ24 K. ductor Nd2ϪxCexCuO4Ϫy is quadratic with temperature Such a robust behavior of the antiferromagnetic order at relatively low temperatures below 200 K and, at with electron-doping level x is in sharp contrast to the higher temperatures, subject to the logarithmic correc- case of hole-doped superconductors, La2ϪxSrxCuO4.In tion characteristic of a 2D Fermi liquid (Tsuei, Gupta, the latter compound, TN decreases at a rate of 160 K/ and Loren, 1989). This is in contrast with the well- at. %, while in the TЈ-phase compounds (as-grown) the known T-linear behavior for optimally hole-doped su- rate of decrease is only Ϸ9 K/at. % (Thurston et al., perconductors. Brinkmann et al. (1995) succeeded in ex- 1990). The slower collapse rate of the antiferromagnetic tending the superconducting composition range to 0.04 order with doping in the TЈ-phase compounds is seem- рxр0.17 for Pr2ϪxCexCuO4Ϫy using an improved re- ingly in accordance with the simple site dilution model duction technique. In those crystals, they observed for (Matsuda et al., 1992), or intuitive scenario that a doped the normal state in the reduced superconducting crystals electron forms the local nonmagnetic Cuϩ1 species and that the resistivity was fitted using the empirical formula 2ϩ n dilutes the original Sϭ1/2 Cu spins. However, such a ␳(T)ϭ␳0ϩbT and that the exponent n changed from simplified model should not in principle be applicable to nϷ1.7 for high Ce concentration to nϷ1.3 for lower Ce the metallic state induced by electron-doping. In fact, concentrations. A similar systematic change is known the minute change in the oxygen content brought about for the overdoped concentration range of the hole- by the reduction procedure seems to drive a weakly lo- doped superconductors (Takagi, Batlogg, et al., 1992a; calized doped (or site-diluted) Mott antiferrromagnet to Kubo et al., 1991). a highly correlated metallic (at low temperatures, super- The long-standing problem is the nature of supercon- conducting) state. The local square symmetry of the Ce ducting carriers in the electron-doped cuprates. The dopant may be broken by disorder, or a small amount of Hall coefficient RH is quite sensitive, not only to the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1186 Imada, Fujimori, and Tokura: Metal-insulator transitions

and without reduction procedures (Arima, Tokura, and Uchida, 1993). Spectroscopically, the reduction proce- dure appears to produce an effect similar to the electron-doping x. With electron-doping, the spectral weight arising from the CT gap transition around 1.5 eV is transferred to the in-gap region (Cooper et al., 1990a), as observed in the hole-doping case, La2ϪxSrxCuO4 (Uchida et al., 1991). However, the broad infrared band (the so-called in-gap state absorption) is far from the conventional Drude spectrum and the charge dynamics appears highly incoherent. The doping-induced change was also observed in the O 2p-Cu 4s interband transi- tion region (Arima, Tokura, and Uchida, 1993). A single band around 5.0 eV in the xϭ0 parent compound is split and the spectral weight is gradually shifted to the lower- lying band around 4.2 eV with doping. This was inter- FIG. 136. Optical conductivity spectra of Pr2ϪxCexCuO4Ϫy with light polarization perpendicular to the c axis (in-plane). preted in terms of the formation of an occupied in-gap state below the Fermi level. Therefore the energy differ- ence between the upper edge of the occupied O 2p state doping x as in the hole-doped superconductors, but also and the in-gap state is about 0.8 eV, nearly half of the to the reduction procedure. In general, the RH is nega- original CT band gap. This is contrary to the case of hole tive and its absolute magnitude scales roughly with x in doping, in which the in-gap state is formed predomi- the lightly doped region, whereas in the nonsupercon- nantly above the Fermi level and hence is observable, ducting overdoped region xϾ0.17 the sign of RH is re- e.g., in the O 1s absorption spectroscopy, as the unoc- versed to positive (Takagi, Uchida, and Tokura, 1989). cupied state (Chen et al., 1991). Apart from the opposite sign, the behavior is analogous Doping-induced changes in the single-particle DOS of to the case of La2ϪxSrxCuO4 (Ong et al., 1987; Takagi, Nd2ϪxCexCuO4 are thus different from those in Ido, et al., 1989). However, it has been occasionally ob- La2ϪxSrxCuO4 in that the in-gap spectral weight appears served for superconducting R2ϪxCexCuO4Ϫy (RϭNd or largely below EF (Anderson et al., 1993). It is then ex- 9 10 Pr) crystals with xϷ0.15 that RH increases steeply to pected that the upper Hubbard band (d d ) loses its positive values with lowering temperature, say below 80 spectral weight, although an increase of→ the spectral K (Wang et al., 1991; Jiang et al., 1994). These were in- weight was indeed observed by oxygen 1s x-ray absorp- terpreted in terms of the two-band model in which both tion spectroscopy (Pellegrin et al., 1993). The position of holes and electrons participate in charge transport for the Fermi level of the doped sample with respect to the the superconducting phase. band gap of Nd2CuO4 has been a long-standing puzzle: Figure 136 displays the in-plane optical conductivity It was pointed out that EF is located below the mid-gap spectra (at room temperature) of Pr2ϪxCexCuO4 with of the parent compound if we consider the band gap of

FIG. 137. Fermi surface of Nd2ϪxCexCuO4 determined by angle-resolved photoemission spectroscopy (King et al., 1993). Solid curves are results of LDA band-structure calculations (Massida et al., 1989). G:(␲,0); X:(␲,␲).

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1187

FIG. 138. Lattice structure of YBa2Cu3O7.

Nd2CuO4 to be the optical gap of 2.0 eV (Allen et al., 1990; Namatame et al., 1990). This is obviously incom- patible with the common notion that the bottom of the conduction band is the affinity level of the parent insu- lator. Angle-resolved photoemission studies of Nd1.85Ce0.15CuO4 revealed a dispersing band which crosses EF (Anderson et al., 1993; King et al., 1993). As shown in Fig. 137, the Fermi surface thus determined is a hole-like one centered at the (␲,␲) point of the Bril- louin zone, in excellent agreement with band-structure calculations for the nonmagnetic metallic state. The re- sults also show that the Fermi surface volume changes following the Luttinger sum rule with varying doping level. There is a saddle point, not in the band dispersions FIG. 139. Fermi surfaces of YBa Cu O determined by angle- near EF , but a few hundred meV below it, unlike the 2 3 6.9 hole-doped cuprate superconductors. The angle- resolved photoemission spectroscopy. (a) Results by Liu et al., resolved photoemission spectra show intense high- 1992a. Open circles denote Fermi level crossings and shaded binding-energy tails and background, indicating that the regions are calculated Fermi surfaces (Pickett et al., 1990). The finite width of the shaded regions represents finite k disper- spectral weight of the coherent quasiparticle Z is very z sion. The measurements on twinned samples have been done small. for the half of the Brillouin zone shown here. X:(␲,0), Y: (0,␲), S:(␲,␲). (b) Another interpretation of the same data by 3. YBa2Cu3O7Ϫy Shen and Dessau (1995). This compound has the same layered structure as La2ϪxSrxCuO4 in the sense that it has a stacking of the 1987). An ordered structure similar to the case of CuO2 planes and the block layers. The lattice structure La2CuO4 was confirmed by subsequent neutron- is shown in Fig. 138. A complexity in this compound is diffraction measurements (Tranquada et al., 1988). In that it has a chain structure of Cu and O embedded in the region 0.6ϽyϽ1.0, holes are mainly doped into the the so-called block layer sandwiched by the CuO2 CuO chain sites, where excess oxygen added to yϭ1.0 is planes. This Cu-O chain structure makes a minor contri- indeed introduced to the layers composed of separated bution to the dc conductivity. CuO chains. Therefore the doping concentration in the 2ϩ The insulating compound of YBa2Cu3O7Ϫy with y CuO2 plane changes little by basically retaining the Cu у0.6 has antiferromagnetic long-range order. The first valence and thereby retaining the antiferromagnetic in- signature of the antiferromagnetic order was detected by sulating phase. Below yϭ0.6, holes start to be trans- the muon-spin-rotation measurements (Nishida et al., ferred to the CuO2 plane, which immediately destroys

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1188 Imada, Fujimori, and Tokura: Metal-insulator transitions

or optimally doped region with Tcϳ90 K. Similarly to the case of La2ϪxSrxCuO4, short-range antiferromag- netic correlations survive in the metallic phase. This en- hancement of antiferromagnetic correlations is seen in the diffuse neutron-scattering measurement around (␲,␲). However, in contrast to the case of La2ϪxSrxCuO4,Im␹(q,␻) resulting from the neutron data does not show clear incommensurate peaks in the metallic phase. Instead, it has a broad peak around (␲,␲), as is shown in Fig. 140 (Rossat-Mignod et al., 1991a; Sternlieb et al., 1994). The origin of this differ- ence in the peak structure between La and Y com- pounds was ascribed to the difference in the shape of the Fermi surface (Furukawa and Imada, 1992; Si et al., 1993; Tanamoto, Kohno, and Fukuyama, 1993, 1994). However, a signature of incommensurate structure was recently detected also in Y compounds (Dai et al., 1998). A single-parameter scaling by ␤␻ for ͐dq Im␹(q,␻), analogous to the case of La2ϪxSrxCuO4 as discussed in Sec. IV.D.1, is observed in the underdoped region (Sternlieb et al., 1993). However this single-parameter scaling seems to break down at lower energies below ϳ12 meV at low temperatures. This may be related to pseudogap formation, as discussed below. Magnetic properties show rather different features be- tween the underdoped (0.6уyу0.3) and the optimally or overdoped (0.3уyу0) samples. The Knight-shift measurement in NMR clearly shows this difference. In overdoped samples with Tcϳ90 K, the Knight shift is more or less constant at TϾTc , indicating that the uni- form magnetic susceptibility ␹(qϭ␻ϭ0) follows the T-independent Pauli susceptibility of standard metals. FIG. 140. Antiferromagnetic correlation observed in peak On the other hand, in the underdoped material, 0.3Ͻy structures of Im ␹(q,␻) by neutron scattering for Ͻ0.6, the uniform magnetic susceptibility derived from YBa2Cu3O6ϩx . From Rossat-Mignod et al., 1991a. the spin part of the Knight shift shows a rapid decrease with decreasing temperatures, as shown in Fig. 141 the antiferromagnetic order and is accompanied by met- (Takigawa et al., 1991). The observed decrease with low- allization. The Ne´el temperature in the insulating phase ering temperatures is in fact much more pronounced is as high as 400 K. Because the electronic structure cal- than that seen in the 2D or 1D spin-1/2 Heisenberg culation using the LDA shows the Fermi surface more model (Bonner and Fisher, 1964; Okabe and Kikuchi, or less tilted by 45° from the case of La2CuO4, as illus- 1988; Makivic´ and Ding, 1991). This decrease appears to trated in Fig. 139 (Liu et al., 1992a), simple nesting con- be continuously and smoothly connected with the de- dition is not satisfied in this case and thus strong corre- crease below Tc , as seen in Fig. 141. Therefore it is lation effects should be crucial in realizing the natural to speculate that formation of the pseudogap antiferromagnetic insulator. A pulsed high-energy neu- structure due to preformed singlet pairing fluctuations tron measurement performed up to 250 meV shows that begins well above Tc . The peak structure of ␹(qϭ␻ the spin-wave theory of the Heisenberg model well de- ϭ0) at TϭT␹ϳ500 K is qualitatively similar to that for scribes the spin-wave dispersion with the in-plane ex- the case of La2ϪxSrxCuO4 in the underdoped region. change constant Jʈϭ125Ϯ5 meV and the inter-layer ex- However, if we look at 1/T1 , we notice that the situ- change JЌϭ11Ϯ2 meV (Hayden et al., 1996). ation is not so simple as naively expected from s-wave Around yϭ0.6, the compound becomes metallic and preformed pairing. In overdoped or optimally doped the hole concentration in the CuO2 layer increases with samples with Tcϳ90 K, 1/T1T at the copper site is more further decreasing y. This compound in the metallic or less similar to the case of La1.85Sr0.15CuO4, although phase shows a unique process of doping due to the the experimentally accessible temperature range is lim- transfer of holes between the Cu-O chain site and the ited to below 400–500 K because of the difficulty of con- CuO2 layer (Tokura, Torrance, et al., 1988). In the re- trolling oxygen concentration above it. A gradual cross- gion 0.3рyр0.6, the hole concentration in the CuO2 over from 1/T1Tϳconst to 1/T1ϳconst appears to take plane is kept low with relatively low superconducting place from low to high temperatures. The underdoped transition temperature Tcϳ60 K. Around yϭ0.3 the samples with Tcϳ60 K show remarkably different be- hole concentration is increased quickly to the overdoped havior, as shown in Fig. 142, where 1/T1T decreases be-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1189

FIG. 141. Temperature dependence of the Knight shift for YBa2Cu3O7Ϫy at yϭ0 and yϭ0.37. Various components of the Cu and O Knight shift are plotted with different vertical scales and origins. Anisotropy obtained from the direction of the applied magnetic field in ab, the ab-plane; c, c-axis; ax, the axial part; iso, the isotropic part (Takigawa et al., 1991). The Knight shift starts decreasing from much higher temperature than Tc . low 150–200 K as if it were under the influence of pseudogap formation (Yasuoka et al., 1989). In fact, these 1/T1T data were the first pioneering indication of pseudogap behavior in the high-Tc cuprates well above Tc . Similar pseudogap formation behavior is also ob- served in other underdoped compounds such as YBa2Cu4O8, and HgBa2CuO4ϩ␦ (Itoh et al., 1996). So far, in La2ϪxSrxCuO4, this pseudogap behavior in 1/T1T is not clearly visible. The possible effect of randomness was discussed for the suppression of the pseudogap in 1/T1T. In any case, the anomaly in the underdoped Y compound appears to show up below TRϳ200 K, which is substantially lower than T␹ in the uniform magnetic susceptibility but markedly higher than Tc . The same type of anomaly has also been confirmed in neutron- scattering experiments (Rossat-Mignod et al., 1991a, 1991b; Sternlieb et al., 1993), where Im ␹(q,␻) shows pseudogap behavior for qϳ(␲,␲) below Tϳ200 K for the underdoped compound with Tcϳ60 K, as shown in Fig. 143. The origin of the difference between T␹ in ␹(qϭ␻ ϭ0) and TR in 1/T1T is not clear enough at the moment and further studies are clearly necessary. Another puz- zling feature of the pseudogap behavior is that the Gaussian decay rate 1/TG does not show a gaplike de- crease either at T␹ or at TR (Itoh et al., 1992). This may be related to the compatibility of antiferromagnetic cor- relations and superconductivity; see Assaad, Imada, and Scalapino (1997), where the above two puzzling features FIG. 142. Temperature dependence of 1/T1T at the Cu site for are reproduced in a theoretical model. several high-Tc compounds (by courtesy of H. Yasuoka, who Other indications of a pseudogap below Tϳ200 K replotted various available data). They show indications of were also suggested from the anomaly of the B2u pho- pseudogap behavior well above Tc . The data are taken from non mode (Harashina et al., 1995) as well as from Machi et al., 1991; T. Imai et al., 1989; Goto et al., 1996; War- anomalies in the Raman mode coupled to ren et al., 1989; Takigawa et al., 1991; Y. Itoh et al., 1996.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1190 Imada, Fujimori, and Tokura: Metal-insulator transitions

weak-coupling BCS superconductors may be continu- ously connected to a strong-coupling region where the picture of Bose condensation by preformed pairs be- comes appropriate (Nozie`res and Schmitt-Rink, 1985). When ␰coh is small, the superconductivity is rather close to the picture of the condensation of bosons. If this is the case, the pairing temperature should be separated from Tc itself. However, the naive picture of Bose con- densation is not appropriate because the preformed pair has d-wave character accompanied by antiferromagnetic correlations with strong momentum dependence. One scenario for preformed singlet pair fluctuations is given by an unusual criticality near the Mott transition, as discussed in Sec. II.F. Because of unusual suppression of the coherence of the single-particle transfer process due to strong spin and charge fluctuations near the Mott transition point, we have to consider a higher-order pro- cess of the transfer. In the Mott insulating phase, this FIG. 143. Im ␹(q,␻), in arbitrary units, as a function of ␻ at indeed makes the superexchange interaction in the or- der of t2/U relevant. If the system is close to the Mott qϭ(␲,␲) for the underdoped compound YBa2Cu3O6.69. The insulator but in the metallic phase, other higher-order pseudogap structure is seen from the temperature above Tc . The inset shows the temperature dependence or fixed ប␻ϭ8 processes of two-particle transfer contributes equally, meV. From Rossat-Mignod et al., 1991b. while the single-particle process is still suppressed due to a large dynamical exponent zϭ4 (Assaad, Imada, and Scalapino, 1996). Near the Mott insulator this two- (Litvinchuk, Thomsen, and Cardona, 1993), where con- particle process can be more relevant than the single- sistency with coupling to d-wave pairing was proposed particle process because of its small dynamical exponent theoretically (Normand, Kohno, and Fukuyama, 1995). zϭ2 (Imada, 1994c, 1995b). This means that, in contrast The energy shift of the phonons due to coupling to the to the single-particle process, the two-particle process is superconducting order parameter has been discussed not anomalously suppressed leading to a gain in the ki- from a more general point of view by Zeyer and Zwick- netic energy. This necessarily yields a pairing effect. An nagl (1988). important point here is a strong momentum dependence The origin of this pseudogap behavior is not fully un- of pairing fluctuations. The precursor of pairing starts derstood. Experimentally, however, it has been con- around (␲,0) and (0,␲) in the momentum space at higher firmed in various cases, including the so-called single- temperatures. For more detailed discussion of this layer compound HgBa2CuO4ϩ␦ . The formation of a mechanism, see Sec. II.F. pseudogap has also been detected by photoemission There are several completely different ways of inter- measurements for underdoped Bi and La compounds, as preting pseudogap behavior as an effect of preformed discussed in Secs. IV.C.1 and IV.C.4. Another important pairing. Pairing through the magnetic polaron or con- indication of pseudogap behavior is a reduction of the ventional lattice polaron effect was studied as a bipo- specific-heat coefficient ␥ at low doping similar to the laron mechanism (Micnas et al., 1990; Alexandrov, Brat- case of La2ϪxSrxCuO4 (Loram et al., 1994; Momono kovsky, and Mott, 1994; Alexandrov and Mott, 1994). In et al., 1994; Liang et al., 1996). the cuprate superconductors, it is not likely that bipo- Theoretically, several different explanations for this laronic mechanisms and local-pairing mechanisms can unusual behavior have been proposed. The possibility of be compatible with the principal features of strong- preformed pairing fluctuations above Tc as the origin of correlation effects. Fluctuating phase-separation effects pseudogap behavior has been examined from various and dynamic charge-ordering effects (Emery and Kivel- approaches. This is a natural idea because the tempera- son, 1995, 1996; Zaanen and Oles´, 1996) were proposed ture for singlet pairing may in general be higher than Tc as sources of preformed pairing. A more general argu- itself when the MIT point is approached. Near a con- ment from the separation of the optimized Bose conden- tinuous MIT point, the coherence temperature Tcoh has sation temperature and the pairing energy, as well as a to decrease to vanish, while the energy scale for the pair- comparison of cuprate superconductors with other cases ing energy may be independent of this reduction. This such as heavy-fermion superconductors and C60 com- idea is supported by the small coherence length of the pounds, was stressed by Uemura et al. (1989, 1991). Cooper pair observed experimentally in the high-Tc cu- Another interpretation of pseudogap behavior is the prates. The short coherence length ␰coh is easily deduced scenario assuming spin-charge separated excitations at from a large critical magnetic field Hc2 and the relation the low-energy level. Pseudogap formation is attributed ␰cohϭ1/ͱ2␲Hc2. In YBa2Cu3O7, ␰coh in the CuO2 plane to the pairing of spinons without Bose condensation of was estimated as ϳ10–20 Å, that is, only a few lattice holons in the picture of the slave-boson formalism dis- spacings, even in the plane direction. It is known that cussed in Sec. II.D.7 (Baskaran, Zou, and Anderson,

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1191

FIG. 144. Im ␹(q,␻), in arbitrary units, as a function of ␻ at qϭ(␲,␲) for the optimally doped compound YBa2Cu3O6.92 (Rossat-Mignod et al., 1991b). A prominent 41-meV peak is observed below Tc . FIG. 145. Temperature dependence of in-plane resistivity for various choices of y for YBa2Cu3O7Ϫy , where the numbers in 1987; Kotliar and Liu, 1988; Suzumura, Hasegawa, and the figure indicate 7Ϫy (Ito, Takenaka, and Uchida, 1993). Fukuyama, 1988; Tanamoto, Kohno, and Fukuyama, The CuO2-plane contribution is obtained by subtracting the 1994). The spinon pairing is basically due to singlet for- chain contribution deduced from the a- and b-axis resistivities mation by the mean-field-type decoupling of the super- shown in the inset. exchange interaction JSi•Sj . From the Fermi-liquid ap- proach the nesting condition has also been proposed as the origin of pseudogap behavior (Miyake and Narikiyo, deviation may be viewed as the reduction of ␳ below the 1994b). temperature T␳ , which increases with decreasing doping In neutron scattering at higher energies, as shown in concentration as seen in Fig. 145 (Ito, Takenaka, and Fig. 144, another unusual behavior has been observed in Uchida, 1993; Takenaka et al., 1994). As compared to Im ␹(q,␻), namely, a remarkable peak around ␻ϭ41 T␹ӍT␳ in the case of La2ϪxSrxCuO4, T␳ in the Y com- pounds appears to be low with weaker effects. Similarly meV with dramatic growth below Tc (Mook et al., 1993; Fong et al., 1995, 1997). The resonance energy decreases to the case of La2Ϫx SrxCuO4, the Hall coefficient shows monotonically with decreasing doping concentration. It a rather sharp increase with decreasing temperature be- is clear that this excitation is strongly coupled to the low a similar temperature T␳ or T␹ (Ito et al., 1993; superconducting order parameter. Demler and Zhang Takeda et al., 1993; Nishikawa, Takeda, and Sato, 1994). (1995) have proposed that this peak is related to a spe- The temperature dependence of the Hall angle de- cial type of triplet pairing excitation, namely, the rota- fined by tional mode connecting the antiferromagnetic order pa- ␴xx rameter and the d-wave superconducting order cot ⌰ ϭ (4.12) H ␴ parameter under SO(5) symmetry (Zhang, 1997). The xy enhancement of the resonance below Tc has some simi- was measured for YBa2Cu3ϪpZnpO7Ϫy (Chien, Wang, larity to the result in La compounds above 7 meV. These and Ong, 1991; Ong et al., 1991). For pϭ0 and ␦ϳ0, all seem to indicate the compatibility of antiferromag- cot ⌰H appears to follow netic correlation in the superconducting state. Similar cot ⌰ Ӎ␣T2. (4.13) and consistent theoretical results were obtained in nu- H merical studies attempting to account for the two- Anderson (1991a) introduced two different carrier re- particle process explicitly (Assaad, Imada, and Scala- laxation rates, ␶tr and ␶H , where ␶tr is the longitudinal pino, 1997; Assaad and Imada, 1997). The peak appears relaxation rate for motion perpendicular to the Fermi to be the remnant of the antiferromagnetic Bragg peak surface while ␶H is the transverse rate for the velocity shifted to a finite frequency. In this context, the component parallel to the Fermi surface. In the usual pseudogap formation widely observed appears to start at Fermi liquid, ␶trϭ␶H must hold, while the above experi- high temperatures as undifferentiated signatures of anti- mental result together with ␳ϳT suggests ␶trϳ1/T and 2 ferromagnetic and d-wave preformed pair fluctuations. ␶Hϳ1/T . Anderson interpreted this unusual tempera- The transport properties of YBa2Cu3O7Ϫy show, ture dependence as coming from a singular interaction roughly speaking, behaviors similar to that of of two electrons located perpendicular to the Fermi sur- La2ϪxSrxCuO4 as described in Sec. IV.C.1 (Iye, 1990; face, and also proposed that it could yield similar behav- Ong, 1990). The in-plane resistivity ␳ab is approximately ior to a 1D Tomonaga-Luttinger liquid. proportional to T near yϭ0, that is, near the optimum Around yϭ0, the c-axis transport is also metallic, in- doping concentration, in the same way as in dicating T-linear behavior of ␳c , although the absolute La1.85Sr0.15CuO4. However, in the underdoped region it value is much larger than ␳ab due to anisotropy. How- shows substantial deviation from T-linear behavior. This ever, it shows insulating temperature dependence in the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1192 Imada, Fujimori, and Tokura: Metal-insulator transitions

state sandwiched between the Fermi liquid and the Anderson localized phases (see, for example, Dobrosav- ljevic´ and Kotliar, 1997). A key point to understand this result may lie in the strong wave-number dependence of carriers, which has been ignored in the above argu- ments. The preformed singlet pair fluctuations around (␲,0) may to some extent contribute to the ‘‘metallic’’ conduction, though they remain rather incoherent well above Tc . This small ‘‘metallic’’ contribution may ex- plain the small reduction in the resistivity upon pseudogap formation. While the single-particle excita- tions may easily be localized by Zn substitution or in- duce the local moment, the Anderson localization may take place predominantly in the single-particle excita- tions because of the underlying large dynamic exponent of the MIT. The residual resistivity ␳0 by Zn doping in the unitar- ity limit depends on the carrier density n and the carrier 2 charge e as ␳0ϰ1/(ne ). Therefore one might think that one could distinguish the following three cases: The first is the usual Fermi-liquid conduction with charge e and nϭ1Ϫx, the second is a complicated situation in which the preformed pair with nϭx/2 and charge 2e makes a minor contribution to the transport while the major con- tribution is from the part around (Ϯ␲/2,Ϯ␲/2). The third is the spin charge separation scenario where the holon conduction gives the charge e with nϭx (Nagaosa and Lee, 1997). Experimentally, the increase of ␳0 is large in FIG. 146. Comparison of in-plane and c-axis resistivity vs tem- the underdoped region and roughly scaled by 1/nϭ1/x perature at several choices of doping concentrations 7Ϫy for (Fukuzumi et al., 1996). However, a complexity exists YBa2Cu3O7Ϫy . From Takenaka et al., 1994. because of the strong wave-number dependence of car- riers in the underdoped region. Carriers in a region low-doping region below certain temperatures. As tem- around (␲/2,␲/2) remain unpaired, while carriers around peratures are lowered, ␳c decreases, with minima at a (Ϯ␲,0) and (0,Ϯ␲) stay strongly damped due to zϭ4 temperature slightly higher but similar to T␳ , and in- with instability to gradual pseudogap formation. If the creases as an insulating behavior below T␳ as in Fig. 146 transport is still dominated by single-particle transport (Takenaka et al., 1994). It should be noted that this mainly coming from a small region around (␲/2,␲/2) anomaly occurs at a similar temperature to that for the over other transport channels, the residual resistivity can deviation of ␳ab from T-linear behavior discussed above be large because the effective carrier number is given by and is again similar to the case of La2ϪxSrxCuO4. In the that small region around (Ϯ␲/2,Ϯ␲/2) as if the Fermi scenario of preformed singlet fluctuations, the reduction volume were small. [This is clearly different from the of ␳ab below T␳ can be ascribed to the increasing coher- scenario of small-pocket formation around (Ϯ␲/2,Ϯ␲/ ence of the fluctuating preformed pairs, while the charge 2)]. This may account for the experimental result. Even excitation in the c direction has a pseudogap because when we consider this complexity, the experimentally interlayer hopping has to break the pair once if the pair- observed large residual resistivity induced by Zn doping ing originates from two-dimensionality, as in the sce- does not seem to be compatible with the first case, while nario of pairing due to large dynamic exponent z of the it seems to be hard to judge the superiority of one to the 2D Mott transition, discussed in Sec. II.F.9. other in the latter two cases. The effects of Zn doping on the anomalies below T␳ The optical conductivity ␴(␻) of untwinned crystals were examined (Mizuhashi, Takenaka, Fukuzumi, and shows a broad tail roughly proportional to 1/␻, as in Fig. Uchida, 1995). There exists a range of Zn concentration 147 (Schlesinger et al., 1990; Azrak et al., 1994). At the around 2% where the metallic and pseudogap-type tem- optimum doping concentration yϳ0, ␴(␻) basically fol- Ϫ␤␻ perature dependence of the resistivity is retained while a lows the scaling ␴(␻)ϳ␴0(1Ϫe )/␻ with a constant Curie-like enhancement in the uniform magnetic suscep- ␴0 indicating the absence of coherent charge dynamics, tibility appears. This implies that the local spin moment as already discussed in Secs. II.E.1, II.F.9, and II.G.2 appears despite the metallic conduction of carriers. This from the theoretical point of view. In untwinned was interpreted from the scenario of spin-charge separa- samples, the broad mid-IR peak structure observed in tion in the slave-boson formalism (Nagaosa and Ng, La compounds is not visible, implying that the peak 1995). structure itself is not an intrinsic property of the high-Tc An alternative scenario is given by a non-Fermi-liquid cuprates. The significance of the absence of the true

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1193

FIG. 147. Optical conductivity ␴(␻) for several different un- FIG. 148. Pseudogap observed in c-axis ␴(␻)ofYBaCu O twinned film samples (denoted as A,B,D,E,F) of YBa Cu O 2 3 7Ϫy 2 3 7 for yϭ0.3. The upper panel (a) shows the data with the by Azrak et al. (1994). Solid lines are fitting curves in the form phonons, while the lower panel shows the data of the elec- of ␻Ϫa. Cross-square points are ␴ by Schlesinger et al. (1990). a tronic part by subtracting the phonon contribution. The inset They roughly follow the scaling 1/␻. in the lower panel shows the conductivity at 50 cmϪ1 normal- ized by the room-temperature conductivity (open circles) and Drude weight, common to all the cuprate superconduct- the normalized Knight shift for Cu(2) (solid line) at yϭ0.37. ors, is also discussed in Sec. IV.C.1. The extended Drude From Homes et al., 1993. analysis of the Y compound also supports the consis- tency of Eq. (2.274) with weakly ␻-dependent and observable in this metallic response, as analyzed by Rot- T-dependent C(␻) (Thomas et al., 1988; Schlesinger ter et al. (1991), Schlesinger et al. (1994), Basov et al. et al., 1990; El Azrak et al., 1994). (1996), and Puchkov et al. (1996). A remarkable feature in the c-axis conductivity ␴c(␻) When the preformed pair is formed by a mechanism is also seen at low doping concentration. Corresponding of in-plane origin below the pseudogap temperature, it is to the insulating temperature dependence of ␳c below expected that in-plane transport will become more co- T␳ , a pseudogap structure grows in ␴c(␻) in the region herent, as is discussed by Tsunetsugu and Imada (1997), ␻Ͻ600 cmϪ1, as shown in Fig. 148 (Homes et al., 1993; while the c-axis resistivity should increase because the Tajima et al., 1995; see also the review by Tajima, 1997). interplane transport has to break preformed pairs unless Even in the ab plane, the pseudogap structure was de- Josephson tunneling of pairs becomes possible in the tected well above Tc , as shown in Fig. 149 (Orenstein superconducting state. Therefore the pseudogap in ␴(␻) et al., 1990; Rotter et al., 1991; Schlesinger et al., 1994; has the opposite effect. The in-plane spectra should Basov et al., 1996; Puchkov et al., 1996). The pseudogap show spectral weight transfer from the incoherent to co- structure seems to be continuously connected with the herent part and thus a decrease in dc resistivity while the superconducting gap observed below Tc . It may be as- c-axis spectra may show transfer to a higher-energy part sociated with the pseudogap observed in angle-resolved and thus an increase in dc resistivity. This is consistent photoemission spectra as discussed in Sec. IV.C.4. This with the observed results. The reduction of the in-plane pseudogap in the ab plane manifests itself as the spec- resistivity below T␳ is gradual. In this context, we note tral weight transfer from the incoherent part roughly the complexity due to strong momentum dependence scaled by 1/␻ to the coherent part. If we considered the between two regions, namely, around (␲,0) and (␲/2, optical spectra only for the coherent part (though it is ␲/2) as described above. actually absent, as discussed above), it would be difficult In YBa2Cu4O8, it was observed that even the c-axis to identify this pseudogap structure because the gap en- conductivity ␴c has a metallic temperature dependence ergy is above the anticipated scale of inverse relaxation in the pseudogap region (Hussey et al., 1997, 1998). This time of charge dynamics (ϳT) as represented by the appears to be associated with extremely good conduc- conventional clean-limit theory. On the other hand, if tion in the double-chain structure specific to this com- we take the incoherent response as the intrinsic and pound. However, the origin of this exceptional behavior dominant part of this ‘‘incoherent metal,’’ in contrast to is not clear enough at the moment. the conventional Drude theory, the pseudogap is indeed The unusual normal-state properties of YBa2Cu3O7Ϫy

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1194 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 150. Schematic phase diagram of pseudogap structure in the plane of temperature and the doping concentration in the high-Tc cuprates. Below T␹ or TR , various indications of pseudogap formation are seen.

Fermi surfaces of YBa2Cu3O7Ϫy (0.1ϽyϽ0.65) were studied by angle-resolved photoemission spectroscopy and generally good agreement with LDA band-structure calculations was obtained for optimally doped (yӍ0.1) samples, as shown in Fig. 139 (Campuzano et al., 1990; FIG. 149. Pseudogap observed in in-plane ␴(␻)of Liu et al., 1992a, 1992b; Tobin et al., 1992): The spectra YBa2Cu3O7Ϫy : (a) Tcϭ93 K at, from top to bottom, Tϭ120, were interpreted as indicating two hole-like Fermi sur- 100, 90, 80, 70, and 30 K; (b) Tcϭ82 K at Tϭ150, 120, 90, 80, faces centered at (␲,␲) (Liu et al., 1992a), as shown in 70, and 20 K; (c) Tcϭ56 K at Tϭ200, 150, 120, 100, 80, 60, 50, Fig. 139(a), or reinterpreted as indicating ‘‘universal’’ and 20 K. From Rotter et al., 1991. Fermi surfaces similar to those of Bi2Sr2CaCu2O8 (Shen and Dessau, 1995). The observed band dispersions near have many aspects in common with La Sr CuO . 2Ϫx x 4 EF are weaker by a factor of ϳ2 than the LDA band These common properties may be summarized as the structure, indicating a moderate mass enhancement: suppressed coherence and the pseudogap behavior. Sup- m*/mbϳ2. Each spectrum shows an intense high- pression of the coherence is typically observed in the binding-energy tail and background, indicating that the T-linear resistivity and the 1/␻ dependence in ␴(␻). The spectral weight of the quasi-particle peak is small: Z pseudogap behavior is observed at low doping below Ӷ1. From these observations, it follows using Eqs. certain temperatures, namely, T␹ in the uniform suscep- (2.125) and (2.126) that mk /mbӶ1, meaning that the tibility ␹ and T␥ in the specific heat. The anomalies in energy bands crossing EF are strongly renormalized by a the Hall coefficient observed below TH , and T␳ in the resistivity, as well as in ␴ (␻) and ␴ (␻), also appear k-dependent self-energy correction. The importance of ab c the k-dependent self-energy is discussed in Sec. II.F.11. to be related. The NMR relaxation rate 1/T1 shows In going from yϭ0.1 (T ϳ90 K) to yϭ0.5 (T pseudogap behavior below TR in the Y compounds, as c c well as in some other compounds, though it is not ob- ϳ50 K), the dispersions and the Fermi surfaces essen- served in La2ϪxSrxCuO4. These crossover temperatures tially do not change. However, while bands crossing EF T␹ , TH , T␳ , TR , and T␥ have quantitatively different along the (␲,␲) direction remain unchanged, those values and they are slightly different from compound to crossing EF along the (␲,0) direction lose their spectral compound. However, dependence on doping concentra- weight significantly [see Fig. 151(a) and 151(b)]. For an tion is similar in all, in the sense that it is more pro- insulating yϭ0.65 sample, whose composition is in the nounced and enhanced in the underdoped region. The vicinity of the MIT, bands do not cross EF and some overall tendency towards pseudogap formation above bands become less dispersive, whereas dispersions of Tc in the underdoped region is clear. We summarize this other bands remain similar to those of the yϭ0.1 basic tendency in a schematic phase diagram with two samples. Interestingly, bands dispersing toward EF and typical crossover temperatures T␹ and TR in Fig. 150. crossing it along the (0,0)-(␲,␲) line remain in the insu- The doping dependence of the electronic structure of lating sample, although their peaks have become YBa2Cu3O7Ϫy is more complicated than that of the broader, as if the Fermi surface remains along this line, other high-Tc materials because of the presence of the as shown in Fig. 151(c). Cu-O chains, which make the number of holes in the A closely related superconductor, YBa2Cu4O8, with CuO2 plane somewhat ambiguous. Electron-energy-loss double chains was studied by angle-resolved photoemis- measurements from the oxygen 1s core level were inter- sion spectroscopy in which band dispersions and Fermi preted as indicating that holes doped in the CuO2 plane surfaces are clarified as shown in Fig. 152 (Gofron et al., having px,y symmetry become itinerant carriers (Nu¨ cker 1994). Notably, a saddle-point behavior was found et al., 1989). around the (␲,0) point of the Brillouin zone, located

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1195

FIG. 151. Angle-resolved photoemission spectra of YBa2Cu3O7Ϫy (Liu et al., 1992a): (a) yϭ0.1 (Tcϳ90 K); (b) yϭ0.5 (Tc ϳ50 K); (c) comparison between yϭ0.5 and yϭ0.65 (insulator).

4 ϳ19 meV below EF ,asinBi2Sr2CaCu2O8 and Sr2RuO4. EF (which is 19 meV for YBa2Cu4O8). The k dispersion 2 The same behavior was found for YBa2Cu3O7Ϫy (Tobin for the single-particle excitation is transformed to a k et al., 1992; Liu et al., 1992a). The experimental band dispersion for a pair excitation with a finite binding en- dispersions are again weaker than calculated by a factor ergy at the MIT edge. of ϳ2. A more remarkable discrepancy from the band- structure calculation is the very flat dispersion along the 4. Bi Sr CaCu O (0,0)-(␲,0) direction, compared to the calculated qua- 2 2 2 8ϩ␦ dratic dispersion. A similar structure is discussed in Sec. Bi2Sr2CaCu2O8 and its derivatives have the strongest IV.C.4. Theoretical aspects of this flat dispersion, which anisotropy in their transport properties between in- appears near the MIT point in 2D systems, are consid- plane and out-of-plane behaviors (Ito et al., 1991) and ered in Sec. II.F.11, and its importance is discussed in therefore are the best 2D conductors among the high-Tc Sec. II.F.11 in terms of the scaling theory. The flatness is cuprates. The idealized crystal structure of qualitatively different from the usual Van Hove singu- Bi2Sr2CaCu2O8 is shown in Fig. 153. It should be noted, larity because the dispersion changes to a k4 law consis- however, that excess oxygen atoms are incorporated in tent with a scaling theory that has a large dynamic ex- the BiO plane, resulting in remarkable superstructure ponent zϭ4. Because the chemical potential (ϭEF)is modulations (Yamamoto et al., 1990). The Tc of determined from the level of the preformed pair, single- Bi2Sr2CaCu2O8 can be controlled by changing the oxy- particle flat dispersion should be separated from this gen stoichiometry (annealing in oxygen or in a vacuum) pair level by a distance equal to the binding energy of or by substituting Y or other rare-earth elements for Ca. the pair. This is why the flat dispersion is shifted from As-grown Bi2Sr2CaCu2O8ϩ␦ samples are close to the op-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1196 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 153. Idealized cystal structure of Bi2Sr2CaCu2O8.

the Y-substituted samples Bi2Sr2Ca1ϪxYxCu2O8ϩ␦ and they become semiconducting (Tcϭ0) at xϳ0.5. The electronic structure of Bi2Sr2CaCu2O8ϩ␦ has been extensively studied by various electron spectroscopic techniques, including angle-resolved photoemission spectroscopy (Takahashi et al., 1989) because of the high quality and stability of cleaved surfaces. According to oxygen 1s electron-energy-loss studies, doped holes in the CuO2 plane have px,y symmetry and occupy the p-d␴ antibonding band of the CuO2 plane (Nu¨ cker et al., 1989). Angle-resolved photoemission studies of optimally doped and overdoped samples show band dispersions and Fermi surfaces similar to those given by band- structure calculations if the energy bands derived from the Bi-O plane are removed from EF . Angle-resolved photoemission spectra along various cuts in the Brillouin zone are shown in Fig. 154 (Dessau et al., 1993). The clearest quasiparticle peak is observed along the (0,0)- (␲,␲) line. The quasiparticle peak disappears approxi- mately at (␲/2,␲/2), unambiguously indicating a Fermi- level crossing. Along the (0,0)-(␲,0) line, on the other hand, the quasiparticle peak disperses towards EF but FIG. 152. Angle-resolved photoemission spectra of never crosses it; it crosses EF along the (␲,0)-(␲,␲) line. YBa2Cu4O8 (a) along the ⌫-Y [(0,0)-(0,␲)] lines; (b) along the Thus the band shows a saddle point at (␲,0) and a very S-Y [(␲,␲)-(0,␲)] lines; (c) band dispersions around the (0,␲) flat dispersion around that point, which is occasionally point. From Gofron et al., 1994. called an extended saddle point or extended Van Hove singularity, although it is much flatter than the usual Van Hove singularity. The flat band is located within timum doping (Tcϳ90 K), while oxygen-annealed ones ϳ50 meV of EF over a large portion of the Brillouin are overdoped and have lower Tc . Holes are depleted in zone around the (␲,0) points. The resulting Fermi sur-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1197

¯ FIG. 154. Angle-resolved photoemission spectra of Bi2Sr2CaCu2O8 taken above 100 K (ϾTcϭ85 K). M:(␲,0), X:(␲,␲). From Shen and Dessau, 1995. faces are shown in Fig. 155, which shows, in addition to formation of a pseudogap due to spin fluctuations or to the hole-like Fermi surface centered at the (␲,␲) point, the preformed pair of binding energy ϳ50 meV. In the an electronlike Fermi surface centered at (0,0). The latter scenario, the single-particle excitation needs a fi- presence of the latter electronlike Fermi surface has nite energy from the chemical potential around (␲,0) been controversial. Ding et al. (1996a) reported only the because of the two-particle bound-state formation. hole Fermi surfaces and attributed other features to the The line shape of the angle-resolved photoemission superstructure modulation of the Bi-O plane or to spectra is unusual as in the other cuprates in that the ‘‘shadow bands,’’ as discussed below. The presence of high-binding-energy tail is intense, the background in- the flat band is in agreement with numerical studies of tensity is high, and consequently the quasiparticle the Hubbard and t-J models (see Sec. II.E.2). The flat weight is anomalously small. The unusually high back- dispersion may constitute a microscopic basis for scaling ground is not due to inelastically scattered secondary theory with a large dynamic exponent z (Sec. II.F.11). electrons, but must be an intrinsically incoherent part of The fact that the Fermi level lies slightly away from the the spectral function (Liu et al., 1991). Indeed, angle- flat band (ϳ50 meV) is interpreted as coming from the resolved photoemission spectra of a conventional 2D (semi)metal TiTe2 show negligibly small background in- tensities (Claessen et al., 1992). The large incoherent spectral weight means that the motion of the conduction electrons in the high-Tc cuprates is highly dressed by electron correlations. Even when the quasiparticle peak is located very close to EF , it does not become a delta- function-like peak as predicted for a Landau Fermi liq- uid, but retains a highly asymmetric peak. In order to describe the unusual photoemission line shape, Varma et al. (1989) proposed the marginal Fermi-liquid phe- nomenology (see Sec. II.G2). Whether the quasiparticle spectral weight Z vanishes at EF or not and whether the quasiparticle lifetime width is proportional to the energy from EF (1/␶ϰ͉␻Ϫ␮͉) as in a marginal Fermi liquid are very subtle problems to be settled experimentally be- cause of the limited energy and momentum resolution of photoemission spectroscopy. Nevertheless, the analysis by Olson et al. (1990) using a Lorentzian function with a Fermi cutoff favors a lifetime width that varies linearly in energy over a relatively large energy scale of a few 100 meV. Although the distinction between Fermi liquid and marginal Fermi liquid can in principle be made only in the low-energy limit, the high Tc’s of ϳ100 K practi-

FIG. 155. Fermi surfaces of Bi2Sr2CaCu2O8 measured by cally make it meaningless to discuss energies lower than angle-resolved photoemission spectroscopy (Dessau et al., ϳ30 meV; the unusual lifetime behavior and large inco- 1993): Filled circles, Fermi surface crossings; striped circles, herent spectral weight seen on the latter energy scale (of locations where the band energy is indistinguishable from EF . the order of the resolution of photoemission) are suffi-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1198 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 156. (a) Angular distribution of photoelectrons from within 10 meV of the Fermi level in Bi2Sr2CaCu2O8 (Aebi et al., 1994); (b) outline of (a), emphasizing the stronger and weaker lines. cient to answer the above questions. Recently, the broad tail was interpreted as the 1/ͱ␻ dependence. This phe- nomenological result is derived from the real part of the self-energy proportional to ͱ␻, which results from the enhanced antiferromagnetic susceptibility in the intermediate-energy region (Chubukov and Schmalian, FIG. 157. Angle resolved photoemission spectra along the 1997; for experimental results, see Norman et al., 1998). (0,0)-(␲,␲) line. Left panel, angle-resolved photoemission The incoherent character of angle-resolved photoemis- spectra of Sr2CuCl2O2 (Wells et al., 1995); right panel, corre- sion spectroscopy may be related to the suppression of sponding spectra of Bi2Sr2CaCu2O8 (Dessau et al., 1993). coherence due to the large dynamic exponent zϭ4of the MIT, as discussed in Sec. II.F.9. the shadow band (Wells et al., 1995; see also Fig. 24): The question of why antiferromagnetic fluctuations Here, the antiferromagnetic Brillouin zone lies along the are strong in the superconducting cuprates has been fre- (␲,0)-(0,␲) line, and therefore the band disperses back quently asked in the context of magnetic mechanisms of beyond the (␲/2,␲/2) point. If the angle-resolved photo- superconductivity. Kampf and Schrieffer (1990) pre- emission spectra of Sr2CuCl2O2 are compared with those dicted that antiferromagnetic fluctuations in the para- of Bi2Sr2CaCu2O8 along the same (0,0)-(␲,␲) line (Fig. magnetic metallic state would induce scattering of elec- 154), a strong similarity between the antiferromagnetic trons by the antiferromagnetic wave vector Q [ϭ(␲,␲) insulator and the metal/superconductor is noted. The for the CuO2 plane] and produce a ‘‘shadow band,’’ flat band around (␲,0) in the metallic samples, on the which is shifted by Q from the original band. Aebi et al. other hand, is shifted well below (ϳ0.3 eV) the valence- (1994) observed ‘‘shadow Fermi surfaces’’ in optimally band maximum at (␲/2,␲/2) (see also Sec. II.E.2). The doped Bi2Sr2CaCu2O8ϩ␦ in the angular distribution of evolution of the electronic structure from insulator to photoelectron intensities with a fixed energy (EF), as metal/superconductor has been systematically studied shown in Fig. 156. According to that study, the original for underdoped Bi2Sr2CaCu2O8ϩ␦ in which a rare-earth Fermi surface is centered at the (␲,␲) point and the element is substituted for Ca or oxygen is reduced shadow Fermi surface is centered at (0,0). Ding et al. (Loeser et al., 1996; Marshall et al., 1996). In under- (1996a) observed corresponding shadow bands in the doped samples, the energy band around the (␲,0) point standard angle-resolved photoemission mode, but also is shifted away from EF even above Tc , and the Fermi- pointed out the possibility that the extra bands are due level crossing along the (0,0)-(␲,0) line disappears, as to superstructure modulation of the Bi-O planes. Theo- shown in Fig. 158(a). That is, a ‘‘normal-state gap’’ retically the existence of the shadow bands and shadow opens around (␲,0) while the Fermi-level crossing near Fermi surfaces has been quite controversial. Calcula- the (␲/2,␲/2) point is not affected. The resulting band tions using the ‘‘fluctuation exchange approximation’’ structure resembles that of a hole-doped antiferromag- showed that the correlation length of antiferromagnetic netic insulator with a small Fermi surface at the (␲/ fluctuation necessary to produce the shadow bands and 2,␲/2) point, although the inner half of the small Fermi shadow Fermi surfaces is only a few lattice spacings surface is missing [Fig. 158(b)]. The latter observation (Langer et al., 1995). On the other hand, shadow Fermi contradicts the appearance of the shadow bands as a surfaces are not visible in numerical studies of the tϪJ result of short-range antiferromagnetic correlation, since and Hubbard models although shadow bands have been the correlation should be more significant in the under- identified in the underdoped regime (Haas et al., 1995; doped regime and the whole small Fermi surface should Moreo et al., 1995; Preuss et al., 1995). be clearly visible. The normal-state gap which opens The photoemission spectra of the antiferromagnetic along the (␲,0) direction but not along the (␲,␲) direc- insulator Sr2CuCl2O2 shown in Fig. 157 clearly indicate tion has the same anisotropy as that of the supercon-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1199

FIG. 159. The temperature dependence of angle-resolved pho- toemission spectra at Fermi surface crossings along the (␲,0)-(␲,␲) line in oxygen-controlled Bi2Sr2CaCu2O8ϩ␦ (Ding et al., 1996b): (a) Tcϭ83 K sample; (b) Tcϭ10 K sample. Spec- tra at the (0,0)-(␲,␲) Fermi-surface crossing are also shown. Broken lines are reference Pt spectra.

the latter correlation would alter the Fermi surface con- siderably. The appearance of a normal-state gap con- spicuous around (␲,0) above Tc is consistent with a pair- FIG. 158. Photoemission results of Bi2Sr2CaCu2O8. (a) Band dispersions of optimally doped and underdoped samples (Mar- ing mechanism due to instability of the flat and shall et al., 1996); (b) Fermi surfaces of slightly overdoped and incoherent dispersion around (␲,0) characterized by the underdoped samples. zϭ4 universality class, as is discussed in Sec. II.F.9. The flat dispersion of single-particle process around (␲,0), ducting gap below T . The temperature dependence of generated from strong-correlation effects near the Mott c insulator and characterized by a large dynamic exponent the normal-state gap was studied by Ding et al. (1996b), zϭ4, may be destabilized, leading to preformed pairs as shown in Fig. 159. The gap closes at a characteristic with k2 dispersion with a binding energy indicated by T* T temperature , which is well above c for underdoped the pseudogap, while the part around (␲/2,␲/2) forms a samples and increases with decreasing hole concentra- more or less usual (but patchy) Fermi surface. There- tion. The doping dependence of T* seems to roughly fore, in the underdoped region with a pseudogap, the follow the temperature TR below which the Cu NMR holes are predominantly doped into the (␲,0) region be- relaxation rate shows a spin-gap behavior in cause of the flat dispersion and they are subject to bind- YBa2Cu3O7Ϫy . Recent tunneling measurements on un- ing to incoherent preformed pairs in this region. derdoped Bi2Sr2CaCu2O8ϩ␦ suggest an opening of the The shift of the chemical potential with hole doping pseudogap at temperatures even higher than that ob- has been studied from the shifts of core levels and the served in the photoemission (Renner et al., 1998). valence band for Bi2Sr2Ca1ϪxYxCu2O8ϩ␦ (van The origin of the normal-state gap has been discussed Veenendaal et al., 1994) and from oxygen-controlled from different points of view, including antiferromag- Bi2Sr2CaCu2O8ϩ␦ (Shen et al., 1991a). The results netic correlation (for example, Preuss et al., 1997), showed a shift of ϳ1 eV/hole in metallic superconduct- spinon pairing (Tanamoto et al., 1992), and preformed ing samples as in the overdoped region of Cooper pairs (Imada, 1993a, 1994b; see also Sec. La2ϪxSrxCuO4, whereas the shift is much larger (ϳ8 eV/ IV.F.11). Ding et al. (1997) found that in underdoped hole) in Y-rich semiconducting samples. The large shift Bi2Sr2CaCu2O8ϩ␦ , the minimum gap locus below T* in the semiconducting samples may be due to a finite coincides with the locus of gapless excitations above T* DOS of Anderson-localized states within the band gap and resembles the large Fermi surface of optimally of the parent insulator. doped and overdoped samples. The same scenario has been considered as an alternative to interpret the Fermi D. Quasi-one-dimensional systems surface of Fig. 158(b) (Marshall et al., 1996). This sug- 1. Cu-O chain and ladder compounds gests that the normal-state gap is due to precursors of singlet pairs (including spinon pairing scenario) above The 1D analog of the hole-doped cuprates may be Tc and not due to antiferromagnetic correlation, since very important, not only because the compounds can be

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1200 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 161. Optical conductivity spectra (at room temperature) for the Cu-O chain compounds. Solid lines, spectra polarized parallel to the chain directions; dashed lines, spectra polarized perpendicular to the chain direction; closed triangles, splitting FIG. 160. Lattice structures of prototypical Cu-O chain com- of the CT transitions due to interchain coupling; open circles, pounds. spin-excitation sideband of the CT transition; closed circles, the 2pϪ4s transitions. From Yasuhara and Tokura, 1998. a touchstone for the theoretical models of high-Tc cu- prates, but also because the 1D systems involve their transfer-type gap as in the layered cuprate compounds own novel properties such as spin-charge separation, a (see Secs. IV.C.1 and IV.C.2). The splitting of the CT spin-gapped metallic state, and associated superconduc- gap peak (indicated by vertical bars), probably due to tivity (see Secs. II.B and II.G.2). In this section, we show interchain coupling, is observed clearly for CaCu2O3 and some examples of the Cu-O chain and ladder com- also as a low-energy shoulder for SrCuO2. Some side- pounds whose electronic structures share many common band structures of the CT transition (marked with open features with those of the 2D CuO2-based compounds circles) are also discernible at about 0.5 eV higher than (Sec. IV.C). the main peak, as in the CuO2 sheet compounds Lattice structures of prototypical Cu-O chain com- (Tokura, Koshihara, et al., 1990), which are supposed to pounds are shown in Fig. 160. Sr2CuO3 or isostructural arise from coupling of the charge-transfer excitation Ca2CuO3 can be viewed as derived from K2NiF4-type with the spin excitation via strong Kondo coupling be- La2CuO4 by removing lines of oxygen in the CuO2 tween the 2p hole and the 3d spin in the final state. The sheet. In SrCuO2 (in the ambient pressure phase), the higher-lying structures are assigned to the transitions Cu-O chains are doubled in the repeating unit, as shown relevant to Cu 4s and Sr (or Ca) 5d (4d) states (Yasu- in the figure, with the edge sharing of the CuO4 square. hara and Tokura, 1998). Among them, the split peaks

The two Cu-O chains are magnetically nearly decou- (closed circles) at 4–5 eV with Eʈ chain are assigned to pled, since the exchange interaction between edge- transitions between the O 2p and Cu 4s states, whose sharing Cu spins is very weak compared with the corner- energy separation may represent the binding energy of sharing ones and gives rise to frustration in the magnetic the Zhang-Rice singlet of the excited O 2p hole coupled interaction between adjacent chains. In CaCu2O3, such with the Cu spin. edge-sharing double chains are connected via the corner The energy of the exchange interaction (J) between oxygen but in a non-planar manner. adjacent Cu spins on the chain is extremely large, 2000– The electronic parameters and structures in these 3000 K, in these Cu-O chain compounds (Motoyama Cu-O chain compounds bear a close analogy to the 2D et al., 1996; Suzuura et al., 1996; Yasuhara and Tokura, CuO2 sheet compounds, that is, the parent compounds 1998). The ␮SR experiment (Keren et al., 1993) and of high-Tc superconductors. Figure 161 shows the opti- neutron-scattering experiment (Yamada, Wada, et al., cal conductivity spectra for these compounds for the 1995) showed that 3D long-range antiferromagnetic or- light E vector parallel and perpendicular to the respec- der takes place only below TNϭ5 K for Sr2CuO3 and tive chain axes (Yasuhara and Tokura, 1998). A clear TNϭ8 K for Ca2CuO3. This means that the interchain gap structure polarized along the chain axis is seen exchange interaction (JЌ) is extremely small, JЌ /J around 1.6–1.8 eV, which is ascribed to the charge- Ϸ10Ϫ5, and hence these Cu-O chain compounds should

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1201 show an ideal 1D behavior for an Sϭ1/2 Heisenberg antiferromagnet over a wide temperature range, from several K to 1000 K [CaCu2O3, shown in Fig. 160(c), shows the spin ordering below TNϷ50 K due to the ap- preciable interchain coupling between the corner- sharing double chains]. Motoyama, Eisaki, and Uchida (1996) estimated the J values for Sr2CuO3 and SrCuO2 by fitting the T dependence of the magnetic susceptibil- ity with the rigorous theoretical result of Eggert, Af- fleck, and Takahashi (1994). In particular, they observed an isotropic drop in susceptibility below 20 K (but obvi- ously above TN), which could be scaled to the asymptotic (ln T)Ϫ1 dependence clarified by the Eggert- Affleck-Takahashi theory but was not recognized in the well-known Bonner-Fisher curve. An alternative method of determining the J value in a 1D antiferromagnet or 1D Mott-Hubbard insulator was developed by Suzuura et al. (1996). The spin excitation is observable as a sideband structure of an optical pho- non in the midinfrared absorption spectrum below the CT gap energy. The absorption spectra of all these 1D FIG. 162. Energy vs momentum curves of SrCuO3. From Kim Cu-O chain compounds show a ␭-like cusp around 0.5 et al., 1997. eV. A similar weak midinfrared absorption due to magnon-phonon coupled excitations is seen in the from analysis of core-level and valence-band photoemis- CuO2-layered compounds (Perkins et al., 1993; Loren- sion spectra and inverse-photoemission spectra using the zana and Sawatzky, 1995), although the spectral profile configuration-interaction cluster model (Maiti et al., is very different for the 1D and 2D cases. The distinct 1997). The deduced charge-transfer energies are indeed cusp structure of the Cu-O chain can be ascribed to the found to be small: ⌬ϭ0.0 eV for SrCu2O3 and ⌬ Van Hove singularity of the interband transition be- ϭ0.5 eV for CaCu2O3. The band gaps deduced from the tween spinon bands (corresponding to the Ϯ␲/2 point of combined photoemission and inverse-photoemission the 1D spin excitation dispersion) and its singular peak spectra are 1.5 and 1.7 eV, respectively, in good agree- position corresponds to the sum of ␲J/2 and the optical ment with the values deduced from optical studies. phonon energy (0.05 eV of the oxygen stretching mode Okada et al. (1996) analyzed the photoemission data us- in the present case). The singularity is so clear in the ing a multi-impurity cluster model and gave slightly absorption spectrum (Suzuura et al., 1996; Yasuhara and larger ⌬’s; the small difference is partly attributed to Tokura, 1998) that one can determine the J value quite intersite charge-transfer screening. accurately to be 3000 K. The difference between the J Spin-charge separation is one of the most outstanding values derived by these methods may be due to (1) an features of 1D systems. While Tomonaga-Luttinger- underestimate of the g value, assumed to be 2 in the liquid behavior has been predicted and experimentally analysis of the susceptibility (Motoyama et al., 1996) or tested in the low-energy excitation spectra of quasi-1D (2) the neglect of second-nearest-neighbor exchange in- metals (Sec. II.G.1), the spin-charge separation is pre- teractions. In any case, the exchange interaction J in the dicted to be more clearly visible in the single-particle 1D compounds is in the range of 2000–3000 K, consid- excitation spectra of insulating 1D systems. If one re- erably larger than in the 2D compounds (Cooper et al., moves an electron from a 1D system in which one elec- 1990b; Tokura, Koshihara, et al., 1990; Tokura, 1992a, tron is localized at each site that is antiferromagnetically 1992b); e.g., JϷ1600 K for La2CuO4 (Aeppli et al., coupled with the nearest neighbors, the created hole can 1989). This is perhaps partly due to the smaller charge- hop from site to site without disturbing the ordering of transfer energy ⌬ in the 1D chain compounds than in the the spin arrangement; or two parallel spins adjacent to 2D sheet compounds because of the different Madelung the initially created hole can propagate through the lat- energy, but still needs to be elucidated theoretically. tice without the motion of the hole. The former case Hole doping or electron doping for these Cu-O chain corresponds to the elementary excitation of a positive compounds has been widely attempted, so far unsuccess- charge (holon) and the latter is that of a spin (spinon), fully, except for the ladder type Cu-O compounds de- indicating the separation of a single hole into charge and scribed below. Nevertheless, the novel charge dynamics spin excitations in the 1D system. This spin-charge sepa- characteristic of correlated 1D systems show up in ration was studied using angle-resolved photoemission angle-resolved photoemission spectra in SrCuO2 which spectroscopy in the two-chain compound SrCuO2 (Kim virtually represent the charge dynamics of a single hole et al., 1996, 1997). The spectra show one band with in the system. strong dispersion and one with weak dispersion, the The magnitudes of the various electronic structure pa- widths of which are 1.2 and 0.3 eV, respectively, as rameters were estimated for SrCu2O3 and CaCu2O3 shown in Fig. 162. The observed bands are explained by

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1202 Imada, Fujimori, and Tokura: Metal-insulator transitions peaks in the spectral function A(k,␻) of the 1D tϪJ model calculated using exact diagonalization (with t ϳ0.6 eV and Jϳ0.15 eV). The peaks in A(k,␻) are as- signed to ‘‘holon’’ and ‘‘spinon’’ bands from comparison with the k- and ␻-dependent charge and spin suscepti- bilities: The energy scales of the holon and spinon bands are set by t and J, respectively. [The coupling between the two Cu-O chains in SrCuO2 is negligible for A(k,␻).] The result for the 1D insulator is contrasted with that for the 2D insulator Sr2CuCl2O2, which shows only a weakly dispersing feature whose width is of order J (Fig. 162). The different behaviors between the 1D and 2D systems have been attributed to strong coupling between the spin and charge excitations in the 2D sys- tem. There is, however, some controversy as to whether there is a spin-charge separation in the single-particle spectra of the 1D system. Suzuura and Nagaosa (1997) have examined the problem using the slave-boson for- malism and found that the holon (boson) excitation and the spinon (fermion) excitation are strongly coupled with each other. Let us turn to the Cu-O ladder compounds, which are natural extensions of the Cu-O chain compounds to- wards the CuO2 sheet structure. Dagotto, Riera, and Scalapino (1992) proposed an electronic mechanism for the formation of spin gaps in the Sϭ1/2 antiferromag- netically coupled double chains, or two-leg ladders, and argued that the ground state upon hole doping should be either superconducting or a charge-density wave. As dis- cussed in Sec. II.B for the category ͓I-2 M-2͔ of the ↔ MIT and also in Sec. II.G.2, the mechanism of supercon- FIG. 163. Temperature dependence of magnetic susceptibility ductivity induced by doping of carriers into a spin- for SrCu2O3 (upper panel) and Sr2Cu3O5 (lower panel): open gapped Mott insulator is similar to that proposed in a circles, observed; closed circles, corrected by subtraction of the dimerized system (Imada, 1991, 1993c). The pairing Curie component from the raw data. The solid line in the up- mechanism proposed in doped ladders shares the same per panel represents the calculated susceptibility assuming a origin as the dimerized case because coupled ladders spin gap of 420 K. From Azuma et al., 1994. and coupled dimerized chains are continuously con- nected in the parameter space of the models (Katoh and The temperature dependence of the magnetic suscep- Imada, 1995). The most representative compounds to tibility (␹) of two-leg SrCu2O3 and three-leg Sr2Cu3O5 is substantiate such an idea are a homologous series of shown in Fig. 163. The data after subtraction of the Cu- ladder systems SrnϪ1Cunϩ1O2n (nϭodd), as shown in rie component, which is due to impurity phases, show a Fig. 67 (Sec. III.B) for nϭ3 (SrCu2O3) and nϭ5 clear spin-gap behavior for SrCu2O3. The observed tem- (Sr2Cu3O5). In SrnϪ1Cunϩ1O2n ,Cunϩ1O2nsheets alter- perature dependence of ␹ in SrCu2O3 follows the theo- nate with SrnϪ1 sheets along the c axes of their ortho- retically predicted form for the two-leg ladder system rhombic cells (Hiroi et al., 1991). In each ladder, strong (Troyer, Tsunetsugu, and Wu¨ rtz, 1994), ␹(T) Ϫ1/2 antiferromagnetic interactions should occur through the ϭ␣T exp(Ϫ⌬s /T), with the spin-gap energy ⌬s linear Cu-O-Cu bonds along the a and b axes. By con- ϭ420 K and the prefactor (relating to the dispersion of trast, Cu-O-Cu bonds across the interface between lad- the excitation) ␣ϭ4ϫ10Ϫ3, as shown by a solid line in ders are formed with the right angle at the oxygen cor- the figure. By contrast, ␹ remains with no spin gap for ner of the Cu-O-Cu bond and hence the exchange Sr2Cu3O5 with a three-leg ladder structure, confirming interaction in these bonds must be much weaker and be the theoretical prediction. frustrated at the interface. The presence or absence of the spin gap depending on Rice and co-workers (Rice et al., 1993; for a review, the number of legs is also confirmed by measurements of see also Dagotto and Rice, 1996) theoretically investi- the nuclear spin-lattice relaxation rate 1/T1 in NMR gated the ground state of Sϭ1/2 multileg spin ladders (Azuma et al., 1994). The Arrhenius plot of 1/T1 for like SrnϪ1Cunϩ1O2n and concluded that the spin gap is SrCu2O3 at temperature TՇ⌬ gave the spin-gap value open or closed according to whether the leg number 680 K. An apparent difference between spin-gap values [ϭ(nϩ1)/2] is even or odd. They further predicted that deduced from the susceptibility (or Knight shift) and the light hole doping into the even-leg ladder keeps the spin NMR relaxation rate is due to the fact that the tempera- gap open and leads to singlet superconductivity. tures of the NMR data are too high to extract the true

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1203

gap by fitting for 1/T1 . When one takes the data at the lowest temperatures, the gap ⌬s extracted from 1/T1 agrees well with ⌬sϳ420 K. Alternatively, the large gap ⌬s deduced from 1/T1 at relatively high temperatures was interpreted in terms of the Majorana Fermion ap- proach by Kishine and Fukuyama (1997), who clarified that higher-energy excitations contribute to 1/T1 . Theo- retically the spin gap is calculated to be about J/2 pro- vided that the inter- and intra-chain exchange interac- tions are equal. Under the same assumption, J for SrCu2O3 is estimated to be about 840 K. However, thus deduced, the value of J in the ladder is considerably smaller than that of the Cu-O chain compounds de- scribed above, and even smaller than that for the CuO2 sheet compounds. This is rather puzzling because the J value of the ladders is expected to be between the chain and sheet compounds and rather close to the chain com- FIG. 164. Perspective view of the crystal structure of pounds due to similarity of the structures. One possible LaCuO2.5. From Hiroi and Takano, 1995. origin of this discrepancy is interladder coupling; an- other is a larger J for leg bonds than for rung bonds in La1ϪxSrxCuO2.5 was studied by photoemission and x-ray the ladder. In fact recent neutron results and susceptibil- absorption (Mizokawa et al., 1997). The evolution of the ity measurements are rather consistent with Jʈϳ2000 K electronic structure with hole doping is qualitatively the for legs and JЌϳ1000 K for rungs (Johnston, 1996; Ec- same as that in La2ϪxSrxCuO4 (see, for example, Chen cleston et al., 1997; Hiroi et al., 1997). A theoretical ex- et al., 1991), but the doping-induced spectral weight just planation of this effective exchange was proposed based above EF—monitored by oxygen 1s absorption on a correction due to next-nearest-neighbor and 4-spin spectroscopy—is distributed at higher energies than in exchange couplings (Nagaosa, 1997). In any case the re- La2ϪxSrxCuO4, resulting in a lower spectral weight at duction of ⌬s drives the system towards the quantum EF . Photoemission spectra show that the DOS at EF is critical point of the antiferromagnetic transition. This substantially lower than that in the 2D La2ϪxSrxCuO4, problem was discussed in the context of Zn substitution but still finite at EF in the metallic samples. The low for Cu in SrCu2O3 by Imada and Iino (1997). DOS at EF may be a characteristic feature of quasi-1D Carrier doping has proved to be very difficult in the metals (Sec. II.G.2). According to an LDA band- Srnϩ1Cunϩ1O2n series, yet some other Cu-O ladder structure calculation of LaCuO2.5 (Mattheiss, 1996a), the compounds are promising for the FC-MIT. p-d␴ band of the Cu-O ladder shows a finite dispersion La1ϪxSrxCuO2.5 was the first system (Hiroi and Takano, width of ϳ1 eV perpendicular to the ladder, compared 1995) in which holes were successfully introduced into to the ϳ2 eV dispersion along the ladder. This results in 2 2 the Cu-O ladder structures and the metallic state (x ͗vʈ ͘/͗vЌ͘Ӎ16, which is substantially smaller than the 2 2 տ0.15) emerged. However, superconductivity was not ͗vʈ ͘/͗vЌ͘ value of Ӎ100 for the 2D system La2CuO4. found in this compound. The crystal structure of Holes can be doped into another ladder compound LaCuO2.5 can be derived from the 3D perovskite struc- with misfit structure, (La, Sr, Ca)14Cu24O41 (McCarron ture (e.g., LaCuO3) by regularly removing oxygens in et al., 1988; Siegrist et al., 1988). The compound, whose rows, as shown in Fig. 164. The ground state of the par- structure is schematically shown in Fig. 165, is composed ent compound LaCuO2.5 was suggested from susceptibil- of Cu2O3 planes with two-leg ladders alternating up the ity measurements to be the singlet spin liquid. The NMR c axis with planes containing CuO2 chains with edge- (Matsumoto et al., 1996) and ␮SR (Kadono et al., 1996) shared links of CuO4, as [(La,Sr,Ca)2Cu2O2]7[CuO2]10 . 2ϩ experiments have, however, shown that the ground state The Cu -based compound La6Ca8Cu24O41 is viewed as is not the spin-gapped state but the antiferromagneti- the parent Mott insulator, and holes can be doped by cally ordered state below TN(Ϸ110 K), probably due to substitution of Sr for La, as in La6ϪySryCa8Cu24O41 . appreciable interladder interaction. Carter et al. (1996) found that the yϭ0 parent com- The hole-doped compounds La1ϪxSrxCuO2.5 exhibit a pound showed a temperature dependence of the suscep- reduction in electrical resistivity with x by several orders tibility consistent with weakly interacting 10(Ϯ0.5) Cu2ϩ of magnitude and become metallic for xу0.15. The spins per formula unit, all contributed from the chain transport properties in the metallic region are those of site. The Cu2ϩ spins on the ladder were virtually inert conventional Fermi liquids (Nozawa et al., 1997): The below 300 K due to the large spin gap. In fact, from the 2 63 resistivity shows a T dependence; the Hall coefficient is NMR measurement of 1/T1 and shift of Cu, Magishi small and negative. As for the magnetic susceptibility, et al. (1998) deduced the spin gap for the ladder unit as upon hole doping, it loses the spin-gap behavior and be- ⌬ladderϭ619Ϯ20 K and for the related compound of comes nearly temperature independent: ␹ Sr14Cu24O41 as 510Ϯ20 K. The ⌬ladder for the above y ϳ10Ϫ4 emu/mol for xу0.15 (apart from Curie contribu- ϭ0 compound is possibly even larger. Introduction of tions from impurities). The electronic structure of one hole by substitution of Sr for La is observed to re-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1204 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 166. The EЌ(ladder) optical conductivity at room tem- perature for single-crystalline samples of Sr14ϪxAxCu24O41 , AϭCa or Y. The anisotropic spectra with a and c polarization are shown for Sr3Ca111Cu24O41 in the inset. From Osafune et al., 1997.

of mobile holes was confirmed by optical spectroscopy with single-crystalline samples of Sr14ϪxCaxCu24O41 (Os- afune et al., 1997). In Fig. 166, the Eʈc (ladder axis) op- tical conductivity spectrum is shown for single crystals of FIG. 165. Crystallographic structure (upper) of Sr14ϪxAxCu24O41 , where A is Ca or Y. The inset com- pares the Eʈc and Eʈa spectra for x(Ca)ϭ11. The Eʈc (La,Y,Sr,Ca)14Cu24O41 (McCarron et al., 1988), in which (a) polarized quasi-Drude absorption below 1 eV gradually the CuO2 chain and (b) the Cu2O3 ladder stack along the b axis (vertical in this figure). The a axis is in the horizontal evolves with increase of x(Ca), indicating the quasim- direction in this figure, while the ladder and chain axes are etallic responses of the doped holes on the ladder site. along the c axis. The x(Y)ϭ3 crystal should show minimal holes among the crystals investigated in Fig. 166, and its conductivity move one Cu spin, which means that the doped hole spectrum shows a clear onset around 1.6 eV due to a (yÞ0) resides on the chain site, producing a nominal charge-transfer-type gap. As expected from the afore- nonmagnetic Cu3ϩ site or Zhang-Rice singlet. mentioned transport and magnetic features, the xϭ0 A rather complicated hole count is needed in the case crystal shows only a small weight of innergap absorp- of the solid solution Sr14ϪxCaxCu24O41 , in which holes tion, ensuring a tiny fraction of holes on the ladder sites appear to be transferred into the ladder with increase of at this composition. However, a strong absorption band x (Kato et al., 1996), although the average Cu valence is peaking around 3 eV emerges with decrease of x(Y) ϩ2.25 and unchanged with x. In the xϭ0 compound and hence has been attributed to excitation of the hole Sr14Cu24O41 , most of the holes seem to reside on the in the chain sites. Osafune et al. (1997) have further chains, forming spin dimers. Below 240 K, the chain ap- done a hole count on the ladder and chain sites as a pears to be subject to charge ordering (4kF CDW), and function of Ca content x, by estimating the spectral at lower temperatures (below Ϸ60 K) undergoes a sin- weight in the inner gap region, using the assumption that glet pairing of the spin sites (Carter et al., 1996; Matsuda the hole concentration is 1/2 on the chain site for the x et al., 1996). The singlet-triplet gap (Ϸ140 K) on the ϭ0(Sr14Cu24O41) compound. This assumption comes chain manifests itself in the temperature dependence of from the aforementioned fact that the CuO2 chains show the ␹ (Carter et al., 1996; Matsuda and Katsumata, 1996; charge ordering perhaps at a rate of one hole per two Cu McElfresh et al., 1989) as well as in the inelastic neutron- sites (Carter et al., 1996; Matsuda et al., 1996). Accord- scattering spectra (Eccleston et al., 1998). As mentioned ing to such an estimate (Osafune et al., 1997), the hole above, the spin gap on the ladder for Sr14Cu24O41 is concentration increases up to Ϸ0.20 for the x(Ca)ϭ11 much larger; ⌬ladderϭ619Ϯ20 K (from the NMR rate compound. The dominant feature of ␴(␻) in the metallic 1/T1)or510Ϯ20 K (from the NMR shift; Magishi et al., phase is similar to that of the high-Tc cuprates—a broad 1998). tail indicating a strongly incoherent contribution in the With increasing Ca content x, the resistivity is much transport. This feature is consistent with recent numeri- reduced (Yamane, Miyazaki, and Hirai, 1990; Kato cal studies of ladder systems (Tsunetsugu and Imada, et al., 1994) yet remains semiconducting. The presence 1997).

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1205

FIG. 168. Electrical resistivity and magnetic susceptibility of BaVS (Graf et al., 1995). The circle in the inset shows where FIG. 167. Crystal structure of BaVS . From Nakamura et al., 3 3 the structural transition occurs. 1994a. sulting in well-separated V chains running along the c Uehara et al. (1996) found that the resistivity ␳ in the axis: The V-V distance along the c axis is 2.84 Å while most Ca-substituted compound (xϭ13.6) still shows an that between the chains is as long as 6.72 Å (Sayetat upturn below 200 K but can be further reduced by the et al., 1982; Ghedira et al., 1986). The formal valence of application of pressure. In fact, they discovered the oc- VisV4ϩand hence the d-band filling is nϭ1. This com- currence of superconductivity at 9–12 K in the limited pound has attracted much interest because it shows a pressure range of 3–4.5 GPa. The quality of polycrystal- temperature-dependent second-order phase transition at line samples prepared under high oxygen pressure is not ϳ74 K between a paramagnetic conducting phase on adequate for a discussion of the charge dynamics in su- the high-temperature side and a paramagnetic semicon- perconducting Sr14ϪxCaxCu24O41 . The measurement of ducting phase on the low-temperature side. Below the anisotropic ␳ in superconducting single crystals of ϳ35 K, antiferromagnetic ordering has been suggested Sr14ϪxCaxCu24O41 under pressure is being attempted by NMR studies (Nishihara and Takano, 1981), neutron (Nagata, Uehara, et al., 1998). spin-flip scattering (Heidemann and Takano, 1980), and It is important to clarify whether the observed super- specific-heat (Imai et al., 1996) studies, although there conductivity in Sr14ϪxCaxCu24O41 is truly relevant to has been so far no experimental evidence for long-range such an exotic mechanism of carrier doping into the order. Indeed, a recent NMR study has shown that the spin-gapped Mott insulator as was proposed by Imada ground state is nonmagnetic with possible orbital order- (1991), Dagotto, Riera, and Scalapino (1992), and Rice ing (Nakamura et al., 1997). As shown in Fig. 168, the et al. (1993). Kumagai et al. (1997) and Magishi et al. resistivity in the high-temperature phase is relatively low Ϫ3 (1998) reported that the ⌬ladder deduced from the NMR (in the 10 ⍀ cm range, which is high for a metal) and Knight shift was reduced with hole doping on the ladder nearly temperature independent; it shows a minimum at site, from 510Ϯ20 K for x(Ca)ϭ0 to 180Ϯ10 K for x ϳ150 K, above which it behaves as metallic (d␳/dT ϭ9. The tentative linear extrapolation of the two (for Ͼ0) (Takano et al., 1977; Massenet et al., 1979; Mats- xϭ0 and 9) experimental points appears to intersect uura et al., 1991; Graf et al., 1995). Below ϳ74 K, the zero around xϷ13. However, later NMR data for high-x resistivity rapidly increases by several orders of magni- (Ϸ13) samples (Magishi et al., 1998) indicated a persis- tude down to the lowest temperature studied (the acti- tent spin gap in such a high-x region where supercon- vation energy being 100–300 K), with an inflection point ductivity was observed in the pressurized sample. Inelas- at ϳ50 K. The Seebeck coefficient is negative and shows tic neutron scattering appears to show that the x metallic temperature dependence above ϳ70 K. Below dependence of the gap is weak (Nagata, Fujino, et al., ϳ70 K, it becomes positive and semiconductor-like, in- 1998), while recent NMR data at xϭ11.5 under a pres- dicating that the 74-K transition is between a p-type sure of 29 kbar suggested collapse of the spin gap (May- semiconductor and a metal (Matsuura et al., 1991). The affre et al., 1997). Much work will be needed to clarify negative Hall coefficient at 100 K corresponds to an this intriguing new phenomenon in the spin-ladder com- electron density of 2.5ϫ1021 cmϪ3 or ϳ0.3 electrons per pounds as well as its possible relevance to high-Tc su- V atom. perconductivity in the CuO2-layered compounds. The magnetic susceptibility above ϳ74 K, displayed in Fig. 168, shows Curie-Weiss behavior with an effec- tive moment of 1.0–1.1 ␮B , which is smaller than the 2. BaVS3 spin-only value of 1.73␮B (for Sϭ1/2) or the free atomic The crystal structure of BaVS3 (Fig. 167) is based on value of 1.55␮B (for Jϭ3/2); it also shows a positive the hexagonal NiAs type such as VS, from which 2/3 of Weiss constant of 20–40 K (Kebler et al., 1980; Naka- the V atoms are removed or replaced by Ba atoms, re- mura et al., 1994a). This indicates that in going from

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1206 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 169. Extra entropy ⌬S due to conduction electrons in BaVS3 obtained by subtracting the specific heat of BaTiS3 from that of BaTiS3. From Imai et al., 1996. high to low temperatures, the spin correlation changes its character from ferromagnetic to antiferromagnetic. The peak at the transition point ϳ74 K is reminiscent of an antiferromagnetic transition, but no magnetic order has been found down to ϳ35 K. Although the simulta- neous decrease of the magnetic susceptibility and the electrical conductivity below ϳ74 K suggests a CDW formation or a spin Peierls transition, no corresponding FIG. 170. Upper panel, orthorhombic lattice parameters of pairing of V atoms has been detected in the structural BaVS3 as functions of temperature (Sayetat et al., 1982); lower studies. If the high-temperature conducting phase can be panel: corresponding changes of the electronic structures (Na- regarded as a metal, then the 74-K transition can be kamura et al., 1994a). The densities of state for the dz2 and thought of as a genuine Mott transition in the sense that dxyϪdx2Ϫy2 bands are shown by thin and thick solid lines, re- it occurs between paramagnetic phases without involv- spectively. ing any long-range magnetic order or lowering of struc- tural symmetry, although the transition is not of first cating that the magnetic correlation becomes less ferro- order. The extra (magnetic and electronic) entropy ⌬S magnetic. of conduction electrons in BaVS3 was obtained by sub- Apart from the electronic and magnetic phase transi- tracting the specific heat of BaTiS3 from that of BaVS3, tions described above, BaVS3 shows a second-order as shown in Fig. 169. The figure shows that ⌬S increases structural phase transition from the high-temperature from ϳ0JKϪ1molϪ1 at ϳ40 K to ϳ4.5JKϪ1molϪ1 at hexagonal phase to the low-temperature orthorhombic ϳ74 K, and then reaches R ln 2 at ϳ170 K. The gradual phase at ϳ250 K, below which the linear V chains pro- increase of the entropy above ϳ35 K implies that short- gressively become zigzag ones (Sayetat et al., 1982; range dynamic order gradually develops above ϳ35 K Ghedira et al., 1986). It is suggested that the zigzag dis- and even above ϳ74 K to a lesser extent. The increase tortion remains as a dynamical fluctuation above in the resistivity with decreasing temperature above ϳ250 K due to a large thermal ellipsoid for the V atom. ϳ74 K (between ϳ74 K and the resistivity minimum at As shown in Fig. 170, the orthorhombic distortion fur- ϳ150 K) would be consistent with this picture, that is, ther increases below ϳ74 K. with the fluctuation of the order parameter above the The 74-K phase transition is sensitive to structural de- second-order phase transition: Such a fluctuation effect fects such as S vacancies and Ti substitution for V. The is characteristic of quasi-1D systems. The appearance of 74-K peak in the magnetic susceptibility disappears for 3D antiferromagnetism below ϳ35 K may be caused by S-deficient samples (typical compositions being a weak but finite interchain interaction, since purely 1D BaVSϳ2.97), which become ferromagnetic with Curie systems should not show long-range order. The unusual temperatures of 16–17 K (Massenet et al., 1978). The magnetic behavior of BaVS3 below ϳ74 K was attrib- electrical resistivity is nearly constant, as in the stoichio- uted to frustrations due to the triangular arrangement of metric BaVS3, above ϳ150 K but gradually increases the V chains (Imai et al., 1996). In addition to the 74-K with decreasing temperature below it. As for Ti substi- and 35-K transitions, the magnetic susceptibility and the tution for V, the 74-K peak in the magnetic susceptibil- specific heat show an anomaly at ϳ160 K: Below ity disappears for such a small Ti concentration as ϳ160 K, the Curie constant slightly increases and the ϳ5%, while the resistivity above ϳ70 K gradually in- Weiss constant decreases from ϳ40 K to ϳ20 K, indi- creases with Ti concentration and becomes semiconduc-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1207

FIG. 172. High-resolution photoemission spectrum of BaVS3 near EF at various temperatures (Nakamura et al., 1994b). The gap starts to open at ϳ90 K, somewhat above the transition temperature of ϳ74 K.

FIG. 171. Valence-band photoemission spectrum of BaVS3 (Nakamura et al., 1994b) compared with the DOS from a octahedra along the hexagonal c axis splits the t2g level band-structure calculation (Itoh, Fujiwara, and Tanaka, 1994). into high-lying dz2 and low-lying dxy and dx2Ϫy2 levels. The sharp Fermi edge is not observed in the photoemission With the orthorhombic distortion below ϳ250 K, the spectrum. degeneracy of the low-lying level is lifted, leading to a population imbalance between the split d Ϫd 2 2 lev- tive; in BaV Ti S , the resistivity becomes 1–3 orders xy x Ϫy 0.8 0.2 3 els. (Here, the z axis is taken along the c direction.) of magnitude higher than that of BaVS over the whole 3 Further increase in the orthorhombic distortion below temperature range (Matsuura et al., 1991). Beyond that composition, the resistivity decreases with Ti concentra- ϳ74 K may indicate that one of the split dxyϪdx2Ϫy2 tion (Massenet et al., 1979). The origin of the increase in levels is fully occupied, resulting in a Mott insulating magnetic susceptibility below ϳ20– 30 K is not clear at state. In this picture, the multiband situation in the me- present. Because its magnitude is sample dependent and tallic phase would favor ferromagnetic coupling between its contribution is not detected in the specific heat, it is the V atoms. According to the band-structure calcula- likely that it is due to a magnetic impurity phase such as tions, near EF , the dxy and dx2Ϫy2 orbitals form narrow ferromagnetic BaVS3Ϫ␦ (Graf et al., 1995). The 74-K energy bands while the dz2 orbitals form relatively wide transition is also sensitive to external pressure: The tran- energy bands dispersing along the c axis. The Fermi sition temperature decreases under pressure and ap- level is pinned by the dxyϪdx2Ϫy2 bands. pears to vanish around 20 kbar (Graf et al., 1995). Although quasi-1D metallic behavior of the dz2 elec- From the systematic chemical trends of the electronic trons is expected from the crystal and electronic struc- structure of transition-metal compounds, it is expected tures, no anisotropic transport has been investigated so that the V 3d-S 3p hybridization will be extremely far on single crystals. On the other hand, it has been strong (the charge-transfer energy being ⌬ϳ0). Indeed, theoretically predicted that a 1D metal is not a Fermi LDA band-structure calculations have shown strongly liquid but behaves as a Tomonaga-Luttinger liquid. The hybridized V 3d andS3pbands. Mattheiss (1995) made single-particle spectral function of such a 1D metal band-structure calculations on the low-temperature should exhibit a characteristic non-Fermi-liquid behav- orthorhombic phase as well as on the high-temperature ior, as described above. Such an observation was re- hexagonal phase without magnetic ordering and found cently made by high-resolution photoemission spectros- no band-gap opening. Itoh, Fujiwara, and Tanaka (1994) copy, as shown in Fig. 172 (Nakamura et al., 1994a): As made calculations on the antiferromagnetic state but in other quasi-1D metals, the spectra exhibit an unusual also failed to open a band gap at EF . These results in- suppression of photoemission intensity at EF . From dicate that electron-electron interaction has to be taken room temperature to ϳ90 K, the spectral function near ␪ into account beyond the level of LDA or LSDA in order EF can be fitted to a power law: ␳(␻)ϰ͉␻Ϫ␮͉ [Eq. to account for the low-temperature insulating state, as in (2.376)] with ␪ϳ0.8– 0.9. This value is well above the the cases of other Mott insulators. Apart from the elec- upper limit of 0.125 for the exponent of the single-band tronic structure near EF , which will be discussed below, Hubbard model (Kawakami and Yang, 1990), indicating the overall photoemission spectra are well described by that the short-range interaction assumed in the Hubbard the DOS given by the band-structure calculations, as model is not sufficient to model electron-electron inter- shown in Fig. 171 (Itti et al., 1991; Nakamura et al., actions in BaVS3. According to Tomonaga-Luttinger 1994a). As shown in Fig. 170, the distortion of the VS6 liquid theory, the large exponent ␪ means that the 4kF

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1208 Imada, Fujimori, and Tokura: Metal-insulator transitions

ideal 2D system La2CuO4 (Mattheiss, 1995). In spite of the small mass anisotropy, the photoemission spectra have exhibited 1D-like characteristics presumably be- cause mass renormalization due to electron correlation is stronger for directions with a heavier band mass.

E. Charge-ordering systems

1. Fe3O4

Magnetite Fe3O4 forms in the spinel structure, in FIG. 173. Crystal structure and charge order in Fe3O4. Left which one-third of the Fe ions are tetrahedrally coordi- panel, the structure of Fe3O4; right panel, the ordered struc- 2ϩ 3ϩ nated with four oxygen ions and the remaining two- ture of Fe and Fe at the B sites proposed by Verwey. thirds are octahedrally coordinated with six oxygen ions, From Tsuda et al., 1991. as shown in Fig. 173. The tetrahedral and octahedral sites are called A and B sites, respectively. In addition to CDW correlation is the longest ranged and therefore a its celebrated magnetic properties, which are utilized in 4kF CDW instability, i.e., localization of one electron permanent magnets, Fe3O4 has attracted much interest per site, is most likely to occur in the real 3D material. for its conductivity transition at TVӍ120 K, called the This is consistent with the occurrence of the 74-K phase Verwey transition (Verwey, 1939). In going from the transition in BaVS3, in which no antiferromagnetic or- high-temperature phase to the low-temperature phase, der (2kF SDW) or Peierls distortion (2kF CDW) has the electrical resistivity increases by two orders of mag- been detected. In the low-temperature phase, a gap of nitude at TV , as shown in Fig. 174. The A-site Fe ions ϳ50 meV is opened at EF in the photoemission spectra are stable trivalent, while the B-site Fe ions are of mixed (Fig. 172). Interestingly, the gap starts to open at ϳ90 K, valency with the ratio Fe2ϩ :Fe3ϩϭ1 : 1. Here, the Fe2ϩ 3ϩ 3 2 3 2 well above the transition temperature of ϳ74 K, prob- and Fe ions have t2g t2g eg (Sϭ2) and t2g eg (S ably reflecting fluctuation effects above the transition ϭ5/2) configurations, respectively.↑ ↓ ↑ The electrical↑ ↑ con- point. It should be noted, however, that the mass anisot- duction and the Verwey transition occur on the B-site 2 2 2ϩ 3ϩ ropy given by the Fermi velocity ratio, ͗vʈ ͘/͗vЌ͘, is only sublattice. The Fe and Fe ions at the B sites are 2 2 ϳ3.8, which is much smaller than ͗vʈ ͘/͗vЌ͘ϳ100 for the ordered below TV .Fe3O4 is ferrimagnetic below TC

FIG. 174. Electrical conductivity of Fe3O4. From Miles et al., 1957.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1209

FIG. 175. Temperatures dependence of optical conductivity spectra of an Fe3O4 crystal. Inset shows the temperature de- pendence of the resistivity for the same crystal. From Park, FIG. 176. Temperature variation of the spectral weight of the Ishikawa, and Tokura, 1998. optical conductivity below the isobaric point energy 0.375 eV in comparison with the dc conductivity change. From Park, Ishikawa, and Tokura, 1998. ϭ858 K: the spins of the A and B sublattices are anti- parallel, resulting in a net magnetization of ϳ4␮B per unit formula. Since there are excellent review articles on spectra showing broad maxima around 0.5 eV are gov- Fe3O4 by Honig (1985) and by Tsuda et al. (1991), we erned by transitions among the B sites, while transitions shall mainly concentrate below on recent developments. between the A-site and B-site Fe show up in a higher- The ordered structure shown in Fig. 173 was first pro- energy region of 1.5–2.2 eV. Below TVϭ121 K (see the posed by Verwey (Verwey and Haayman, 1941; Verwey inset for the resistive transition on the same crystal), the et al., 1947). Since then, however, there has been much conductivity spectrum undergoes an abrupt change char- controversy about the ordered structure from both ex- acteristic of a first-order phase transition and shows a perimental and theoretical points of view, and no con- clear gap structure in the low-energy part. Further de- sensus has been reached till now. Detailed descriptions crease of temperature causes minimal change of the of the structural studies are beyond the scope of this spectral shape, and the gap magnitude at the lowest tem- article and we refer the reader to the literature (e.g., perature (10 K) is Ϸ0.12 eV, about 10 times kBTV . Kita et al., 1983). As for the charge-disordered phase Above TV , the optical conductivity remains finite in above TV , Anderson (1956) pointed out the importance the lower-energy region below 0.1 eV, but shows no con- of the short-range order, in which the total charge within ventional Drude features typical of metals. At tempera- 2ϩ aFe4 tetrahedron is conserved, i.e., two Fe and two tures between TV and 300 K, a conspicuous spectral 3ϩ Fe ions should form the Fe4 octahedron. The electrical weight transfer is observed from the peak region around Ϫ2 resistivity above TV remains rather high (Շ10 ⍀ cm) 0.5 eV to the lower-energy region with the increase of and nonmetallic ͓d␳(T)/dTϽ0͔ up to ϳ300 K (Fig. temperature, accompanying the isosbetic (equiabsorp- 174), which may be due to the short-range order above tion) point at 0.375 eV. With further increase of tem- TV . The entropy change at the Verwey transition is perature, the broad spectrum peaking at 0.5–0.6 eV close to ⌬SVϭR ln 2 per mole Fe3O4, half of the value gradually decreases and transfers spectral weight to the ⌬SVϭ2R ln 2 expected for a random distribution of higher energy above 1.2 eV. The transferred spectral Fe2ϩ and Fe3ϩ at the B sites (Shepherd et al., 1985). The weight below the isosbetic point energy, as deduced by role of the electron-phonon interaction or local lattice Eq. (4.4), is plotted in Fig. 176 as a function of tempera- distortion in the high-temperature phase has been a ture together with the dc conductivity (Park, Ishikawa, matter of controversy. Yamada, Mori, et al. (1979) pro- and Tokura, 1998). The transfer of spectral weight or the posed, based on their diffuse neutron-scattering data, evolution of the pseudogap feature is nearly parallel the existence of bipolarons above TV , which they called with the change in conductivity. Such a spectral shape 2ϩ ‘‘molecular polarons.’’ In a molecular polaron, two Fe and conspicuous weight transfer above TV up to near ions induce the displacement of surrounding oxygen 300 K may be interpreted in terms of pseudogap forma- ions. tion due to the short-range charge order or polaron (or A gap opening below TV , as well as possible short- bipolaron) formation (Ihle and Lorentz, 1986). range charge order, can be observed in the optical con- When considering the charge ordering, transport, and 5 3ϩ 6 ductivity spectra. Figure 175 (Park, Ishikawa, and optical properties of Fe3O4, only the d (Fe ) and d Tokura, 1998) displays the temperature dependence of (Fe2ϩ) configurations are relevant because the on-site the optical conductivity spectra below 1.5 eV, which was d-d Coulomb interaction is sufficiently large (U deduced from reflectance spectra on a single crystal. The ϳ5 eV) to prevent double occupancy of the B site by

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1210 Imada, Fujimori, and Tokura: Metal-insulator transitions the extra electrons. Then the transport properties of down-spin and up-spin bands from the A-site Fe are Fe3O4 are governed by the intersite Coulomb interac- fully occupied and empty, respectively (corresponding to tion, which is orders of magnitude smaller than U. Be- the high-spin Fe3ϩ ion), and the Fermi level is located cause the spins are fully polarized below TCϭ858 K, the within the t2g band of the B-site Fe, resulting in the ↓ extra electrons at the B site enter the t2g level. There- metallic state. In the LDAϩU calculations, two sublat- fore a ‘‘spinless’’ Fermion model with intersite↓ Coulomb tices for the B sites (B1 and B2 corresponding to the 3ϩ 2ϩ interaction (Cullen and Callen, 1973; Ihle and Lorenz, Fe and Fe sites, respectively) were assumed in a 1980) has often been employed: Verwey-type charge ordering, and the resulting DOS is shown in Fig. 177(b). While the local DOS of the B1 † 1 sublattice is similar to that of the A site (with the spin ϭϪt͚ ci cjϩ ͚ Uijnjnj . (4.14) H ͗i,j͘ 2 iÞj directions interchanged), the down-spin t2g band of the B2 sublattice is split into a nondegenerate occupied Here Uij (ӶU) is the nearest-neighbor (ϵU1) or next- band and a twofold degenerate unoccupied band due to nearest-neighbor [ϵU2 (ӶU1)] Coulomb energy, and the on-site U, as in Mott insulators. Thus a small gap of the orbital degeneracy of the t2g level has been ne- 0.34 eV is opened between the B2-site t2g band below glected for simplicity. The parameter region U ӷU ↓ 1 2 EF and the B1-site t2g band above EF . Although the ↓ with U1ϳ0.1 eV is considered to be realistic for Fe3O4. large on-site Coulomb interaction is a necessary condi- The Hall coefficient (Siratori et al., 1988) and Seebeck tion for opening the gap in Fe3O4, the magnitude of the coefficient (Kuipers and Brabers, 1976) above TV are gap is determined by the smaller intersite d-d Coulomb negative, consistent with the small (1/6) filling of the t2g interaction. The intersite Coulomb energy was estimated band. ↓ using the constrained LDA method to be 0.18 eV, which Below TV ,ifU2ϭ0, U1 determines only short-range is an order of magnitude smaller than the bare point- order of the Anderson type, and a large number of con- charge value. This reduction is due to p-d covalency and figurations are degenerate in the ground state (Ihle and dielectric polarization of the constituent ions. Lorenz, 1980). This degeneracy is lifted if U2Ͼ0, and the The overall line shape of the valence-band photoemis- Verwey order is stabilized. A finite transfer t or contri- sion spectra of Fe3O4 can be described as a superposi- butions from electron-lattice coupling, which arises from tion of that of the Fe2ϩ ion (at octahedral sites) and the different sizes of Fe2ϩ and Fe3ϩ ions, can stabilize those of the Fe3ϩ ions (at both octahedral and tetrahe- other structures (Lorenz and Ihle, 1979, 1982). Recently, dral sites) (Alvarado et al., 1976; Siratori et al., 1986; Lad the relative stability of the Verwey structure and the and Henrich, 1989; Park et al., 1997a). To be more exact, more complicated ‘‘Mizoguchi structure’’ (Mizoguchi, each of the Fe2ϩ and Fe3ϩ spectra consists of the main 1978), which was determined through analysis of NMR structure, which arises primarily from the dnLគ final data, has been studied by tight-binding band-structure state, and a high-binding energy satellite, which arises calculations (Mishra and Satpathy, 1993). This study primarily from the dnϪ1 final state, as described by the showed that the inclusion of the kinetic energy (band configuration-interaction cluster model (Sec. III.A.2; energy or transfer energy) favors the Mizoguchi struc- Fujimori, Saeki, et al., 1986; Fujimori, Kimizuka, et al., ture, which has a higher second-neighbor Coulomb en- 1987). Here, nϭ6 and 5 for the Fe2ϩ and Fe3ϩ compo- ergy. Below TV ,iftis negligibly small, a gap of 2(U1 nents, respectively. According to this picture, the spec- ϩU2) is opened between the occupied and empty B tral feature just below EF comes from the removal of a sites (Ihle and Lorenz, 1980). Above TV , partial de- t2g electron from the B2 site, which well corresponds to struction of the short-ranged order creates new states in the↓ band picture presented above. Detailed electronic the middle of the gap and these states are partially filled structure near EF and its temperature dependence were by electrons. Ihle and Lorenz (1980) calculated the elec- investigated by high-resolution photoemission studies. trical conductivity using this model and obtained an ex- According to Park et al. (1997a), the spectra show no cellent fit to the experimental data (with tϭ0 and U1 intensity at EF either above or below TVϳ120 K, but ϭ0.11 eV). The optical conductivity was calculated to the magnitude of the gap is reduced above TV . They show a peak at ␻ϭU1 , which may qualitatively explain interpreted the result as an insulator-to-insulator transi- the experimental data. The Seebeck coefficient above tion. The resistivity jump at TV is quantitatively ex- TV was temperature independent, consistent either with plained by a change in the magnitude of the gap. Chai- the localized (tϭ0) model or with strong correlation in nani et al. (1995) reported that, above TV , the spectra Hubbard-like models. show a finite DOS at EF , as shown in Fig. 178, and First-principles calculations of the electronic structure concluded that the transition is indeed a metal-insulator of Fe3O4 in the ferrimagnetic state were performed us- transition. However, the broad peak at ϳ1 eV below 6 3 2 ing the LSDA (Zhang and Satpathy, 1991; Yanase and EF , which was assigned to the A1g (t2g eg ) final state Siratori, 1984). Although the spin-polarized calculations at the Fe2ϩ site in the low-temperature phase,↑ ↑ survives in were able to explain the magnetic structure, it was not the high-temperature phase. This implies that short- possible to open a band gap within the t2g band. In range order persists above TV . The photoemission in- ↓ order to explain the insulating state of Fe3O4, LDAϩU tensity at EF is extremely weak and therefore, if the calculations were performed by Anisimov et al. (1996b). spectra are compared with the band DOS, an extremely According to the LSDA calculation [Fig. 177(a)], the small mk /mbӶ1 and hence ZӶ1 are implied (Sec.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1211

FIG. 177. Density of states of Fe3O4 in the ferrimagnetic state, calculated using (a) the LDA and (b) LDAϩU methods (Anisimov et al., 1996). The Fermi level is at 0 eV.

II.D.1). This is consistent with the poor/incoherent me- change proved to be the simultaneous charge and spin tallic state implied by the transport and optical studies. ordering in a superlattice structure depicted in Fig. 180 (Battle et al., 1990). 2. La1ϪxSrxFeO3 The charge disproportionation in the sample with near-xϭ2/3 was first detected by Mo¨ ssbauer spectros- A unique type of charge-ordering phenomenon is ob- served in La1ϪxSrxFeO3 below about 210 K around x ϭ2/3, involving valence skipping of Fe sites as well as antiferromagnetic spin ordering (Takano and Takeda, 1983; Battle et al., 1988, 1990). One end compound 5 LaFeO3 is an antiferromagnetic insulator with 3d (S ϭ5/2) configuration of Fe that is characterized by high Ne´el temperature (TNϭ738 K). The other end com- pound SrFeO3 is an antiferromagnetic metal with TN ϭ134 K. In the solid solution system La1ϪxSrxFeO3, the Ne´el temperature decreases monotonically with x (Park, Yamaguchi, and Tokura, 1998). (To be precise, the an- tiferromagnetic state shows spin canting with weak fer- romagnetism.) For xϾ0.1, the high-temperature para- magnetic phase above TN is semiconducting or conducting, while the resistivity shows a steep increase at temperatures below TN and insulating behavior in the antiferromagnetic phase, at least for xՇ0.7. When the doping level x is around 2/3 (0.67), the resistivity shows an abrupt increase at around 207 K with a narrow ther- mal hysteresis, as shown in Fig. 179, signaling the occur- rence of a first-order phase transition, that is, the charge- FIG. 178. Photoemission spectra of Fe3O4 near EF (Chainani ordering (CO) phase transition with simultaneous et al., 1995). The gap opens below the Verwey transition of antiferromagnetic spin ordering. This electronic phase 121 K.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1212 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 180. Schematic structure of the charge- and spin-ordered state with charge disproportion in La1ϪxSrxFeO3 (xϭ0.67).

non modes due to folding of the phonon dispersion branch, caused by lattice distortion at the CO transition, as well as the opening of a small (Ͻ0.1 eV) charge gap (Ishikawa et al., 1998). The electronic structures of the end members, LaFeO3 and SrFeO3, are quite contrasting, as expected from the opposite consequences of multiplet effects in 5 4 d and d compounds predicted in Sec. III.A.4. LaFeO3 is a rather ionic insulator with a large band gap. Doped holes have relatively pure p-hole character, as demon- FIG. 179. Temperature dependence of spontaneous magneti- strated by O 1s x-ray absorption spectroscopy (Abbate zation (MS), resisitivity (␳), and magnetic susceptibility (␹) for et al., 1992). On the other hand, SrFeO is strongly co- La Sr FeO with xϭ0.67 (Park, Yamaguchi, and Tokura, 3 1Ϫx x 3 valent and the band gap is closed (see Fig. 56): the 1998). The small spontaneous magnetization observed in the ground state is dominated by the d5Lគ configuration charge-ordered phase is due to the minimal spin canting in the 4 essentially antiferromagnetic state. rather than the d configuration (Bocquet et al., 1992b). Therefore the net d-electron number is not much differ- ent in the nominally Fe3ϩ and Fe4ϩ (and probably Fe5ϩ) copy (Takano and Takeda, 1983). According to a profile compounds. Because of this, the electrostatic energy of analysis of temperature-dependent Mo¨ ssbauer spectra, the system does not increase significantly with charge two kinds of Fe-site species with nominal valences of disproportionation, which is nominally denoted by Fe3ϩ and Fe5ϩ were identified with number ratio of 2:1. Fe4ϩϩFe4ϩ Fe3ϩϩFe5ϩ, compared to other com- → The Fe5ϩ valence is unusually high and seldom realized. pounds. More importantly, the small charge-transfer en- 5ϩ It is worth noting that Fe shows the full spin-up occu- ergy ⌬ϳ0 and the resulting small ‘‘band gap’’ (Egap 4ϩ pancy of t2g states, benefited by maximal Hund’s-rule ϳ0) for the Fe oxides would favor charge dispropor- coupling energy without loss of crystal-field splitting (10 tionation, as observed experimentally for La1ϪxSrxFeO3 Dq) energy. Such an anomalous valence state as well as with xϳ0.7 and for CaFeO3 (Takano and Takeda, 1983; the real-space ordering of the valence-skipping sites was Takano et al., 1991). confirmed by neutron-scattering measurements (Battle et al., 1990). The disproportionated charges are con- densed within the (111) plane in the cubic perovskite setting. From the magnitudes of the respective staggered 3. La Sr NiO moments on the Fe sites, estimated by neutron diffrac- 2Ϫx x 4 tion, the Fe valence was estimated to be ϩ3.4 and ϩ4.2, Hole-doped La2NiO4 also shows cooperative ordering respectively, perhaps corresponding to the Fe3ϩ and of holes and Ni spins below some critical temperature 5ϩ Fe species detected by Mo¨ ssbauer spectroscopy. The (Tco). The behavior of doped holes in a 2D antiferro- structural change of the lattice upon the CO transition, magnetic correlated insulator has been the subject of perhaps associated with breathing-type distortions of vital theoretical investigations concerned with the FeO6 octahedra, was revealed by electron diffraction anomalous normal-state properties of the high- measurements (Li et al., 1997), indicating a consistent temperature superconducting cuprates. The charge- superlattice structure, as shown in Fig. 180. The optical ordering (CO) phenomena of doped nickelates are one spectra also show a distinct splitting of the optical pho- such example and may be viewed as a prototype of the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1213

FIG. 181. Progression of spin and charge modulation in NiO2 planes with increasing hole concentration proposed by Tran- quada et al. (1995). microscopic phase separation that involves localization of holes in domain walls with an antiferromagnetic back- FIG. 182. Temperature dependence of normalized integrated ground. intensities of g1 and g⑀ peaks, corresponding, respectively, to Charge-ordering phenomena are seen at almost any the oxygen and spin ordering, in La2NiO4.125. From Tranquada doping level (pϭxϩ2y)inLa2ϪxSrxNiO4ϩy as long as et al., 1995. the compound remains insulating (pϽ0.7). The progres- sion of spin and charge modulations in NiO2 planes with increasing hole concentration p, which was proposed by that in the charge-ordered state in La2NiO4.125 the mag- Tranquada, Lorenzo, et al. (1995) on the basis of netic moments and the Ni-O bond lengths are both si- neutron-scattering results, is schematically shown in Fig. nusoidally modulated within the NiO2 plane. The maxi- 181. In this model, it is assumed that the doped hole is of mum value of the Ni moment is comparable to that of p orbital character, residing on the O site. In these spin- the undoped compound (La2NiO4), and the moments and charge-ordered states, the modulation wave vectors point transverse to the modulation direction, as shown in Fig. 181, with nearest neighbors antiparallel. There- are, with use of the orthorhombic cell specified by &aជ fore the ordering can be viewed as alternating stripes of ជ ជ ϫ&bϫc relative to the tetragonal, g⑀ϭ(1Ϫ⑀,0,0) for antiferromagnetic and hole-rich regions. The hole-rich the magnetic ordering and g2⑀ϭ(2⑀,0,1) for the charge regions act as antiphase-domain boundaries. The posi- ordering. The charge-ordering superlattice peaks are of tion of the domain walls is staggered in neighboring lay- course not from the charge-density-modulation itself but ers for the case of La2NiO4.125 . likely to arise from the accompanied modulation of the Inelastic neutron scattering has been used to measure Ni-O bond lengths. Here, the modulation wave number the low-energy spin excitations in the charge-stripe ⑀ is expected to coincide with the doping level p, which phase of La2NiO4.133 . Spin-wave-like excitations were was partly confirmed experimentally (Sachan et al., observed to disperse away from the incommensurate 1995). As shown below, however, the commensurate magnetic superlattice points with a velocity ϳ60% of locking at a particular modulation vector (mostly ⑀ that in La2NiO4 (Tranqnada, Wochner, and Buttery, ϭ1/3 and 1/2 for Sr-doped compounds) is observed, sug- 1997). gesting mesoscopic hole-segregation behavior (Chen As described in Sec. II.H.3, the stability of charge or- et al., 1993; Cheong et al., 1994). dering has been studied within the unrestricted Hartree- The most thoroughly characterized case is for ⑀ϭ1/4, Fock approximation (Poilblanc and Rice, 1989; Zaanen which was investigated by a single-crystal neutron- and Gunnarsson, 1989; Schulz, 1990a). The observed di- diffraction study for La2NiO4.125 (pϭ0.25) (Tranquada agonal domain-wall pattern in the Ni compound is in et al., 1993; Tranquada, Lorenzo, et al., 1995). The inter- accord with the results of calculations by Zaanen and stitial oxygens form a superlattice with unit cell of 3aជ Littlewood (1994) which were derived from a multiband ϫ5bជ ϫ5cជ, which gives an ideal interstitial density of y Hubbard model with parameters appropriate to the ϭ2/15 per unit formula. The compound shows a simul- NiO2 plane. They showed that Peierls-type electron- taneous ordering of dopant-induced holes and Ni spins, phonon coupling further stabilizes the formation of as modeled in the left panel of Fig. 181. Figure 182 dis- charged domain walls. plays the temperature dependence of the normalized in- Charge ordering or formation of charged diagonal do- tensity of superlattice peaks arising from interstitial oxy- main walls at higher doping levels, as shown in the cen- gen ordering (1ϩ1/3,1,0) below 310 K and Ni spin ter and right panels of Fig. 181, has been observed in ordering (1Ϫ⑀,0,0) below 110 K. The modulation wave Sr-doped La2NiO4 (Chen et al., 1993; Cheong et al., number is ideally ⑀ϭ3/11 but actually temperature de- 1994). Chen et al. (1993) observed superlattice spots at pendent, varying from 0.295 at 110 K to 0.271 at 10 K. the positions (1,Ϯ␦,0) or (Ϯ␦,1,0) in the [001] zone- It was proved from a detailed analysis of neutron- axis electron-diffraction patterns. This coincides with the scattering patterns by Tranquada, Lorenzo, et al. (1995) charge-ordering peaks of the aforementioned neutron

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1214 Imada, Fujimori, and Tokura: Metal-insulator transitions diffraction in the (hk0) zone of the reciprocal space. (Note that, even with the different choices of modula- tion vector in the neutron and electron-diffraction stud- ies, the relation ␦ϩ2⑀ϭ1 still holds.) In measurements of the [100] zone-axis electron-diffraction pattern, dif- fuse streaks are observed along the (00l) direction, peaked at lϭodd with a coherence length р60 Å. The modulation g is then (1Ϯ␦,0,1) or equivalently (2⑀,0,1), which is in accordance with the charge-ordering peaks in the neutron scattering for LaNiO4.125 . The modulation wave vector ⑀ϭ(1Ϫ␦)/2 increases generally as Sr doping increases (Chen et al., 1993). The value of ⑀ was found to increase rapidly from Ϸ1/4 near xϭ0toϷ1/3 for xϾ0.075. Near xϭ0 the ordering is perhaps induced by excess oxygen, as typically seen for La2NiO4.125 crystals. The modulation vector stays close to ⑀ϭ1/3 for a fairly wide x region of 0.075ϽxϽ0.4 and around ⑀ϭ1/2 between xϭ0.5 and 0.6. However, the value of ⑀ is known to be sensitive to oxygen stoichiom- etry (Sachan et al., 1995). The ⑀ϭ1/2 (or ␦ϭ0) ordering is the most fundamental FIG. 183. Temperature dependence of resistivity for one, where one hole is present for every other Ni atom La2ϪxSrxNiO4 with representative Sr concentrations (Cheong et al., 1994). The inset displays the temperature derivative of (or on its surrounding oxygen sites; see the right panel of the logarithmic resistivity. Vertical arrows indicate features in ϭ Fig. 181). Chen et al. (1993), observed that this ⑀ 1/2 the derivative for xϭ0.3 and 0.5, indicative of the charge- CO is essentially two dimensional without any intensity ordering phase transitions. modulation of the (␲,␲) point superlattice peak along the c axis. This is perhaps because two breathing-type shows a sharp dip at 240 K. A weaker and much broader distortions cannot sit right next to each other and the feature for xϭ0.5 is also evident at 340 K, suggesting coupling along the c axis is frustrated. In the picture of that the ⑀ϭ1/2 CO transition occurs around this tem- holes residing on Ni sites, the ⑀ϭ1/3 polaron lattice can perature. The truly 2D nature of the ⑀ϭ1/2 state is re- be derived from the ⑀ϭ1/2 lattice by first removing holes sponsible for the blurred nature of this transition. (polarons) along the diagonal of the NiO2 square lattice In the following, let us focus on the ⑀ϭ1/3 CO transi- (the x direction) and then repeating this process for ev- tion in La2ϪxSrxNiO4 (xϭ0.33) and related compounds. ery third polaron lattice unit cell along the y direction The CO transition for ⑀ϭ1/3 is robust against the choice (Chen et al., 1993). In the case of the ⑀ϭ1/2 state, intu- of divalent ions (Sr,Ba,Ca) (Cheong et al., 1994) and of itively and theoretically (Zaanen and Littlewood, 1994), trivalent ions (La,Nd) (Katsufuji, Tanabe, and Tokura, a ferromagnetic spin arrangement with alternating S 1998), ensuring that the transition is stabilized by the ϭ1 and Sϭ1/2 spins is expected to occur in the NiO2 commensurate value of the hole number. Figure 184 plane. However, there is no sign of ferrimagnetism in shows the change in the sound velocity (top panel) due the observed magnetic susceptibility (Cheong et al., to the phase transition at 240 K, together with the spe- 1994). For La2ϪxSxNiO4 (0.1ϽxϽ0.6), spin-glass-like cific heat and temperature derivatives of resistivity and behavior and weak ferromagnetism were observed be- magnetic susceptibility for La1.67Sr0.33NiO4 (Ramirez low 150 K, presumably due to antiferromagnetic (short- et al., 1996). The temperature dependence of the sound range) ordering. However, the susceptibility above 200 velocity and the specific heat show a pronounced K was field independent. Around xϭ1/3, the susceptibil- anomaly at the same temperature as observed in the ity decreased at 240 K, implying the onset of an elec- resistivity and the susceptibility. This CO phase transi- tronic phase transition, that is, a charge-ordering transi- tion at 240 K is continuous in nature. The change in the tion. sound velocity was observed to be insensitive to mag- A charge-ordering transition is clearly manifested in netic fields up to 9T, not as in a typical antiferromagnet. the transport and optical properties of La2ϪxSrxNiO4 This observation suggests that the CO transition for x with xϭ1/3 and 1/2. Figure 183 displays the temperature ϭ0.33 at Tcoϭ240 K does not involve long-range order- dependence of the resistivity in polycrystalline samples ing of the Ni spins. The entropy involved in the CO of La2ϪxSrxNiO4 with the inset showing the temperature transition is obtained by subtracting a smooth back- derivative of the logarithmic resistivity (Cheong et al., ground from the specific-heat data. It is estimated to be 1994). The ␳(T) exhibits semiconducting behavior be- ⌬˜Sϵ⌬S/Rϳ0.25(ln 2)ϭ0.17. The observed value is low room temperature and weak metallic behavior considerably smaller than that for the configurational above 600 K for x between 0.2 and 0.5. For xϭ0.3, ␳(T) ˜ increases abruptly at 240 K, indicative of a charge- (real-space ordering) hole entropy (⌬Sϭ0.64) or for the ordering transition. The temperature derivative of the Sϭ1 spin ordering (⌬˜Sϭ0.73), but obviously an order logarithmic resistivity, plotted in the inset of Fig. 183, of magnitude larger than that of the weak-coupling

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1215

FIG. 185. Temperature dependence of normalized integrated intensities of g⑀ (0.67,0,1) and g2⑀ (3.33,0,1) peaks, correspond- ing, respectively, to spin and charge ordering, in La2ϪxSrxNiO4 (xϭ0.33). From Kakeshita et al., 1997.

sponding to Ϸ1 hole carrier per Ni site. However, the Hall coefficient decreases rapidly with cooling for tem- peratures (ϳ300 K) near and above Tco and takes a FIG. 184. The normalized sound velocity ⌬␯/␯, specific heat negative value with a large absolute magnitude on en- capacity (C), and temperature derivatives of resistivity (␳) and tering the charge-ordered state below Tco (Tanabe et al., magnetic susceptibility (␹)inLa2ϪxSrxNiO4, with xϭ0.33 1998). The hole-type carrier conduction well above Tco From Ramirez et al., 1996. is characterized by a large Fermi surface. As the charge- ordered state is approached with decreasing tempera- CDW/SDW transition. Ramirez et al. (1996) argued this ture, however, a small electron pocket is produced due springs from independent hole-spin ordering. to charge-ordering fluctuations and subsequent static or- Composition-tuned (xϭ0.33) single crystals have sub- dering. For xՇ0.30, on the other hand, the Hall coeffi- sequently become available. According to recent neutron-scattering studies (Lee and Cheong, 1997; Kakeshita et al., 1998), the (static) charge and spin or- dering in an x(ϭ⑀)ϭ0.33 crystal, which were probed by the (4Ϫ2⑀,0,1) and (3ϩ⑀,0,1) superlattice peaks, takes place below Tcoϭ230 K and TNϭ180 K, respectively, as shown in Fig. 185. The temperature dependence of an- isotropic transport and optical conductivity spectra were investigated by Katsufuji et al. (1996). The temperature dependence of the in-plane (␳ab) and out-of-plane (␳c) resistivity is displayed in Fig. 186 together with the inset for the conductivity (inversed resistivity) on a linear scale. The resistivity of the xϭ0.33 crystal shows a large anisotropy (Ͼ102) between the in-plane and c-axis com- ponents, yet both show a simultaneously abrupt increase at Tco , as indicated by arrows in Fig. 186. An important point to be noticed is that the in-plane resistivity is me- tallic, i.e., d␳c /dTϾ0 above 500 K and then shows an upturn below 500 K, which is a much higher tempera- ture than Tco , while the c-axis resistivity is semicon- ducting or insulating, namely, d␳c /dTϽ0, over the whole temperature region up to 800 K, as can be clearly FIG. 186. Temperature dependence of in-plane and out-of- seen in the inset. The Hall coefficient above 300 K is plane resistivity in single crystalline La2ϪxSrxNiO4 (xϭ0.33). nearly constant and takes a small positive value, corre- From Katsufuji et al., 1996.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1216 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 187. Temperature-variation of the in-plane optical con- FIG. 188. Temperature dependence of resistivity (␳) for single crystals of La ϪxSr ϩxMnO with various hole concentrations ductivity spectra in the gap region for La2ϪxSrxNiO4 (x 1 1 4 ϭ0.33) (Katsufuji et al., 1996). The inset shows the variation of (x) (Moritomo et al., 1995); Solid curves, in-plane (␳ab) com- the charge gap in the charge-ordered state. ponents; dashed curves, out-of-plane (␳c) components. A steep increase in ␳ as indicated by upward arrows is due to the charge-ordering transitions. cient rises steeply as the temperature is lowered below TCO , indicating the presence of a small hole pocket The optical conductivity spectra show considerable (Tanabe et al., 1998). temperature dependence even above Tco : As signaled The CO instability crucially affects the electronic by the presence of the isosbetic (equi-absorption) point structure over a fairly large energy scale, at least up to 2 at 0.4 eV for the spectral change above Tco(ϭ240 K), eV in the case of a La2ϪxSrxNiO4 (xϭ0.33) crystal. The the spectral weight below 0.4 eV is transferred to a temperature variation of the low-energy part of the in- higher-lying energy region (at least up to 2 eV, not plane optical conductivity spectra (obtained by shown here) with decreasing temperature, apparently Kramers-Kronig analysis of the reflectivity spectra) is producing a pseudogap feature in the low-energy side shown in Fig. 187 (Katsufuji et al., 1996). Below Tco ,a (Katsufuji et al., 1996). This is contrasted with a nearly gradual opening of the charge gap as well as change in rigid shift of the optical conductivity spectrum below the gap magnitude 2⌬ with temperature is clearly ob- Tco . This effect is rather surprising but in accordance served. At the lowest temperature (ϭ10 K) shown here, with the semiconducting behavior of the resistivity be- 2⌬ is estimated to be about 0.26 eV by a linear extrapo- low 500 K. A charge-ordering fluctuation might be re- lation from the onset of conductivity as indicated by a sponsible for the anomalous spectral change and weight long-dashed line in the figure. There appears to be an transfer above TCO . appreciable amount of residual spectral weight below this estimated gap, which may arise from a slight devia- 4. La Sr MnO tion of the nominal hole concentration (p) from the 1Ϫx 1ϩx 4 commensurate value (pϭ1/3) or from a microscopic im- The manganite system La1ϪxSr1ϩxMnO4 with an n perfection of the charge-ordered lattice. The tempera- ϭ1 layered perovskite (or K2NiF4-type) structure (see ture dependence of the gap 2⌬, shown in the inset, is in Fig. 66) also shows a charge-ordered state near xϭ1/2. accordance with a BCS-like function represented by a The composition x represents the nominal hole number 3ϩ solid line. Examples include the charge-density-wave per Mn site, since the Mn valence state (LaSrMnO4; (CDW) or spin-density-wave (SDW) transitions, which xϭ0) is viewed as the parent Mott insulator. Figure 188 show BCS-like behavior both in wave amplitude and in displays the temperature dependence of the in-plane re- the energy gap. However, the ratio of the energy gap at sistivity (␳ab)inLa1ϪxSr1ϩxMnO4 crystals with x Tϭ0K[2⌬(0)] to the transition temperature (kBTco)is ϭ0 – 0.7 (Moritomo et al., 1995). In the figure, the tem- ϳ13 in this compound, which is significantly larger than perature dependence of the c-axis resistivity (␳c , dotted the theoretical value for a conventional weak-coupling curves) is displayed for xϭ0.3 and 0.5, showing highly CDW/SDW transition, 2⌬(0)/kBTcϳ3.5. Such a large anisotropic charge dynamics. In contrast to the case of value of 2⌬(0)/Tco implies that the charge gap in the the pseudocubic and nϭ2 layered perovskites (see Sec. charge-ordered state is affected by strong electron cor- IV.F), no ferromagnetic phase emerges, perhaps due to relation effects or that the Tco is suppressed due to fluc- the reduced double-exchange interaction. It is to be tuations arising from two dimensionality. noted in Fig. 188 that both ␳ab and ␳c curves for the x

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1217

FIG. 189. The MnO2 spin configuration in the charge-ordered state in La1ϪxSr1ϩxMnO4 (xϭ0.5). From Sternlieb et al., 1996.

ϭ0.5 crystal increase steeply at 220 K, as shown by up- ward arrows. A similar resistivity anomaly is also dis- FIG. 190. Temperature dependence of the (a) charge and (b) cernible for an xϭ0.6 crystal around 260 K. magnetic order observed in a neutron-scattering study (Stern- Corresponding to the anomalies in the resistivity as lieb et al., 1996). Inset in (a) shows the resitivity vs T. well as in the magnetic susceptibility for xϭ0.5, electron-diffraction studies (Moritomo et al., 1995; Bao, is in accordance with the steep increase in resistivity Chen, et al., 1996) revealed (1/4,1/4,0) and (3/4,3/4,0) su- shown in Fig. 188. Below the Ne´el temperature, TN perlattice peaks which are thought to arise from the ϭ110 K, the magnetic moments form an ordered struc- charge-ordered state. On the basis of the electron- ture, shown in Fig. 189, with dimensions 2&aϫ2&b diffraction patterns, Bao, Chen, et al. (1996) hypoth- ϫ2c, relative to the underlying tetragonal chemical cell. esized the presence of diagonally ordered charged It is worth noting that the MnO2 charge/spin order stripes, as in the case of La2NiO4.125 (⑀ϭ1/4, Sec. model for this nϭ1 layered perovskite is identical with IV.F.3), with modulation wave vector (1/4,1/4,0). This the (001) face projection of the well-known CE-type state can be viewed as a stripe-type ordering of Mn2ϩ charge/spin ordered pattern for pseudocubic (ortho- 3 2 4ϩ 3 (3dt2geg ; Sϭ5/2) and Mn (3dt2g ; Sϭ3/2). By con- rhombic Pbnm) perovskite manganites, R1/2A1/2MnO3 trast, Moritomo et al. (1995) assumed a much simpler (Goodenough, 1955; Wollan and Kohler, 1955). CE-type (1/2,1/2,0) ordering (or ⑀ϭ1/2), as in La2ϪxSrxNiO4 [x charge-ordering transitions and their modification under ϭ0.5; see Fig. 181(c)], associated with the additional a magnetic field will be discussed in Sec.IV.F. Suffice it fourfold periodic distortion perhaps due to orbital or- to say here that such a spin ordering must accompany dering. ordering of the orbital. Apparently, the degeneracy of More recently, a single-crystalline neutron-scattering the eg orbital is lifted by the tetragonal crystal field in a study was performed by Sternlieb et al. (1996), in which layered perovskite. In fact, the xϭ0 compound the aforementioned ⑀ϭ1/4 type scattering seen in the LaSrMnO4 shows coherent elongation of the out-of- electron-diffraction studies was not observed, though it plane Mn-O(apical) bond length due to the Jahn-Teller is observed in x-ray scattering. Sternlieb et al. (1996) effect. However, near xϭ1/2, the in-plane and out-of- viewed the (1/4,1/4,l) peaks as spurious (tentatively at- plane Mn-O bond lengths are nearly equal (Ϸ1.9 Å), tributing them to nonstoichiometry at the crystal sur- indicating a minimal static Jahn-Teller effect in heavily face) and presented the model of charge and spin order hole-doped manganites, despite the layered structure. in the MnO2 plane shown in Fig. 189. The temperature Orbital degrees of freedom should play an important dependence of the charge and spin order as measured by role in producing a rather large magnetic unit cell, as superlattice intensities is also shown in Fig. 190. The shown in Fig. 189. In accordance with this expectation, charge shows an alternating Mn3ϩ/Mn4ϩ pattern with the concomitant ordering of the orbital and charge with wave vector of (1/2,1/2,0), i.e., the ⑀ϭ1/2 charge order- a charge-exchange-type projection onto the ab plane ing, below Tcoϭ217 K. The onset of the charge ordering was recently demonstrated by Murakami et al. (1998)

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1218 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 191. Electronic phase diagrams in the plane of the doping concentration x and temperature for representative distorted perovskites of R1ϪxAxMnO3: (a) La1ϪxSrxMnO3; (b) Nd1ϪxSrxMnO3; (c) La1ϪxCaxMnO3; and (d) Pr1ϪxCaxMnO3. States de- noted by abbreviations: PI, paramagnetic insulating; PM, paramagnetic metallic; CI, spin-canted insulating; COI, charge-ordered insulating; AFI, antiferromagnetic insulating (in the COI); CAFI, canted antiferromagnetic insulating (in the COI). through x-ray resonant diffraction. Thus the ferromagnetic metallic state is stabilized by maximizing the kinetic energy of the conduction elec- F. Double-exchange systems trons for ␪ϭ0. As was discussed in Sec. II.H.1, this ef- fective transfer was treated using the average ͗cos ␪/2͘

1. R1ϪxAxMnO3 by the Hartree-Fock approach (Anderson and Hase- gawa, 1955). Giant negative magnetoresistance (MR) in the perov- To understand the important features observed ex- skite manganites near the Curie temperature (TC) was perimentally, we need to be aware of some factors that first reported by Searle and Wang (1969) for a crystal of are missing from the simplified double-exchange model La Pb MnO . Soon after, Kubo and Ohata (1972) 1Ϫx x 3 above. In the following, we focus on some experimental gave an essential account of this phenomenon using the observations that may be relevant to anomalous metallic double-exchange Hamiltonian given in Eqs. (2.9a)– features, although the recent extensive research on the (2.9c). They calculated the magnetoresistance in the so-called ‘‘colossal’’ magnetoresistance has generated a Hartree-Fock approximation with additional consider- huge number of reports in the scientific and technologi- ation of the lifetime effect of the quasiparticles scattered cal literature. by spin fluctuations. This model contains the essential We show in Fig. 191 electronic phase diagrams in the ingredient of the double-exchange mechanism elabo- plane of the doping concentration x and temperature for rated by Zener (1951), Anderson and Hasegawa (1955), representative distorted perovskites of R A MnO and de Gennes (1960). 1Ϫx x 3 (Schhiffer et al., 1995; Tokura, Tomioka, et al., 1996b); AMn3ϩ ion shows the electronic configuration of these include (a) La Sr MnO , (b) Nd Sr MnO , t3 e1 . The e -state electrons, hybridized strongly with 1Ϫx x 3 1Ϫx x 3 2g g g (c) La Ca MnO , and (d) Pr Ca MnO . As the tol- p 1Ϫx x 3 1Ϫx x 3 theO2 state, are subject to the correlation effect but erance factor or equivalently the average ionic radius of can be itinerant when hole-doped, while the t2g elec- the perovskite A site decreases from (La,Sr) to (Pr,Ca) trons, less hybridized with 2p states and stabilized by through (Nd,Sr) or (La,Ca), orthorhombic distortion of crystal-field splitting, are viewed as forming the local the GdFeO type increases, resulting in the decrease of ϭ 3 spin (S 3/2). In Eqs. (2.9a)–(2.9c), t stands for the the one-electron bandwidth W, as described in detail in transfer of eg electrons between nearest-neighbor pairs Sec. III.C.1. With reduced transfer tij or W, other elec- (i,j) of sites, and JH the on-site ferromagnetic (Hund’s- tronic instabilities, such as Jahn-Teller-type electron- rule) coupling between the eg electron spin (s) and the lattice coupling, charge/orbital-ordering, and antiferro- t2g local spin (␴). According to estimates from photo- magnetic superexchange interactions, may become emission studies (Sarma, Shanthi, et al., 1996) and opti- important and compete with the ferromagnetic double- cal spectroscopy (Arima and Tokura, 1995; Okimoto exchange interaction. In fact, such a competition pro- et al., 1995a), JH is about 2 eV and probably larger than duces a number of intriguing magnetoresistive and mag- the one-electron bandwidth W(ϭ2zt, where z is the co- netostructural phenomena, some of which will be ordination number). introduced later in this section. In the limit of large JH /W, with classical treatment of The compound La1ϪxSrxMnO3 is the most canonical S, the absolute magnitude of the effective transfer de- double-exchange system with the maximal one-electron pends on the relative angle ␪ between neighboring t2g bandwidth W. The parent compound LaMnO3, with the i j 3 1 spins at sites and (Zener, 1951; Anderson and Hase- t e configuration of the Mn3ϩ ion, shows the ordering gawa, 1955; Millis et al., 1995), 2g g of the two eg orbitals, d3x2Ϫr2 and d3y2Ϫr2, alternating teffϭt cos͑␪/2͒. (4.15) from site to site on the (001) plane, in an ‘‘antiferro-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1219

FIG. 193. Magnetic-field effect on the resistivity of a La1ϪxSrxMnO3 (xϭ0.175) crystal (Tokura et al., 1994): (a) T dependence of resistivity; (b) isothermal magnetoresistance.

netic field, as reported by Asamitsu et al. (1995, 1996). Various structual phase changes manifest themselves in anomalous resistivity curves, as indicated by triangles in FIG. 192. Temperature dependence of resisitivity in crystals of Fig. 192. The possibility of phase separation in a simple double-exchange model was also examined theoretically La1ϪxSrxMnO3 (Urushibara et al., 1995): Arrows, the critical temperature for the ferromagnetic transition; triangles, critical by Yunoki et al. (1998). temperature for the structural (rhombohedral-orthorhombic) The magnetic-field effect on the resistivity is shown in transition. Fig. 193 for an xϭ0.175 crystal (Tokura, Urushibara, et al., 1994; Urushibara et al., 1995). The magnetic field orbital’’ fashion, as was discussed in Sec. II.H.1. This greatly reduces the resistivity near TC by shifting the orbital ordering is accompanied by collective Jahn-Teller resistivity maximum to a higher temperature due to sup- distortions (Kanamori, 1959). Such an orbital ordering pression of the spin-scattering of eg-state carriers. The produces ferromagnetic spin ordering within the (001) change in resistivity appears to be scaled with the plane and antiferromagnetic coupling along the c-axis change in the magnetization of the system if the magne- [A-type ordering in Fig. 38(c)]. Anisotropic dispersion tization is small. Figure 194 shows the isothermal mag- of spin-wave excitations in LaMnO3 was observed by netoresistance near and above TC as well as the tem- single-crystal neutron scattering where the dispersion in perature variation of the resistivity as a function of the (001) plane was ferromagnetic while that alone in normalized magnetization, m(T)ϵM/Ms (where Ms is the (001) axis was antiferromagnetic (Hirota et al., the saturation magnetization, about 3.8␮B), for 1996). La1ϪxSrxMnO3 (xϭ0.175). The good scaling of the The LaMnO3 undergoes a phase transition to the magnetoresistance curves with m(T) in the paramag- spin-canted phase for xϽ0.15 (Kawano et al., 1996), al- netic state as a function of the ferromagnetic magnetiza- though for xϾ0.10 the spin-ordered phase is almost fer- tion indicates that the change in resistivity on the romagnetic. With further doping, a ferromagnetic phase present scale is governed by temperature- or magnetic- appears below TC which steeply increases with x up to field-dependent spin scattering of the eg-state carriers. 0.3 and then saturates. Figure 192 shows the tempera- Solid lines in the figure represent the result calculated by ture dependence of the resisitivity in crystals of Furukawa (1995a, 1995b, 1996), who applied the dy- La1ϪxSrxMnO3 (Urushibara et al., 1995). Arrows in the namic mean-field approximation, that is, the infinite- figure indicate the ferromagnetic transition temperature dimensional approach discussed in Sec. II.C.6, to the TC . The ferromagnetic insulating (FI) phase is present double-exchange model of Eqs. (2.9a)–(2.9c) with an in a fairly narrow x region (xϭ0.10– 0.17), in which the additional approximation for the t2g local spin, namely, double-exchange carrier is subject to (Anderson) local- a classical treatment in the limit Sϭϱ. In this limit, the ization but can still mediate the ferromagnetic interac- t2g spins provide as an external magnetic field in which tion between neighboring sites and realize the ferromag- one can solve for the dynamics of the eg electrons are netic state in a bond-percolation manner. Around x described in infinite dimensions. Although the spatial in- ϭ1/8, a possible ordering of polarons was observed in a tersite correlations and the Coulomb repulsion effect of single-crystal neutron-diffraction study (Y. Yamada the eg electrons responsible for the Mott insulator are et al., 1996). Furthermore, the orthorhombic-to- ignored, this approximation improves upon the Hartree- rhombohedral structural transition also couples with the Fock-type treatment of Kubo and Ohata by taking ac- magnetic transition around xϭ0.17, giving rise to a count of dynamic fluctuations. The result for JH /Wϭ4 unique magnetostructural transition or switching of the well reproduces the experimental results, confirming crystal structure upon application of an external mag- that the double-exchange mechanism is a primary source

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1220 Imada, Fujimori, and Tokura: Metal-insulator transitions

orbital degeneracy in the limit dϭϱ by taking account of the coupling to dynamic but classical local lattice dis- tortions. From this treatment an enhancement of the re- sistivity around TC with suppressions both above and below it is obtained, in agreement with the experimental results. In fact, the presence of a dynamic Jahn-Teller distortion has been supported indirectly in some experi- ments, in particular for the narrower-bandwith system La1ϪxCaxMnO3 (Dai et al., 1996). Another possible origin of the increased resistivity near and above TC as well as of the effective suppres- sion of resistivity by an external magnetic field is Ander- son localization of the double-exchange carriers arising from the random potential which is inevitably present in the solid solution system (Varma, 1996), or to antiferro- magnetic fluctuations which compete with the double- exchange interaction (Kataoka, 1996). One of the im- portant issues relating to all these instabilities is how we

FIG. 194. Isothermal magnetoresistance near and above Tc as take into account the orbital degrees of freedom of the well as temperature variation of resistivity as a function of eg-state electrons and their possibly strong intersite cor- normalized magnetization, m(T)ϵM/Ms (where Ms is the relation or coupling to the lattice degrees of freedom for saturation magnetization, about 3.8␮B), for La1ϪxSrxMnO3 at metals near the Mott insulator phase. In the Mott insu- xϭ0.175 (Tokura et al., 1994): solid line, calculated results by lating phase, antiferromagnetic order, orbital order, and the mean-field approach (Furukawa, 1995a); dashed line, the Jahn-Teller distortions are all present, while the primary J 0 (weak-coupling) limit. Here J is the Hund’s-rule coupling → origin of the insulating phase is the strong-correlation energy normalized by the bandwidth. effect due to a charge gap of the order of 1 eV. There- fore the metallic phase near this insulator is under the of the observed large negative magnetoresistance as well critical fluctuation of the Mott transition even when the as of the temperature dependence of the resistivity saturated ferromagnetic moment appears. Strong fluc- around TC . tuations toward orbital, spin, and Jahn-Teller ordering However, fluctuations around the Curie temperature have to be considered. The importance of orbital degen- as well as the strongly incoherent character of the charge eracy was stressed in studies in infinite dimensions (Ro- dynamics even at low temperatures, discussed below, zenberg, 1996b) and in the slave-boson approximation cannot be well accounted for in the dϭϱ approach (Fre´sard and Kotliar, 1997). In particular, the former within single-orbital and rigid lattice systems. A contro- emphasized the proximity of the Mott insulating phase versial issue is the semiconducting or insulating behavior to the experimentally observed ferromagnetic metal above TC in the rather low-x region, e.g., xϭ0.15– 0.20 phase within the treatment in dϭϱ. for La1ϪxSrxMnO3, as shown in Fig. 192, or in a wider x Strong on-site coupling betweeen the eg-electron spin region for narrower-W systems such as La1ϪxCaxMnO3 and t2g local spin and the resultant critical dependence [see Fig. 191 where the ferromagnetic metallic phase is of the electronic structure on spin polarization shows up very limited (Schiffer et al., 1995)]. In such a region of x, in the temperature variation of the optical conductivity the negative magnetoresistance effect is most pro- spectra (Okimoto et al., 1995a, 1997). We show in Fig. nounced around TC and hence the origin of the semi- 195 the temperature dependence of the optical conduc- conducting transport is of great interest. Millis et al. tivity spectrum for a La1ϪxSrxMnO3(xϭ0.175) crystal (1995) pointed out that the resistivity of the low-doped (Okimoto et al., 1995a). At high temperatures near and crystals above TC is too high to be interpreted in terms above TC , the spectrum is characterized by an of the simple double-exchange model and ascribed its interband-like transition centered on 0.9 eV, while at origin to a dynamic Jahn-Teller distortion. Static and low enough temperatures it shows an intraband-like coherent Jahn-Teller distortions disappear above x transition. The spectral weight appears to be transferred ϭ0.1 in La1ϪxSrxMnO3, judging from the crystal struc- from interband to intraband region with decrease of tural phase changes between orthorhombic forms temperature. This unconventional temperature depen- (Kawano et al., 1996). However, Jahn-Teller coupling dence of the optical conductivity spectrum below about energy is large (of order 1 eV), and the distortion should 3 eV is ascribed to a change in the electronic structure of remain finite when carrier itinerancy is reduced by dis- the eg electrons with the spin polarization. The eg con- ordered spin configurations above TC . Therefore the duction band is split into two bands due to JH being mean amplitude of the Jahn-Teller distortion is critically larger than the bandwidth W (Furukawa, 1995a; Ha- reduced as the temperature is lowered below TC . This mada et al., 1995; Pickett and Singh, 1995). Above TC , idea was examined theoretically by Millis et al. (1996) in in the absence of net spin polarization, the up-spin and a classical approximation of a Jahn-Teller fluctuation. down-spin bands overlap, with identical density-of-states Millis et al. solved the double-exchange model with an profiles and equal populations. For TӶTC , however,

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1221

FIG. 196. Temperature dependence for intraband excitations: closed circles, dependence of the Drude weight D; open circles, dependence of the total low-energy spectral weight Neff (Okimoto et al., 1995a, 1997). The change in the square of 2 the magnetization (M/Ms) (where Ms is the saturation mag- netization) is about 3.8␮B .

site exchange coupling originating from the 3d Hund’s- FIG. 195. Temperature dependence of the optical conductivity rule coupling (Okimoto et al., 1997). It is worth pointing out here some puzzling features of spectra of a La1ϪxSrxMnO3 (xϭ0.15) crystal (Okimoto et al., 1995a). The hatched curve represents the temperature- the ferromagnetic metallic state at low temperatures. independent part mainly arising from 2p-3d interband transi- From the perfect spin polarization in the metallic tions. The inset shows magnified spectra in the low-energy re- ground state, we might expect the simplest spinless (per- gion. fectly spin-polarized) Fermi-liquid behavior. However, the optical conductivity spectra at low temperatures are the only up-spin band is filled with saturated (100%) far from the simple Drude type: As an example, a low- polarization, as shown in the inset. (This situation is in energy spectrum of the optical conductivity is shown in contrast with the case of conventional ferromagnetic the inset of Fig. 195 for the ferromagnetic ground state transition metals; for example, in Ni metal the exchange (taken at 9 K) of a barely metallic La1ϪxSrxMnO3 (x splitting of the conduction band is incomplete and the ϭ0.175) crystal. The coherent contribution is present maximum spin polarization of the conduction electrons only in the far-infrared region, say below 20 meV, as a is only 11%.) In accordance with such a sensitive depen- sharp peak centered at zero energy, while most of the dence of electronic structure on the spin-polarization, low-energy spectral weight is dominated by the slowly low-energy optical conductivity spectra are expected to ␻-dependent incoherent contribution. With decreasing show a drastic change over the energy scale of JH . The temperatures these low-energy spectral weights in total excitation band around 1.7 eV observed at higher tem- are transferred from the aforementioned interband tran- peratures corresponds to an interband transition be- sitions between the exchange-split bands. The main tween the exchange-split conduction bands, while the panel of Fig. 196 shows the temperature dependence of lower-energy excitation observed at low temperatures is the Drude weight D and of the total low-energy spectral assigned to an intraband excitation within the up-spin weight Neff for intraband excitations (Okimoto et al., band. Thus the maximum of the interband transition 1995a, 1997). The Neff , keeps on changing, even in the should correspond to the exchange splitting of the con- low-temperature regions, say below 50 K, implying the duction band or to SJH (where Sϭ3/2 is the magnitude presence of a persistent carrier-scattering mechanism of t2g local spins) which is about 1.7 eV for even in the almost fully spin-polarized state. This is in La1ϪxSrxMnO3 at xϭ0.175. The observed value is in ac- contradiction with the expectation of the simple double- cordance with an estimate (SJHϳ2 eV) obtained by exchange model because NeffϳD should be scaled by photoemission (Saitoh, Bocquet, et al., 1995b; Park, M2 in the treatment by Kubo and Ohata (1972) as well Chen, et al., 1996) and other optical spectroscopies as in that by Furukawa (1995a). (Arima and Tokura, 1995). On the other hand, a similar Importantly, the Drude weight D (i.e., the spectral analysis of the spectrum for a La1ϪxSrxMnO3 (xϭ0.3) weight of the coherent contribution) also increases with crystal (Okimoto et al., 1997) gives a small value, about decrease of temperature, but remains about 1/5 of the 0.9 eV, indicating that the apparent value is dependent total low-energy spectral weight. This means that carrier on the doping level x. One plausible reason for this is motion in the fully spin-polarized ground state is mostly that the oxygen 2p hole character of the doped carrier incoherent. The ordinary Hall coefficient observed for increases with x and hence effectively reduces the on- the ferromagnetic metallic state in La1ϪxSrxMnO3 (x

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1222 Imada, Fujimori, and Tokura: Metal-insulator transitions

ϭ0.175) gives the carrier number nϳ1 hole/Mn site (Asamitsu and Tokura, 1998; Matl et al., 1997). The nominal value of m*/n is given as the inverse of D in terms of the Drude model, that is, about 80 in the present case, and hence m*/m0 should be as large as 80. This is, however, obviously in contradiction with the fairly small value (Ϸ3 – 5 mJ/mol K2) of the electronic specific-heat constant (Woodfield, 1997), which corre- sponds to the unrenormalized band mass between 2m0 and 3m0 . The conventional Drude model is not appli- cable to the ferromagnetic metallic state in La1ϪxSrxMnO3 for small x, where carrier motion is mostly incoherent. The broad structure of ␴(␻) is similar to many other transition-metal compounds near the Mott insulator, as we notice from the cases in other com- pounds discussed in this section, although Fig. 195 shows even broader and less structured features than the other cases. (Compare, for example, with Figs. 75, 86, 94, 102, 107, 114, 130, 136, 147, 161, 166, 175, 187, 210, 221, and 228.) In the case of La1ϪxSrxMnO3, discussed in Secs. II.H.1 and II.H.2, strong scattering in the fully spin- polarized ground state may be ascribed to the orbital FIG. 197. CE-type charge ordering with orbital and antiferro- degrees of freedom in the eg state. Interband transitions magnetic spin ordering. within two eg bands were examined as a possible origin of the broad structure at relatively large ␻ (Shiba, ties competing with the double-exchange interaction Shiina, and Takahashi, 1997). An exotic framework for complicate the phase diagram. In particular, when the the spin-charge separation in 3D was applied, and the doping x is close to the commensurate value xϭ1/2, a flat structure of the band dispersion leading to an orbit- charge-ordering instability occasionally shows up. In the ally disordered state was claimed to be the origin of a case of Nd1ϪxSrxMnO3 [Fig. 191(b)], the ferromagnetic broad incoherent response (Ishihara, Yamanaka, and metallic phase emerges for xϾ0.3, yet in the immediate Nagaosa, 1997). However, the weak mass renormaliza- vicinity of xϭ1/2 the ferromagnetic state changes into a tion observed in the small specific-heat coefficient ␥ is charge-ordered insulating (COI) state with decrease of not easily explained when we simply assume an orbitally temperature below TCOϭ160 K. This charge-ordered disordered state, because the entropy due to orbital de- insulating state accompanies a concomitant orbital and grees of freedom would remain large at low tempera- antiferromagnetic spin ordering of the CE type (Good- tures in the orbital disordered phase, as discussed in the enough, 1955; Wollan and Koehler, 1955; Jirak et al., case of Y1ϪxCaxTiO3 in Sec. IV.B.1. In any case, a cru- 1985; Yoshizawa et al., 1995, 1997) as illustrated in Fig. cial puzzling feature is the extremely small Drude 197. The nominal Mn3ϩ and Mn4ϩ species with 1:1 ratio weight, with the nominal effective mass of m*/m0ϳ80. show the real-space ordering in the (001) plane of the We need further studies, both experimental and theoret- orthorhombic lattice (Pbnm), while the eg orbital shows ical, to elucidate this issue. Recently, the incoherent the 1ϫ2 superlattice on the same (001) plane. Reflecting charge dynamics of La1ϪxAxMnO3 with small Drude such an orbital ordering, the spin ordering shows a com- weight and small ␥ were derived from the scaling theory plicated sublattice structure extending over a larger unit by anisotropic hyperscaling generated from the critical- cell. ity of the orbital fluctuation of two eg orbitals (Imada, Such a concomitant charge and spin ordering is also 1998). This approach, derived from strongly wave- seen in other transition-metal oxides, as described in number-dependent renormalization, appears to be im- Sec. IV.G, yet the most notable feature in the perovskite portant for an understanding of the universally observed manganites is that the charge-ordered state can be re- tendency towards incoherence in a variety of d-electron laxed to a ferromagnetic state by application of an ex- and other compounds near the Mott insulator. ternal magnetic field. Figure 198 displays the tempera- A large-energy-scale reshuffle like that observed in ture dependence of the resistivity under various the temperature and magnetic-field dependence of the magnetic fields in Nd1ϪxSrxMnO3 (xϭ0.5) (Kuwahara optical spectra near TC (Kaplan et al., 1996) can also be et al., 1995; Tokura, Kuwahara, et al., 1996a). The field- due to a change of the Jahn-Teller coupling, which dependent change in resistivity around TC(ϭ250 K) forms an intense mid-infrared polaronic band (Kaplan represents the conventional magnetoresistance effect, as et al., 1996; Millis et al., 1996; Okimoto et al., 1997) described above. A much more notable feature is that whose spectral weight and energy positions vary with the MIT due to the charge ordering transition around temperature and applied magnetic field. 160 K at zero field, which accompanies the thermal hys- Let us come back to the electronic phase diagram teresis characteristic of a first-order phase transition, is shown in Fig. 191. With a decrease of W, other instabili- shifted to the lower-temperature side, and above 7T the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1223

FIG. 198. Temperature dependence of the resistivity under various magnetic fields in Nd1ϪxSrxMnO3 (xϭ0.5). From Ku- wahara et al., 1995, 1997. crystal eventually remains metallic down to the lowest temperature without a trace of a charge-ordered phase. The metal-insulator phase diagram derived in the plane of H (magnetic field) and T (temperature) shows a large hysteretic region arising from the metastability of the ferromagnetic metal and charge-ordered insulating FIG. 199. MI phase diagrams for Pr Ca MnO (x states at low temperatures in the course of the first-order 1Ϫx x 3 ϭ0.3– 0.5) in the H-T plane (Tomioka et al., 1996, Tokunaga et al. phase transition (Kuwahara , 1995; Tokura, Kuwa- et al., 1997). hara, et al., 1996a). A quite similar behavior was ob- served for La1ϪxCaxMnO3 with xϳ0.5 (Schiffer et al., 1995). 191). These features were interpreted in terms of the Increasing x beyond 0.5 for R1ϪxSrxMnO3 (RϭLa, role of the double-exchange electrons extending along Pr, and Nd; xϭ0.5– 0.6) seems to produce a unique fer- the c-axis direction, i.e., the charged-stripe direction romagnetic state, which is an A-type layered antiferro- (Jirak et al., 1985). magnet like LaMnO3 but is conducting within the ferro- The metal-insulator phase diagram for magnetic basal plane (Kawano et al., 1997). The Pr1ϪxCaxMnO3 (xϭ0.3– 0.5) is shown in Fig. 199 in the resultant confinement of the spin-polarized carriers H-T plane (Tomioka et al., 1996; Tokunaga et al., 1998). within the ferromagnetic sheets gives rise to large an- The ferromagnetic metal/charge-ordered insulator phase isotropy in the resistivity, in spite of the nearly cubic boundaries, accompanied by a metamagnetic change in lattice structure (Kuwahara et al., 1998). When x is in- the magnetization, were determined by studies of iso- creased further above 0.6, e.g., in La1ϪxCaxMnO3, a new thermal magnetoresistance in which a change in resistiv- pattern of charge ordering different from the CE type is ity of several orders of magnitude was observed (To- observed (Chen, Cheong, and Hwang, 1997). mioka et al. 1995, 1996). A larger external field is needed The charge-ordered insulating state appears to be ex- to destroy the charge-ordered insulating state in tended over a wider doping region (x) when W is fur- Pa1ϪxCaxMnO3 (xϭ0.5) crystals than in the corre- ther narrowed, as can be seen for xϾ0.3 in the case of sponding Nd-Sr compound (see Fig. 198) with a larger Pr1ϪxCaxMnO3 [Fig. 191(c); Jirak et al., 1980, 1985]. The W. The field hysteresis (hatched region) expands at low charge and orbital ordering take place at first in the temperatures, which represents the supercooling (super- same commensurate pattern, as shown in Fig. 197 below heating) phenomena characteristic of a first-order phase TCOϭ220– 250 K when xϾ0.3. The onset of spin order- transition. With deviation of x from 0.5, however, the ing emerges at a lower temperature than charge order- ferromagnetic metal/charge-ordered insulator phase ing. The crystal at xϭ0.5 undergoes a spin-collinear an- boundary is shifted to the low-magnetic-field side, in tiferromagnetic transition of the CE type at 160 K. With particular at low temperatures, reflecting perhaps a spin- deviation of x from 0.5, the coupling of the spins along canting tendency. The reentrant behavior of the charge- the c axis becomes ferromagnetic in the antiferromag- ordered transition, e.g., in the field-cooled run at x netically ordered state, and then the spin-canted phase Ͻ0.5, may arise from large entropy in the discommen- arises in the lower temperature region. The spin-canting surate charge-ordered state. In particular, for xϭ0.3 the transition temperature (TCA) and the spin-canting angle field-induced transition from ferromagnetic metal to tend to increase with deviation of x from 0.5 (see Fig. charge-ordered insulator in the isothermal process be-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1224 Imada, Fujimori, and Tokura: Metal-insulator transitions comes irreversible below 40 K, accompanying a persis- the band gap already exists in a hypothetically cubic tent change of resistivity by more than 10 orders of mag- LaMnO3. Starting from the Hartree-Fock or LDAϩU nitude. band structure, inclusion of correlation effects through It is worth noting that the irreversible transition of the the self-energy correction reduces the band gap, result- metastable charge-ordered insulator state in ing in better agreement with experiment (Mizokawa and Pr1ϪxCaxMnO3 (xϭ0.3) crystals can be triggered not Fujimori, 1996b). The charge-ordering phenomena in only by an external magnetic field or pressure (Tokura, R1ϪxAxMnO3 have been studied theoretically using the Tomioka, et al., 1996b; Moritomo et al., 1997) but also LADϩU and Hartree-Fock methods. The correct CE- by x-ray (Kiryukhin et al., 1997) and light irradiations type magnetic structure with charge ordering, observed (Miyano et al., 1997) and by an electric field (Asamitsu, for Pr0.5Ca0.5MnO3, can indeed be reproduced in these Tomioka, et al., 1997). These field- or photo-induced calculations in the presence of Jahn-Teller distortion transitions have potential applications in new magneto- (Anisimov et al., 1997; Mizokawa and Fujimori, 1997). electronic devices. However, different types of ordered structures (CE- The electronic structure parameters for RMnO3 and type, antiferromagnetic, A-type antiferromagnetic, and AMnO3 have been derived from configuration- ferromagnetic) are nearly degenerate in energy, and interaction cluster-model analysis of photoemission therefore it is a subtle problem which of these structures spectra (Saitoh, Bocquet, et al., 1995b; Chainani, is realized under which conditions. There has also been Mathew, and Sarma, 1993) and from first-principles controversy as to whether the Jahn-Teller distortion is band-structure calculations (Satpathy, Popovic´, and necessary for the correct prediction of magnetically and Vukajlovic´, 1996). The deduced values for LaMnO3 are orbitally ordered structures. Although band-structure Uϳ7 eV and ⌬ϳ4.5 eV, which characterize RMnO3 as calculations have shown that the Jahn-Teller distortion a charge-transfer-type insulator and hence mean that is necessary, mean-field solutions (Ishihara, Inoue, and doped holes enter primarily oxygen p-like orbitals. The Maekawa, 1997; Maezono et al., 1998; Shiina, Nishitani, doped p hole is coupled with the high-spin (Sϭ2) Mn and Shiba, 1997) as well as exact diagonalization studies 4 3 d (t2g eg ) configuration antiferromagnetically to form of small clusters (with 2ϫ2ϫ2 Mn ions; Koshibae et al., ↑ ↑ a d4Lគ (Sϭ3/2) object (Saitoh, Bocquet, et al., 1995b), 1997) have also shown that correct spin-orbital struc- analogously to the Zhang-Rice singlet (d9Lគ with Sϭ0) tures can be obtained by including correlation effects in the high-Tc cuprates. The p hole has eg symmetry without the lattice distortion for a certain parameter with respect to the Mn site, and therefore the orbital and range. spin symmetry of the d4Lគ state is identical to the high- The changes in electronic structure with composition 3 3 spin Mn d (t2g ) configuration. This justifies the use of in the La1ϪxSrxMnO3 systems can be seen in Fig. 200. ↑ an effective degenerate Hubbard model for the Mn eg As holes are doped into LaMnO3, spectral weight trans- bands (Ishihara, Inoue, and Maekawa, 1997) instead of fer occurs from the occupied eg band, whose peak is ↑ the full d-p multiband Hubbard model (Mizokawa and located ϳ1.5 eV below EF , to the unoccupied eg band. ↑ Fujimori, 1995), as long as values for the effective pa- The spectral intensity at EF remains low for all x, how- rameters are properly chosen. The exchange interaction ever. Therefore it can be said that the overall eg spec- 3 ↑ between the localized t2g spin and the itinerant eg elec- tral distribution exhibits a wide (ϳ1 eV) pseudogap at eff↑ tron is estimated to be JH ϳ1 eV. The Jahn-Teller split- EF . The intensity at EF is an order of magnitude ting of the eg orbital is rather difficult to estimate em- smaller than in another ferromagnetic metal, pirically because not only the change in the p-d transfer La1ϪxSrxCoO3, indicating an unusually small quasiparti- integrals but also the change in the Madelung potential cle spectral weight Z. This nonrigid band behavior is caused by lattice distortion contributes to this splitting. qualitatively similar to that of R1ϪxAxTiO3, although The crystal-field splitting is large enough to open a band the spectral intensity at EF is much weaker in the Mn gap in the LSDA calculation. In SrMnO3, the eg band is compounds. Even more unusual is the strong tempera- ↑ empty and the highest occupied states are t2g -like. ture dependence of the intensity at EF (Sarma, Shanthi, Jahn-Teller distortion and orbital ordering, which↑ play et al., 1996), as shown in Fig. 201, a behavior that has not essential roles in the physical properties of been seen in any other metals so far. The EF intensity R1ϪxAxMnO3, are well described by Hartree-Fock decreases with temperature and almost disappears band-structure calculations on the d-p multiband Hub- somewhat below TC . Essentially the same behavior was bard model, where the parameters derived from photo- observed for La1ϪxCaxMnO3 (Park, Chen, et al., 1996). emission studies are used (Mizokawa and Fujimori, The extremely low intensity at EF (mk /mbӶ1 and 1995). The LDAϩU method is also successful because hence ZӶ1 because the effective mass m* is not so the self-interaction has been eliminated by applying great) may correspond to the extremely small Drude Hartree-Fock-type corrections to the LDA potential weight (Okimoto et al., 1997). The strong deviation of (Satpathy, Popovic´, and Vukajlovic´, 1996). The LSDA is the spectral function from that of the Hubbard model, successful in predicting the insulating nature of LaMnO3 for which hole doping creates a rather intense peak if the Jahn-Teller distortion is taken into account (So- around EF , indicates that interactions neglected in the lovyev, Hamada, and Terakura, 1996a). In the LSDA Hubbard model are probably important. A possible ori- calculations, however, the band gap is underestimated, gin is the electron-phonon interaction associated with whereas in the Hartree-Fock and LDAϩU calculations the local, dynamic Jahn-Teller distortion in doped com-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1225

FIG. 200. Photoemission and oxygen 1s core-level x-ray ab- sorption spectra of La1ϪxSrxMnO3 near EF as a function of x FIG. 202. Temperature dependence of in-plane resistivity (Saitoh, Bocquet, et al., 1995b). The shaded area represents (␳ab) and c-axis resistivity (␳c) as well as inverse of the mag- the occupied part of the eg component estimated as the dif- netization for a La2ϪxSr1ϩxMn2O7 (xϭ0.3) crystal. From ↑ ference from SrMnO3. Kimura et al. 1996.

2. La2Ϫ2xSr1ϩ2xMn2O7 pounds discussed above (Millis, 1996). Alternatively, an- other type of atomic displacement that is symmetric The title compounds correspond to the nϭ2 layered around the electron, as discussed for lightly doped d0 perovskite structures (Ruddlesden-Popper series) (Fig. systems such as La Sr TiO with xϳ1 (Sec. IV.B.1), 66) with the MnO2 bilayer in a repeated unit. Taking the 1Ϫx x 3 3ϩ may produce a pseudogap over an energy range of Mn -based La2Sr1Mn2O7 (xϭ0) as the parent Mott in- ϳ1 eV. sulator, the composition x for La2Ϫ2xSr1ϩ2xMn2O7 stands for the nominal hole number per Mn site. We show in Fig. 202 the temperature dependence of the in- plane resistivity (␳ab) and the c-axis resistivity (␳c)as well as the inverse of the magnetization for an xϭ0.3 single crystal (Kimura et al., 1996). ␳c shows a sharp c maximum at Tmaxϳ100 K and a semiconducting behav- c ior above Tmax , with metallic-like behavior below c Tmax . Compared with the magnetization data, the steep c drop of ␳c below Tmax has an intimate connection to the 3D spin ordering, where the weak antiferromagnetic in- terlayer coupling was confirmed by magnetization and neutron scattering studies (Kimura et al., 1997; Perring et al., 1998). By contrast, ␳ab shows a broad maximum at ab Tmaxϳ270 K, where the slope of inverse magnetization deviates from the Curie-Weiss law (a dotted line in Fig. 202). This implies that the in-plane 2D ferromagnetic ab correlation occurs for temperatures below Tmax , which c is far above Tmax . In between these two temperatures, the xϭ0.3 compound shows a 2D ferromagnetic metal behavior. This of course arises from the fact that the ferromagnetic interaction is mediated by hole motion, which is strongly confined within the MnO2 bilayers. The anisotropy of the resistivity ␳c /␳ab is already as large as ϳ103 at room temperature, and further in- creases with decrease of temperature, reaching as large a 4 c value as ϳ10 at Tmax . Below TC , ␳c steeply decreases with decreasing temperature. Such a change of the inter-

FIG. 201. Photoemission spectra of La0.6Sr0.4MnO3 near EF at plane charge dynamics from incoherent to coherent has different temperatures. From Sarma, Shanthi, et al., 1996. rarely been encountered even among the highly aniso-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1226 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 203. Magnetic-field effect on ␳ab and ␳c for a La2ϪxSr1ϩxMn2O7 crystal with xϭ0.3. From Kimura et al., 1996. tropic compounds, except for the high-Tc cuprates (see Sec. IV.C.) and Sr2RuO4 (see Sec. IV.H.). FIG. 204. Isothermal MR curves (with application of magnetic In Fig. 203, the effect of the magnetic field on ␳ab and ␳ is shown for an xϭ0.3 crystal (Kimura et al., 1996). field along the c axis) for a La2ϪxSr1ϩxMn2O7 crystal with x c ϭ0.3. From Kimura et al., 1996. The magnetoresistance (MR) was found to depend weakly on the relative orientation of the field. The mag- netoresistance for the in-plane component is enhanced plane magnetization but can be removed by an external ab c in the vicinity of both Tmax and Tmax , while for the magnetic field. c interplane component Tmax is apparently shifted to a In the low-temperature case at 4.2 K, ␳c drastically higher temperature with a magnetic field, and the maxi- decreases in the low-field region, in accordance with the mum resistivity value is remarkably suppressed. An- magnetization process, and becomes constant when the other important feature is seen in anisotropic low-field magnetization saturates at an external field of about c magnetoresistance effects: Below Tmax , the effect of 0.5 T. Kimura et al. (1996) interpreted this behavior in magnetoresistance in ␳c is appreciably large, even at 1T, terms of effects of interplane tunneling upon the magne- c but saturated at higher fields. By contrast, the magne- toresistance: Below Tmax , the interplane antiferromag- toresistance in ␳ab is much smaller than that in ␳c . netic correlation is established, producing many mag- The isothermal magnetoresistance curves (with appli- netic boundaries lying on the (La,Sr)2O2 layers (see Fig. cation of a magnetic field along the c axis) are shown in 66). Such boundaries block interplane tunneling of spin- Fig. 204 (Kimura et al., 1996) for the in-plane and inter- polarized electrons. By applying a magnetic field, how- plane components together with the magnetization ever, one can increase the volume of the ferromagnetic ab curves at characteristic temperatures; around Tmax (273 domains and finally make a single domain with a field c K), Tmax (100 K), and at a rather low temperature (4.2 larger than the saturation field for magnetization, in K) with almost full spin polarization. Large magnetore- which the carrier-blocking boundaries should disappear. sistance is observed for both components due to the en- It is worth noting that such a removal of the interlayer hanced in-plane and interplane ferromagnetic correla- blocking domain boundary also affects the in-plane tions at 273 K and more conspicuously at 100 K. In charge dynamics, as can be seen in the low-field magne- particular, the interplane magnetoresistance at 100 K is toresistance for the in-plane component (top panel of Ϫ4 extremely large (␳c(0)/␳c(H)ϳ10 at 5T) due to the Fig. 204). field-induced incoherent-coherent transition for c-axis The charge-transport properties, including magne- charge transport at this temperature. In other words, toresistance characteristics in these bilayer manganites, magnetic 2D confinement of highly spin-polarized elec- are observed to be quite sensitive to the band-filling or trons (or holes) is realized due to uncorrelated inter- hole-doping level, as in the aforementioned case of the

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1227

FIG. 205. Temperature dependence of ␳ab and ␳c for a La2ϪxSr1ϩxMn2O7 (xϭ0.4) crystal under various magnetic fields (applied perpendicular to the c axis). From Moritomo et al., 1996. pseudocubic perovskite manganites. Figure 205 shows the temperature dependence of ␳ab and ␳c for La2ϪxSr1ϩxMn2O7 (xϭ0.4) crystals under various mag- netic fields (applied perpendicular to the c axis; Mori- tomo et al., 1996). By contrast with the case of the x ϭ0.3 crystal, both the in-plane and interplane resistivi- ties show a semiconducting increase with decrease of temperature down to TC(ϭ126 K) and with activation FIG. 206. Temperature dependence for a La2ϪxSr1ϩxMn2O7 energy of 30–40 meV, though the anisotropy ratio of the (xϭ0.4) crystal of (a) resistivity; (b) the integrated intensity of 2 the (110) Bragg peak; (c) ferromagnetic critical scattering; and resistivity is still as large as 10 . At around TC (d) antiferromagnetic cluster scattering. From Perring et al., (ϭ126 K), both ␳ab and ␳c show a steep decrease by more than two orders of magnitude, showing a metallic 1997a. behavior below TC (but not below about 20 K due to some localization effect). By application of magnetic peak and to find others, as well as from an analysis of fields, the resistivity above TC is appreciably suppressed, putting the resistivity maximum towards higher tem- momentum and energy transfer profiles, they concluded perature. that this peak originated from the antiferromagnetic An inelastic neutron-scattering study was performed spin cluster depicted in Fig. 206(d), with a correlation (Perring et al., 1997b) on an xϭ0.4 crystal. Figure 206 length of about 9 Å and a lifetime of about 0.04 ps at 142 shows the temperature dependence of (a) the resistivity, K, just above TC . This diffuse scattering intensity, inte- (b) the integrated intensity of the (110) Bragg peak, (c) grated over the energy-transfer range of Ϫ5 meV to the ferromagnetic critical scattering, and (d) the antifer- ϩ5 meV and momentum-transfer range 0.30 to 0.63 romagnetic cluster scattering. Due to growth in the mag- (2␲/a0), as shown in Fig. 206, gradually increased with netization below TC , the (110) Bragg peak, which is also decreasing temperature down to TC and then steeply a nuclear reflection, increases in intensity as in the con- decreased below TC . The behavior was parallel to that ventional ferromagnet. In accord with this, the ferro- of the ferromagnetic critical scattering shown in Fig. magnetic critical scattering displayed in Fig. 206, that is, 206(c). Thus the paramagnetic state near TC and just the integrated intensity over an energy transfer range above it consisted of a slowly fluctuating mixture of from Ϫ5 meV to ϩ0.5 meV and over a momentum ferro- and antiferromagnetic microdomains. The insulat- transfer range from 0.09 to 0.24ϫ(2␲/a0) along [110], is ing behavior is perhaps due to the breaking of local sym- enhanced around TC . This critical scattering disappears metry by this anisotropic antiferromagnetic spin cluster again below TC , as it should. In addition to the above and to the resultant localization effect. The magnetic features expected for a conventional ferromagnet, Per- field obviously suppresses carrier scattering by such an ring et al. (1997) found diffuse magnetic scattering that antiferromagnetic spin fluctuation, giving rise to the ob- peaked at the zone boundary point (1/2,0,0). From an served negative magnetoresistance of colossal magni- extensive survey of reciprocal space to characterize this tude (Fig. 205).

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1228 Imada, Fujimori, and Tokura: Metal-insulator transitions

The above observation has significant implications for the origin of the colossal magnetoresistance (CMR) ob- served generally for doped manganites. As mentioned in the previous section, the scenario of local lattice distor- tions such as the dynamic Jahn-Teller distortion has been a favored potential source of colossal magnetore- sistance in addition to the double-exchange mechanism. However, the orbital degree of freedom in the eg state may give rise not only to Jahn-Teller electron-lattice coupling but also to strong and anisotropic antiferro- magnetic fluctuations: The latter can arise from the or- bital correlation between sites, as typically seen in the charge-exchange type charge-ordered state (see Sec. IV.F.1), and can compete with the double-exchange in- teraction. The anisotropic antiferromagnetic cluster above TC foraLa2ϪxSr1ϩxMn2O7 (xϭ0.4) crystal may also be a consequence of intersite orbital correlations. In fact, the static Jahn-Teller distortion, which is coherent along the c axis for the xϭ0 parent crystal, is minimal in such a heavily doped crystal as xϭ0.4: Typical bond FIG. 207. Crystal structure of FeSi. From Jaccarino et al., length ratios at 300 K are Mn-O(1)/Mn-O(3)Ϸ1.001 and 1967. Mn-O(2)/Mn-O(3)Ϸ1.03, O(1),O(3),O(2), representing the inner and outer apex and equatorial oxygen, respec- tively (Mitchell et al., 1997). However, the Jahn-Teller (1992) pointed out the close similarity between the mag- distortion was observed to increase in the low- netic and transport properties of FeSi and those of temperature ferromagnetic-metallic phase (Mitchel Kondo insulators with rare-earth 4f or actinide 5f elec- et al., 1997), implying the onset of orbital ordering at trons, reviving interest in FeSi. Since then there has TC . been controversy whether FeSi can be viewed as a The electronic structure of La2ϪxSr1ϩxMn2O7 was d-electron Kondo insulator or not. studied by angle-resolved photoemission spectroscopy FeSi crystallizes in a complicated cubic structure, by Dessau et al. (1998). Energy bands of d3z2Ϫr2 and shown in Fig. 207, in which the unit cell contains four dx2Ϫy2 character showed dispersions, consistent with the molecules. The magnetic susceptibility, shown in Fig. LDA band-structure calculation. Although the observed 208, has been obtained and can be fitted with a Weiss dispersions and Fermi-surface crossings were consistent temperature of Ϫ(150– 200) K and an effective moment with band-structure calculations, the spectral intensity of 2.6Ϯ0.1␮B /Fe. The crystal is semiconducting below was drastically reduced near EF and there was practi- ϳ300 K and weakly metallic above it, as shown in Fig. cally no intensity as the bands crossed EF (for the me- 209 (Wolfe et al., 1965). These data can be fitted to an tallic xϭ0.4). Strong electron-phonon coupling (includ- activation law ␳(T)ϰexp(⌬tr/2kT) at relatively high ing Jahn-Teller-type coupling) was proposed as a temperatures (between ϳ100 K and ϳ150 K) with ⌬tr possible origin of this very unusual behavior. The re- Ӎ550– 700 K, implying that the intrinsic gap is duced intensity at EF was also observed for 3D Mn ox- ϳ50– 60 meV. However, this temperature range may be ides (Sec. IV.F.1), but was more dramatic here in that too small to justify the activation fit; at lower tempera- the spectral weight at EF was totally suppressed in spite tures (below a few tens of K), the resistivity has been of the metallic conductivity. fitted to power laws (see Fig. 209) or 3D variable-range hopping within Anderson-localized states, ␳(T) G. Systems with transitions between metal ϰexp(T /T)1/4 (S. Takagi et al., 1981; Hunt et al., 1994). and nonmagnetic insulators 0 FeSi shows a gigantic thermoelectric power of positive sign below ϳ150 K (with a maximum of S 1. FeSi ϳ250– 480 ␮V/K at ϳ50 K); it becomes negative above FeSi is a nonmagnetic insulator at low temperatures ϳ150 K and then positive above ϳ300 K (Wolfe et al., (TӶ500 K). With increasing temperature, however, it 1965). Because the Hall coefficient is negative and in- gradually becomes a Curie-Weiss paramagnet (Benoit, creases with decreasing temperature (Kaidanov et al., 1955) with poor metallic conductivity at TϾ500 K 1968), the semiconducting charge carriers must be pre- (Wolfe et al., 1965). In spite of the ϳ500-K maximum in dominantly n type. The different signs of the Hall coef- the magnetic susceptibility, no evidence for antiferro- ficient and the Seebeck coefficient and the temperature magnetic ordering has been found in neutron diffraction dependence of the latter imply two types of carriers: (Watanabe et al., 1963), NMR, and Mo¨ ssbauer (Wer- n-type and p-type. theim et al., 1965) studies. Because of its unusual mag- Optical measurements by Schlesinger et al. (1993) netic and transport properties, FeSi has attracted much showed a gap of ϳ60 meV, as shown in Fig. 210. The interest for many decades. Recently, Aeppli and Fisk gap is gradually filled with increasing temperature up to

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1229

FIG. 208. Magnetic suceptibility of FeSi after subtraction of the low-temperature paramagnetic contribution by Jaccarino et al. (1967). The curve has been calculated using the local- moment model with parameters Sϭ1/2, gϭ3.92, and ⌬ ϭ750 K (for definition, see text). FIG. 210. Optical reflectivity R, optical conductivity ␴(␻), and effective electron number n(␻)/m* of FeSi (Schlesinger et al., 1993): solid curve, Tϭ20 K; dashed curve, Tϭ100 K; dotted curve, Tϭ150 K; dot-dashed curve, Tϭ200 K; solid curve, T ϭ250 K. DC conductivity values are also shown at ␻ϭ0. The inset shows the difference in n(␻)/m* between Tϭ250 and 20 K.

250 K. The ␻ϭ0 peak in the 250-K spectrum is very broad and is that of a ‘‘poor metal’’ or a dirty metal, far from being a typical Drude peak. It was pointed out that the gap starts to be filled at temperatures far below the temperature corresponding to the band gap (ϳ600 K). Simulations assuming a rigid band and the Fermi-Dirac distribution failed to predict the complete filling of the optical gap at high temperatures, indicating the unusual nature of the optical gap (Ohta et al., 1994). The n(␻)/m* curves show that spectral weight integrated up to ␻ϭ2000 cmϪ1 is not conserved between different temperatures, meaning that spectral weight transfer oc- curs up to much higher energies (of order 1 eV). This picture was questioned, however, by a more recent in- frared optical study (Degiorgi et al., 1994), according to which the spectral weight sum rule is satisfied up to ϳ3000 cmϪ1. The latter study confirmed the Anderson- localized nature of carriers at low frequencies, ␻ Ͻ10 cmϪ1. In order to explain the magnetic susceptibility of FeSi within the local-moment picture, Jaccarino et al. (1967) considered a model in which the nonmagnetic (Sϭ0) FIG. 209. Electrical resistivity of FeSi. From Wolfe et al., 1965. ground state and an Sϭ1/2 or 1 excited state were sepa-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1230 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 212. Temperature dependence in FeSi: Open circles, in- tegrated paramagnetic scattering intensity; closed circles, the paramagnetic moment M(0)2 (qϭ0) determined by fitting analysis using Eq. (4.16). The solild curve is 3kT␹(0), where ␹(0) is the uniform magnetic susceptibility. From Tajima et al., 1988.

quires the inclusion of electron correlation effects for the bands near EF . Takahashi and Moriya (1979) incor- porated correlation effects using spin-fluctuation theory into the above band structure (with some simplification) and explained the peculiar magnetic and thermal prop- erties of FeSi. They attributed the nonmagnetic ground state and the temperature-induced paramagnetism to a negative mode-mode coupling, and predicted the satura- FIG. 211. Quasielastic magnetic scattering spectra of FeSi for tion of the temperature-induced local moment at high different q’s (Tajima et al., 1988). The reciprocal-lattice points temperatures. are (1ϩ␨,1ϩ␨,0) in units of 2␲/a. Solid curves are the result The question of whether local moments are induced of fitting using Eq. (4.16). 10M corresponds to 10 min counting at high temperatures according to the prediction of spin- time at ប␻ϭ0. fluctuation theory was addressed by inelastic neutron- scattering experiments, but these studies failed to ob- rated by a spin gap ⌬. The solid curve in Fig. 208 shows serve paramagnetic moments (Kohgi and Ishikawa, a best fit with Sϭ1/2, ⌬ϭ750 K, and a g factor of 3.92. 1981). Tajima et al. (1988) made a polarized neutron- The magnetic contribution to the specific heat, which scattering study on single crystals and indeed detected a was estimated as the difference between FeSi and CoSi, temperature-induced local moment. The results show was also well fitted with the same model, although there paramagnetic (quasielastic) scattering as described by was a large ambiguity in the difference curve. Because ប/k T 1 ⌫ q 2 B ͑ ͒ the local-moment model is not compatible with the S͑q,␻͒ϭM ͑q͒ 2 2 , semiconducting properties, Jaccarino et al. also exam- 1Ϫexp͑Ϫប/kBT͒ ␲ ⌫͑q͒ ϩ␻ ined a band model in which conduction and valence (4.16) bands of widths w were separated by a band gap of 2⌬. where M(q) is the q-dependent paramagnetic moment. They had to assume, however, that 2wӶ2⌬ (with 2⌬ The ␻ϳ0 scattering intensity shows a sharp increase to- ϭ760 K) in order to reproduce the magnetic susceptibil- wards the reciprocal-lattice points, as shown in Fig. 211, ity. Such a narrow bandwidth is obviously unrealistic. indicating ferromagnetic correlation. The ␻ϳ0 intensity LDA band-structure calculations were performed by as well as the integrated intensity show a temperature Mattheiss and Hamann (1993) and more recently by Fu dependence which closely follows that of the static mag- et al. (1994) and predicted a small indirect gap of netic susceptibility, as shown in Fig. 212, a clear indica- ϳ0.1 eV, which agrees with the experimental value of tion of temperature-induced paramagnetism. The ϳ0.05 eV rather well. The band structure consists of unique feature of the scattering function of FeSi is that wide bands of strongly hybridized Fe 3d-Si 3p character the ␻-integrated intensity is q independent, i.e., M(q)is (extending from ϳϪ12 eV to ϳ10 eV with respect to qindependent, while the ␻ϳ0 intensity is q dependent, the Fermi level) and relatively narrow (ϳ1 eV) bands of as can be seen from the simultaneous sharpening of the purely Fe 3d character. The ϳ1 eV widths of the energy spectrum and the increase in the ␻ϳ0 intensity as one bands near EF is again inconsistent with the very narrow approaches a reciprocal-lattice point. M(q) is found to 2 (Ӷ0.1 eV) bands proposed by Jaccarino et al. and re- be as large as (3.5Ϯ2)␮B at 300 K.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1231

FIG. 213. Photoemission spectrum of FeSi (Saitoh, Sekiyama FIG. 214. Photoemission spectrum of FeSi near E compared et al., 1995) compared with the band DOS calculated using the F with the band DOS calculated using the LDA (Mattheiss and LDA (Mattheiss and Hamann, 1993) and that modified by the Hamann, 1993) and that modified by the self-energy correction self-energy correction. The employed local self-energy ⌺(␻)is (Saitoh, Sekiyama, et al., 1995): (a) Spectral function calculated also shown. using ⌺(␻) shown in the lower panel; (b) spectral function cal- culated using only the high-energy part ⌺h(␻); (c) the LDA The prediction of temperature-induced paramagnet- density of states; (d) density of quasiparticles. ism was also tested by Fe core-level photoemission spec- troscopy (Oh et al., 1987). Fe 3s core-level spectra, which show an exchange splitting when the Fe atom has Fig. 214. Because of the increase in the negative slope of a local moment, did not change between low and high Re ⌺l(␻) toward EF , the quasiparticle DOS N*(␻)is temperatures. Valence-band photoemission studies were enhanced near the band edge. The angle-resolved pho- made by Kakizaki et al. (1982) and more recently by Sai- toemission study by C. H. Park et al. (1995) has found a toh, Sekiyama, et al. (1995), Chainani et al. (1994) and C. sharp peak ϳ20 meV below E for a certain part of the H. Park et al. (1995) with improved energy resolution. F Brillouin zone. The peak disperses only very weakly, Since the Fe 3d and Si 3sp states are strongly hybrid- again indicating the narrow quasiparticle band at the top ized and form wide bands, the configuration-interaction of the valence band. cluster model is certainly not a good starting point for Thus the presence of the narrow bands at the band FeSi. If one compared the band-theoretical DOS with the DOS ␳(␻) measured by photoemission, however, edges, originally suggested by Jaccarino et al. (1967), significant discrepancies would be found in the overall was corroborated by the microscopic experimental line shape, as shown in Fig. 213, as well as in the high- probe. The narrow bandwidths W*ϳ20 meV naturally explain the poor metallic behavior as seen in the optical resolution spectra near EF , as shown in Fig. 214. There- fore a local self-energy model was introduced to incor- conductivity (Fig. 210), since charge transport should be porate correlation effects in a phenomenological way: incoherent at TӷW*/kB . In the sense that the physical The model self-energy has the form ⌺(␻)ϵ⌺h(␻) properties of f-electron Kondo insulators would have ϩ⌺l(␻), where ⌺h(␻) is a self-energy correction of the their origin in the presence of narrow f bands, which type g/(␻ϩi␥)2 and is effective on the large energy would be strongly renormalized, especially in the Kondo scale of 1–10 eV, while ⌺l(␻) is a self-energy correction limit (Susaki et al., 1996), FeSi and the f-electron Kondo whose absolute magnitude is small but has a significant insulators can be grouped in the same class of materials, effect on the low-energy electronic structure on the scale although the underlying (high-energy) electronic struc- of 10–100 meV (Saitoh, Sekiyama, et al., 1995). ⌺(␻)is tures are very different. assumed to have the Fermi-liquid property (Im ⌺(␻)ϰ As for the microscopic origin of temperature-induced Ϫ␻2)inthe␻ 0 limit. Figure 213 shows that the appli- , on the basis of first-principles electronic cal- → cation of ⌺h(␻) causes the narrowing of the Fe culations, Anisimov, Ezhov, et al. (1996) predicted a 3d-derived DOS, the broadening of spectral features at first-order nonmagnetic semiconductor-to-ferromagnetic higher binding energies, and the transfer of spectral metal transition as a function of magnetic field using the weight towards higher binding energies. The same self- LDAϩU method. The transition was predicted to occur energy alone cannot reproduce the sharp rise of the if the d-d Coulomb repulsion U was large enough (Ͼ3.2 band edge and the relatively flat line shape from the eV). The ferromagnetic moment was predicted to be as band edge to ϳ0.5 eV below it. The addition of ⌺l(␻), large as 1␮B per Fe. The magnetic and thermal proper- whose real part has a large negative slope near the band ties at finite temperatures were then calculated using a edge, introduces further band narrowing near EF ; simplified model based on the first-principles results, through the Kramers-Kronig relation, the ͉Im ⌺l(␻)͉ rap- yielding good agreement with experiment. This picture idly increases as one moves from EF , smearing out the appears quite different from the correlated narrow-band structures in ␳(␻). The quasiparticle density of states picture, but the proximity of the ferromagnetic state to N*(␻), calculated using Eq. (2.72), is also plotted in the ground state may be reflected in the ferromagnetic

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1232 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 216. Schematic energy diagram of the 3d bands around the Fermi level for VO2. From Shin et al., 1990.

The schematic electronic structures in metallic and in- sulating VO2, corresponding to Goodenough’s model (Goodenough, 1971), are shown in Fig. 216 (Shin et al., 1990), in which some gap energies were derived by spec- troscopic studies (Ladd and Paul, 1969; Shin et al., 1990). In the R phase, the t2g levels in the octahedral crystal

field are further split into dʈ and ␲* levels in the R FIG. 215. Comparison of V-V pairing in the three phases, R, phase, comprising the electronic states near the Fermi level of the metallic state. Here, the d orbitals are M1 , and M2 ,inVO2.InM1 (open circles) all the vanadium ʈ atoms both pair and twist from the rutile positions. In M2 rather nonbonding, while the ␲* orbitals are strongly (filled circles) one-half of the vanadium atoms pair but do not hybridized with the O 2p␲ state and hence lie higher twist and the other half form unpaired zigzag chains. (The dis- than the dʈ level. In the insulating M1 phase, the pairing tortions are exaggerated by a factor of 2 for clarity.) From of the V atoms along the cr axis (see Fig. 215) promotes Marezio et al., 1972. 3d-2p hybridization and upshifts the ␲* band off the Fermi level, as well as causing bonding-antibonding correlations found in the neutron scattering study. splitting of the dʈ band (Goodenough, 1971), as shown Misawa and Tate (1996) explained the magnetic suscep- in the right panel of Fig. 216. tibility of FeSi using a Fermi-liquid theory, according to The importance of the electron correlation for this which a singular contribution from the Fermi level (from lattice-coupled MIT has long been disputed. The issue is the band edge in the case of FeSi) causes the suscepti- whether the essential nature of the insulating state is bility maximum as well as and ferro- that of a Peierls insulator with the character of an ordi- magnetic instability in other systems (Misawa, 1995). nary band insulator or otherwise a Mott-Hubbard insu- Recently, the transport properties of substituted com- lator. Recently, Wentzcovitch et al. (1994) pointed out pounds FeSi1ϪxAlx have been studied and have been that the LDA calculation can find a monoclinic distorted successfully interpreted in the conventional doped semi- (M1 phase) ground state in agreement with experiment conductor picture (DiTusa et al., 1997). How the corre- and an almost opening of gap in charge excitations. Al- lation effects manifest themselves in FeSi still remains though it is conceptually difficult to distinguish the band an open question, calling for further investigations. and Mott insulators when there is a gap in the spin sec- tor, Wentzcovitch et al., (1994) concluded that VO2 may be more bandlike than correlated. Strong opposition to the band view (Rice et al., 1994) 2. VO 2 is based on observation of the Mott-Hubbard insulator, Many of the vanadium oxides exhibit temperature- that is, the insulator with no gap in the spin sector, in Ϫ3 induced MITs, as observed in V2O3 (Sec. IV.B.1) and V1ϪxCrxO2 with very small x (у3ϫ10 ) (Pouget et al., the Magneli phases VnO2nϪ1 (4рnр8). Among them 1974) or in pure VO2 when a small uniaxial pressure is 1 VO2, with a 3d configuration, differs from others, applied along the (110)r direction (Pouget et al., 1975). showing the nonmagnetic ground state without an anti- The TϪx phase diagram for V1ϪxCrxO2 (Marezio et al., ferromagnetic phase. The MIT in VO2 takes place at 1972; Villeneuve, Drillon, and Hagenmuller, 1973) is around 340 K, accompanying the lattice structural tran- presented in Fig. 217. Two other distinct phases emerge, sition from a high-temperature rutile (TiO2-type) struc- the insulating monoclinic M2 and the triclinic T phase, ture (R phase) to a low-temperature monoclinic struc- in addition to the aforementioned metallic R and insu- ture (M1 phase), as illustrated in Fig. 215 (Marezio et al., lating (nonmagnetic) M1 phases. Uniaxial pressure also 1972). V-V pairing occurs in the M1 phase together with gives rise to the M2 and T phases and leads to a similar tilting of the V-V pair. phase diagram (Pouget et al., 1975).

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1233

FIG. 217. The TϪx phase diagram for V1ϪxCrxO2 (Marezio FIG. 218. Photoemission spectra of VO2 in the insulating and et al., 1972; Villeneuve, Drillon, and Hagenmuller, 1973). metallic states (Shin et al., 1990). The solid curve was taken at 298 K in the insulating phase and the dashed curve at 298 K in the metallic phase. The dotted curves are spectra after decon- The V-V pairing pattern for the M2 phase is also volution with instrumental resolution. shown in Fig. 215, which is composed of two kinds of V 4ϩ chains. In M2 one half of the V ions form equally spaced V chains, and NMR and EPR experiments lation plays an important role in the MIT or whether the (Pouget et al., 1974) showed that they behave magneti- one-electron band picture is sufficient. The V 3d band cally as Sϭ1/2 Heisenberg chains (JϷ300 K). Thus it is spectrum in the metallic phase is obviously much broader than predicted by the band-structure calculation clear that these V chains in M2 are Mott-Hubbard insu- lators. The insulating triclinic phase can be viewed as a and even shows a sign of the remnant of the lower Hub- bard band, although weak, implying that electron corre- spin Peierls state of the M2 phase, although the transi- lation cannot be neglected in the metallic state (Fuji- tion between M2 and T is weakly of the first order. In mori, Hase, et al., 1992a). In going from the metallic to fact, further cooling leads to a continuous M2 M1 tran- sition through the intermediate T phase, where→ the pair- the insulating state, the quasiparticle spectral weight dis- ing (or dimerization) on one set of V chains grows. The appears, and only the ‘‘lower Hubbard band’’ survives. Although no serious comparison has been made be- Mott-Hubbard insulating M2 phase is obviously a local tween the spectra of the insulating phase and the density minimum for VO2, since the M2 can be stabilized by minimal perturbations such as very light Cr-doping or a of states of the LDA band (Wentzcovitch, Schulz, and small uniaxial pressure. Rice et al. (1994) stressed again Allen, 1994), it seems that the discrepancy between the measured spectra and the band DOS is substantial and that all the insulating phases of VO2, M1 , M2 , and T, are of the same type and should be classified as Mott- that electron correlation is again important for realizing Hubbard and not band insulators. the insulating phase of VO2. The electron correlation effect should also show up in the metallic state in VO2. The high-temperature metallic 3. Ti O state in the R phase (Tу330 K) shows T-linear behav- 2 3 ior up to 840 K (Allen et al., 1996). The mean free path Corundum-type Ti2O3 undergoes a gradual metal- l (800 K)ϭ3.3 Å is too short when the Bloch-Boltzman insulator transition around 400–600 K (for a review and interpretation is applied for a metal such as is described comprehensive guide to the literature before 1975, see by the LDA calculation. Allen et al. (1996) interpreted Honig and Van Zandt, 1975), as displayed in Fig. 219 this in terms of either a sample problem, such as crack- together with the temperature dependence of the resis- ing at the phase transition, or of an indication of anoma- tivity for the V-doped compound (Chandrashekhar lous behavior of the metal close to the Mott-Hubbard et al., 1970). In contrast to the perovskite analogs (e.g., 3ϩ 1 insulator, which differs from the conventional Fermi liq- LaTiO3) of the Ti (3d )-based oxide (Sec. IV.C.1) as uid. To pursue the electron correlation effect in the me- well as to V2O3 with a similar corundum-type structure tallic VO2, however, it is desirable to realize the barely (Sec. IV.B.1), Ti2O3 has a low-temperature insulating metallic state down to low enough temperatures, on state that is nonmagnetic. In fact, the magnetic suscepti- which more detailed spectroscopic and transport inves- bility increases to the paramagnetic value at tempera- tigations are possible. tures nearly corresponding to the MIT (Pearson, 1958). A clear gap opening below the MIT temperature has The conventional explanation for the nonmagnetic insu- been observed by photoemission spectroscopy (Sa- lating ground state (Van Zandt, Honig, and Good- watzky et al., 1979; Shin et al., 1990), as shown in Fig. enough, 1968) is based on the splitting of the lower-lying 218. The results give some hint whether electron corre- t2g level, depicted in Fig. 220. In a corundum-type struc-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1234 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 221. Optical spectra for Ti2O3. From Lucovsky et al., 1979.

In accord with the above scenario, the two types of FIG. 219. Temperature dependence of resistivity for pure and gap transitions with corresponding polarization charac- V-doped Ti O . From Chandrashekhar et al., 1970. 2 3 teristics were observed in the optical spectra shown in Fig. 221 (Lucovsky et al., 1979). The onset of the ⑀2 ture (see Fig. 68), the lower-lying t manifold for the spectrum at 0.2 eV with light polarization perpendicular 2g to the c axis (EЌc) was assigned to the aforementioned TiO6 octahedron splits into bonding and antibonding ␲ semiconductor band gap, while the Eʈc band around 0.9 states of a1g and eg levels. The a1g orbitals, which have eV was assigned to the a1g interband transition. The d3z2Ϫr2 character, form strong bonds between pairs of Ti band peak corresponding to the a1gϪa1*g transition atoms within the face-sharing octahedra along the c axis. shifts to lower energy (down to 0.6 eV) with increasing * Then the resulting bonding (a1g) and antibonding (a1g) temperature, indicating a loosened c-axis Ti-Ti bond ␲ ␲ bands are split and bracket the eg and eg * bands. The and hence supporting the Van Zandt-Honig- minimum charge gap, or the semiconductor gap, is ex- Goodenough model. ␲ pected to occur between the filled a1g and empty eg All the experimental results so far obtained appear to subbands. According to this model, known as the Van be consistent with the thermal closure of the semicon- Zandt-Honig-Goodenough model, the increase in the ductor band gap, yet little is known about the electronic ratio c/a with increase of temperature reduces the structures and properties in the high-temperature metal- bonding-antibonding splitting of the a1g bands and pro- lic state, which is likely to show strongly correlated elec- motes the collapse of the semiconductor gap between tron behavior. Concerning this point, Mattheiss (1996b) ␲ recently pointed out, on the basis of an LDA band- a1g and e . g structure calculation, the importance of the electron cor- relation effect on the MIT. The LDA calculation with the use of the LAPW (linearized argumented plane wave) method shows a partially filled t2g complex near ␲ EF which originates from overlapping a1g and eg sub- bands. Decreasing the c/a ratio, even beyond the ob- served low-temperature value, reduces but does not ␲ eliminate this a1gϪeg overlap as postulated in the model of Fig. 220. For example, to open a semiconduc- tor gap in this system, an unphysically small Ti-Ti sepa- ration of about 2.2 Å along the c axis would be required. This may preclude the simplest band explanation for the MIT. The electron correlation effect may show up explicitly when the MIT is driven by filling control and a low- temperature metallic state emerges. A Ti deficiency or excess oxygen in the chemical form of Ti2O3ϩy may cor- respond to hole doping, as in the case of V2Ϫ␦O3 (Sec. IV.B.1). In fact, a nonstoichiometric sample up to y ϭ0.03 could be prepared that would show a greatly re- duced resistivity value but still remain semiconducting FIG. 220. Van Zandt–Honig–Goodenough model (1968) of or insulating at low temperatures, with fairly large Hall electronic structure of Ti2O3. From Mattheis, 1994. coefficient (Honig and Reed, 1968). Another method of

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1235

FIG. 223. Possible spin states in LaCoO3: (a) low-spin (LS, S ϭ0); (b) intermediate-spin (IS, Sϭ1); (c) high-spin (HS, S ϭ2); (d) diagram of the eg orbital-ordered state in the IS state. FIG. 222. Temperature dependence of resistivity (␳, upper From Korotin et al., 1996. panel) and magnetic susceptibility (␹, lower panel) in LaCoO3 6 and lightly hole-doped LaCoO3 (Tokura et al., 1997). A solid insulator with a predominant configuration of 3d elec- 3ϩ line in the lower panel represents the calculated curve based trons of Co fully occupying the t2g level (Fig. 223), on the low-spin (Sϭ0) and intermediate-spin (Sϭ1) transi- perhaps due to crystal-field splitting (10Dq), which tion model (see text). slightly exceeds the Hund’s-rule coupling energy. With increase of temperature, however, the compound under- filling control is V doping, as exemplified in Fig. 219. goes a spin-state transition from a nonmagnetic (Sϭ0) The metal-site substitution may inherently contain com- to a paramagnetic state, as signaled by a steep increase plexity in assigning the effective valence of the doped in the magnetic susceptibility around 100 K, shown in and host metal cations. Recalling the parallel behavior the lower panel of Fig. 222. Knight-shift measurements of Ti doping with oxygen offstoichiometry in the case of (Itoh et al., 1995) and polarized neutron scattering (Asai et al., 1989) have confirmed such a spin-state transition, V2O3, one may consider the V-doped Ti2O3 as an electron-doped system. As can be seen in Fig. 219, x arising from a change in the electronic state of the Co տ0.02 doping induces metallic conduction even at low site. The nature of the high-temperature spin state, that temperatures (Chandrashekhar et al., 1970; Dumas and is, the problem of whether it arises from the Schlenker, 1976, 1979). However, the Ti V O is a intermediate-spin (Sϭ1) or the high-spin (Sϭ2) state 2Ϫ2x 2x 3 3ϩ at low temperatures for above xϭ0.005 (Du- of the Co ion (see Fig. 223), has not yet been settled, mas and Schlenker, 1976, 1979; Dumas et al., 1979). A although the intermediate-spin-state model appears to large T-linear term in the specific heat (50–80 mJ/mol be prevailing at the moment. Korotin et al. (1996) dem- K) was observed for xϭ0.02– 0.10 samples (Sjostrand onstrated, using the LDAϩU method, the relative sta- and Keesom, 1973). This was first interpreted as an in- bility of the Sϭ1 state rather than the Sϭ2 state, as a dication of a narrow conduction bandwidth or correlated result of strong p-d hybridization and orbital ordering in contrast to the behavior anticipated by the simple ionic electron behavior, as in Ti-doped V2O3. However, the large T-linear term was later reinterpreted in terms of model. Their proposed orbital ordering pattern is shown the high entropy of the spin-glass state (Dumas and in Fig. 223. A recent finding of appreciable local lattice Schlenker, 1976, 1979; Dumas et al., 1979). Nevertheless, distortion during the spin-state transition (Yamaguchi correlated electron dynamics is evident in the enhanced et al., 1997) suggests the presence of Jahn-Teller distor- temperature-independent (Pauli-like) component of the tions in the high-temperature spin state, supporting the susceptibility, and some supplemental experimental intermediate-spin-state model. An example of the fitting work would be desirable for a more complete under- procedures for the magnetic susceptibility with the local- standing of the Mott transition. ized spin model (Yamaguchi et al., 1997) is shown in Fig. 222. The Ising model molecular field calculation was done with assumption of a two-state model, i.e., Sϭ0 4. LaCoO 3 and Sϭ1 states with an energy gap of ⌬ and the antifer- The thermally induced insulator-to-metal crossover romagnetic exchange interaction J between thermally phenomenon in LaCoO3, as shown in the upper panel of excited neighboring Sϭ1 spins. An appropriate choice Fig. 222, is quite unique because of its nonmagnetic (⌬ϭ230 K and Jϭ9.7 K) gives a reasonable account of ground state, distinguishing it from a conventional Mott the temperature dependence, although the high-spin insulator. The ground state of LaCoO3 is a nonmagnetic model shows similar reproducibility, for example, with a

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1236 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 224. Temperature dependence of La1ϪxSrxCoO3 with various doping levels. Closed triangles represent the ferromag- netic transition temperature.

FIG. 225. Variations of optical conductivity spectra of LaCoO3 set of parameters, ⌬ϭ299 K and Jϭ26.7 K. Whichever with temperature (upper panel) and with doping x (by Sr sub- model would be better, the spin-gap energy is 0.02–0.03 stitution; lower panel) at 290 K. From Tokura et al., 1997. eV, and above room temperature an intermediate or high spin density exceeds 80% of the Co sites. As can be seen in the lower panel of Fig. 222, a small spectrum shows a gap feature at low temperatures and amount of hole doping gives rise not only to a decrease undergoes minimal change even when temperature is in- in resistivity but also to a large Curie tail in the low-spin creased up to 293 K and the spin-state transition is al- region at low enough temperatures. By measurements of most completed. The gap energy estimated from the on- the magnetization and fitting with the modified Brillouin set of conductivity is more than 0.1 eV (Yamaguchi function, it was demonstrated (Yamaguchi et al., 1996a) et al., 1996b), in accord with the transport data, although that a doped single hole shows an extremely high spin the accurate determination of the gap energy is difficult number, Sϭ10– 16. This means that a hole with O because of the blurred onset of conductivity, perhaps 2p-state character causes a local spin-state transition due to the indirect-gap nature (Sarma et al., 1995; Ha- (low-spin to intermediate-spin) over 10–16 Co sites mada, Sawada, and Terakura, 1995). With change of around itself, perhaps due to strong 2p␴-3deg hybrid- temperature around 500 K, the spectral weight is ob- ization. Such a gigantic spin polaron can be a precursor served to be rapidly transferred from the higher-lying state for the ferromagnetic metallic state in interband transition region to the lower-energy region, La1ϪxSrxCoO3 (xտ0.2; see Fig. 224, in which the ferro- accompanying the well-defined isosbetic (equal- magnetic transition is indicated by closed triangles). absorption) point at 1.4 eV. The resistivity of LaCoO3 shows no anomaly around A surprisingly similar change in the conductivity spec- this spin-state transition near 100 K, with a nearly con- trum is observed for the case of hole doping, i.e., for stant activation energy of about 0.2 eV. However, a La1ϪxSrxCoO3, when the doping level x is changed as large reduction in the resistivity takes place around 500 shown in the lower panel of Fig. 225. Nominal hole dop- K (upper panel of Fig. 222), where the spin-state transi- ing by partial substitution of La with Sr drives the FC- tion is almost complete. Furthermore, this continuous MIT as shown in Fig. 224. The optical conductivity spec- insulator-to-metal change is least affected by a small tra at 290 K (the lower panel of Fig. 224) show a similar amount of hole-doping as shown in Fig. 222 and Fig. 224 spectral weight transfer with increase of x, accompany- (Yamaguchi and Tokura, 1998). ing the isosbetic point around 1.4 eV as well. Most of the In the upper panel of Fig. 225, we show the tempera- low-energy spectral weight is borne not by the conven- ture dependence of the optical conductivity spectra of tional Drude component but by the incoherent part, LaCoO3 in the range of 0–4 eV (Tokura, Okimoto, both for the high-temperature LaCoO3 and for et al., 1998). With increasing temperature up to 800 K La1ϪxSrxCoO3. This nearly parallel behavior for tem- through the metal-insulator crossover temperature perature increasing and hole doping indicates that the around 500 K, the spectra show conspicuous change electronic structural change in both the insulator-metal over a large energy region up to several eV. Such a huge transitions is also nearly identical in nature. spectral change cannot be accounted for by a simple Figure 226 displays the inverse-temperature depen- band closing like that in a narrow-gap band insulator dence of the Hall coefficient (RH ; lower panel) together (semiconductor), in spite of the nonmagnetic nature of with that of the resistivity coefficient in LaCoO3 and the the insulating ground state. The optical conductivity lightly hole-doped compounds La1ϪxSrxCo3 (xϭ0.005

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1237

FIG. 227. Temperature (T) dependence of resistivity (␳) and Ϫ1 d log ␳/d(T ) for crystals of RCoO3 (RϭLa, Pr, Nd, Sm, Eu, FIG. 226. Inverse-temperature dependence of resistivity (␳, and Gd). From Yamaguchi et al., 1996b). upper panel) and Hall coefficient (RH , lower panel) in LaCoO3 and lightly hole-doped La1ϪxSrxCoO3. From Tokura et al., 1997. dependence of d log ␳/d(TϪ1) in the above series of RCoO3 crystals. This quantity, if constant, would repre- and 0.01; Tokura, Okimoto, et al., 1998). The RH is posi- sent a thermal activation energy. The gradual MIT tive, indicating hole-type conduction, over the whole shows up as a broad peak for the respective compounds, temperature region and shows as steep a change as the which systematically shifts to higher temperature with resistivity with temperature. In particular, the carrier decrease of the ionic radius of R or equivalently with density nϭ1/RH in LaCoO3 shows a rapid increase at decrease of W. The nearly constant value of 400–600 K. This should be interpreted as a rapid de- dlog ␳/d(TϪ1) at low temperatures represents the ther- crease in the gap energy itself, with increase of tempera- mal activation energy, or half of the transport energy ture above 400 K, that is, after completion of the spin- gap ⌬a , which also increases with decrease of W. state transition. The variation of the charge gap is directly seen in the The light hole-doping effect manifests itself in the optical conductivity spectra for the ground state (low- saturation of the increase of the resistivity and Hall co- spin state, e.g., at 9 K) for RCoO3 shown in Fig. 228 efficient below 400 K. The flat temperature dependence (Yamaguchi, Okimoto, and Tokura, 1996b). The con- of the RH in doped crystals implies the saturation behav- ior of impurity conduction. By contrast, the RH value above 400 K shows approximately common values and temperature dependence irrespective of doping. The fully conducting state, e.g., at 800 K, shows a small RH value corresponding to three holes per Co site. These facts clearly indicate the presence of a large Fermi sur- face in the high-temperature conducting phase, which is least affected by light doping as observed. Thus the ob- served insulator-metal crossover shows the essential fea- tures characteristic of a Mott transition at constant band filling. The thermally induced MIT is also critically affected by changes of the one-electron bandwidth (W). The standard method of bandwidth control (BC) is to utilize the orthorhombic distortion of the perovskite, which de- pends on the A-site ionic radius, as described in detail in Sec. III.B. In the present case, RCoO3 with R changing from La to Pr, Nd, Sm, Eu, and Gd causes an increase in FIG. 228. Optical conductivity spectra ␴(x) for RCoO3. Inset the Co-O-Co bond distortion in the GdFeO3-type ortho- gives the overall features of NdCoO3. Spiky peaks below 0.1 rhombic lattice (apart from the rhombohedral LaCoO3) eV are due to optical phonons. Dashed lines are the best-fit and hence a decrease of W in this order. Figure 227 results with the model function for the indirect-gap plus direct- (Yamaguchi, Okimoto, and Tokura, 1996b) shows the T gap transitions. From Yamaguchi et al., 1996b.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1238 Imada, Fujimori, and Tokura: Metal-insulator transitions

large compared with the change in tpd . Such a critical change in the charge gap with tpd , as well as an anoma- lously low TIM as compared with the charge-gap value, suggests again the important role of electron correlation in charge-gap formation below TIM as well as in the MIT itself. The electronic structure of LaCoO3 has also been studied by photoemission spectroscopy and electronic structure calculations. These studies have invariably shown strong hybridization between Co 3d andO2p. An LDA band-structure calculation gave a semimetallic band structure, in which the bottom of the conduction band and the top of the valence band overlap (Hamada, Sawada, and Terakura, 1995). This is due to a common tendency of the local density approximation to underes- timate band gaps; a finite gap was opened in a Hartree- Fock calculation of the multiband Hubbard model using parameters derived from photoemission spectroscopy FIG. 229. Characteristic energies vs the tolerance factor for a (Mizokawa and Fujimori, 1996b). The density of states series of RCoO3. Solid circles, gap energies for the optical given by the LDA calculation and that from the photo- (direct-gap) transition (Eg,d); solid squares, gap energies for emission spectra are in good agreement with each other the charge transport (2⌬a) in the low-temperature insulating (Sarma et al., 1995), implying that LaCoO3 is an ordi- state. Open circles indicate temperatures (kBTIM) for the nary semiconductor. Nevertheless, the presence of a insulator-metal transition as well as their magnification weak but clear charge-transfer satellite indicates that (10kBTIM) for a better comparison. From Yamaguchi et al., electron correlation is also important (Saitoh et al., 1996b. 1997a). Indeed, the on-site d-d Coulomb energy is found to be as large as Uϭ5 – 7 eV from cluster-model ductivity shows a gradual onset at 1–1.5 eV perhaps due analysis of the Co 2p core-level photoemission spectra to the indirect nature of the band gap (Sarma et al., (Chainani, Mathew, and Sarma, 1992; Saitoh et al., 1995), while a residual broad tail is present at 0.1–1 eV. 1997a). With R substitution from La to Gd, the spectral shape Concerning the phase transition from nonmagnetic and the onset shift as a whole to higher energy. Broken semiconductor to paramagnetic semiconductor at curves in the figure represent the fitting using the model ϳ90 K, changes in the photoemission spectra in going functions for direct and indirect band-gap transitions. from below to above ϳ90 K are insignificant, ruling out The lower-lying indirect gap is difficult to pinpoint due the possibility that the transition is a low-spin–to–high- to the blurred feature, while the shift of the direct gap spin transition. While within the ligand-field theory the with R ions is quite clear. intermediate-spin state should be higher than the low- Figure 229 (Yamaguchi, Okimoto, and Tokura, spin and high-spin states for any parameter set, the 1996b) shows the observed gaps and insulator-metal intermediate-spin state is found to be close to and in crossover temperature versus the tolerance factor (see some cases lower than the low-spin state according to Sec. III.B.1), which is related to the distortion of the the cluster-model and band-structure calculations. Koro- perovskite lattice and hence the degree of p-d hybrid- tin et al. (1996) studied homogeneous solutions of the ization or W. Solid circles represent the optical gap Eg,d for the direct transition (see Fig. 228), while solid band model for LaCoO3, that is, solutions with all the squares and open circles the transport gap (2⌬ ) and Co atoms in the identical electronic state, using the a LDAϩU method. They found that, for an expanded lat- the metal-insulator crossover temperature kBTIM (as tice at high temperatures, the intermediate-spin state well as 10kBTIM for reference), respectively, as derived (2.11␮ ) could be stabilized while the high-spin state from Fig. 227. The energy scale of kBTIM is by far B (3.16␮ ) always remained high in energy. Although the smaller than 2⌬a for all the RCoO3 crystals, and the B variation of kBTIM with R or the tolerance factor is not intermediate-spin state is an antiferromagnetic metal in so significant as that of 2⌬a or Eg,d . The Eg,d and 2⌬a the homogeneous solution, Korotin et al. (1996) also show nearly parallel behavior as a function of tolerance studied inhomogeneous solutions and found indications factor. The change in the p-d transfer interaction tpd that orbital ordering with antiferromagnetism almost should approximately scale with cos ␪, ␪ being the opens a gap and further lowers the energy. Hartree- Co-O-Co bond angle, as described in Sec. III.B.1. This Fock calculations (Mizokawa and Fujimori, 1996b) means that tpd changes almost linearly with the toler- showed that the intermediate-spin state is ferromagnetic ance factor in this region and by about 10% between with antiferro-orbital ordering but a band gap is not RϭLa and Gd in RCoO3. The observed changes in both opened. Since there is no long-range order in the high- the transport and the optical gap magnitudes is quite temperature state, ferromagnetic correlation rather than

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1239

FIG. 231. Crystal structure of La1.17ϪxAxVS3.17. From Nish- ikawa et al., 1996.

with the d2 configuration. The system becomes nonmag- netic below ϳ280 K, indicating a spin-singlet formation of some form at low temperatures (Yasui et al., 1995). If the d2 configuration is localized, it should be in the high- spin (Sϭ1) state. Therefore the Sϭ1 spins should somehow be coupled to form a singlet ground state. A most likely scenario is the formation of Sϭ0V3trimers FIG. 230. Temperature dependence of the oxygen 1s x-ray as in LiVO2 (Onoda, Naka, and Nagasawa, 1991); in- absorption spectra of LaCoO3 (Abbate et al., 1993). The low- deed, a structural deformation is involved in the 280-K energy peak which appears above ϳ500 K may be assigned transition, as revealed by thermal dilatation and ultra- either to the t2g band or to the coherent spectral weight of the sound velocity measurements (Nishikawa et al., 1996). If metallic state. the d2 configuration is not localized but forms band states, then the spin-gap behavior may be simply be- ferromagnetic ordering is expected, in accordance with cause the distorted LaA0.17VS3.17 is a band insulator. In the results of a neutron-scattering study by Asai et al. the band-insulator–to–metal transition, the disappear- (1997). ance of the spin gap and the appearance of metallic be- Concerning the 500–600-K insulator-to-(poor) metal havior should occur simultaneously. The latter possibil- transition, Abbate et al. (1993) observed a significant ity cannot be excluded from the resistivity and magnetic change in the oxygen 1s x-ray absorption spectra across susceptibility data alone because the electrical resistivity the transition, as shown in Fig. 230, which is consistent is relatively low above ϳ280 K and nearly temperature with the large change expected for a transition from the independent, as shown in Fig. 232 (Nishikawa, Yasui, low-spin ground state to the high-spin excited state but may also be consistent with an insulator-to-metal transi- tion. According to both the LADϩU and Hartree-Fock calculations, however, the high-spin state is an insulator. Korotin et al. (1996) suggested that, above ϳ600 K, or- bitals are disordered within the intermediate-spin state, driving the system metallic. La1ϪxSrxCoO3 is a unique filling-control system in that doped holes not only act as metallic carriers but also induce a ferromagnetic moment. The doping- induced changes in the photoemission spectra are rela- tively small and indicate that the system is in the intermediate-spin state (Saitoh et al., 1997b), unlike the high-spin La1ϪxSrxMnO3.

5. La1.17ϪxAxVS3.17

La1.17ϪxAxVS3.17 , where A is the divalent Sr or Pb ion, is a ‘‘misfit-layer’’ compound consisting of alternat- ing stacks of CdI2-type VS2 layers and rock-salt-type La1ϪxAxS layers, as shown in Fig. 231: (La1ϪyAyS)1.17VS2.InLaA0.17VS3.17 (xϭ0.17), the V FIG. 232. Electrical resistivity of La1.17ϪxSrxVS3.17 (Nishikawa atoms are trivalent and the compound is an insulator et al., 1996). It becomes the largest for xϭ0.17.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1240 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 234. Temperature dependence of the Hall coefficient for La1.17ϪxSrxVS3.17. From Yasui et al., 1995.

posite to the 3D Ti perovskite oxides (Tokura et al., 1993).

FIG. 233. Seebeck coefficient of La1.17ϪxSrxVS3.17 (Yasui et al., Band-structure calculations have not been performed 1995). It is negative for xϽ0.17 while it becomes positive for on this system because of the incommensurate crystal xу0.17, indicating electron and hole doping, respectively. structure. The electronic structure of the VS2 layer can be envisaged from the band structure of the layered ma- and Sato, 1994; Nishikawa et al., 1996). terial VS2 (Myron, 1980), according to which the La1.17ϪxAxVS3.17 is a unique filling-control system in partially-filled V 3d band is located above the filled S 3p that it can be doped with either holes (xϾ0.17) or elec- band and these bands are strongly hybridized with each trons (xϽ0.17). Its crystal structure implies that the other. Photoemission studies confirmed this electronic electronic properties are highly two dimensional as in structure [Fig. 235(a); Ino et al., 1997b], meaning that the cuprates. As shown in Fig. 232, the highest electrical thecompound can be viewed as a Mott-Hubbard-type resistivity indeed occurs at xϳ0.17 and the material be- system (although it must be located near the boundary comes more conductive with either decreasing or in- between the Mott-Hubbard and charge-transfer re- creasing x with respect to xϳ0.17. Thermoelectric power gimes). The doping dependence of the photoemission is positive for xу0.17 and negative for xϽ0.17, and its spectra reveals an interesting behavior of the spectral magnitude decreases as x deviates from 0.l7, indicating weight redistribution and chemical potential shift as that holes and electrons, respectively, are doped into the functions of band filling. As shown in Fig. 235(a), the V Mott insulator LaSr0.17VS3.17 . Samples with the highest 3d band (within ϳ1.5 eV of EF) is shifted toward EF doping (xϳ0 for n-type doping and xϳ0.35 for p-type with electron doping and away from EF with hole dop- doping) still show a resistivity upturn at low tempera- ing, but the shift is slower than that of the S 3p band tures, but the temperature dependence of the thermo- (located 1.5–1.7 eV below EF). If the shift of the S 3p electric power (Fig. 233) indicates metallic behavior (S band reflects a chemical potential shift, the slower shift ϰT) below 50–100 K (Yasui et al., 1995). Thermoelectric of the V 3d band would be interpreted as due to the power, which is not influenced by intergrain transport, repulsion between quasiparticles of V 3d character, as should reflect more intrinsic properties of the materials. described in Sec. II.D.1. The shift of the chemical poten- n, as reflected in the shift of the S 3p band, isץ/␮ץ It becomes semiconducting (dS/dTϽ0) above ϳ100 K tial and again metallic above ϳ250 K with smaller slope. slow for low hole/electron concentration and fast for The Hall coefficient (Fig. 234) is positive at high tem- high hole/electron concentration, as shown in Fig. -n near the Mott insulaץ/␮ץ peratures for all doping concentrations x, but in 235(b). The suppression of electron-doped samples it becomes negative at low tem- tor may imply that the effective mass is enhanced, which peratures (Ӷ100 K). Its strong temperature depen- excludes the simplest band-insulator picture for the un- dence, especially in lightly doped samples, is similar to doped compound. The suppression of the chemical po- that found in the high-Tc cuprates. tential shift is apparently inconsistent with the decreas- The electronic specific heat decreases as x 0.17, that ing electronic specific heat toward xϭ0.17, if one is, with decreasing carrier concentration→ (Nishikawa assumes the Fermi-liquid relations (2.74) and (2.75) with et al., 1996), which means that the effective mass of the nonsingular behavior of the Landau parameters (Fu- n 0ץ/␮ץ conduction electrons becomes lighter or the carrier rukawa and Imada, 1993). The observations number decreases as the system approaches the un- and ␥ 0 for x 0.17 can be reconciled with each other→ → → doped insulator. This is analogous to the high-Tc cu- if the metallic state is not a Fermi liquid or if there is a prates (Loram et al., 1989; Momono et al., 1994) but op- pseudogap in the vicinity of the chemical potential. An-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1241

FIG. 236. Temperature dependence of anisotropic resistivity of Sr2RuO4 (Lichtenberg et al., 1992; Maeno et al., 1997).

across the Mott-insulating gap. This gap is much larger than the temperature at which the spin gap is formed in the undoped compound, again disfavoring the band- insulator picture for undoped LaA0.17VS3.17 .

H. 4d systems

1. Sr2RuO4 The ruthenium oxide compounds are among those showing the Ruddlesden-Popper series (layered and cu- bic perovskite) phase illustrated in Fig. 64. Sr2RuO4 has the K2NiF4-type structure and is isostructural with the La2ϪxSrxCuO4, apart from the absence of orthorhombic distortion. Maeno et al. (1994) discovered the TcϷ1 K superconductivity in this com- pound, which has stimulated studies on its anisotropic charge dynamics and electronic structures as well as on related layered ruthenates both for their possible rel- evance to the mechanism of high-Tc superconductivity and for the occurrence of exotic symmetry of supercon- ductivity (Rice and Sigrist, 1996; Mazin and Singh, 1997; Agterberg, Rice, and Sigrist, 1997). The temperature dependence of the in-plane and c-axis resistivity (␳ab and ␳c , respectively) is shown in Fig. 236 (Maeno et al., 1997). The in-plane conduction is typically metallic with a low residual resistivity just above Tc , reaching Ϸ1 ␮⍀ cm. By contrast, the c-axis transport is nonmetallic above TMϷ130 K, but shows metallic behavior at lower temperatures (Lichtenberg et al., 1992), indicating a crossover from a 2D to a 3D metal. The resistivity anisotropy ␳c /␳ab reaches as large as 200 around TM . 2 FIG. 235. Photoemission results of La1.17ϪxPbxVS3.17 (Ino Below about 25 K, the resistivity follows the T de- et al., 1997b). (a) Photoemission spectra; (b) shift of the chemi- pendence both for ␳ab and ␳c , indicating the conven- cal potential (deduced from the shift of the S 3p band) as a tional Fermi-liquid behavior (Maeno, 1996). In fact, de function of x. The dashed line is the shift expected from the Haas–van Alphen oscillations were observed by Mack- band structure density of states. enzie et al. (1996) and revealed some important Fermi surface parameters. The results could be well inter- other important feature of the chemical potential in the preted in terms of an almost 2D Fermi-liquid model, La1.17ϪxAxVS3.17 system is a discontinuous jump of which is consistent with the LDA band structure (Ogu- ϳ0.08 eV between n-type and p-type doped samples, chi, 1995; Singh, 1995). The observed (and LDA- giving the first evidence for a chemical potential jump calculated) Fermi surface consists of two large electron

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1242 Imada, Fujimori, and Tokura: Metal-insulator transitions

FIG. 238. Photoemission spectrum of Sr2RuO4 (␳exp(␻); Inoue et al., 1996) compared with the DOS from a band-structure calculation [D(␻); Oguchi, 1995]. The solid curve represents the DOS modified by the self-energy correction ⌺(k,␻) ϭg Ϫi Ϫi ϩa gϭ FIG. 237. Energy-dependent effective mass and scattering rate ␻͓⌫/(␻ ⌫)͔͓⌬/(␻ ⌬)͔ ␧k , where 10.1, ⌫ ϭ ϭ aϭ for the c-axis conduction, derived by extended Drude analysis 0.102 eV, ⌬ 0.246 eV, and 1.5. of c-axis polarized optical conductivity spectra. From Katsu- fuji, Kasai, and Tokura, 1996. URu2Si2 (Bonn, Garrett, and Timusk, 1988). By con- trast, mab* (␻)/mab is nearly independent of ប␻, though a cylinders (␤ and ␥) centered on the ⌫Z line and one small mass enhancement is observed in a T-dependent narrow hole cylinder (␣) running around the corner of manner below 0.3 eV. These results indicate that the 3D the Brillouin zone (see dashed lines of Fig. 238 below). metallic conduction below TMϷ130 K is brought about The metallic c-axis transport at low temperatures is by the onset of coherent motion of carriers for which the dominated by ␤ and ␥ sheets, for which the c-axis mean mass and scattering rate along the c axis are strongly free path lc (Ϸ30 Å at 1 K) can be substantially longer renormalized. Such anisotropic mass renormalization than the separation between adjacent RuO2 sheets, c/2 may be considered as a generic feature of quasi-2D met- ϭ6.4 Å (Mackenzie et al., 1996). als with strong electron correlation, but it is obviously The carrier electrons in Sr2RuO4 appear to be consid- distinct from the case of the cuprate superconductors. erably renormalized due to the electron correlation. The The LDA band structure of Sr2RuO4 was calculated electronic specific-heat coefficient (␥) is 37.5 mJ/K2 mol by Oguchi (1995) and Singh (1995). Their calculations (Maeno et al., 1997), which gives rise to a density of predict that energy bands which cross the EF are derived states at the Fermi level enhanced by a factor of 3.6 over from the strong hybridization of the Ru 4dxy ,4dyz , and the LDA result. The Wilson ratio, RWϭ␹/␥, is 1.7–1.9, 4dzx orbitals with the O 2p orbitals. The photoemission also typical of correlated metals. spectra are generally consistent with this picture, but the A strong mass renormalization for the c-axis charge observed Ru 4d band feature is distributed over a wider dynamics was observed in the strongly temperature- energy range than the calculation, as shown in Fig. 238 dependent c-axis spectra of the optical conductivity (Inoue et al., 1996). This may be interpreted as due to (Katsufuji, Kasai, and Tokura, 1996). In the c-axis spec- electron correlation effects within the Ru 4d band, as in tra, the Drude band evolves below TM with an apparent the case of the V and Ti oxides described above, that is, plasma edge around 0.01 eV at TӶTM . Figure 237 (Kat- as due to a transfer of spectral weight from the coherent sufuji, Kasai, and Tokura, 1996) shows the result of an quasiparticle band to the incoherent part of the spectral analysis of the c-axis spectra in terms of the extended function. Also, the spectral weight at EF is suppressed Drude model, in which the energy dependence of the compared to the band DOS as in the V and Ti oxides. mass and relaxation time of the conduction carriers was Thus Sr2RuO4 may be viewed as one of the strongly taken into account. Below TMϷ130 K, mc* for ប␻ correlated electron systems like the 3d transition-metal Ͻ0.02 eV is enhanced and reaches about 30 times the oxides, and any interpretation of its various physical unrenormalized c-axis mass mc for ប␻Ͻ0.01 eV at 15 K. properties should take into account electron correlation. On the other hand, ប/␶ is suppressed for ប␻Ͻ0.012 eV Dispersions of the Ru 4d-derived bands were investi- but increases for ប␻Ͼ0.012 eV with a decrease in tem- gated by angle-resolved photoemission studies (Yokoya perature. Such a conspicuous temperature dependence et al., 1996a, 1996b; Lu et al., 1996). In spite of the wide is similar to that of a heavy-fermion compound, e.g., energy distribution of the coherent plus incoherent spec-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1243

FIG. 239. Fermi surfaces of Sr2RuO4 measured by angle- resolved photoemission spectroscopy (Yokoya et al., 1996a). The filled circles denote electronlike Fermi surfaces and the open and gray circles denote holelike Fermi surfaces. Fermi FIG. 240. Effective moment peff and Weiss temperature ⌰ of surfaces derived from band-structure calculations (Oguchi, Ca1ϪxSrxRuO3 (Fukunaga and Tsuda, 1994). The open and 1995) are given by the dashed lines. filled symbols correspond to values deduced at high and low temperatures. tral weight, the dispersions near EF (reflecting only the coherent part) were generally weak compared with the ϭ2.83␮ (Callagan et al., 1966; Longo et al., 1968). The band-structure calculations. In particular, they revealed B Rhodes-Wolfarth ratio is thus ϳ1.3, a value close to that a flat band with a saddle point, i.e., a so-called ‘‘ex- of the itinerant ferromagnet Ni metal, implying that tended van Hove singularity’’ around the (␲,0) point of SrRuO is an ‘‘intermediately localized’’ ferromagnet the Brillouin zone, as in the high-T cuprates. The posi- 3 c (Fukunaga and Tsuda, 1994). The T decreases under tion of the flat band was ϳ20 meV below E . The cal- C F hydrostatic pressure at the rate of dT /dP culations predicted an electron-like Fermi surface C Ӎ Ϫ(6–8) K GPaϪ1 (Neumeier et al., 1994; Shikano et al., around the ⌫ point and two hole-like Fermi surfaces • 1994), as in other itinerant ferromagnets. The negative around the X [(␲,␲)] point. Although the mass is more dT /dP is contrasted with the positive dT /dP in enhanced in the de Haas–van Alphen measurement C N local-moment systems such as LaTiO (Okada et al., than the LDA result, the volumes of the calculated 3 1992) and the positive dT /dP in double-exchange sys- Fermi surfaces are in agreement with those derived from C et al. the de Haas-van Alphen measurements to within 1% tems such as La1ϪxSrxMnO3 (Moritomo , 1995). (Mackenzie et al., 1996). On the other hand, the Fermi CaRuO3 is also metallic. The crystal structure of surfaces obtained by the angle-resolved photoemission CaRuO3 is more distorted than SrRuO3: the Ru-O-Ru studies are in disagreement with the band-structure cal- bond angle is reduced from ϳ165° to ϳ150°, in going culations and the de Haas–van Alphen measurements, from SrRuO3 to CaRuO3 (Kobayashi et al., 1994). The as shown in Fig. 239. It was found that the agreement is Weiss temperature is negative, indicating antiferromag- netic interaction between magnetic moments, but it re- improved if the EF is shifted upward by ϳ70 meV com- pared to the calculated band structure (Lu et al., 1996), mains paramagnetic down to the lowest temperature although it is difficult to justify such a procedure. Since (Gibb et al., 1974). In the solid solution Ca1ϪxSrxRuO3, this is the only case where ARPES experiments and the Curie temperature TC decreases with Ca content LDA calculations have given different Fermi surfaces and is completely suppressed at xӍ0.4 (Kanbayashi, for a 2D metal, a more critical experimental check 1978) [while a recent study on single crystals showed should be made in order to exclude, e.g., surface effects. that TC remains finite up to xՇ1 (Cao et al., 1997)]. The decrease of TC with Ca substitution may appear consis- tent with the decrease of TC under pressure, in the sense 2. Ca1ϪxSrxRuO3 that the lattice constant shrinks in both cases. It should SrRuO3 has a slightly distorted perovskite be remembered, however, that the reduction in the Ru- (GdFeO3-type) structure and is metallic. It shows ferro- O-Ru bond angle with Ca substitution would normally magnetic order below TCӍ160 K (Callagan et al., 1966). reduce the d-band width and would therefore have an The saturation magnetization is 1.1– 1.3 ␮B /Ru, and a effect opposite to that of increased pressure. More stud- neutron-diffraction study gave the ordered moment of ies are needed to explain both the pressure and Ca sub- 1.4Ϯ0.4 ␮B /Ru, which is the largest ordered magnetic stitution effects in a consistent way. moment among 4d transition-metal compounds (Longo With Ca substitution, the effective moment peff in- et al., 1968). While those values are smaller than that creases and the Weiss temperature ␪ decreases, as 4ϩ (2␮B) expected for the Ru ion in the low-spin state shown in Fig. 240 (Fukunaga and Tsuda, 1994): ␪ 4 Ϫ1 (t2g , Sϭ1), the effective moment above TC , ϳ2.6␮B , changes its sign at xӍ0.4. The ␹ -T plot in a wide is rather close to the low-spin value of 2ͱS(Sϩ1) temperature range (TϽ100 K) changes its slope above

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1244 Imada, Fujimori, and Tokura: Metal-insulator transitions

tering mechanism dominates in the high-temperature range (Fukunaga and Tsuda, 1994). With Ca substitu- tion, the resistivity below TC increases as TC decreases towards xϳ0.4, and with further substitution it de- creases again: In particular, CaRuO3 shows an unusual resistivity drop below ϳ50 K, which may indicate the disappearance of scattering by magnetic fluctuations, possibly due to the opening of an excitation gap in the magnetic fluctuation spectrum. The critical behaviors around TC give insight into the coupling between mag- netic and transport properties. Anomalous critical trans- port behavior has been found, but is attributed to the ‘‘bad metallic’’ behavior and not to the coupling be- tween the magnetic and transport properties, because the critical magnetic behavior is conventional (Klein et al., 1996). The band structure of ferromagnetic SrRuO3 was cal- FIG. 241. Resistivity of Ca1ϪxSrxRuO3 normalized at 300 K culated within the LSDA (Singh, 1996; Allen et al., (Fukunaga and Tsuda, 1994). ␳ is the residual resistivity and 0 1996). The results show very strong Ru 4d-O 2p hybrid- for SrRuO ranges from ␳ ϳ2ϫ10Ϫ5 ⍀ cm (Bouchard and 3 0 ization as in Sr RuO . Both electronlike and hole-like Gillson, 1972; Klein et al., 1996) to 2.3ϫ10Ϫ6 ⍀ cm (Izumi 2 4 et al., 1997). Here ␳(300 K)ϭ2 – 3ϫ10Ϫ4 ⍀ cm and ␣ is a scal- Fermi surfaces are indeed present. Flat bands cross EF , ing parameter. giving rise to a high DOS at EF and hence a ferromag- netic ground state through the Stoner mechanism. The obtained ordered moment of ϳ1.5 ␮B /Ru is in good and below Tϳ600 K for all x’s: In the low-temperature agreement with the experimental value of 1.4 region, ␪ is a little larger, indicating that ferromagnetic Ϯ0.4 ␮B /Ru. The overall line shape of the photoemis- sion spectra (Cox et al., 1983; Fujioka et al., 1997) agrees coupling increases at low temperatures, and peff is smaller, indicating that the magnetic electrons are more itinerant at low temperatures. These observations mean that metallic transport favors ferromagnetic coupling. The increased localized character at high temperatures may be a general property of narrow-band metals, in which there is no coherent transport above T ϳW*/3kB , where W* is the renormalized bandwidth. The resistivity value of ϳ2ϫ10Ϫ4 ⍀ cm at high tem- peratures (ϳ300 K) indicates kFlϳ1, but no saturation is found up to 1000 K (Allen et al., 1996). This indicates the breakdown of conventional metallic transport; Klein et al. (1996) attributed this to the ‘‘bad metallic behav- ior’’ discussed by Emery and Kivelson (1995). The opti- cal spectra of CaxSr1ϪxRuO3 show great similarity to those of the high-Tc cuprates (Bozovic et al., 1994). Nearly ␴(␻)ϳ1/␻ behavior is found up to ␻ϳ1 eV; the electronic Raman scattering shows an energy- independent continuum up to ϳ1 eV. These optical be- haviors are characteristic of incoherent metals. Coupling between the magnetic and transport proper- ties is an interesting issue in metallic ferromagnets. SrRuO3 shows negative magnetoresistance over the whole temperature range (Gasusepohl et al., 1995; Izumi et al., 1997). The Hall coefficient changes its sign from negative to positive at ϳ50 K with increasing tempera- ture (Gausepohl et al., 1996), which implies that there are both electronlike and hole-like Fermi surfaces and that a large portion of the Fermi surfaces are very flat. Figure 241 shows that the resistivity drops at tempera- tures below TC due to the disappearance of scattering FIG. 242. Photoemission and O 1s x-ray absorption spectra of from spin disorder. Above TC , the temperature depen- SrRuO3 compared with the band DOS in the ferromagnetic dence of the normalized resistivity is nearly identical for state (Fujioka et al., 1997). For the XAS spectra, the oxygen p all x’s, probably because the same (spin-disorder) scat- partial DOS is presented.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1245 well with the band DOS, as shown in Fig. 242. The Ru 4d band, however, is spread over a wider energy range than the calculation, and the spectral weight at EF is suppressed compared to the band DOS, as in the case of Sr2RuO4. This may be interpreted as due to transfer of spectral weight from the coherent part to the incoherent part of the spectral function arising from electron corre- lation. From the reduced spectral intensity at EF , one obtains mk /mbϳ0.3 (Sec. II.D.1). Combining this value with the mass enhancement factor of m*/mbϭ3.7 de- duced from the electronic specific heat ␥ and the band- structure calculation (Allen et al., 1996), one obtains the quasiparticle weight Zϭ(m /m )/(m*/m )m ϳ0.1. FIG. 243. The chemical potential ␮ and the doping concentra- k b b b tion ␦ to illustrate the width of the critical region. The dotted Such a small weight should be related to the bad/ line is the ␮ vs ␦ for a noninteracting system and the solid incoherent metallic behavior discussed above. curve for an interacting system with the Mott insulating phase. Evolution of the spectral function in going from me- tallic to insulating Ru oxides was studied for various Ru width. Therefore the entropy due to the component re- oxides, including the Pauli-paramagnetic metal mains unreleased above such low energy scales. When Bi2Ru2O7, paramagnetic metal CaRuO3, ferromagnetic the system is slightly metallized, the coherent part is still metal SrRuO3, and antiferromagnetic insulator small and the entropy related to the dominant incoher- Y2Ru2O7 (Cox et al., 1983). These oxides show qualita- ent part is also not released above the temperature of tively similar behavior to that of the V and Ti oxides 1 the intersite exchange interactions. This provides the with d configuration (Sec. I.A.5), in that there is a background for the anomalous behavior of metals at low transfer of spectral weight from higher binding energies temperatures near the Mott insulator. (the ‘‘incoherent part’’) to lower binding energies (the With this underlying residual entropy, in the lightly ‘‘coherent part’’) with increasing U/W. From the shal- doped metallic region, more happens. At integer fillings, lower position of the incoherent peak, the Coulomb en- the states near the Fermi level for the noninteracting ergy U is estimated to be somewhat smaller (ϳ3 eV) system are kicked out of the Mott gap region by strong than that of the Ti and V oxides (ϳ4 eV). Coulomb repulsion. The region reconstructed out of the Mott gap forms a flat band as in Fig. 243, causing degen- V. CONCLUDING REMARKS eracy in momentum space, and constitutes another ori- gin of degeneracy. The mobile carriers under this flat Almost localized electrons near the metal-insulator dispersion further strongly couple to the component de- transitions are well described neither by a simple itiner- grees of freedom and disturb the process of entropy re- ant picture in momentum space nor by a localized pic- lease for the spin and orbital degrees of freedom, when ture in real space. Theoretical and experimental efforts the carrier motion destroys the ordering of spins and to make bridges between the itinerant and localized or, orbitals. As a counteraction, the spin and orbital corre- in other words, coherent and incoherent pictures have lations also disturb the coherent carrier motion to make been made for years and have been fruitful, as we have a flatter band with strong damping in a self-consistent seen in this article. However, this is still a challenging fashion. Then the coupling of charge dynamics to the subject and many open questions have yet to be an- component destabilizes the charge coherence as well as swered. Marginally retained metals near the metal- the component ordering and compresses the process of insulator transition have offered interesting and rich residual entropy release into a much lower energy scale. phenomena for decades and will continue to do so in the This combined effect is the main origin of the various future, presenting fundamental problems in condensed anomalous features of correlated metals near the metal- matter physics as well as potential sources of applica- insulator transition that have been the main subject of tions. this review article. As we have seen in various examples, We first summarize the basic understanding of corre- the residual entropy and related fluctuations cause rich lation effects of almost localized electrons near the phenomena which are quite different from those of stan- metal-insulator transitions. In the Mott insulating state, dard metals—for example, mass enhancement, unusual a charge gap is formed and electrons are localized below temperature dependence of the resistivity and spin sus- the gap energy scale. Because of the local nature of the ceptibility, obstinate incoherent responses in transport, electron wave function in the insulator, the internal de- optical, photoemission, and magnetic studies, and en- grees of freedom (component degrees of freedom) such hanced spin and orbital fluctuations. as spin and orbital fluctuations give rise to degeneracy, These anomalous features are most distinctly de- which, along with the resultant residual entropy, is re- scribed by the quantum critical phenomena of the Mott leased at low temperatures by intersite interactions such transition. In particular, coupling to the component de- as superexchange and orbital exchange. In the strong- grees of freedom is able to cause a change even in the correlation regime, these exchange interactions can be universality class of the metal-insulator transition, as in much smaller than the charge gap and the bare band- zϭ4 in the 2D systems. In terms of the critical expo-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1246 Imada, Fujimori, and Tokura: Metal-insulator transitions nents of the universality class, the degree of coherent erly taken into account. In this context, the Fermi-liquid carrier motion is characterized by the dynamic exponent approaches must include strong wave-number and fre- z, which represents the dispersion of the charge carrier quency dependences in the self-energy correction be- near the metal-insulator transition. Coupling to a com- yond the present level. It would be desirable for this ponent with large quantum fluctuations drives the ordi- Fermi-liquid framework to include a program for pre- nary exponent zϭ2 to a higher value, which qualita- dicting the breakdown of the Fermi liquid itself as the tively changes the character of a metal to a more Mott insulator is approached in a continuous fashion. incoherent liquid. Whether a metal can be described in The same requirement applies to first-principles calcula- terms of quantum critical phenomena depends on the tions and spin-fluctuation theories. When spatial fluctua- width of the critical region. If the region is narrow, it will tions are ignored, temporal fluctuations are correctly ex- not influence experimental results. The critical region of pressed by the dynamic mean-field theory, in contrast to a metal-insulator transition for filling control is roughly other theoretical approaches such as the Gutzwiller, estimated from a simple argument. As discussed above, Hubbard, Hartree-Fock, and slave-particle approxima- when the Mott gap ⌬c opens at an integer filling, the tions. When both spatial and temporal fluctuations be- states of the noninteracting counterpart contained come important, a prescription in low dimensions is es- within the gap are reconstructed and pushed out from tablished by the scaling theory. Further understanding the gap region to the outer gap edge Eϭ⌬c roughly along the lines of the scaling description, especially the within the energy range of the gap, namely ⌬cрE singular wave-number dependence of the renormaliza- р2⌬c . This energy range is transformed to the wave- tion that it implies, has to be pursued. The flat band number scale kϳmin(2␲,2␲⌬c /t), which determines the structure around (␲,0) in 2D relevant in the cuprates area of flat dispersion due to many-body effects and offers a particular example. Numerical methods have characterizes the critical region (see Fig. 243). This is proven to be powerful when combined with an appro- interpreted as the critical region of the doping concen- priate low-energy description of the scaling behavior. It tration extended up to ␦ϳmin(1,⌬c /t). In the strong- would be desirable to further improve the method to correlation regime, ⌬c could have the same order of reduce the limitation of finite system size and also to magnitude as the bare bandwidth, which makes the criti- allow the calculation of dynamical quantities in large cal region of the doping concentration wide. This is why systems. The criticality of the Mott phenomena has been Mott phenomena in the strong-correlation regime have recognized as a subject to be studied in greater detail in a wide critical region and why treatments in terms of real materials. Systematic studies have only just begun quantum critical phenomena are appropriate in experi- recently, and further experiments that fine-tune and mentally observed metallic regions, in contrast to the control the parameters will be needed. case of simple magnetic transitions. As mentioned above, the appearance of the strong There are several different views of each anomalous correlation effects in real materials is accompanied by feature of metals near the Mott insulator, in which dif- various other degrees of freedom. These secondary ef- ferent and individual aspects can be emphasized. They fects inevitably occur to some extent because of slow include Anderson localization due to disorder, localiza- and incoherent charge dynamics generated by the Mott tion transitions due to coupling to the lattice (polaron phenomena. Studies on these repercussion effects are all effects), instabilities to charge ordering and other local at a primitive stage of understanding at the moment and orders, as well as instabilities to superconductivity. We should be elucidated in the future. The first issue is first note that these additional aspects are both repercus- Anderson localization near the Mott insulator. It re- sion effects of the Mott transition and at the same time mains to be clarified how the disorder effects appear just inevitable products of it. When the Mott transition is before a mass-diverging Mott transition. This is clearly approached from the metallic side, anomalously sup- beyond the presently available scenarios of the Ander- pressed coherence with the vanishing renormalization son localization. The second issue is the effect of po- factor generates slower and slower charge dynamics, larons, which may be particularly important when the which leads to the high sensitivity of charged carriers to lattice degrees of freedom are strongly coupled through various external fluctuations, and the system becomes the component, as in the case of dynamic Jahn-Teller subject to instabilities to localization or some type of distortion coupled to orbital fluctuations. Another sec- symmetry breaking. This sensitive region is illustrated in ondary effect is the instability of incoherent metal near Fig. 1 in the shaded area. the Mott insulator to charge, spin, or orbital order and To understand the global features and the principal competition between these types of order. Charge order- driving force of the Mott transition, we first have to un- ing or stripe structure may happen more easily near the derstand its intrinsic nature. In this article, we have Mott insulator due to the mass enhancement. In particu- sketched various theoretical approaches for that pur- lar, it has been proposed that the inherent instability of pose. In conjunction with experimental results, various incoherent metal to the superconducting state is driven mean-field approaches have succeeded in explaining by the change in universality class connected with the some aspects of the anomalies. However, for unusual mechanism of high-Tc superconductivity. This idea de- metals with increasingly incoherent responses toward serves elucidation in more detail. the metal-insulator transition, it has been shown that For a better understanding of the d-electron systems, temporal as well as spatial fluctuations have to be prop- we need to study orbital degrees of freedom more sys-

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1247 tematically. Mean-field phase diagrams have been stud- ACKNOWLEDGMENTS ied for the coupled spin and orbital systems and com- pared to the experimental results, but fluctuation effects The authors thank G. Aeppli, P. W. Anderson, S. An- are not well studied at the moment. zai, A. Asamitsu, F. F. Assaad, A. E. Bocquet, E. Dag- Here, we briefly summarize the effort to determine otto, D. S. Dessau, H. Ding, N. Furukawa, J. M. Honig, the overall electronic model parameters of correlated I. H. Inoue, N. Katoh, T. Katsufuji, T. Kimura, K. Kishio, K. Kitazawa, G. Kotliar, Y. Kuramoto, H. Ku- metals. The global electronic structure of 3d transition- wahara, K. Matho, M. Matoba, T. Mizokawa, T. Moriya, metal compounds on the eV scale is now well under- Y. Motome, Y. Okimoto, J.-H. Park, T. M. Rice, D. D. stood from spectroscopic and first-principles studies. Sarma, T. Sasagawa, G. A. Sawatzky, H. Tabata, H. The electronic structure parameters such as the on-site Takagi, Y. Tomioka, H. Tsunetsugu, J. Wilkins, and H. Coulomb energy U, the charge-transfer energy ⌬, and Yasuoka for their help at various stages in the writing of the transfer integrals T have been deduced and their this article, including fruitful discussions and preparation systematic changes with chemical composition ranging of figures. The authors thank J. Wilkins for his encour- from the Mott-Hubbard regime to the charge-transfer agement in writing this article. The authors are grateful regime have been clarified. The LDAϩU method pro- for financial support by a Grant-in-Aid for Scientific Re- vides a practical way to incorporate electron-electron in- search on the Priority Area ‘‘Anomalous Metallic States teractions into first-principles band-structure calcula- near the Mott Transition’’ from the Ministry of Educa- tions on the Hartree-Fock level. However, in many cases tion, Science and Culture, Japan, and are also indebted the U value has been determined not from first prin- to members of this project for valuable discussions. The ciples but empirically. That is, a true first-principles ap- authors also acknowledge support from Japan Society proach to strongly correlated systems is lacking so far, for Promotion of Science under the project JSPS- even on the mean-field level. To understand the spectro- RFTP97P01103. The authors appreciate K. Fujii and N. scopic properties, effects of interactions that are not in- Sasaki for their devotion and help during the prepara- cluded in the standard models such as the Hubbard and tion of this article. d-p models, namely, the long-range Coulomb interac- tion, electron-phonon interaction, impurity potentials, etc. need to be studied further in the future. REFERENCES Experimentally, in this decade, many d-electron com- pounds have been studied with the idea of systematically Preprint numbers refer to the archives [email protected] controlling the parameters to tune the distance from the Abbate, M., F. M. F. de Groot, J. C. Fuggle, A. Fujimori, O. metal-insulator transition. Filling control, bandwidth Strebel, F. Lopez, M. Domke, G. Kaindl, G. A. Sawatzky, M. Takano, Y. Takeda, H. Eisaki, and S. Uchida, 1992, Phys. control, and dimensionality control have been exten- Rev. B 46, 4511. sively studied. The compounds studied indeed show di- Abbate, M., F. M. F. de Groot, J. C. Fuggle, A. Fujimori, Y. verse properties of transport, optical and magnetic re- Tokura, Y. Fujishima, O. Strebel, M. Domke, G. Kaindl, J. sponses while theoretical analyses have provided useful van Elp, B. T. Thole, G. A. Sawatzky, M. Sacchi, and N. schemes for classifying the anomalous properties sur- Tsuda, 1991, Phys. Rev. B 44, 5419. veyed in this article. In the last decade, interesting and Abbate, M., J. C. Fuggle, A. Fujimori, L. H. Tjeng, C. T. Chen, surprising phenomena have been found in the region R. Potze, G. A. Sawatzky, H. Eisaki, and S. Uchida, 1993, where the metallic state is retained close to the Mott Phys. Rev. B 47, 16124. insulator when instability to electron localization is sup- Abrahams, E., P. W. Anderson, D. C. Licciardello, and T. V. pressed. These include high-Tc superconductivity and Ramakrishnan, 1979, Phys. Rev. Lett. 42, 673. colossal magnetoresistance. In future experimental stud- Abrahams, E., and P. A. Lee, 1986, Phys. Rev. B 33, 683. ies, it is certainly hoped that metallic states as close as Aebi, P., J. Osterwalder, P. Schwaller, L. Schlapbach, M. Shi- possible to the Mott insulator can be further realized in moda, T. Mochiku, and K. Kadowaki, 1994, Phys. Rev. Lett. a variety of materials and that various instabilities can 72, 2757. be controlled to obtain strongly nonlinear responses of Aeppli, G., et al., 1988, Phys. Rev. Lett. 61, 203. charge, spin, orbital, and lattice properties, with the goal Aeppli, G., and Z. Fisk, 1992, Comments Condens. Matter of developing useful devices for applications. Nonlinear Phys. 16, 155. responses also offer a rich array of dynamic properties in Aeppli, G., S. M. Hayden, H. A. Mook, Z. Fisk, S-W. Cheong, D. Rytz, J. P. Remeika, G. P. Espinosa, and A. S. Cooper, both equilibrium and nonequilibrium states, with meta- 1989, Phys. Rev. Lett. 62, 2052. stabilities and slow relaxations as in the Mn perovskite Affleck, I. A., and J. B. Marston, 1988, Phys. Rev. B 37, 3774. compounds. These show great promise for useful optical Agterberg, D. F., T. M. Rice, and M. Sigrist, 1997, Phys. Rev. or magnetic applications. It is certainly clear that our Lett. 78, 3374. knowledge on the controllability of carrier dynamics Aiura, Y., H. Bando, I. Hase, Y. Nishihara, Y. Haruyama, and close to the Mott insulator is still at a primitive stage H. Suzuki, 1996, J. Electron Spectrosc. Relat. Phenom. 78, compared to that in semiconductor devices. Correlated 199. electron devices will have broad potential applications Akimitsu et al., 1998, unpublished. when properly developed, because of the cooperative Alexandrov, A. S., A. M. Bratkovsky, and N. F. Mott, 1994, and collective nature of electrons. Phys. Rev. Lett. 72, 1734.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1248 Imada, Fujimori, and Tokura: Metal-insulator transitions

Alexandrov, A. S., and N. F. Mott, 1994, Rep. Prog. Phys. 57, Arima, T., Y. Tokura, and S. Uchida, 1993, Phys. Rev. B 48, 1197. 6597. Allen, P. B., H. Berger, O. Chauvet, L. Forro, T. Jarlborg, A. Arovas, D. P., and A. Auerbach, 1988, Phys. Rev. B 38, 316. Junod, B. Revaz, and G. Santi, 1996, Phys. Rev. B 53, 4393. Aryasetiawan, F., 1992, Phys. Rev. B 46, 13051. Allen, J. W., S-J. Oh, L. E. Cox, W. P. Ellis, M. S. Wire, Z. Aryasetiawan, F. A., and O. G. Gunnarsson, 1997, Fisk, J. L. Smith, B. B. Pate, I. Lindau, and A. J. Arko, 1985, cond-mat/9712013. Phys. Rev. Lett. 54, 2635. Asai, K., P. Gehring, H. Chou, and G. Shirane, 1989, Phys. Allen, J. W., C. G. Olson, M. B. Maple, J.-S. Kang, L. Z. Liu, Rev. B 40, 10982. J.-H. Park, R. O. Anderson, W. P. Ellis, J. T. Markert, Y. Asai, K., O. Yokokura, N. Nishimori, H. Chou, J. M. Tran- Dalichaouch, and R. Liu, 1990, Phys. Rev. Lett. 64, 595. quada, G. Shirane, S. Higuchi, Y. Okajima, and K. Kohn, Allen, P. B., W. E. Pickett, and H. Krakauer, 1987, Phys. Rev. 1994, Phys. Rev. B 50, 3025. B 36, 3926. Asai, K., A. Yoneda, O. Yokokura, J. M. Tranquada, G. Alvarado, S. F., M. Erbudak, and P. Munz, 1976, Phys. Rev. B Shirane, and K. Kohn, 1997, unpublished. 14, 2740. Asamitsu, A., Y. Moritomo, R. Kumai, Y. Tomioka, and Y. An, G., and J. M. J. van Leeuwen, 1991, Phys. Rev. B 44, 9410. Tokura, 1996, Phys. Rev. B 54, 1716. Anderson, P. W., 1952, Phys. Rev. 86, 694. Asamitsu, A., Y. Moritomo, Y. Tomioka, T. Arima, and Y. Anderson, P. W., 1956, Phys. Rev. 102, 1008. Tokura, 1995, Nature (London) 373, 407. Anderson, P. W., 1959, Phys. Rev. 115,2. Asamitsu, A., and Y. Tokura, 1998, Phys. Rev. B 58, 47. Anderson, P. W., 1963a, in Magnetism, edited by G. T. Rado Asamitsu, A., Y. Tomioka, H. Kuwahara, and Y. Tokura, and H. Suhl (Academic, New York), Vol. I, p. 25. 1997, Nature (London) 388, 50. Anderson, P. W., 1963b, in Solid State Physics, edited by H. Ashcroft, N. W., and N. D. Mermin, 1976, Solid State Physics Ehrenreich, F. Seitz, and D. Turnbull (Academic, New (Saunders, Philadelphia). York), Vol. 14, p. 99. Ashkenazi, J., and M. Weger, 1973, Adv. Phys. 22, 207. Anderson, P. W., 1984, Basic Notions of Condensed Matter Assaad, F. F., and M. Imada, 1995, Phys. Rev. Lett. 74, 3868. Physics (Benjamin/Cummings, London). Assaad, F. F., and M. Imada, 1996a, J. Phys. Soc. Jpn. 65, 189. Anderson, P. W., 1987, Science 235, 1196. Assaad, F. F., and M. Imada, 1996b, Phys. Rev. Lett. 76, 3176. Anderson, P. W., 1990, Phys. Rev. Lett. 64, 1839. Assaad, F. F., and M. Imada, 1998, Phys. Rev. B 58, 1845. Anderson, P. W., 1991a, Phys. Rev. Lett. 67, 2092. Assaad, F. F., M. Imada, and D. Scalapino, 1996, Phys. Rev. Anderson, P. W., 1991b, Phys. Rev. Lett. 67, 3844. Lett. 77, 4592. Anderson, P. W., 1992, Science 256, 1526. Assaad, F. F., M. Imada, and D. Scalapino, 1997, Phys. Rev. B Anderson, P. W., 1997, The Theory of Superconductivity in the 56, 15001. High-Tc Cuprate Superconductors (Princeton University, Azrak, A. El, et al., 1994, Phys. Rev. B 49, 9846. Princeton, New Jersey). Azuma, M., Z. Hiroi, M. Takano, K. Ishida, and Y. Kitaoka, Anderson, P. W., and H. Hasegawa, 1955, Phys. Rev. 100, 675. 1994, Phys. Rev. Lett. 73, 3463. Anderson, R. O., R. Claessen, J. W. Allen, C. G. Olson, C. Bagno, P. G., O. Jepsen, and O. Gunnarsson, 1989, Phys. Rev. Janowitz, L. Z. Liu, J.-H. Park, M. B. Maple, Y. Dalichaouch, B 40, 1997. M. C. de Andrade, R. F. Jardim, E. A. Early, S.-J. Oh, and W. Bao, W., C. Broholm, S. A. Carter, T. F. Rosenbaum, G. Aep- P. Ellis, 1993, Phys. Rev. Lett. 70, 3163. pli, S. F. Trevino, P. Metcalf, J. M. Honig, and J. Spalek, 1993, Ando, Y., G. S. Boebinger, A. Passner, T. Kimura, and K. Phys. Rev. Lett. 71, 766. Kishio, 1995, Phys. Rev. Lett. 75, 4662. Bao, W., C. Broholm, J. M. Honig, P. Metcalf, and S. F. Tre- Anisimov, V. I., I. S. Elfimov, N. Hamada, and K. Terakura, vino, 1996, Phys. Rev. B 54, R3726. 1996, Phys. Rev. 54, 4387. Bao, W., C. H. Chen, S-W. Cheong, and S. A. Carter, 1996, Anisimov, V. I., I. S. Elfimov, M. A. Korotin, and K. Terakura, Solid State Commun. 98, 55. 1997, Phys. Rev. B 55, 15494. Bao, W., et al., 1997, Phys. Rev. Lett. 78, 507. Anisimov, V. I., S. Yu. Ezhov, I. S. Elifimov, I. V. Solovyev, Barker, A. S., Jr., B. I. Halperin, and T. M. Rice, 1968, Phys. and T. M. Rice, 1996, Phys. Rev. Lett. 76, 1735. Rev. Lett. 20, 384. Anisimov, V. I., M. A. Korotin, J. Zaanen, and O. K. Ander- Barker, A. S., Jr. and J. P. Remeika, 1974, Phys. Rev. B 10, sen, 1992, Phys. Rev. Lett. 68, 345. 987. Anisimov, V. I., A. I. Poteryaev, M. A. Korotin, A. O. Barman, S. R., A. Chainani, and D. D. Sarma, 1994, Phys. Rev. Anokhin, and G. Kotliar, 1997, J. Phys.: Condens. Matter 9, B 49, 8475. 7359. Barman, S. R., N. Shanti, and D. D. Sarma, 1996, Europhys. Anisimov, V. I., J. Zaanen, and O. Andersen, 1991, Phys. Rev. Lett. 36, 307. B 44, 943. Barnes, S. E., 1976, J. Phys. F 6, 1375. Annett, J. F., N. Goldenfeld, and A. J. Leggett, 1996, J. Low Barnes, S. E., 1977, J. Phys. F 7, 2637. Temp. Phys. 105, 473. Barthelemy, E., C. Chavant, G. Collin, and O. Gorochov, 1976, Anzai, S., T. Futami, and T. Sawa, 1981, J. Appl. Phys. 52, J. Phys. (Paris) 37, C4-17. 6270. Barthelemy, E., O. Gorochov, and H. McKinzie, 1973, Mater. Anzai, S., M. Matoba, M. Hatori, and H. Sakamoto, 1986, J. Res. Bull. 8, 1401. Phys. Soc. Jpn. 55, 2531. Baskaran, G., Z. Zou, and P. W. Anderson, 1987, Solid State Anzai, S., and K. Ozawa, 1968, J. Phys. Soc. Jpn. 24, 271. Commun. 63, 973. Arima, T., and Y. Tokura, 1995, J. Phys. Soc. Jpn. 64, 2488. Basov, D. N., et al., 1996, Phys. Rev. Lett. 77, 4090. Arima, T., Y. Tokura, and J. B. Torrance, 1993, Phys. Rev. B Battle, P. D., T. C. Gibb, and P. Lightfoot, 1990, J. Solid State 48, 17006. Chem. 84, 271.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1249

Battle, P. D., T. C. Gibb, and S. Nixon, 1988, J. Solid State Caffarel, M., and W. Krauth, 1994, Phys. Rev. Lett. 72, 1545. Chem. 77, 124. Callagan, A., C. W. Moeller, and R. Ward, 1966, Inorg. Chem. Beard, B. B., and U.-J. Wiese, 1996, Preprint No. 5 1573. cond-mat/9602164. Campuzano, J. C., G. Jennings, M. Faiz, L. Beaulaigue, B. W. Becke, A. D., 1988, Phys. Rev. A 38, 3098. Veal, J. Z. Liu, A. P. Paulikas, K. Vandervoort, H. Claus, R. Bednorz, J. G., and K. A. Mu¨ ller, 1986, Z. Phys. B 64, 189. S. List, A. J. Arko, and R. J. Bartlett, 1990, Phys. Rev. Lett. Belitz, D., and T. R. Kirkpatrick, 1994, Rev. Mod. Phys. 66, 64, 2308. 261. Canfield, P. C., J. D. Thompson, S-W. Cheong, and L. W. Benoit, R., 1955, J. Chem. Phys. 52, 119. Rupp, 1993, Phys. Rev. B 47, 12357. Bethe, H., 1928, Ann. Phys. (Leipzig) 87, 55. Cao, G., S. McCall, M. Shepard, J. E. Crow, and R. P. Guertin, Bickers, N. E., and D. J. Scalapino, 1989, Ann. Phys. (N.Y.) 1997, Phys. Rev. B 56, R5740. 193, 206. Capellmann, H., 1974, J. Phys. F 4, 1466. Bickers, N. E., and S. R. White, 1991, Phys. Rev. B 43, 8044. Capellmann, H., 1979, Z. Phys. B 34, 29. Birgeneau, R. J., A. Aharony, N. R. Belk, F. C. Chou, Y. En- Carter, S. A., B. Batlogg, R. J. Cava, J. J. Krajewski, W. F. doh, M. Greven, S. Hosoya, M. A. Kastner, C. H. Lee, Y. S. Peck Jr., and T. M. Rice, 1996, Phys. Rev. Lett. 77, 1378. Lee, G. Shirane, S. Wakimoto, B. O. Wells, and K. Yamada, Carter, S. A., T. F. Rosenbaum, J. M. Honig, and J. Spalek, 1995, J. Phys. Chem. Solids 56, 1913. 1991, Phys. Rev. Lett. 67, 3440. Birgeneau, R. J., D. R. Gabbe, H. P. Jenssen, M. A. Kastner, Carter, S. A., T. F. Rosenbaum, M. Lu, H. M. Jaeger, P. Met- P. J. Picone, T. R. Thurston, G. Shirane, Y. Endoh, M. Sato, calf, J. M. Honig, and J. Spalek, 1994, Phys. Rev. B 49, 7898. K. Yamada, Y. Hidaka, M. Oda, Y. Enomoto, M. Suzuki, and Carter, S. A., T. F. Rosenbaum, P. Metcalf, J. M. Honig, and J. T. Murakami, 1988, Phys. Rev. B 38, 6614. Spalek, 1993, Phys. Rev. B 48, 16841. Bishop, D. J., E. G. Spencer, and R. C. Dynes, 1985, Solid- Castellani, C., C. Di Castro, and W. Metzner, 1992, Phys. Rev. State Electron. 28, 73. Lett. 69, 1703. Blankenbecler, R., D. J. Scalapino, and R. L. Sugar, 1981, Castellani, C., C. Di Castro, and W. Metzner, 1998, Adv. Phys. Phys. Rev. D 24, 2278. 47, 317. Bloch, F., 1929, Z. Phys. 57, 545. Castellani, C., C. R. Natoli, and J. Ranninnger, 1978a, Phys. Bocquet, A. E., A. Fujimori, T. Mizokawa, T. Saitoh, H. Rev. B 18, 4945. Namatame, and A. Fujimori, 1992a, Phys. Rev. B 46, 3771. Castellani, C., C. R. Natoli, and J. Ranninnger, 1978b, Phys. Bocquet, A. E., A. Fujimori, T. Mizokawa, T. Saitoh, H. Rev. B 18, 4967. Namatame, S. Suga, N. Kimizuka, Y. Takeda, and M. Ta- Castellani, C., C. R. Natoli, and J. Ranninnger, 1978c, Phys. kano, 1992b, Phys. Rev. B 45, 1561. Rev. B 18, 5001. Bocquet, A. E., T. Mizokawa, K. Morikawa, A. Fujimori, S. R. Cava, R. J., et al., 1990, Nature (London) 345, 602. Barman, K. B. Maiti, D. D. Sarma, Y. Tokura, and M. Cava, R. J., et al., 1991, Phys. Rev. B 43, 1229. Onoda, 1996, Phys. Rev. B 53, 1161. Ceperley, D. M., and B. J. Alder, 1980, Phys. Rev. Lett. 45, Bocquet, A. E., T. Saitoh, T. Mizokawa, and A. Fujimori, 566. 1992c, Solid State Commun. 83, 11. Chainani, A., M. Mathew, and D. D. Sarma, 1992, Phys. Rev. Bonn, D. A., J. D. Garrett, and T. Timusk, 1988, Phys. Rev. B 46, 9976. Lett. 61, 1305. Chainani, A., M. Mathew, and D. D. Sarma, 1993, Phys. Rev. Bonner, J. C., and M. E. Fisher, 1964, Phys. Rev. 135, A640. B 47, 15397. Bouchard, R. J., and Gillson, 1972, Mater. Res. Bull. 7, 873. Chainani, A., T. Yokoya, T. Morimoto, T. Takahashi, and S. Bouchard, R. J., et al., 1973, Mater. Res. Bull. 8, 489. Todo, 1995, Phys. Rev. B 51, 17976. Bozovic, I., J. H. Kim, J. S. Harris, Jr., C. B. Eom, J. M. Phil- Chainani, A., T. Yokoya, T. Morimoto, T. Takahashi, S. lips, and J. T. Cheung, 1994, Phys. Rev. Lett. 73, 1436. Yoshii, and M. Kasaya, 1994, Phys. Rev. B 50, 8915. Brandow, B. H., 1977, Adv. Phys. 26, 651. Chakravarty, S., B. I. Halperin, and D. R. Nelson, 1989, Phys. Brandt, U., and C. Mielsch, 1989, Z. Phys. 75, 365 (Part I); Rev. B 39, 2344. 1990, 79, 295 (Part II). Chandrasekhar, Q. Won Choi, J. Moyo, and J. M. Honig, 1970, Brinkmann, M., T. Rex, H. Bach, and K. Westerholt, 1995, Mater. Res. Bull. 5, 999. Phys. Rev. Lett. 74, 4927. Chao, K. A., and M. C. Gutzwiller, 1971, J. Appl. Phys. 42, Brinkmann, M., T. Rex, M. Stief, H. Bach, and K. Westerholt, 1420. 1996, Physica C 269, 76. Chao, K. A., J. Spalek, and A. M. Oles, 1977, J. Phys. C 10, Brinkman, W. F., and T. M. Rice, 1970, Phys. Rev. B 2, 4302. L271. Bru¨ ckner, W., H. Oppermann, W. Reichelt, J. I. Terukow, F. Chayes, J., L. Chayes, D. S. Fisher, and T. Spencer, 1986, Phys. A. Tschudnowski, and E. Wolf, 1983, Vanadiumoxide Rev. Lett. 57, 2999. Darstellung, Eigenschaften, Anwendung (Akademie, Berlin). Chen, C. H., S-W. Cheong, and A. S. Cooper, 1993, Phys. Rev. Brusetti, R., J. M. D. Coey, G. Czjzek, J. Fink, F. Gomp, and Lett. 71, 2461. H. Schmidt, 1980, J. Phys. F 10, 33. Chen, C. H., S-W. Cheong, and H. Y. Hwang, 1997, J. Appl. Bulut, N., D. J. Scalapino, and S. R. White, 1994a, Phys. Rev. Phys. 81, 4326. Lett. 72, 705. Chen, C. T., F. Sette, Y. Ma, M. S. Hybertsen, E. B. Stechel, Bulut, N., D. J. Scalapino, and S. R. White, 1994b, Phys. Rev. W. M. C. Foulkes, M. Schluter, S.-W. Cheong, A. S. Cooper, B 50, 7215. L. W. Rupp Jr., B. Batlogg, Y. L. Soo, Z. H. Ming, A. Krol, Buttrey, D. J., J. D. Sullivan, and A. L. Reingold, 1990, J. Solid and Y. H. Kao, 1991, Phys. Rev. Lett. 66, 104. State Chem. 88, 291. Chen, C. T., L. H. Tjeng, J. Kwo, H. L. Kao, P. Rudolf, F. Byers, N., and C. N. Yang, 1961, Phys. Rev. Lett. 7, 46. Sette, and R. M. Fleming, 1992, Phys. Rev. Lett. 68, 2543.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1250 Imada, Fujimori, and Tokura: Metal-insulator transitions

Cheong, S-W., et al., 1991, Phys. Rev. Lett. 67, 1791. Dardel, B., D. Malterre, M. Grioni, P. Weibel, Y. Baer, J. Voit, Cheong, S-W., H. Y. Hwang, C. H. Chen, B. Batlogg, L. W. and D. Je´roˆ me, 1992b, Europhys. Lett. 24, 687. Rupp, and S. A. Carter, 1994, Phys. Rev. B 49, 7088. Dare´, A.-M., Y. M. Vilk, and A.-M. S. Tremblay, 1996, Phys. Chien, T. R., Z. Z. Wang, and N. P. Ong, 1991, Phys. Rev. Rev. B 53, 14236. Lett. 67, 2088. Davis, L. C., 1982, Phys. Rev. B 25, 2912. Cho, J. H., F. C. Chou, and D. C. Johnston, 1993, Phys. Rev. de Boer, J. H., and E. J. W. Verway, 1937, Proc. Phys. Soc. Lett. 70, 222. London, Sect. A 49, 59. Chou, F. C., N. R. Belk, M. A. Kastner, R. J. Birgeneau, and de Gennes, P. G., 1960, Phys. Rev. 118, 141. A. Aharony, 1995, Phys. Rev. Lett. 75, 2204. Degiorgi, L., M. B. Hunt, H. R. Ott, M. Dressel, B. J. Feentrra, Chou, F. C., F. Borsa, J. H. Cho, D. C. Johnston, A. Lascial- G. Gru¨ ner, Z. Fisk, and P. Canfield, 1994, Europhys. Lett. 28, fari, D. R. Torgeson, and J. Ziolo, 1993, Phys. Rev. Lett. 71, 341. 2323. Deisz, J. J., D. W. Hess, and J. W. Serene, 1996, Phys. Rev. Chubukov, A. V., and J. Schmalian, 1998, Phys. Rev. B 57, Lett. 76, 1312. R11085. Demler, E., and S.-C. Zhang, 1995, Phys. Rev. Lett. 75, 4126. Claessen, R., R. O. Anderson, J. W. Allen, C. G. Olson, C. Dessau, D. S., C.-H. Park, T. Saitoh, Z.-X. Shen, P. Villela, N. Janowitz, W. P. Ellis, S. Harm, M. Kalning, R. Manzke, and Hamada, Y. Moritomo, and Y. Tokura, 1998, Phys. Rev. Lett. M. Skibowski, 1992, Phys. Rev. Lett. 69, 808. 81, 192. Clarke, D. G., S. P. Strong, and P. W. Anderson, 1994, Phys. Dessau, D. S., Z.-X. Shen, D. M. King, D. S. Marshall, L. W. Rev. Lett. 72, 3218. Lombardo, P. H. Dickinson, J. DiCarlo, C.-H. Park, A. G. Coey, J. M. D., R. Brusetti, A. Kallel, J. Schweizer, and H. Loeser, A. Kapitulnik, and W. E. Spicer, 1993, Phys. Rev. Fuss, 1974, Phys. Rev. Lett. 32, 1257. Lett. 71, 2781. Coey, J. M. D., and H. Roux-Buisson, 1979, Mater. Res. Bull. Ding, H., A. F. Bellman, J. C. Campuzano, M. Randeria, M. R. 14, 711. Norman, T. Yokoya, T. Takahashi, H. Katayama-Yoshida, T. Coleman, P., 1984, Phys. Rev. B 29, 3035. Mochiku, K. Kadowaki, G. Jennings, and G. P. Brivio, 1996a, Collins, R. T., et al., 1989, Phys. Rev. B 39, 6571. Phys. Rev. Lett. 76, 1533. Coluzza, C., H. Berger, P. Alme´ras, F. Gozzo, G. Margari- Ding, H., M. R. Norman, T. Yokoya, T. Takeuchi, M. Rand- tondo, G. Indlekofer, L. Forro, and Y. Hwu, 1993, Phys. Rev. eria, J. C. Campuzano, T. Takahashi, T. Mochiku, and K. B 47, 6625. Kadowaki, 1997, Phys. Rev. Lett. 78, 2628. Continentino, M. A., 1992, Phys. Rev. B 45, 11312. Ding, H., T. Yokoya, J. C. Campuzano, T. Takahashi, M. Ran- Continentino, M. A., 1994, Phys. Rep. 239, 179, and references deria, M. R. Norman, T. Mochiku, K. Kadowaki, and J. Gi- therein. apintzakis, 1996b, Nature (London) 382, 51. Cooper, S. L., G. A. Thomas, J. Orenstein, D. H. Rapkine, A. DiTusa, J. F., K. Friemelt, E. Bucher, G. Aeppli, and A. P. J. Millis, S-W. Cheong, A. S. Cooper, and Z. Fisk, 1990a, Ramirez, 1997, Phys. Rev. Lett. 78, 2831. Phys. Rev. B 41, 11605. Dobrosavljevic´, V., E. Abrahams, E. Miranda, and S. Chakra- Cooper, S. L., et al., 1990b, Phys. Rev. B 42, 10785. varty, 1997, Phys. Rev. Lett. 79, 455. Cooper, S. L., et al., 1992, Phys. Rev. B 45, 2549. Dobrosavljevic´, V., and G. Kotliar, 1997, Phys. Rev. Lett. 78, Cox, P. A., R. G. Edgell, J. B. Goodenough, A. Kamnett, and 3943. C. C. Naishi, 1983, J. Phys. C 16, 6221. Dopf, G., A. Muramatsu, and W. Hanke, 1992, Europhys. Lett. Crandles, D. A., T. Timusk, J. D. Garrett, and J. E. Greedan, 17, 559. 1994, Phys. Rev. B 49, 16207. Dopf, G., J. Wagner, P. Dieterich, A. Muramatsu, and W. Cullen, J. R., and E. R. Callen, 1973, Phys. Rev. B 7, 397. Hanke, 1992, Phys. Rev. Lett. 68, 2082. Czjzek, G., J. Fink, H. Schmidt, G. Krill, M. F. Lapierre, P. Dougier, P., and A. Casalot, 1970, J. Solid State Chem. 2, 396. Paissod, F. Gautier, and C. Robert, 1976, J. Magn. Magn. Dougier, P., J. C. C. Fan, and J. B. Goodenough, 1975, J. Solid Mater. 3, 58. State Chem. 14, 247. Czyzyk, M. T., and G. A. Sawatzky, 1997, unpublished. Dougier, P., and P. Hagenmuller, 1975, J. Solid State Chem. Dagotto, E., 1994, Rev. Mod. Phys. 66, 763. 15, 158. Dagotto, E., A. Moreo, F. Ortolani, D. Poilblanc, and J. Riera, Dumas, J., and C. Schlenker, 1976, J. Phys. (Paris), Colloq. 37, 1992, Phys. Rev. B 45, 10741. C4-41. Dagotto, E., A. Moreo, F. Ortolani, J. Riera, and D. J. Scala- Dumas, J., and C. Schlenker, 1979, unpublished. pino, 1991, Phys. Rev. Lett. 67, 1918. Dumas, J., C. Schlenker, J. L. Tholence, and R. Tournier, 1979, Dagotto, E., A. Nazarenko, and M. Boninsegni, 1994, Phys. Phys. Rev. B 20, 3913. Rev. Lett. 73, 728. Eccleston, R., M. Uehara, J. Akimitsu, H. Eisaki, N. Mo- Dagotto, E., and T. M. Rice, 1996, Science 271, 618. toyama, and S. Uchida, 1998, Phys. Rev. Lett. 81, 1702. Dagotto, E., J. Riera, and D. J. Scalapino, 1992, Phys. Rev. B Edgell, R. G., M. R. Harrison, M. D. Hill, L. Porte, and G. 45, 5744. Wall, 1984, J. Phys. C 17, 2889. Dai, P., H. A. Mook, and F. Dogˇan, 1998, Phys. Rev. Lett. 80, Edwards, P. P., and C. N. R. Rao, 1985, The Metallic and Non- 1738. Metallic States of Matter (Taylor & Francis). Dai, P., J. Zhang, H. A. Mook, S.-H. Liou, P. A. Dowben, and Eggert, S., I. Affleck, and M. Takahashi, 1994, Phys. Rev. Lett. E. W. Plummer, 1996, Phys. Rev. B 54, R3694. 73, 332. Dai, P., Y. Zhang, and M. P. Sarachik, 1992, Phys. Rev. B 45, Eisaki, H., 1991, Ph.D. thesis (University of Tokyo). 3984. Emery, V. J., 1979, in Highly Conducting One-Dimensional Dardel, B., D. Malterre, M. Grioni, P. Weibel, Y. Baer, D. Solids, edited by J. T. Devreese et al. (Plenum, New York), p. Schlenker, and Y. Pe´troff, 1992a, Europhys. Lett. 19, 525. 327.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1251

Emery, V. J., 1987, Phys. Rev. Lett. 58, 2794. Fujimori, A., and F. Minami, 1984, Phys. Rev. B 30, 957. Emery, V. J., and S. A. Kivelson, 1995, Nature (London) 374, Fujimori, A., F. Minami, and S. Sugano, 1984, Phys. Rev. B 29, 434. 5225. Emery, V. J., and S. A. Kivelson, 1996, Physica C 263, 44, and Fujimori, A., M. Saeki, N. Kimizuka, M. Taniguchi, and S. references therein. Suga, 1986, Phys. Rev. B 34, 7318. Emery, V. J., S. A. Kivelson, and H. Q. Lin, 1990, Phys. Rev. Fujimori, A., S. Takekawa, E. Takayama-Muromachi, Y. Lett. 64, 475. Uchida, A. Ono, T. Takahashi, Y. Okabe, and H. Katayama- Eskes, H., and A. Oles´, 1994, Phys. Rev. Lett. 73, 1279. Yoshidda, 1989, Phys. Rev. B 39, 2255. Eskes, H., A. Oles´, M. B. Meinders, and W. Stephan, 1994, Fujimori, A., K. Terakura, M. Taniguchi, S. Ogawa, S. Suga, Phys. Rev. B 50, 17980. M. Matoba, and S. Anzai, 1988, Phys. Rev. B 37, 3109. Eskes, H., L. H. Tjeng, and G. A. Sawatzky, 1990, Phys. Rev. B Fujioka, K., J. Okamoto, T. Mizokawa, A. Fujimori, I. Hase, 41, 288. M. Abbate, H. J. Lin, C. T. Chen, Y. Takeda, and M. Takano, Evertz, H. G., G. Lana, and M. Marcu, 1993, Phys. Rev. Lett. 1997, Phys. Rev. B 56, 6380. 70, 875. Fukuda, K., S. Shamoto, M. Sato, and K. Oda, 1988, Solid Evertz, H. G., and M. Marcu, 1993, Nucl. Phys. B (Proc. State Commun. 65, 1323. Suppl.) 30, 277. Fukunaga, F., and N. Tsuda, 1994, J. Phys. Soc. Jpn. 63, 3798. Fabrizio, M., A. Parola, and E. Tosatti, 1992, Phys. Rev. B 46, Fukushima, A., F. Iga, I. H. Inoue, K. Murata, and Y. Nishi- 3159. hara, 1994, J. Phys. Soc. Jpn. 63, 409. Finkelstein, A. M., 1983, Zh. Eksp. Teor. Fiz. 84, 168 [Sov. Fukuzumi, Y., K. Mizuhashi, K. Takenaka, and S. Uchida, Phys. JETP 57, 97 (1983)]. 1996, Phys. Rev. Lett. 76, 684. Finkelstein, A. M., 1984, Zh. Eksp. Teor. Fiz. 86, 367 [Sov. Furukawa, N., 1995a, J. Phys. Soc. Jpn. 64, 2734. Phys. JETP 86, 367 (1984)]. Furukawa, N., 1995b, J. Phys. Soc. Jpn. 64, 2754. Fisher, M. P. A., P. B. Weichman, G. Grinstein, and D. S. Furukawa, N., 1995c, J. Phys. Soc. Jpn. 64, 3164. Fisher, 1989, Phys. Rev. B 40, 546. Furukawa, N., 1996, J. Phys. Soc. Jpn. 65, 1174. Folkerts, W., G. A. Sawatzky, C. Haas, R. A. de Groot, and F. Furukawa, N., and M. Imada, 1991a, J. Phys. Soc. Jpn. 60, 810. U. Hillebrecht, 1987, J. Phys. C 20, 4135. Furukawa, N., and M. Imada, 1991b, J. Phys. Soc. Jpn. 60, Fong, H. F., B. Keimer, P. W. Anderson, D. Reznik, F. Dog˘an, 3604. and I. A. Aksay, 1995, Phys. Rev. Lett. 75, 316. Furukawa, N., and M. Imada, 1991c, J. Phys. Soc. Jpn. 60, 3669. Fong, H. F., B. Keimer, D. L. Milius, and I. A. Aksay, 1997, Furukawa, N., and M. Imada, 1992, J. Phys. Soc. Jpn. 61, 3331. Phys. Rev. Lett. 78, 713. Furukawa, N., and M. Imada, 1993, J. Phys. Soc. Jpn. 62, 2557. Fowler, R. H., 1933a, Proc. R. Soc. London, Ser. A 140, 505. Futami, T., and S. Anzai, 1984, J. Appl. Phys. 56, 440. Fowler, R. H., 1933b, Proc. R. Soc. London, Ser. A 141, 56. Fye, R. M., M. J. Martins, D. J. Scalapino, J. Wagner, and W. Frahm, H., and V. E. Korepin, 1990, Phys. Rev. B 42, 10553. Hanke, 1991, Phys. Rev. B 44, 6909. Fre´sard, R., and G. Kotliar, 1997, Phys. Rev. B 56, 12909. Gansepohl, S. C., M. Lee, K. Char, R. A. Rao, and C. B. Eom, Fu, C., M. P. C. M. Krijn, and S. Doniach, 1994, Phys. Rev. B 1995, Phys. Rev. B 52, 3495. 49, 2219. Garcı´a-Mun˜oz, J. L., J. Rodriguez-Carvajal, and P. Lacorre, Fujimori, A., 1992, J. Phys. Chem. Solids 53, 1595. 1992, Europhys. Lett. 20, 241. Fujimori, A., A. E. Bocquet, K. Morikawa, K. Kobayashi, T. Garcı´a-Mun˜oz, J. L., J. Rodrı´guez-Carvajal, and P. Lacorre, Saitoh, Y. Tokura, I. Hase, and M. Onoda, 1996, J. Phys. 1994, Phys. Rev. B 50, 978. Chem. Solids 57, 1379. Georges, A., and G. Kotliar, 1992, Phys. Rev. B 45, 6479. Fujimori, A., A. E. Bocquet, T. Saitoh, and T. Mizokawa, 1993, Georges, A., G. Kotliar, W. Krauth, and M. J. Rozenberg, J. Electron Spectrosc. Relat. Phenom. 62, 141. 1996, Rev. Mod. Phys. 68, 13. Fujimori, A., I. Hase, H. Namatame, Y. Fujishima, Y. Tokura, Georges, A., and W. Krauth, 1992, Phys. Rev. Lett. 69, 1240. H. Eisaki, S. Uchida, K. Takegahara, and F. M. F. de Groot, Georges, A., and W. Krauth, 1993, Phys. Rev. B 48, 7167. 1992a, Phys. Rev. Lett. 69, 1796. Ghedira, M., M. Anne, J. Chenavas, and F. Sayetat, 1986, J. Fujimori, A., I. Hase, M. Nakamura, H. Namatame, Y. Phys. C 19, 6489. Fujishima, Y. Tokura, M. Abbate, F. M. F. de Groot, M. T. Giamarchi, T., and C. Lhuillier, 1990, Phys. Rev. B 42, 10641. Czyzyk, J. C. Fuggle, O. Strebel, F. Lopez, M. Domke, and G. Gibb, T. C., R. Greatrex, N. N. Greenwood, D. C. Puxley, and Kaindl, 1992b, Phys. Rev. B 46, 9841. K. G. Snowdon, 1974, J. Solid State Chem. 11, 17. Fujimori, A., A. Ino, T. Mizokawa, C. Kim, Z.-X. Shen, T. Gofron, K., J. C. Campuzano, A. A. Abrikosov, M. Lindroos, Sasagawa, T. Kimura, K. Kishio, M. Takaba, K. Tamasaku, A. Bansil, H. Ding, D. Koelling, and B. Dabrowski, 1994, H. Eisaki, and S. Uchida, 1998, J. Phys. Chem. Solids, in Phys. Rev. Lett. 73, 3302. press. Goodenough, J. B., 1955, Phys. Rev. 100, 564. Fujimori, A., K. Kawakami, and N. Tsuda, 1988, Phys. Rev. B Goodenough, J. B., 1971, J. Solid State Chem. 3, 490. 38, 7889. Goral, J. P., J. E. Greedan, and D. A. MacLean, 1982, J. Solid Fujimori, A., N. Kimizuka, T. Akahane, T. Chiba, S. Kimura, State Chem. 43, 244. F. Minami, K. Siratori, M. Taniguchi, S. Ogawa, and S. Suga, Goto, A., H. Yasuoka, and Y. Ueda, 1996, J. Phys. Soc. Jpn. 1990, Phys. Rev. B 42, 7580. 65, 3043. Fujimori, A., N. Kimizuka, M. Taniguchi, and S. Suga, 1987, Graf, T., D. Mandrus, J. M. Lawrence, J. D. Thompson, P. C. Phys. Rev. B 36, 6691. Canfield, S.-W. Cheong, and L. W. Rupp, Jr., 1995, Phys. Fujimori, A., K. Mamiya, T. Mizokawa, T. Miyadai, T. Sekigu- Rev. B 51, 2037. chi, H. Takahashi, N. Moˆ ri, and S. Suga, 1996, Phys. Rev. B Granados, X., J. Fontcuberta, X. Obradors, and J. B. Torrance, 54, 16329. 1992, Phys. Rev. B 46, 15683.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1252 Imada, Fujimori, and Tokura: Metal-insulator transitions

Grant, J. B., and A. K. McMahan, 1991, Phys. Rev. Lett. 66, Hirota, K., N. Kaneko, A. Nishizawa, and Y. Endoh, 1996, J. 488. Phys. Soc. Jpn. 65, 3736. Greeff, C. W., H. R. Glyde, and B. E. Clements, 1992, Phys. Hirsch, J. E., 1983, Phys. Rev. B 28, 4059. Rev. B 45, 7951. Hirsch, J. E., 1985a, Phys. Rev. Lett. 54, 1317. Greywall, D. S., 1990, Phys. Rev. B 41, 1842. Hirsch, J. E., 1985b, Phys. Rev. B 31, 4403. Grilli, M., and G. Kotliar, 1990, Phys. Rev. Lett. 64, 1170. Hirsch, J. E., and R. M. Fye, 1986, Phys. Rev. Lett. 56, 2521. Gunnarsson, O., O. K. Andersen, O. Jepsen, and J. Zaanen, Hohenberg, P., and W. Kohn, 1964, Phys. Rev. 136, 864B. 1989, Phys. Rev. B 39, 1708. Homes, C. C., T. Timusk, R. Liang, D. A. Bonn, and W. N. Gutzwiller, M. C., 1965, Phys. Rev. A 137, A1726. Hardy, 1993, Phys. Rev. Lett. 71, 1645. Haas, S., A. Moreo, and E. D. Dagotto, 1995, Phys. Rev. Lett. Honig, J., 1985, in Metallic and Non-Metallic States of Matter, 74, 4281. edited by P. P. Edwards and C. N. Rao (Taylor & Francis, Haldane, F. D. M., 1980, Phys. Rev. Lett. 45, 1358. London), p. 226. Haldane, F. D. M., 1981, Phys. Rev. Lett. 47, 1840. Honig, J. M., and T. B. Reed, 1968, Phys. Rev. 174, 1020. Haldane, F. D. M., 1983a, Phys. Lett. A 93, 464. Honig, J. M., and L. L. van Zandt, 1975, Annu. Rev. Mater. Haldane, F. D. M., 1983b, Phys. Rev. Lett. 50, 1153. Sci. 5, 225. Haldane, F. D. M., 1994, Proceedings of the International Hubbard, J., 1963, Proc. R. Soc. London, Ser. A 276, 238. School of Physics Enrico Fermi Course CXXI, edited by R. Hubbard, J., 1964a, Proc. R. Soc. London, Ser. A 277, 237. A. Broglia and J. R. Schrieffer (North Holland, Amsterdam). Hubbard, J., 1964b, Proc. R. Soc. London, Ser. A 281, 401. Hamada, N., 1993, J. Phys. Chem. Solids 54, 1157. Hu¨ fner, S., 1985, Z. Phys. B 61, 135. ¨ Hamada, N., H. Sawada, and K. Terakura, 1995, in Spectros- Hufner, S., 1995, Photoelectron Spectroscopy: Principles and copy of Mott Insulators and Correlated Metals, edited by A. Applications, Springer Series in Solid-State Sciences No. 82, 2nd. Ed. (Springer, Berlin). Fujimori and Y. Tokura (Springer-Verlag, Berlin), p. 95. Hu¨ fner, S., T. Riesterer, and F. Hulliger, 1985, Solid State Harashina, H., K. Kodama, S. Shamoto, M. Sato, K. Kakurai, Commun. 54, 689. and M. Nishi, 1995, J. Phys. Soc. Jpn. 64, 1462. Hu¨ fner, S., and G. K. Wertheim, 1973, Phys. Lett. A 44, 133. Harrison, W. A., 1989, Electronic Structure and the Properties Hunt, M. B., M. A. Chernikov, E. Felder, H. R. Ott, Z. Fisk, of Solids (Dover, New York). and P. Canfield, 1994, Phys. Rev. B 50, 14933. Hase, M., I. Terasaki, and K. Uchinokura, 1993, Phys. Rev. Husmann, A., D. S. Jin, Y. V. Zastavker, T. F. Rosenbaum, X. Lett. 70, 3651. Yao, and J. M. Honig, 1996, Science 274, 1874. Hasegawa, H., 1997a, J. Phys. Soc. Jpn. 66, 1391. Hussey, N. E., M. Kibune, M. Nakagawa, N. Miura, Y. Iye, H. Hasegawa, H., 1997b, Phys. Rev. B 56, 1196. Takagi, S. Adachi, and K. Tanabe, 1998, Phys. Rev. Lett. 80, Hasenfratz, P., and F. Niedermayer, 1991, Phys. Lett. B 268, 2909. 231. Hussey, N. E., K. Nozawa, H. Takagi, S. Adachi, and K. Hasenfratz, P., and F. Niedermayer, 1993, Z. Phys. B 92, 91. Tanabe, 1997, Phys. Rev. B 56, R11423. Hatano, N., and M. Suzuki, 1993, in Quantum Monte Carlo Hutchings, M. T., G. Parisot, and D. Tocchett, 1978, in Neu- Methods in , edited by M. Suzuki tron Inelastic Scattering 1977 (IAEA, Vienna), Vol. II, p. 123. (World Scientific, Singapore), p. 13. Hwang, H. Y., et al., 1994, Phys. Rev. Lett. 72, 2636. Hayden, S. M., G. Aeppli, H. Mook, D. Rytz, M. F. Hundley, Hybertsen, M. S., M. Schlu¨ ter, and N. E. Christensen, 1989, and Z. Fisk, 1991a, Phys. Rev. Lett. 66, 821. Phys. Rev. B 39, 9028. Hayden, S. M., G. Aeppli, R. Osborn, A. D. Taylor, T. G. Hybertsen, M. S., E. B. Stechel, M. Schluter, and D. R. Jenni- Perring, S-W. Cheong, and Z. Fisk, 1991b, Phys. Rev. Lett. son, 1990, Phys. Rev. B 41, 11068. 67, 3622. Idemoto, Y., K. Fueki, and T. Shinbo, 1990, Physica C 166, Hayden, S. M., G. Aeppli, T. G. Perring, H. A. Mook, and F. 513. Dogˇan, 1996, unpublished. Igarashi, J., 1983, J. Phys. Soc. Jpn. 52, 2827. Hayden, S. M., et al., 1996, Phys. Rev. Lett. 76, 1344. Ihle, D., and B. Lorenz, 1980, Philos. Mag. B 42, 337. Hedin, L., and S. Lundqvist, 1969, Solid State Phys. 23,1. Ihle, D., and B. Lorenz, 1986, J. Phys. C 19, 5239. Heidemann, A., and M. Takano, 1980, Phys. Status Solidi B Iino, Y., and M. Imada, 1995, J. Phys. Soc. Jpn. 64, 4392. 100, 343. Imada, M., 1991, J. Phys. Soc. Jpn. 60, 1877. Hellberg, C. S., and E. Manousakis, 1997, Phys. Rev. Lett. 78, Imada, M., 1992, J. Phys. Soc. Jpn. 61, 423. 4609. Imada, M., 1993a, J. Phys. Soc. Jpn. 62, 1105. Henrich, V. E., and R. L. Kurtz, 1981, J. Vac. Sci. Technol. 18, Imada, M., 1993b, J. Phys. Soc. Jpn. 62, 2557. 416. Imada, M., 1993c, Phys. Rev. B 48, 550. Herring, C., 1952, Phys. Rev. 87, 60. Imada, M., 1993d, in Quantum Monte Carlo Methods in Con- Herring, C., 1963, in Magnetism, edited by G. T. Rado and H. densed Matter Physics, edited by M. Suzuki (World Scientific, Suhl (Academic, New York), Vol. II.B, p. 2). Singapore), p. 299. Hertel, P., and J. Appel, 1986, Phys. Rev. B 33, 2098. Imada, M., 1994a, J. Phys. Soc. Jpn. 63, 851. Hertz, J. A., 1976, Phys. Rev. B 14, 1165. Imada, M., 1994b, J. Phys. Soc. Jpn. 63, 3059. Hida, K., 1992, Phys. Rev. B 45, 2207. Imada, M., 1994c, J. Phys. Soc. Jpn. 63, 4294. Hida, K., M. Imada, and M. Ishikawa, 1983, J. Phys. C 16, 4945. Imada, M., 1995a, J. Low Temp. Phys. 99, 437. Hiroi, Z. et al., 1997, private communication. Imada, M., 1995b, J. Phys. Soc. Jpn. 64, 2954. Hiroi, Z., M. Azuma, M. Takano, and Y. Bando, 1991, J. Solid Imada, M., 1995c, in Spectroscopy of Mott Insulators and Cor- State Chem. 95, 230. related Metals, edited by A. Fujimori and Y. Tokura Hiroi, Z., and M. Takano, 1995, Nature (London) 377, 41. (Springer-Verlag, Berlin), p. 3.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1253

Imada, M., 1995d, in Computational Physics as a New Frontier Itoh, Y., H. Yasuoka, Y. Fujiwara, Y. Ueda, T. Machi, I. in Condensed Matter Research, edited by H. Takayama et al. Tomeno, K. Tai, N. Koshizuka, and S. Tanaka, 1992, J. Phys. (The Physical Society of Japan, Tokyo), p. 11. Soc. Jpn. 61, 1287. Imada, M., 1998, J. Phys. Soc. Jpn. 67, 45. Itoh, Y., H. Yasuoka, A. Hayashi, and Y. Ueda, 1994, J. Phys. Imada, M., F. F. Assaad, H. Tsunetsugu, and Y. Motome, 1998, Soc. Jpn. 63, 22. cond-mat/9808040. Itti, R., T. Wada, K. Matsuura, T. Itoh, K. Ikeda, H. Yamauchi, Imada, M., N. Furukawa, and T. M. Rice, 1992, J. Phys. Soc. N. Koshizuka, and S. Tanaka, 1991, Phys. Rev. B 44, 2306. Jpn. 61, 3861. Iye, Y., 1990, in Properties of High-Temperature Superconduct- Imada, M., and Y. Hatsugai, 1989, J. Phys. Soc. Jpn. 58, 3752. ors, edited by. P. M. Ginsberg (World Scientific, Singpore), Imada, M., and Y. Iino, 1997, J. Phys. Soc. Jpn. 66, 568. Vol. 2, p. 459. Imada, M., and M. Takahashi, 1986, J. Phys. Soc. Jpn. 55, 3354. Izumi, M., K. Nakazawa, Y. Bando, Y. Yoneda, and H. Terau- Imai, H., H. Wada, and M. Shiga, 1996, J. Phys. Soc. Jpn. 65, chi, 1997, J. Phys. Soc. Jpn. 66, 3893. 3460. Jaccarino, V., G. K. Wertheim, J. H. Wernich, L. R. Walker, Imai, T., C. P. Slichter, K. Yoshimura, and K. Kosuge, 1993, and S. Arajs, 1967, Phys. Rev. 160, 476. Phys. Rev. Lett. 70, 1002. Jaklicˇ, J., and P. Prelovsˇek, 1994, Phys. Rev. B 49, 5065. Imai, T., H. Yasuoka, T. Shimizu, Y. Ueda, K. Yoshimura, and Jaklicˇ, J., and P. Prelovsˇek, 1995a, Phys. Rev. B 52, 6903. K. Kosuge, 1989, Physica C 162-164, 169. Jaklicˇ, J., and P. Prelovsˇek, 1995b, Phys. Rev. Lett. 74, 3411. Inaba, F., T. Arima, T. Ishikawa, T. Katsufuji, and Y. Tokura, Jaklicˇ, J., and P. Prelovsˇek, 1996, Phys. Rev. Lett. 77, 892. 1995, Phys. Rev. B 52, R2221. Jarrell, M., 1992, Phys. Rev. Lett. 69, 168. Inagaki, S., and R. Kubo, 1973, Int. J. Magn. 4, 139. Jarrell, M., H. Akhlaghpour, and T. Pruschke, 1993, Phys. Rev. Inagaki, S., and R. Kubo, 1975, J. Phys. Soc. Jpn. 39, 596. Lett. 70, 1670. Ino, A., T. Mizokawa, A. Fujimori, K. Tamasaku, H. Eisuki, S. Jarrell, M., J. K. Fredericks, and T. Pruschke, 1995, Phys. Rev. Uchida, T. Kimura, T. Sasagawa, and K. Kishio, 1997a, Phys. B 51, 11704. Rev. Lett. 79, 2101. Jarrell, M., and T. Pruschke, 1993, Z. Phys. B 90, 187. Ino, A., T. Mizokawa, A. Fujimori, Y. Yasui, T. Nishikawa, Jarrell, M., and T. Pruschke, 1994, Phys. Rev. B 49, 1458. and M. Sato, 1997b, unpublished. Jarrett, H. S., et al., 1973, Mater. Res. Bull. 8, 877. Ino, A., T. Mizokawa, K. Kobayashi, A. Fujimori, T. Sasa- Jiang, Wu, S. N. Mao, X. X. Xi, Xiuguang Jiang, J. L. Peng, T. gawa, T. Kimura, K. Kishio, K. Tamasaku, H. Eisaki, and S. Venkatesan, C. J. Lobb, and R. L. Greene, 1994, Phys. Rev. Uchida, 1997c, unpublished. Lett. 73, 1291. Inoue, I. H., Y. Aiura, Y. Nishihara, Y. Haruama, S. Nishizaki, Jichu, H., T. Matsuura, and Y. Kuroda, 1989, J. Phys. Soc. Jpn. Y. Maeno, T. Fujita, J. G. Bednorz, and F. Lichtenberg, 1996, 58, 4280. J. Electron Spectrosc. Relat. Phenom. 78, 175. Jirak, Z., S. Krupicka, V. Nekvasild, E. Pollert, G. Villeneuve, Inoue, I. H., I. Hase, Y. Aiura, A. Fujimori, Y. Haruyama, T. and F. Zounova, 1980, J. Magn. Magn. Mater. 15-18, 519. Maruyama, and Y. Nishihara, 1995, Phys. Rev. Lett. 74, 2539. Jirak, Z., S. Krupicka, Z. Simsa, M. Dlouha, and Z. Vratislav, Inoue, I. H., H. Makino, I. Hase, M. Ishikawa, N. E. Hussey, 1985, J. Magn. Magn. Mater. 53, 153. and M. J. Rozenberg, 1997a, unpublished. Johnston, D. C., 1989, Phys. Rev. Lett. 62, 957. Inoue, I. H., et al., 1997b, Physica B 237-238, 61. Johnston, D. C., 1996, Phys. Rev. B 54, 13009. Inoue, J., and S. Maekawa, 1990, J. Phys. Soc. Jpn. 59, 2110. Kadono, R., H. Okajima, A. Yamashita, K. Ishii, T. Yokoo, J. Inui, M., and P. B. Littlewood, 1991, Phys. Rev. B 44, 4415. Akimitsu, N. Kobayashi, Z. Hiroi, M. Takano, and K. Naga- Ioffe, L. B., and G. Kotliar, 1990, Phys. Rev. B 42, 10348. mine, 1996, Phys. Rev. B 54, R9628. Ioffe, L. B., and A. Larkin, 1989, Phys. Rev. B 39, 8988. Kadowaki, K., and S. B. Woods, 1986, Solid State Commun. Ioffe, L. B., and P. B. Wiegmann, 1990, Phys. Rev. Lett. 65, 58, 507. 653. Kaidanov, V. I., V. A. Tselishchev, I. K. Iesalniek, L. D. Dud- Ishihara, S., J. Inoue, and S. Maekawa, 1997, Phys. Rev. B 55, kin, B. K. Voronov, and N. N. Trusova, 1968, Sov. Phys. 8280. Semicond. 2, 382. Ishihara, S., Y. Yamanaka, and N. Nagaosa, 1997, Phys. Rev. B Kajueter, H., and G. Kotliar, 1996, Preprint No. 56, 686. cond-mat/9609176. Ishikawa, T., S. K. Park, and Y. Tokura, 1998, Phys. Rev. B (in Kajueter, H., and G. Kotliar, 1997, Int. J. Mod. Phys. B 11, 729. press). Kakeshita, T., R. Kajimoto, H. Yoshizawa, T. Tanabe, T. Kat- Isobe, M., and Y. Ueda, 1996, J. Phys. Soc. Jpn. 65, 1178. sufuji, and Y. Tokura, 1998, unpublished. Ito, T., H. Takagi, T. Ido, S. Ishibashi, and S. Uchida, 1991, Kakizaki, A., H. Sugawara, I. Nagakura, Y. Ishikawa, T. Ko- Nature (London) 350, 596. matsubara, and T. Ishii, 1982, J. Phys. Soc. Jpn. 51, 2597. Ito, T., K. Takenaka, and S. Uchida, 1993, Phys. Rev. Lett. 70, Kakizaki, A., K. Sugeno, T. Ishii, H. Sugawara, I. Nagakura, 3995. and S. Shin, 1983, Phys. Rev. B 28, 1026. Itoh, M., M. Sugahara, I. Natori, and K. Motoya, 1995, J. Phys. Kambe, S., H. Yasuoka, H. Takagi, S. Uchida, and Y. Tokura, Soc. Jpn. 64, 3967. 1991, J. Phys. Soc. Jpn. 60, 400. Itoh, S., K. Yamada, M. Arai, Y. Endoh, Y. Hidaka, and S. Kamimura, H., and Y. Suwa, 1993, J. Phys. Soc. Jpn. 62, 3368. Hosoya, 1994, J. Phys. Soc. Jpn. 63, 4542. Kampf, A., and J. R. Schrieffer, 1990, Phys. Rev. B 41, 6399. Itoh, T., T. Fujiwara, and S. Tanaka, 1994, unpublished. Kanamori, J., 1959, J. Phys. Chem. Solids 10, 87. Itoh, Y., 1994, J. Phys. Soc. Jpn. 63, 3522. Kanamori, J., 1963, Prog. Theor. Phys. 30, 275. Itoh, Y., T. Machi, A. Fukuoka, K. Tanabe, and H. Yasuoka, Kanbayashi, A., 1978, J. Phys. Soc. Jpn. 44, 108. 1996, J. Phys. Soc. Jpn. 65, 3751. Kanoda, K., 1997, Hyperfine Interact. 104, 235.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1254 Imada, Fujimori, and Tokura: Metal-insulator transitions

Kaplan, S. G., S. Wu, H.-T. S. Lihn, H. D. Drew, Q. Li, D. B. King, D. M., Z.-X. Shen, D. S. Dessau, B. O. Wells, W. E. Fenner, J. M. Phillips, and S. Y. Hou, 1996, Phys. Rev. Lett. Spicer, A. J. Arko, D. S. Marshall, J. DiCarlo, A. G. Loeser, 76, 696. C. H. Park, E. R. Ratner, J. L. Peng, Z. Y. Li, and R. L. Karlsson, K., O. Gunnarsson, and O. Jepsen, 1992, J. Phys.: Greene, 1993, Phys. Rev. Lett. 70, 3159. Condens. Matter 4, 895. Kirkpatrick, T. R., and D. Belitz, 1995, Phys. Rev. Lett. 73, Kasowski, R. V., 1973, Phys. Rev. B 8, 1378. 862. Kastner, M. A., R. J. Birgeneau, T. R. Thurston, P. J. Picone, Kirkpatrick, T. R., and D. Belitz, 1997, cond-mat/9707001. H. P. Jenssen, D. R. Gabbe, M. Sato, K. Fukuda, S. Shamoto, Kiryukhin, V., D. Cax, J. P. Hill, B. Keimer, A. Vigliante, Y. Y. Endoh, K. Yamada, and G. Shirane, 1988, Phys. Rev. B 38, Tomioka, and Y. Tokura, 1997, Nature (London) 386, 813. 6636. Kishine, J., and H. Fukuyama, 1997, J. Phys. Soc. Jpn. 66, 26. Kasuya, H., Y. Tokura, T. Arima, H. Eisaki, and S. Uchida, Kita, E., Y. Tokuyama, A. Tasaki, and K. Siratori, 1983, J. 1993, Phys. Rev. B 47, 6197. Magn. Magn. Mater. 31-34, 787. Katano, S., J. A. Fernandez-Baca, S. Funahashi, N. Mori, Y. Kitaoka, Y., K. Ishida, S. Hiramatu, and K. Asayama, 1988, J. Ueda, and K. Koga, 1993, Physica C 214, 64. Phys. Soc. Jpn. 57, 734. Kataoka, M., 1996, Czech. J. Phys. 46, Suppl. F4. Kittel, C., 1968, in Solid State Physics 22 (Academic, New Katayama, M., S. Akimoto, and K. Asayama, 1977, J. Phys. York), p. 1. Soc. Jpn. 42, 97. Klein, L., J. S. Dodge, C. H. Ahn, G. J. Snyder, T. H. Geball, Kato, M., H. Chizawa, Y. Koike, T. Toji, and Y. Saito, 1994, M. R. Beasley, and A. Kapitulnik, 1996, Phys. Rev. Lett. 77, Physica (Amsterdam) 235C-244C, 1327. 2774. Kato, M., K. Shiota, and Y. Koike, 1996, Physica C 258, 284. Kobayashi, H., M. Nagata, R. Kanno, and Y. Kawamoto, 1994, Kato, T., and M. Imada, 1998, J. Phys. Soc. Jpn. 67, 2828. Mater. Res. Bull. 29, 1271. Katoh, N., and M. Imada, 1993, J. Phys. Soc. Jpn. 62, 3728. Koehler, R. F., Jr., and R. L. White, 1973, J. Appl. Phys. 44, Katoh, N., and M. Imada, 1994, J. Phys. Soc. Jpn. 63, 4529. 1682. Katoh, N., and M. Imada, 1995, J. Phys. Soc. Jpn. 64, 4105. Kohgi, M., and Y. Ishikawa, 1981, Solid State Commun. 37, Katoh, N., and M. Imada, 1998, J. Phys. Soc. Jpn. 67, 564. 833. Katsufuji, T., M. Kasai, and Y. Tokura, 1996, Phys. Rev. Lett. Kohn, W., 1964, Phys. Rev. A 133, 171. 76, 126. Kohn, W., and L. J. Sham, 1965, Phys. Rev. 140, A1133. Katsufuji, T., Y. Okimoto, T. Arima, Y. Tokura, and J. B. Kohno, H., and K. Yamada, 1991, Prog. Theor. Phys. 85, 13. Torrance, 1995, Phys. Rev. B 51, 4830. Kohno, M., 1997, Phys. Rev. B 55, 1435. Katsufuji, T., Y. Okimoto, and Y. Tokura, 1995, Phys. Rev. Kopietz, P., V. Meden, and K. Scho¨ nhammer, 1995, Phys. Rev. Lett. 75, 3497. Lett. 74, 2997. Katsufuji, T., Y. Taguchi, and Y. Tokura, 1997, Phys. Rev. B Korenman, V., J. L. Murray, and R. E. Prange, 1977a, Phys. 56, 10145. Rev. B 16, 4032. Katsufuji, T., T. Tanabe, T. Ishikawa, Y. Fukuda, T. Arima, Korenman, V., J. L. Murray, and R. E. Prange, 1977b, Phys. and Y. Tokura, 1996, Phys. Rev. B 54, R14230. Rev. B 16, 4048. Katsufuji, T., T. Tanabe, and Y. Tokura, 1998, unpublished. Korenman, V., J. L. Murray, and R. E. Prange, 1977c, Phys. Katsufuji, T., and Y. Tokura, 1994, Phys. Rev. B 50, 2704. Rev. B 16, 4058. Kawabata, A., 1972, Prog. Theor. Phys. 48, 1793. Korotin, M. A., S. Yu. Ezhov, I. V. Solovyev, V. I. Anisimov, Kawabata, A., 1975, Prog. Theor. Phys. 54, 45. D. I. Khomskii, and G. A. Sawatzky, 1996, Phys. Rev. B 54, Kawakami, N., and S.-K. Yang, 1990, Phys. Lett. A 148, 359. 5309. Kawakami, N., and S.-K. Yang, 1991, Phys. Rev. Lett. 67, 2493. Koshibae, W., Y. Kawamura, S. Ishihara, S. Okamoto, J. In- Kawano, H., R. Kajimoto, M. Kubota, and H. Yoshizawa, oue, and S. Maekawa, 1997, J. Phys. Soc. Jpn. 66, 957. 1996, Phys. Rev. B 53, R14709. Koshibae, W., Y. Ohta, and S. Maekawa, 1993, Phys. Rev. B Kawano, H., R. Kajimoto, H. Yoshizawa, Y. Tomioka, H. Ku- 47, 3391. wahara, and Y. Tokura, 1997, Phys. Rev. Lett. 78, 4253. Koshibae, W., Y. Ohta, and S. Maekawa, 1994, Phys. Rev. B Kebler, J., A. T. Aldred, G. H. Lander, M. H. Mueller, O. 50, 3767. Massenet, and G. D. Stucky, 1980, J. Solid State Chem. 351, Kotani, T., 1997, cond-mat/9711062. 351. Kotliar, G., 1993, in Correlated Electron Systems, edited by V. Keimer, B., N. Belk, R. J. Birgeneau, A. Cassanho, C. Y. J. Emery (World Scientific, Singapore), p. 118. Chen, M. Greven, M. A. Kastner, A. Aharony, Y. Endoh, R. Kotliar, G., and H. Kajueter, 1996, Phys. Rev. B 54, 14221. W. Erwin, and G. Shirane, 1992, Phys. Rev. B 46, 14034. Kotliar, G., and J. Liu, 1988, Phys. Rev. B 38, 5142. Keren, A., L. P. Le, G. M. Luke, B. J. Sternlieb, W. D. Wu, Y. Koyama, T., and M. Tachiki, 1986, Phys. Rev. B 34, 3272. J. Uemura, S. Tajima, and S. Uchida, 1993, Phys. Rev. B 48, Kramers, H. A., 1934, Physica (Amsterdam) 1, 1934. 12926. Kravchenko, S. V., W. E. Mason, G. E. Bowker, J. E. Fur- Kim, C., A. Y. Matsuura, Z.-X. Shen, N. Motoyama, H. Eisaki, neaux, V. M. Pudalov, and M. D’Iorio, 1995, Phys. Rev. B 51, S. Uchida, T. Tohyama, and S. Maekawa, 1996, Phys. Rev. 7038. Lett. 77, 4054. Krill, G., and A. Amamou, 1980, J. Phys. Chem. Solids 41, 531. Kim, C., Z.-X. Shen, N. Motoyama, H. Eisaki, S. Uchida, T. Ku¨ bert, C., and A. Muramatsu, 1996, Int. J. Mod. Phys. B 10, Tohyama, and S. Maekawa, 1997, Phys. Rev. B 56, 15589. 3807. Kimura, T., Y. Tomioka, A. Asamitsu, and Y. Tokura, 1997, Kubo, K., and T. Kishi, 1988, Phys. Rev. Lett. 61, 2585, and unpublished (submitted to Phys. Rev. Lett.). references therein. Kimura, T., Y. Tomioka, H. Kuwahara, A. Asamitsu, M. Kubo, K., and N. Ohata, 1972, J. Phys. Soc. Jpn. 33, 21. Tamura, and Y. Tokura, 1996, Science 274, 1698. Kubo, R., 1952, Phys. Rev. 87, 568.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1255

Kubo, R., 1957, J. Phys. Soc. Jpn. 12, 570. Liu, R., B. W. Veal, A. P. Paulikas, J. W. Downey, H. Shi, C. Kubo, Y., Y. Shimakawa, T. Manako, and H. Igarashi, 1991, G. Olson, C. Gu, A. J. Arko, and J. J. Joyce, 1992b, Phys. Phys. Rev. B 43, 7875. Rev. B 45, 5614. Kugel, K. I., and D. I. Khomskii, 1973, Sov. Phys. JETP 37, Lloyd, P., and P. R. Best, 1975, J. Phys. C 8, 3752. 725. Loeser, A., Z.-X. Shen, D. S. Dessau, D. S. Marshall, C.-H. Kugel, K. I., and D. I. Khomskii, 1982, Sov. Phys. Usp. 25, 231. Park, P. Founier, and A. Kapitulnik, 1996, Science 273, 325. Kuipers, A. J. M., and V. A. M. Brabers, 1976, Phys. Rev. B 14, Loh, E. Y., Jr., J. E. Gubernatis, R. T. Scalettar, S. R. White, 1401. D. J. Scalapino, and R. L. Sugar, 1990, Phys. Rev. B 41, 9301. Kumagai, K., K. Kawano, I. Watanabe, K. Nishiyama, and K. Longo, J. M., P. M. Raccah, and J. B. Goodenough, 1968, J. Nagamine, 1994, J. Supercond. 7, 63. Appl. Phys. 39, 1327. Kumagai, K., H. Matoba, N. Wada, M. Okaji, and K. Nara, Loram, J. W., K. A. Mirza, J. R. Cooper, N. Athamassopoulou, 1991, J. Phys. Soc. Jpn. 60, 1448. and W. Y. Liang, 1997, unpublished. Kumagai, K. I., T. Suzuki, Y. Taguchi, Y. Okada, Y. Fujishima, Loram, J. W., K. A. Mirza, J. R. Cooper, and W. Y. Liang, and Y. Tokura, 1993, Phys. Rev. B 48, 7636. 1994, Physica C 235-240, 134. Kumagai, K., S. Tsuji, M. Kato, and Y. Koike, 1997, Phys. Rev. Loram, J. W., K. A. Mirza, W. Y. Liang, and J. Osborne, 1989, Lett. 78, 1992. Physica C 162-164, 498. Kuramoto, Y., 1985, in Theory of Heavy Fermions and Valence Lorenz, B., and D. Ihle, 1979, Phys. Status Solidi B 96, 659. Fluctuations, edited by T. Kasuya and T. Saso (Springer, Ber- Lorenz, B., and D. Ihle, 1982, Phys. Status Solidi B 113, 601. lin), p. 152. Lorenzana, J., and G. A. Sawatzky, 1995, Phys. Rev. Lett. 74, Kuramoto, Y., and N. Fukushima, 1998, J. Phys. Soc. Jpn. 67, 1867. 583. Lu, D. H., M. Schmidt, T. R. Cummins, S. Schuppler, F. Lich- tenberg, and J. G. Bednorz, 1996, Phys. Rev. Lett. 76, 4845. Kuramoto, Y., and K. Miyake, 1990, J. Phys. Soc. Jpn. 59, 2831. Lu, J. P., 1994, Phys. Rev. B 49, 5687. Kuramoto, Y., and T. Watanabe, 1987, Physica B 148, 80. Lucovsky, G., J. W. Allen, and R. Allen, 1979, in Physics of Kuroki, K., T. Kimura, and H. Aoki, 1996, Phys. Rev. B 54, Semiconductors, Conference Series No. 43, edited by B. L. H. R15641. Wilson (Institute of Physics, London), p. 465. Kuwahara, H., Y. Tomioka, A. Asamitsu, Y. Moritomo, and Luke, G. M., B. Sternlieb, Y. J. Uemura, J. H. Brewer, R. Y. Tokura, 1995, Science 270, 961. Kadono, R. F. Kiefl, S. R. Kreitzman, T. M. Riseman, J. Go- Kuwahara, H., et al., 1998, unpublished. palakrishman, A. W. Sleight, M. A. Subramanian, S. Uchida, Kuwamoto, H., J. M. Honig, and J. Appel, 1980, Phys. Rev. B H. Takagi, and Y. Tokura, 1989, Nature (London) 338, 49. 22, 2626. Lusher, C. P., B. P. Cowan, and J. Saunders, 1993, Phys. Rev. Kwizera, P., M. S. Dresselhaus, and D. Adler, 1980, Phys. Rev. Lett. 67, 2497. B 21, 2328. Luther, A., and V. J. Emery, 1974, Phys. Rev. Lett. 33, 589. Lad, R. J., and V. E. Henrich, 1989, Phys. Rev. B 39, 13478. Luttinger, J. M., 1963, J. Math. Phys. 4, 1154. Ladd, L., and W. Paul, 1969, Solid State Commun. 7, 425. Lyons, K. B., P. A. Fleury, J. P. Remeika, A. S. Cooper, and T. Landau, L. D., 1957a, Sov. Phys. JETP 3, 920. J. Negran, 1988, Phys. Rev. B 37, 2353. Landau, L. D., 1957b, Sov. Phys. JETP 5, 101. Ma, S.-K., 1976, Modern Theory of Critical Phenomena, Fron- Landau, L. D., 1958, Sov. Phys. JETP 8, 70. tiers in Physics Series No. 46 (Addison-Wesley, Reading, Langer, M., J. Schmalian, S. Grabowsky, and K. H. Benne- MA). mann, 1995, Phys. Rev. Lett. 75, 4508. Machi, T., I. Tomeno, T. Miyatake, N. Koshizuka, S. Tanaka, Langreth, D. C., and M. J. Mehl, 1983, Phys. Rev. B 28, 1809. T. Imai, and H. Yasuoka, 1991, Physica C 173, 32. Lee, P. A., and N. Nagaosa, 1992, Phys. Rev. B 46, 5621. Machida, K., 1989, Physica C 158, 192. Lee, P. A., and T. V. Ramakrishnan, 1985, Rev. Mod. Phys. 57, Mackenzie, A. P., S. R. Julian, A. J. Diver, G. J. McMullan, M. 287. P. Ray, G. G. Lonzarich, Y. Maeno, S. Nishizaki, and T. Lee, S.-H., and S-W. Cheong, 1997, Phys. Rev. Lett. 79, 2514. Fujita, 1996, Phys. Rev. Lett. 76, 3786. Li, E. K., K. H. Johnson, D. E. Eastman, and J. L. Freeouf, MacKinnon, A., and B. Kramer, 1993, Rep. Prog. Phys. 56, 1974, Phys. Rev. Lett. 32, 470. 1469. Li, J. Q., Y. Matsui, S. K. Park, and Y. Tokura, 1997, Phys. MacLean, D. A., Hok-Nam Ng, and J. E. Greedan, 1979, J. Rev. Lett. 79, 297. Solid State Chem. 30, 35. Li, Q., J. Callaway, and L. Tan, 1991, Phys. Rev. B 44, 10256. Maeno, Y., 1996, in Advances in Superconductivity VIII, edited Liang, W. Y., J. W. Loram, K. A. Mirza, N. Athanassopoulou, by H. Hayakawa and Y. Enomoto (Springer, Tokyo), p. 15. and J. R. Cooper, 1996, Physica C 263, 277. Maeno, Y., H. Hashimoto, K. Yoshida, S. Nishizaki, T. Fujita, Lichtenberg, F., A. Catana, J. Mannhart, and D. G. Schlom, J. G. Bednorz, and F. Lichtenberg, 1994, Nature (London) 1992, Appl. Phys. Lett. 60, 1138. 372, 532. Liechtenstein, A. I., V. I. Anisimov, and J. Zaanen, 1995, Phys. Maeno, Y., K. Yoshida, H. Hashimoto, S. Nishizaki, S. Ikeda, Rev. B 52, 5467. M. Nohara, T. Fujita, A. P. Machenzie, N. E. Hussey, J. G. Litvinchuk, A. P., C. Thomsen, and M. Cardona, 1993, Solid Bednorz, and F. Lichtenberg, 1997, J. Phys. Soc. Jpn. 66, State Commun. 83, 343. 1405. Liu, L.-Z., R. O. Anderson, and J. W. Allen, 1991, J. Phys. Maezono, R., S. Ishihara, and N. Nagaosa, 1998, Phys. Rev. B Chem. Solids 52, 1473. 57, R13993. Liu, R., B. W. Veal, A. P. Paulikas, J. W. Downey, P. J. Kostic´, Magishi, K., S. Matsumoto, Y. Kitaoka, K. Ishida, K. Asayama, S. Fleshler, U. Welp, C. G. Olson, X. Wu, A. J. Arko, and J. M. Uehara, T. Nagata, and J. Akimitsu, 1998, Phys. Rev. B J. Joyce, 1992a, Phys. Rev. B 46, 11056. 57, 11533.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1256 Imada, Fujimori, and Tokura: Metal-insulator transitions

Mahadevan, P., N. Shanthi, and D. D. Sarma, 1996, Phys. Rev. Mattheiss, L. F., 1994, J. Phys.: Condens. Matter 6, 6477. B 54, 11199. Mattheiss, L. F., 1995, Solid State Commun. 93, 791. Mahajan, A. V., D. C. Johnston, P. R. Torgeson, and F. Borsa, Mattheiss, L. F., 1996a, Solid State Commun. 97, 751. 1992, Phys. Rev. B 46, 10973. Mattheiss, L. F., 1996b, J. Phys.: Condens. Matter 8, 5987. Maiti, K., D. D. Sarma, T. Mizokawa, and A. Fujimori, 1997, Mattheiss, L. F., and D. R. Hamann, 1993, Phys. Rev. B 47, Europhys. Lett. 37, 359. 13114. Makivic´, M. S., and H.-Q. Ding, 1991, Phys. Rev. B 43, 3562. Mayaffre, H., P. Auban-Senzier, D. Je´rome, D. Poilblanc, C. Mamiya, K., T. Mizokawa, A. Fujimori, N. Chandrasekaran, S. Bourbonnais, U. Ammerrahl, G. Dhalenne, and A. Rev- R. Krishnakumar, D. D. Sarma, T. Miyadai, H. Takahashi, N. colevschi, 1997, cond-mat/9710063. Moˆ ri, and S. Suga, 1997a, Phys. Rev. B, in press. Mazin, I. I., and D. J. Singh, 1997, Phys. Rev. Lett. 79, 733. Mamiya, K., T. Mizokawa, A. Fujimori, T. Miyadai, H. Taka- McCarron, E. M., M. A. Subramanian, J. C. Calabrese, and R. hashi, N. Moˆ ri, and S. Suga, 1997b, unpublished. L. Harlow, 1988, Mater. Res. Bull. 23, 1355. Marezio, M., D. B. McWhan, J. P. Remeika, and P. D. McElfresh, M. W., J. M. D. Coey, P. Strobel, and S. von Mol- Dernier, 1972, Phys. Rev. B 5, 2541. nar, 1989, Phys. Rev. B 40, 825. Marezio, M., J. P. Remeika, and P. D. Dernier, 1970, Acta McKane, A. J., and M. Stone, 1981, Ann. Phys. (N.Y.) 131, 36. Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 26, McMahan, A. K., J. F. Annett, and R. M. Martin, 1990, Phys. 2008. Rev. B 42, 6268. Marshall, D. S., D. S. Dessau, A. G. Loeser, C.-H. Park, A. Y. McMahan, A. K., R. M. Martin, and S. Satpathy, 1989, Phys. Matsuura, J. N. Eckstein, I. Bozovic, P. Fournier, A. Kapit- Rev. B 38, 6650. ulnik, W. E. Spicer, and Z.-X. Shen, 1996, Phys. Rev. Lett. 76, McWhan, D. B., M. Marezio, J. P. Remeika, and P. D. 4841. Dernier, 1972, Phys. Rev. B 5, 2552. Mason, T. E., G. Aeppli, and H. A. Mook, 1992, Phys. Rev. McWhan, D. B., A. Menth, J. P. Remeika, W. F. Brinkman, Lett. 68, 1414. and T. M. Rice, 1973, Phys. Rev. B 7, 1920. Mason, T. E., et al., 1993, Phys. Rev. Lett. 71, 919. McWhan, D. B., and J. P. Remeika, 1970, Phys. Rev. B 2, 3734. Mason, T. E., A. Schro¨ der, G. Aeppli, H. A. Mook, and S. M. McWhan, D. B., J. P. Remeika, T. M. Rice, W. F. Brinkman, J. Hayden, 1996, Phys. Rev. Lett. 77, 1604. P. Maita, and A. Menth, 1971, Phys. Rev. Lett. 27, 941. Massida, S., N. Hamada, J. Yu, and A. J. Freeman, 1989, ¨ Physica C 157, 571. Meden, V., and K. Schonhammer, 1993a, Phys. Rev. B 46, Massenet, O., R. Buder, J. J. Since, C. Schlenker, J. Mercier, J. 15753. Kebler, and D. G. Stucky, 1978, Mater. Res. Bull. 13, 187. Meden, V., and K. Scho¨ nhammer, 1993b, Phys. Rev. B 47, Massenet, O., J. J. Since, J. Mercier, M. Avignon, R. Buder, V. 16205. D. Ngyuyen, and J. Kebler, 1979, Solid State Commun. 40, Metzner, W., and D. Vollhardt, 1989, Phys. Rev. Lett. 62, 324. 573. Metzner, W., P. Schmit, and D. Vollhardt, 1992, Phys. Rev. B Matl, P., Y. F. Yan, N. P. Ong, Y. Q. Li, D. Studebaker, T. 45, 2237. Baum, and G. Doubinina, 1997, unpublished. Micnas, R., J. R. Ranninger, and S. Robaszkiewicz, 1990, Rev. Matoba, M., and S. Anzai, 1987, J. Magn. Magn. Mater. 70, Mod. Phys. 62, 113. 213. Mila, F., and T. M. Rice, 1989, Physica C 157, 561. Matoba, M., S. Anzai, and A. Fujimori, 1991, J. Phys. Soc. Jpn. Miles, P. A., W. B. Westphal, and A. von Hippel, 1957, Rev. 60, 4230. Mod. Phys. 29, 279. Matoba, M., S. Anzai, and A. Fujimori, 1994, J. Phys. Soc. Jpn. Millis, A. J., 1993, Phys. Rev. B 48, 7183. 63, 1429. Millis, A. J., 1996, Phys. Rev. B 53, 8434. Matsuda, M., Y. Endoh, Y. Yamada, K. Kojima, I. Tanaka, R. Millis, A. J., P. B. Littlewood, and B. I. Shraiman, 1995, Phys. J. Birgeneau, M. A. Kastner, and G. Shirane, 1992, Phys. Rev. Rev. Lett. 74, 5144. B 45, 12548. Millis, A. J., H. Monien, and D. Pines, 1990, Phys. Rev. B 42, Matsuda, M., and K. Katsumata, 1996, Phys. Rev. B 53, 12201. 167. Matsuda, M., K. Katsumata, H. Eisaki, N. Motoyama, S. Millis, A. J., B. I. Shraiman, and R. Mueller, 1996, Phys. Rev. Uchida, S. M. Shapiro, and G. Shirane, 1996, Phys. Rev. B 54, Lett. 77, 175. 12199. Mills, D. L., A. A. Maradudin, and E. Burstein, 1970, Ann. Matsuda, M., K. Yamada, Y. Endoh, T. R. Thurston, G. Phys. (N.Y.) 56, 504. Shirane, R. J. Birgeneau, M. A. Kastner, I. Tanaka, and H. Misawa, S., 1995, Physica B 211, 158. Kojima, 1994, Phys. Rev. B 49, 6958. Misawa, S., and A. Tate, 1996, J. Magn. Magn. Mater. 157-158, Matsumoto, S., Y. Kitaoka, K. Ishida, K. Asayama, Z. Hiroi, 617. N. Kobayashi, and M. Takano, 1996, Phys. Rev. B 53, Mishra, S. K., and S. Satpathy, 1993, Phys. Rev. B 47, 5564. R11942. Mitchell, J. F., D. N. Argyriou, J. D. Jorgensen, D. G. Hinks, Matsumoto, S., et al., 1997, unpublished. C. D. Potter, and S. D. Bader, 1997, Phys. Rev. B 55, 63. Matsuura, A. Y., Z.-X. Shen, D. S. Dessau, C.-H. Park, T. Miyadai, T., K. Kikuchi, and H. Ito, 1978, Phys. Lett. A 67, 61. Thio, J. W. Bennett, and O. Jepsen, 1996, Phys. Rev. B 53, Miyadai, T., S. Sudo, Y. Tazuke, N. Moˆ ri, and Y. Miyao, 1983, R7584. J. Magn. Magn. Mater. 31-34, 337. Matsuura, K., T. Wada, T. Nakamizo, H. Yamauchi, and S. Miyake, K., 1996, unpublished. Tanaka, 1991, Phys. Rev. B 43, 13118. Miyake, K., and O. Narikiyo, 1994a, J. Phys. Soc. Jpn. 63, 2042. Mattheiss, L. F., 1972, Phys. Rev. B 5, 290 (Part I); 5, 306 (Part Miyake, K., and O. Narikiyo, 1994b, J. Phys. Soc. Jpn. 63, 3821. II). Miyano, K., T. Tanaka, Y. Tomioka, and Y. Tokura, 1997, Mattheiss, L. F., 1974, Phys. Rev. B 10, 995. Phys. Rev. Lett. 78, 4257.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1257

Miyasaka, S., H. Takagi, Y. Sekine, H. Takahashi, N. Moˆ ri, Motome, Y., and M. Imada, 1998, cond-mat/9808124. and R. J. Cava, 1998, Phys. Rev. B (in press). Motoyama, N., H. Eisaki, and S. Uchida, 1996, Phys. Rev. Lett. Mizoguchi, M., 1978, J. Phys. Soc. Jpn. 44, 1512. 76, 3212. Mizokawa, T., and A. Fujimori, 1995, Phys. Rev. B 51, 12880. Mott, N. F., 1937, Proc. Phys. Soc. London, Ser. A 49, 72. Mizokawa, T., and A. Fujimori, 1996a, Phys. Rev. B 53, R4201. Mott, N. F., 1949, Proc. Phys. Soc. London, Ser. A 62, 416. Mizokawa, T., and A. Fujimori, 1996b, Phys. Rev. B 54, 5368. Mott, N. F., 1956, Can. J. Phys. 34, 1356. Mizokawa, T., and A. Fujimori, 1997, Phys. Rev. B 56, R493. Mott, N. F., 1961, Philos. Mag. 6, 287. Mizokawa, T., A. Fujimori, T. Arima, Y. Tokura, N. Moˆ ri, and Mott, N. F., 1990, Metal-Insulator Transitions (Taylor and J. Akimitsu, 1995, Phys. Rev. B 52, 13865. Francis, London/Philadelphia). Mizokawa, T., A. Fujimori, H. Namatame, K. Akeyama, and Mu¨ ller-Hartman, E., 1989, Z. Phys. B 74, 507. N. Kosugi, 1994, Phys. Rev. Lett. 49, 7193. Murakami, Y., H. Kawada, H. Kawata, M. Tanaka, T. Arima, Mizokawa, T., H. Namatame, A. Fujimori, K. Akeyama, H. Y. Moritomo, and Y. Tokura, 1998, Phys. Rev. Lett. 80, 1932. Kondoh, H. Kuroda, and N. Kosugi, 1991, Phys. Rev. Lett. Muramoto, S., Y. Kitaoka, K. Magishi, K. Ishida, K. Asayama, 67, 1638. M. Uehara, T. Nagata, and J. Akimitsu, 1997, unpublished. Mizokawa, T., K. Ootomo, T. Konishi, A. Fujimori, Z. Hiroi, Mutou, T., and D. S. Hirashima, 1996, Phys. Rev. B 54, 9058. N. Kobayashi, and M. Takano, 1997, Phys. Rev. B 55, Mutou, T., H. Takahashi, and D. S. Hirashima, 1997, J. Phys. R13373. Soc. Jpn. 66, 2781. Mizuhashi, K., K. Takenaka, Y. Fukuzumi, and S. Uchida, Myron, H. W., 1980, Physica B 99, 243. 1995, Phys. Rev. B 52, R3884. Nagano, T., Y. Tomioka, Y. Nakayama, K. Kishio, and K. Ki- Moeller, G., Q. Si, G. Kotliar, M. J. Rozenberg, and D. S. tazawa, 1993, Phys. Rev. B 48, 9689. Fisher, 1995, Phys. Rev. Lett. 74, 2082. Nagaoka, Y., 1966, Phys. Rev. 147, 392. Mohan Ram, R. A., L. Ganpathi, P. Ganguly, and C. N. Rao, Nagaosa, N., 1997, unpublished. 1984, J. Solid State Chem. 63, 139. Nagaosa, N., and P. A. Lee, 1990, Phys. Rev. Lett. 64, 2450. Momono, N., M. Ido, T. Nakano, M. Oda, Y. Okajima, and K. Nagaosa, N., and P. A. Lee, 1997, cond-mat/9702103. Yamaya, 1994, Physica C 233, 395. Nagaosa, N., and T-K. Ng, 1995, Phys. Rev. B 51, 15588. Momono, N., M. Ido, M. Oda, N. Yamada, and A. Onodera, Nagata, T., H. Fujino, J. Akimitsu, M. Nishi, K. Kakurai, and 1991, Physica C 183, 241. Mook, H. A., M. Yethiraj, G. Aeppli, T. E. Mason, and T. A. S. Katano, 1998, private communication. Armstrong, 1993, Phys. Rev. Lett. 70, 3490. Nagata, T., M. Uehara, J. Goto, J. Akimitsu, N. Motoyama, H. Moreo, A., and E. Dagotto, 1990, Phys. Rev. B 42, 4786. Eisaki, S. Uchida, H. Takahashi, T. Nakanishi, and N. Moˆ ri, Moreo, A., E. Dagotto, T. Jolicoeur, and J. Riera, 1990, Phys. 1998, Phys. Rev. Lett. 81, 1090. Rev. B 42, 6283. Nakamura, M., A. Sekiyama, H. Namatame, A. Fujimori, H. Moreo, A., S. Haas, A. W. Sandvik, and E. Dagotto, 1995, Yoshihara, T. Ohtani, A. Misu, and M. Takano, 1994a, Phys. Phys. Rev. B 51, 12045. Rev. B 49, 16191. Moreo, A., D. J. Scalapino, and E. Dagotto, 1991, Phys. Rev. B Nakamura, M., A. Sekiyama, H. Namatame, H. Kino, A. Fuji- 43, 11442. mori, A. Misu, H. Ikoma, M. Matoba, and S. Anzai, 1994b, Moreo, A., D. J. Scalapino, R. L. Sugar, S. R. White, and N. E. Phys. Rev. Lett. 73, 2891. Bickers, 1990, Phys. Rev. B 41, 2313. Nakamura, M., H. Imai, and M. Shiga, 1997, Phys. Rev. Lett. Mori, M., H. Fukuyama, and M. Imada, 1994, J. Phys. Soc. Jpn. 79, 3779. 63, 1639. Nakano, T., M. Oda, C. Manabe, N. Momono, Y. Miura, and Morikawa, K., T. Mizokawa, A. Fujimori, Y. Taguchi, and Y. M. Ido, 1994, Phys. Rev. B 49, 16000. Tokura, 1996, Phys. Rev. B 54, 8446. Namatame, H., A. Fujimori, Y. Tokura, M. Nakamura, K. Morikawa, K., T. Mizokawa, K. Kobayashi, A. Fujimori, H. Yamaguchi, A. Misu, H. Matsubara, S. Suga, H. Eisaki, T. Eisaki, S. Uchida, F. Iga, and Y. Nishihara, 1995, Phys. Rev. Ito, H. Takagi, and S. Uchida, 1990, Phys. Rev. B 41, 7205. B 52, 13711. Narikiyo, O., 1996, Physica C 267, 204. Moritomo, Y., A. Asamitsu, H. Kuwahara, and Y. Tokura, Narikiyo, O., 1997, unpublished. 1996, Nature (London) 380, 141. Narikiyo, O., and K. Miyake, 1994, J. Phys. Soc. Jpn. 63, 4169. Moritomo, Y., H. Kuwahara, Y. Tomioka, and Y. Tokura, Nazarenko, A., K. J. E. Vos, S. Haas, E. Dagotto, and R. J. 1997, Phys. Rev. B 55, 7549. Gooding, 1995, Phys. Rev. B 51, 8676. Moritomo, Y., Y. Tomioka, A. Asamitsu, Y. Tokura, and Y. Negele, J. W., and H. Orland, 1988, Quantum Many-Particle Matsui, 1995, Phys. Rev. B 51, 3297. Systems (Addison Wesley, Redwood City, CA). Moriya, T., 1979, J. Magn. Magn. Mater. 14,1. Nersesyan, A. A., A. Luther, and F. V. Kusmartsev, 1993, Moriya, T., 1985, Spin Fluctuations in Itinerant Electron Mag- Phys. Lett. A 176, 363. netism, Springer Series in Solid-State Sciences No. 56 Neumeier, J. J., A. L. Cornelius, and S. Schilling, 1994, Physica (Springer, Berlin). B 198, 324. Moriya, T., 1995, in Spectroscopy of Mott Insulators and Cor- Neves, E. J., and J. F. Perez, 1986, Phys. Lett. A 114, 331. related Metals, edited by A. Fujimori and Y. Tokura Newns, D. M., and N. Read, 1987, Adv. Phys. 36, 799. (Springer, Berlin), p. 66. Nimkar, S., D. D. Sarma, and H. R. Krishnamurthy, 1993, Moriya, T., and H. Hasegawa, 1980, J. Phys. Soc. Jpn. 48, 1490. Phys. Rev. B 47, 10927. Moriya, T., Y. Takahashi, and K. Ueda, 1990, J. Phys. Soc. Jpn. Nishida, N., et al., 1987, Jpn. J. Appl. Phys., Part 2 26, L1856. 59, 2905. Nishihara, H., and M. Takano, 1981, J. Phys. Soc. Jpn. 50, 426. Motome, Y., and M. Imada, 1996, J. Phys. Soc. Jpn. 65, 2135. Nishikawa, T., S. Shamoto, M. Sera, M. Sato, S. Ohsugi, Y. Motome, Y., and M. Imada, 1997, J. Phys. Soc. Jpn. 66, 1872. Kitaoka, and K. Asayama, 1993, Physica C 209, 553.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1258 Imada, Fujimori, and Tokura: Metal-insulator transitions

Nishikawa, T., J. Takeda, and M. Sato, 1993, J. Phys. Soc. Jpn. Olson, C. G., R. Liu, D. W. Lynch, R. S. List, A. J. Arko, B. W. 62, 2568. Veal, Y. C. Chang, P. Z. Jiang, and A. P. Paulikas, 1990, Phys. Nishikawa, T., J. Takeda, and M. Sato, 1994, J. Phys. Soc. Jpn. Rev. B 42, 381. 63, 1441. Olson, C. G., R. Liu, A.-B. Yang, D. W. Lynch, A. J. Arko, R. Nishikawa, T., Y. Yasui, Y. Kobayashi, and M. Sato, 1996, J. S. List, B. W. Veal, Y. C. Chang, P. Z. Jiang, and A. P. Pau- Phys. Soc. Jpn. 65, 2543. likas, 1989, Science 245, 731. Nishikawa, T., Y. Yasui, and M. Sato, 1994, J. Phys. Soc. Jpn. Ong, N. P., 1990, in Properties of the High Temperature Super- 63, 3218. conductors, edited by D. M. Ginsberg (World Scientific, Sin- Noack, R. M., N. Bulut, D. J. Scalapino, and M. G. Zacher, gapore), Vol. 2, p. 459. 1997, Phys. Rev. B 56, 7162. Ong, N. P., T. W. Jing, T. R. Chien, Z. Z. Wang, T. V. Ra- Norman, M. R., et al., 1997, cond-mat/9710163. makrishnan, J. M. Tarascon, and K. Remschnig, 1991, Physica Normand, B., H. Kohno, and H. Fukuyama, 1995, J. Phys. Soc. C 185-189, 34. Jpn. 64, 3903. Ong, N. P., Z. Z. Wang, J. Clayhold, J. M. Tarascon, L. H. Nozaki, A., H. Yoshikawa, T. Wada, H. Yamauchi, and S. Greene, and W. R. McKinnon, 1987, Phys. Rev. B 35, 8807. Tanaka, 1991, Phys. Rev. B 43, 181. Onoda, M., T. Naka, and H. Nagasawa, 1991, J. Phys. Soc. Jpn. Nozawa, K., M. Nohara, H. Takagi, Z. Hiroi, and M. Takano, 60, 2550. 1997, in Advances in Superconductivity IX, edited by S. Na- Onoda, M., H. Ohta, and H. Nagasawa, 1991, Solid State Com- kajima and M. Murakami (Springer-Verlag, Tokyo, 1997), p. mun. 79, 281. 69. Orenstein, J., G. A. Thomas, A. J. Millis, S. L. Cooper, D. H. Nozie`res, P., and S. Schmitt-Rink, 1985, J. Low Temp. Phys. Rapkine, T. Timusk, L. F. Schneemeyer, and J. V. Waszczak, 59, 195. 1990, Phys. Rev. B 42, 6342. Osafune, T., N. Motoyama, H. Eisaki, and S. Uchida, 1997, Nu¨ cker, N., H. Romberg, X. X. Xi, J. Fink, B. Gegenheimer, Phys. Rev. Lett. 78, 1980. and Z.-X. Zhao, 1989, Phys. Rev. B 39, 6619. Otsuka, H., 1990, J. Phys. Soc. Jpn. 59, 2916. Obradors, X., L. M. Paulius, B. Maples, J. B. Torrance, A. I. Overhauser, A. W., 1962, Phys. Rev. 128, 1437. Nazzal, J. Fontcuberta, and Y. Granados, 1993, Phys. Rev. B Panissod, P., G. Krill, C. Vettier, and R. Madar, 1979, Solid 47, 12353. State Commun. 29, 67. Ogata, M., M. U. Luchini, and T. M. Rice, 1991, Phys. Rev. B Park, C.-H., Z.-X. Shen, A. G. Loeser, D. S. Dessau, D. G. 44, 12083. Mandrus, A. Migliori, J. Sarrao, and Z. K. Fisk, 1995, Phys. Ogata, M., and H. Shiba, 1990, Phys. Rev. B 41, 2326. Rev. B 52, R16981. Ogawa, S., 1979, J. Appl. Phys. 50, 2308. Park, J.-H., and J. W. Allen, 1997, unpublished. Ogawa, S., et al., 1974, Int. J. Magn. 5, 349. Park, J.-H., C. T. Chen, S.-W. Cheong, W. Bao, G. Meigs, V. Oguchi, T., 1995, Phys. Rev. B 51, 1385. Chakarian, and Y. U. Idzerda, 1996, Phys. Rev. Lett. 76, 4215. Oh, S.-J., J. W. Allen, and J. M. Lawrence, 1987, Phys. Rev. B Park, J.-H., L. H. Tjeng, J. W. Allen, P. Metcalf, and C. T. 35, 2267. Chen, 1997a, Phys. Rev. B 55, 12813. Oh, S.-J., J. W. Allen, I. Lindau, and J. C. Mikkelsen, 1982, Park, J.-H., L. H. Tjeng, J. W. Allen, P. Metcalf, C. T. Chen, F. Phys. Rev. B 26, 4845. M. F. de Groot, and G. A. Sawatzky, 1997b, unpublished. Ohata, N., and R. Kubo, 1970, J. Phys. Soc. Jpn. 28, 1402. Park, K. T., K. Terakura, T. Oguchi, A. Yanase, and M. Ikeda, Ohta, H., S. Kimura, E. Kulatov, S. V. Halilov, T. Nanba, M. 1988, J. Phys. Soc. Jpn. 57, 3445. Motokawa, M. Sato, and K. Nagasaka, 1994, J. Phys. Soc. Jpn. Park, S. K., T. Ishikawa, and Y. Tokura, 1998, Phys. Rev. B 58, 63, 4206. 3717. Ohta, Y., T. Tohyama, and S. Maekawa, 1991a, Phys. Rev. B Park, S. K., S. Yamaguchi, and Y. Tokura, 1998, unpublished. 43, 2968. Pearson, A. D., 1958, J. Phys. Chem. Solids 5, 316. Ohta, Y., T. Tohyama, and S. Maekawa, 1991b, Phys. Rev. Peierls, R., 1937, Proc. Phys. Soc. London, Ser. A 49, 72. Lett. 66, 1228. Pellegrin, E., N. Nu¨ cker, J. Fink, S. L. Molodtsov, A. Gutie´r- Ohtani, T., 1974, J. Phys. Soc. Jpn. 37, 701. rez, E. Navas, O. Strebel, Z. Hu, M. Domke, G. Kaindl, S. Ohtani, T., K. Kosuge, and S. Kachi, 1970, J. Phys. Soc. Jpn. Uchida, Y. Nakamura, J. Markl, M. Klauda, G. Saemann- 28, 1588 (Part I); 29, 521 (Part II). Ischenko, A. Krol, J. L. Peng, Z. Y. Li, and R. L. Greene, Okabe, Y., and M. Kikuchi, 1988, J. Phys. Soc. Jpn. 57, 4351. 1993, Phys. Rev. B 47, 3354. Okada, K., A. Kotani, K. Maiti, and D. D. Sarma, 1996, J. Pen, H. F., M. Abbate, A. Fujimori, Y. Tokura, H. Eisaki, S. Phys. Soc. Jpn. 65, 1844. Uchida, and G. A. Sawatzky, 1998 (unpublished). Okada, K., A. Kotani, and B. T. Thole, 1992, J. Electron Spec- Penn, D., 1966, Phys. Rev. 142, 350. trosc. Relat. Phenom. 58, 325. Perdew, J. P., 1986, Phys. Rev. B 33, 8822. Okada, K., T. Uozumi, and A. Kotani, 1994, J. Phys. Soc. Jpn. Perdew, J. P., and Y. Wang, 1986, Phys. Rev. B 33, 8800. 63, 3176. Perdew, J. P., and A. Zunger, 1981, Phys. Rev. B 23, 5048. Okimoto, Y., T. Katsufuji, T. Ishikawa, T. Arima, and Y. Perkins, J. D., J. M. Graybeal, M. A. Kastner, R. J. Birgeneau, Tokura, 1997, Phys. Rev. B 55, 4206. J. P. Falck, and M. Greven, 1993, Phys. Rev. Lett. 71, 1621. Okimoto, Y., T. Katsufuji, T. Ishikawa, A. Urushibara, T. Perring, T., G. Aeppli, T. Kimura, and Y. Tokura, 1998, un- Arima, and Y. Tokura, 1995a, Phys. Rev. Lett. 75, 109. published. Okimoto, Y., T. Katsufuji, Y. Okada, T. Arima, and Y. Perring, T., G. Aeppli, Y. Moritomo, and Y. Tokura, 1998, Tokura, 1995b, Phys. Rev. B 51, 9581. Phys. Rev. Lett. 78, 3197. Okuno, Y., O. Narikiyo, and K. Miyake, 1997, J. Phys. Soc. Pethick, C. J., and G. M. Carneiro, 1973, Phys. Rev. A 7, 304. Jpn. 66, 2389. Pickett, W. E., 1989, Rev. Mod. Phys. 61, 433.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1259

Pickett, W. E., R. E. Cohen, and H. Krakauer, 1990, Phys. Rozenberg, M. J., G. Kotliar, H. Kajueter, G. A. Thomas, D. Rev. B 42, 8764. H. Rapkine, J. M. Honig, and P. Metcalf, 1995, Phys. Rev. Pickett, W. E., S. C. Erwin, and E. C. Ethridge, 1996, con-mat/ Lett. 75, 105. 9611225. Rozenberg, M. J., G. Kotliar, and X. Y. Zhang, 1994, Phys. Pickett, W. E., and D. J. Singh, 1995, Europhys. Lett. 32, 759. Rev. B 49, 10181. Pines, D., and P. Nozie`res, 1989, The Theory of Quantum Liq- Rozenberg, M. J., X. Y. Zhang, and G. Kotliar, 1992, Phys. uids (Benjamin, New York). Rev. Lett. 69, 1236. Poilblanc, D., and T. M. Rice, 1989, Phys. Rev. B 39, 9749. Sachan, V., D. J. Buttrey, J. M. Tranquada, J. E. Lorenzo, and Pouget, J. P., H. Launois, J. P. D’Haenens, P. Merender, and G. Shirane, 1995, Phys. Rev. B 51, 12742. T. M. Rice, 1975, Phys. Rev. Lett. 35, 873. Saitoh, T., A. E. Bocquet, T. Mizokawa, and A. Fujimori, Pouget, J. P., H. Launois, T. M. Rice, P. Dernier, A. Gossard, 1995a, Phys. Rev. B 52, 7934. G. Villeneuve, and P. Hagenmuller, 1974, Phys. Rev. B 10, Saitoh, T., A. E. Bocquet, T. Mizokawa, H. Namatame, A. 1801. Fujimori, M. Abbate, Y. Takeda, and M. Takano, 1995b, Preuss, R., W. Hanke, C. Gro¨ ber, and H. G. Evertz, 1997, Phys. Rev. B 51, 13942. Phys. Rev. Lett. 79, 1122. Saitoh, T., T. Mizokawa, A. Fujimori, M. Abbate, Y. Takeda, Preuss, R., W. Hanke, and W. von der Linden, 1995, Phys. and M. Takano, 1997a, Phys. Rev. B 55, 4257. Rev. Lett. 75, 1344. Saitoh, T., T. Mizokawa, A. Fujimori, M. Abbate, Y. Takeda, Pruschke, Th., D. L. Cox, and M. Jarrell, 1993, Phys. Rev. B and M. Takano, 1997b, Phys. Rev. B 56, 1290. 47, 3553. Saitoh, T., A. Sekiyama, T. Mizokawa, A. Fujimori, K. Ito, H. Puchkov, A. V., P. Fournier, D. N. Basov, T. Timusk, A. Ka- Nakamura, and M. Shiga, 1995, Solid State Commun. 95, 307. pitulnik, and N. N. Kolesnikov, 1996, Phys. Rev. Lett. 77, Sakai, O., and Y. Kuramoto, 1994, Solid State Commun. 89, 3212. 307. Radaelli, P. G., D. G. Hinks, A. W. Mitchell, B. A. Hunter, J. Sakai, T., G. Adachi, and J. Shiokawa, 1977, J. Appl. Phys. 48, L. Wagner, B. Dabrowski, K. G. Vandervoort, H. K. 379. Viswanathan, and J. D. Jorgensen, 1994, Phys. Rev. B 49, Sarma, D. D., S. R. Barman, H. Kajueter, and G. Kotliar, 1996, 4163. Europhys. Lett. 36, 307. Radaelli, P. G., J. D. Jorgensen, A. J. Schultz, J. L. Peng, and Sarma, D. D., N. Shanthi, S. R. Barman, N. Hamada, H. R. L. Greene, 1994, Phys. Rev. B 49, 15322. Sawada, and K. Terakura, 1995, Phys. Rev. Lett. 75, 1126. Rado, G. T., and H. Suhl, 1963, Eds., Magnetism (Academic, Sarma, D. D., N. Shanthi, S. R. Krishnakumar, T. Saitoh, T. New York) Vol. I, p. 25. Mizokawa, A. Sekiyama, K. Kobayashi, A. Fujimori, E. We- Ramirez, A. P., P. L. Gammel, S-W. Cheong, D. J. Bishop, and schke, R. Meier, G. Kaindl, Y. Takeda, and M. Takano, 1996, P. Chandra, 1996, Phys. Rev. Lett. 76, 447. Phys. Rev. B 53, 6873. Reger, J. D., and A. P. Young, 1988, Phys. Rev. B 37, 5978. Satpathy, S., Z. S. Popovic´, and F. R. Vukajlovic´, 1996, Phys. Renner, C., et al., 1998, Phys. Rev. Lett. 80, 149. Rev. Lett. 76, 960. Rice, T. M., 1995, in Spectroscopy of Mott Insulators and Cor- Sawada, H., N. Hamada, K. Terakura, and T. Asada, 1996, rrelated Metals, edited by A. Fujimori and Y. Tokura Phys. Rev. B 53, 12742. (Springer, Berlin), p. 221. Sawatzky, G. A., and J. W. Allen, 1984, Phys. Rev. Lett. 53, Rice, T. M., and W. R. Brinkman, 1972, Phys. Rev. B 5, 4350. 2339. Rice, T. M., S. Gapalan, and M. Sigrist, 1993, Europhys. Lett. Sawatzky, G. A., and D. Post, 1979, Phys. Rev. B 20, 1546. 23, 445. Sayer, M., R. Chen, R. Fletcher, and A. Mansingh, 1975, J. Rice, T. M., H. Launois, and J. P. Pouget, 1994, Phys. Rev. Phys. C 8, 2053. Lett. 73, 3042. Sayetat, F., M. Ghedira, J. Chenavas, and M. Marezio, 1982, J. Rice, T. M., and M. Sigrist, 1996, J. Phys.: Condens. Matter 7, Phys. C 15, 1627. L643. Scalapino, D. J., 1990, in High Temperature Superconductivity: Rice, T. M., and F. C. Zhang, 1989, Phys. Rev. B 39, 815. Proceedings—The Los Alamos Symposium 1989, edited by Robey, S. W., V. E. Henrich, C. Eylem, and B. W. Eichhorn, K. S. Bedell, D. Coffey, D. E. Meltzer, David Pines, and Rob- 1995, Phys. Rev. B 52, 2395. ert Schrieffer (Addison-Wesley, Redwood City, CA), p. 400. Robey, S. W., L. T. Hudson, C. Eylem, and B. W. Eichorn, Scalapino, D. J., 1994, in Proceedings of the International 1993, Phys. Rev. B 48, 562. School of Physics, Enrico Fermi, Course CXXI, edited by R. Romberg, H., M. Alexander, N. Nu¨ cker, P. Adelmann, and J. A. Broglia and J. R. Schrieffer (North-Holland, Amsterdam), Fink, 1990, Phys. Rev. B 42, 8768. p. 95. Rossat-Mignod, J., et al., 1991a, Physica B 169, 58. Scalapino, D. J., 1995, Phys. Rep. 250, 329. Rossat-Mignod, J., L. P. Regnault, C. Vettier, P. Bourges, P. Scalapino, D. J., S. R. White, and S. Zhang, 1993, Phys. Rev. B Burlet, J. Bossy, J. Y. Henry, and G. Lapertot, 1991b, Physica 47, 7995. C 185-189, 86. Scalettar, R. T., D. J. Scalapino, R. L. Sugar, and D. Toussaint, Roth, L. M., 1966, Phys. Rev. 149, 306. 1989, Phys. Rev. B 39, 4711. Roth, L. M., 1967, J. Appl. Phys. 38, 1065. Schiffer, P., A. P. Ramirez, W. Bao, and S.-W. Cheong, 1995, Rotter, L. D., et al., 1991, Phys. Rev. Lett. 67, 2741. Phys. Rev. Lett. 75, 3336. Rozenberg, M. J., 1996a, Phys. Rev. B 55, 4855. Schiller, A., and K. Ingersent, 1995, Phys. Rev. Lett. 75, 113. Rozenberg, M. J., 1996b, Preprint No. cond-mat/9612089. Schlesinger, Z., R. T. Collins, F. Holtzberg, C. Feild, S. H. Rozenberg, M. J., I. H. Inoue, H. Makino, F. Iga, and Y. Nishi- Blanton, U. Welp, G. W. Crabtree, Y. Fang, and J. Z. Liu, hara, 1996, Preprint No. cond-mat/9603051. 1990, Phys. Rev. Lett. 65, 801.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1260 Imada, Fujimori, and Tokura: Metal-insulator transitions

Schlesinger, Z., R. T. Collins, L. D. Rotter, F. Holtzberg, C. Smith, K. E., and V. E. Henrich, 1994, Phys. Rev. B 50, 1382. Feild, U. Welp, G. W. Crabtree, J. Z. Liu, Y. Fang, K. G. Solovyev, I., N. Hamada, and K. Terakura, 1996a, Phys. Rev. Vandervoort, and S. Fleshler, 1994, Physica C 235-240, 49. Lett. 76, 4825. Schlesinger, Z., Z. Fisk, H.-T. Zhang, M. B. Maple, J. F. Di- Solovyev, I., N. Hamada, and K. Terakura, 1996b, Phys. Rev. Tusa, and G. Aeppli, 1993, Phys. Rev. Lett. 71, 1748. B 53, 7158. Schultz, A. J., J. D. Jorgensen, J. L. Peng, and R. L. Greene, Solyom, J., 1979, Adv. Phys. 28, 201. 1996, Phys. Rev. B 53, 5157. Sommerfeld, A., 1928, Z. Phys. 47,1. Schulz, H. J., 1989, J. Phys. (Paris) 50, 2833. Sorella, S., R. Car, S. Baroni, and M. Parrinello, 1989, Euro- Schulz, H. J., 1990a, Phys. Rev. Lett. 64, 1445. phys. Lett. 8, 663. Schulz, H. J., 1990b, Phys. Rev. Lett. 64, 2831. So”rensen, E. S., M. Wallin, S. M. Girvin, and A. P. Young, Schulz, H. J., 1990c, Phys. Rev. Lett. 65, 2462. 1992, Phys. Rev. Lett. 69, 828. Schulz, H. J., 1991, Int. J. Mod. Phys. B 5, 57. Sparks, J. T., and T. Komoto, 1967, Phys. Lett. 25, 398. Searle, C. W., and S. T. Wang, 1969, Can. J. Phys. 47, 2703. Sparks, J. T., and T. Komoto, 1968, Rev. Mod. Phys. 40, 752. Sega, I., and P. Prelovsˇek, 1990, Phys. Rev. B 42, 892. Sreedhar, K., J. M. Honig, M. Darwin, M. McElfresh, P. M. Sekiyama, A., A. Fujimori, S. Aonuma, H. Sawa, and R. Kato, Shand, J. Xu, B. C. Crooker, and J. Spalek, 1992, Phys. Rev. 1995, Phys. Rev. B 51, 13899. B 46, 6382. Shastry, B. S., B. I. Shraiman, and R. R. P. Singh, 1993, Phys. Stafford, C. A., and A. J. Millis, 1993, Phys. Rev. B 48, 1409. Rev. Lett. 70, 2004. Stechel, E. B., and D. R. Jennison, 1998, Phys. Rev. B 38, 4632. Shastry, B. S., and B. Sutherland, 1990, Phys. Rev. Lett. 65, Stephan, W., and P. Horsch, 1990, Phys. Rev. B 42, 8736. 243. Stephan, W., and P. Horsch, 1991, Phys. Rev. Lett. 66, 2258. Shen, Z.-X., J. W. Allen, J. J. Yeh, J.-S. Kang, W. Ellis, W. Stephan, W., and P. Horsch, 1992, Int. J. Mod. Phys. B 6, 589. Spicer, I. Lindau, M. B. Maple, Y. D. Dalichaouch, M. S. Sternlieb, B. J., J. P. Hill, U. C. Wildgruber, G. M. Luke, B. Torikachvili, J. Z. Sun, and T. H. Geballe, 1987, Phys. Rev. B Nachumi, Y. Moritomo, and Y. Tokura, 1996, Phys. Rev. 36, 8414. Lett. 76, 2169. Shen, Z.-X., and D. S. Dessau, 1995, Phys. Rep. 253,1. Sternlieb, B. J., G. Shirane, J. M. Tranquada, M. Sato, and S. Shen, Z.-X., R. S. List, D. S. Dessau, B. O. Wells, O. Jepsen, Shamoto, 1993, Phys. Rev. B 47, 5320. A. J. Arko, R. Barttlet, C. K. Shih, F. Parmigiani, J. C. Sternlieb, B. J., J. M. Tranquada, G. Shirane, M. Sato, and S. Huang, and P. A. P. Lindberg, 1991a, Phys. Rev. B 44, 3604. Shamoto, 1994, Phys. Rev. B 50, 12915. Shen, Z.-X., D. S. Dessau, B. O. Wells, C. G. Olson, D. B. Sugano, S., Y. Tanabe, and H. Kamimura, 1970, Multiplets of Mitzi, L. Lombardo, R. S. List, and A. J. Arko, 1991b, Phys. Transition-Metal Ions in Crystals (Academic, New York). Rev. B 44, 12098. Susaki, T., A. Sekiyama, K. Kobayashi, T. Mizokawa, A. Fuji- Shepherd, J. P., J. W. Koenitzer, R. Arago´ n, C. J. Sanberg, and mori, M. Tsunekawa, T. Muro, T. Matsushita, S. Suga, H. J. M. Honig, 1985, Phys. Rev. B 31, 1107. Ishii, T. Hanyu, A. Kimura, H. Namatame, M. Taniguchi, T. Shiba, H., R. Shiina, and A. Takahashi, 1997, J. Phys. Soc. Jpn. Miyahara, F. Iga, M. Kasaya, and H. Harima, 1996, Phys. 66, 941. Rev. Lett. 77, 4269. Shiina, R., T. Nishitani, and H. Shiba, 1997, J. Phys. Soc. Jpn. Suzuki, K., K. Kishio, T. Hasegawa, and K. Kitazawa, 1990, 66, 3159. Physica C 166, 357. Shikano, M., T.-K. Huang, Y. Inaguma, M. Itoh, and T. Naka- Suzuki, M., 1976, Prog. Theor. Phys. 56, 1454. mura, 1994, Solid State Commun. 90, 115. Suzumura, Y., Y. Hasegawa, and H. Fukuyama, 1988, J. Phys. Shin, S., S. Suga, M. Taniguchi, M. Fujisawa, H. Kanzaki, A. Soc. Jpn. 57, 2768. Fujimori, H. Daimon, Y. Ueda, K. Kosuge, and S. Kachi, Suzuura, H., and N. Nagaosa, 1997, Phys. Rev. B 56, 3548. 1990, Phys. Rev. B 41, 4993. Suzuura, H., H. Yasuhara, A. Furusaki, N. Nagaosa, and Y. Shin, S., Y. Tezuka, T. Kinoshita, T. Ishii, T. Kashiwakura, M. Tokura, 1996, Phys. Rev. Lett. 76, 2579. Takahashi, and Y. Suda, 1995, J. Phys. Soc. Jpn. 64, 1230. Svane, A., and O. Gunnarsson, 1988a, Europhys. Lett. 7, 171. Si, Q., and G. Kotliar, 1993, Phys. Rev. B 48, 13881. Svane, A., and O. Gunnarsson, 1988b, Phys. Rev. B 37, 9919. Si, Q., Y. Zha, K. Levin, and J. P. Lu, 1993, Phys. Rev. B 47, Svane, A., and O. Gunnarsson, 1990, Phys. Rev. Lett. 65, 1148. 9055. Swendsen, R. H., and J. S. Wang, 1987, Phys. Rev. Lett. 58, 86. Siegrist, T., L. F. Schneemeyer, S. A. Sunshine, and J. V. Tachiki, M., T. Koyama, and S. Takahashi, 1994, Phys. Rev. B Waszczak, 1988, Mater. Res. Bull. 23, 1429. 50, 7065. Singh, D. J., 1994, Phys. Rev. B 49, 1580. Taguchi, Y., Y. Tokura, T. Arima, and F. Inaba, 1993, Phys. Singh, D. J., 1995, Phys. Rev. B 52, 1358. Rev. B 48, 511. Singh, D. J., 1996, J. Appl. Phys. 79, 4818. Tajima, K., Y. Endoh, J. E. Fischer, and G. Shirane, 1988, Singh, D. J., and W. E. Pickett, 1991, Phys. Rev. B 44, 7715. Phys. Rev. B 38, 6954. Siratori, K., S. Suga, M. Taniguchi, K. Soda, S. Kimura, and A. Tajima, S., 1997, Superconductivity Review (Gordon and Yanase, 1986, J. Phys. Soc. Jpn. 55, 690. Breach Science Publishers, in press). Siratori, K., S. Todo, and S. Kimura, 1988, J. Phys. Soc. Jpn. 57, Tajima, S., J. Schutzmann, and S. Miyamoto, 1995, Solid State 2093. Commun. 95, 759. Sjostrand, M. E., and P. H. Keesom, 1973, Phys. Rev. B 7, Takagi, H., B. Batlogg, H. L. Kao, J. Kwo, R. J. Cava, J. J. 3558. Krajewski, and W. F. Peck, Jr., 1992a, Phys. Rev. Lett. 69, Slater, J. C., 1951, Phys. Rev. 82, 538. 2975. Slater, J. C., 1974, Quantum Theory of Molecules and Solids Takagi, H., R. J. Cava, M. Marezio, B. Batlogg, J. J. Krajewski, (McGraw-Hill, New York). W. F. Peck, Jr., P. Bordet, and D. E. Cox, 1992b, Phys. Rev. Smith, K. E., and V. E. Henrich, 1988, Phys. Rev. B 38, 5965. Lett. 68, 3777.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1261

Takagi, H., H. Eisaki, S. Uchida, and R. J. Cava, 1995, in Spec- Thomas, G. A., J. Orenstein, D. H. Rapkine, M. Capizzi, A. J. troscopy of Mott Insulator and Correlated Metals, edited by Millis, R. N. Bhatt, L. F. Schneemeyer, and J. V. Waszczak, A. Fujimori and Y. Tokura (Springer, Berlin), p. 185. 1988, Phys. Rev. Lett. 61, 1313. Takagi, H., T. Ido, S. Ishibashi, M. Uota, S. Uchida, and Y. Thomas, G. A., D. H. Rapkine, S. A. Carter, A. J. Millis, T. F. Tokura, 1989, Phys. Rev. B 40, 2254. Rosenbaum, P. Metcalf, and J. M. Honig, 1994, Phys. Rev. Takagi, H., S. Uchida, K. Kitazawa, and S. Tanaka, 1987, Jpn. Lett. 73, 1529. J. Appl. Phys., Part 2 26, L123. Thomas, G. A., D. H. Rapkine, S. L. Cooper, S-W. Cheong, Takagi, H., S. Uchida, and Y. Tokura, 1989, Phys. Rev. Lett. and A. S. Cooper, 1991, Phys. Rev. Lett. 67, 2906. 62, 1197. Thouless, D. J., 1974, Phys. Rep., Phys. Lett. 13C, 94. Takagi, S., H. Yasuoka, S. Ogawa, and J. H. Wernick, 1981, J. Thuler, M. R., R. L. Benbow, and Z. Hurych, 1983, Phys. Rev. Phys. Soc. Jpn. 50, 2539. B 27, 2082. Takahashi, M., 1972, Prog. Theor. Phys. 47, 69. Thurston, T. R., R. J. Birgeneau, M. A. Kastner, N. W. Preyer, Takahashi, M., 1989, Phys. Rev. B 40, 2494. G. Shirane, Y. Fujii, K. Yamada, Y. Endoh, K. Kakurai, M. Takahashi, M., and J. Igarashi, 1996, Ann. Phys. (N.Y.) 5, 247. Matsuda, Y. Hidaka, and T. Murakami, 1989, Phys. Rev. B Takahashi, M., and J. Kanamori, 1991, J. Phys. Soc. Jpn. 60, 40, 4585. 3154. Thurston, T. R., P. M. Gehring, G. Shirane, R. J. Birgeneau, Takahashi, T., H. Matsuyama, H. Katayama-Yoshida, Y. Ok- M. A. Kastner, Y. Endoh, M. Matsuda, K. Yamada, H. abe, S. Hosoya, K. Seki, H. Fujimoto, M. Sato, and H. Inoku- Kojima, and I. Tanaka, 1992, Phys. Rev. B 46, 9128. chi, 1989, Phys. Rev. B 39, 6636. Thurston, T. R., M. Matsuda, K. Kakurai, K. Yamada, Y. En- Takahashi, Y., and T. Moriya, 1979, J. Phys. Soc. Jpn. 46, 1452. doh, R. J. Birgeneau, P. M. Gehring, Y. Hidaka, M. A. Kast- Takano, M., H. Kosugi, N. Nakanishi, M. Shimada, T. Wada, ner, T. Murakami, and G. Shirane, 1990, Phys. Rev. Lett. 65, 263. and M. Koizumi, 1977, J. Phys. Soc. Jpn. 43, 1101. Tobin, J. G., C. G. Olson, C. Gu, J. Z. Liu, F. R. Solal, M. J. Takano, M., S. Nasu, T. Abe, K. Yamamoto, S. Endo, Y. Fluss, R. H. Howell, J. C. O’Brien, H. B. Radousky, and P. A. Takeda, and J. B. Goodenough, 1991, Phys. Rev. Lett. 67, Sterne, 1992, Phys. Rev. B 45, 5563. 3267. Tohyama, T., and S. Maekawa, 1992, Physica C 191, 193. Takano, M., and Y. Takeda, 1983, Bull. Inst. Chem. Res., Tokunaga, M., N. Miura, Y. Tomioka, and Y. Tokura, 1997, Kyoto Univ. 61, 406. Phys. Rev. B 57, 5259. Takeda, J., T. Nishikawa, and M. Sato, 1993, J. Phys. Soc. Jpn. Tokura, Y., 1992a, Jpn. J. Appl. Phys. Ser. 7, 14. 62, 2571. Tokura, Y., 1992b, Comments Condens. Matter Phys. 16, 69. Takeda, J., et al., 1994, J. Phys. Soc. Jpn. 63, 3564. Tokura, Y., 1994, Physica C 235-240, 138. Takegahara, K., 1994, J. Electron Spectrosc. Relat. Phenom. Tokura, Y., 1997, unpublished. 66, 303. Tokura, Y., and T. Arima, 1990, Jpn. J. Appl. Phys., Part 2 29, Takenaka, K., K. Mizuhashi, H. Takagi, and S. Uchida, 1994, 2388. Phys. Rev. B 50, 6534. Tokura, Y., S. Koshihara, T. Arima, H. Takagi, S. Ishibashi, T. Takigawa, M., E. T. Ahrens, and Y. Ueda, 1996, Phys. Rev. Ido, and S. Uchida, 1990, Phys. Rev. B 41, 11657. Lett. 76, 283. Tokura, Y., H. Kuwahara, Y. Moritomo, Y. Tomioka, and A. Takigawa, M., P. C. Hammel, R. H. Heffner, Z. Fisk, J. L. Asamitsu, 1996a, Phys. Rev. Lett. 76, 3184. Smith, and R. B Schwarz, 1989, Phys. Rev. B 39, 300. Tokura, Y., Y. Okimoto, S. Yamaguchi, H. Taniguchi, T. Takigawa, M., and D. B. Mitzi, 1994, Phys. Rev. Lett. 73, 1287. Kimura, and H. Takagi, 1998, Phys. Rev. B 58, R1699. Takigawa, M., A. P. Reyes, P. C. Hammel, J. D. Thompson, R. Tokura, Y., Y. Taguchi, Y. Okada, Y. Fujishima, T. Arima, K. H. Heffner, Z. Fisk, and K. C. Ott, 1991, Phys. Rev. B 43, 247. Kumagai, and Y. Iye, 1993, Phys. Rev. Lett. 70, 2126. Tamasaku, K., Y. Nakamura, and S. Uchida, 1992, Phys. Rev. Tokura, Y., H. Takagi, and S. Uchida, 1989, Nature (London) Lett. 69, 1455. 337, 345. Tanabe, T., A. Katsufuji, and Y. Tokura, 1998, unpublished. Tokura, Y., Y. Tomioka, H. Kuwahara, A. Asamitsu, Y. Mori- Tanamoto, T., H. Kohno, and H. Fukuyama, 1992, J. Phys. tomo, and M. Kasai, 1996b, J. Appl. Phys. 79, 5288. Soc. Jpn. 61, 1886. Tokura, Y., J. B. Torrance, T. C. Huang, and A. I. Nazzal, Tanamoto, T., H. Kohno, and H. Fukuyama, 1993, J. Phys. 1988, Phys. Rev. B 38, 7156. Soc. Jpn. 62, 717. Tokura, Y., A. Urushibara, Y. Moritomo, T. Arima, A. Tanamoto, T., H. Kohno, and H. Fukuyama, 1994, J. Phys. Asamitsu, G. Kido, and N. Furukawa, 1994, J. Phys. Soc. Jpn. Soc. Jpn. 63, 2739. 63, 3931. Taniguchi, S., T. Nishikawa, Y. Yasui, Y. Kobayashi, M. Sato, Tomioka, Y., A. Asamitsu, H. Kuwahara, Y. Moritomo, and T. Nishioka, M. Kontani, and K. Sano, 1995, J. Phys. Soc. Jpn. Y. Tokura, 1996, Phys. Rev. B 53, R1689. 64, 2758. Tomioka, Y., A. Asamitsu, Y. Moritomo, and Y. Tokura, 1995, Temmerman, W. M., Z. Szotek, and H. Winter, 1993, Phys. J. Phys. Soc. Jpn. 64, 3626. Rev. B 47, 11533. Tomonaga, S., 1950, Prog. Theor. Phys. 5, 349. Terakura, K., 1987 (unpublished). Torrance, J. B., P. Lacorre, C. Asavaroengchai, and R. M. Terakura, K., T. Oguchi, A. R. Williams, and J. Ku¨ bler, 1984, Metzger, 1991, Physica C 182, 351. Phys. Rev. B 30, 4734. Torrance, J. B., P. Lacorre, A. I. Nazzal, E. J. Ansaldo, and C. Thio, T., and J. W. Bennett, 1994, Phys. Rev. B 50, 10574. Niedermayer, 1992, Phys. Rev. B 45, 8209. Thio, T., T. R. Thurston, N. W. Preyer, P. J. Picone, M. A. Tranquada, J. M., et al., 1988, Phys. Rev. Lett. 60, 156. Kastner, H. P. Jenssen, D. R. Gabbe, C. Y. Chen, R. J. Bir- Tranquada, J. M., D. J. Buttrey, V. Sachan, and J. E. Lorenzo, geneau, and A. Aharony, 1988, Phys. Rev. B 38, 905. 1993, Phys. Rev. Lett. 73, 1003.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 1262 Imada, Fujimori, and Tokura: Metal-insulator transitions

Tranquada, J. M., J. E. Lorenzo, D. J. Buttrey, and V. Sachan, van Santen, J. H., and G. H. Jonker, 1950, Physica (Amster- 1995, Phys. Rev. B 52, 3581. dam) 16, 337; 16, 599. Tranquada, J. M., B. J. Sternlieb, J. D. Axe, Y. Nakamura, and van Veenendaal, M. A., G. A. Sawatzky, and W. A. Groen, S. Uchida, 1995, Nature (London) 375, 561. 1994, Phys. Rev. B 49, 1407. Tranquada, J. M., P. Wochner, and D. J. Buttrey, 1997, Phys. Van Zandt, L. L., J. M. Honig, and J. B. Goodenough, 1968, J. Appl. Phys. 39, 594. Rev. Lett. 79, 2133. Varma, C. M., 1996a, Phys. Rev. B 54, 7328. Trotter, H. F., 1959, Proc. Am. Math. Soc. 10, 545. Varma, C. M., 1996b, Preprint No. cond-mat/9606086. ¨ Troyer, M., H. Tsunetsugu, and D. Wurtz, 1994, Phys. Rev. B Varma, C. M., P. B. Littlewood, S. Schmitt-Rink, E. Abra- 50, 13515. hams, and A. E. Ruckenstein, 1989, Phys. Rev. Lett. 63, 1996; Tsuda, N., K. Nasu, A. Yanase, and K. Siratori, 1991, Elec- 64, 497(E). tronic Conduction in Oxides (Springer, Berlin). Varma, C. M., S. Schmitt-Rink, and E. Abrahams, 1987, Solid Tsuei, C. C., A. Gupta, and G. Koren, 1989, Physica C 161, State Commun. 62, 681. 415. Vasanthacharya, N. Y., P. Ganguly, J. B. Goodenough, and C. Tsujioka, T., T. Mizokawa, A. Fujimori, and Y. Tokura, 1997, N. R. Rao, 1984, J. Phys. C 17, 2745. unpublished. Verge´s, J. A., E. Louis, P. S. Lomdahl, F. Guinea, and A. R. Tsunetsugu, H., and M. Imada, 1997, J. Phys. Soc. Jpn. 66, Bishop, 1991, Phys. Rev. B 43, 6099. 1876. Verwey, E. J. W., 1939, Nature (London) 144, 327. Tsunetsugu, H., and M. Imada, 1998, J. Phys. Soc. Jpn. 67, Verwey, E. J. W., and P. W. Haaymann, 1941, Physica (Am- 1864. sterdam) 8, 979. Tsunetsugu, H., M. Troyer, and T. M. Rice, 1994, Phys. Rev. B Verwey, E. J. W., P. W. Haaymann, and F. C. Romeijn, 1947, 49, 16078. J. Chem. Phys. 15, 174, 181. Vilk, Y. M., L. Chen, and A.-M. S. Tremblay, 1994, Phys. Rev. Tyler, J. M., and J. L. Fry, 1971, Phys. Rev. B 1, 4601. B 49, 13267. Uchida, S., T. Ido, H. Takagi, T. Arima, Y. Tokura, and S. Villeneuve, G., M. Drillon, and P. Hagenmuller, 1973, Mater. Tajima, 1991, Phys. Rev. B 43, 7942. Res. Bull. 8, 1111. Uchida, S., H. Takagi, Y. Tokura, N. Koshihara, and T. Arima, Virosztek, A., and J. Ruvalds, 1990, Phys. Rev. B 42, 4064. Strong Correlation and Superconductivity 1989, in , edited by Visscher, P. B., 1974, Phys. Rev. B 10, 943. H. Fukuyama, S. Maekawa, and A. P. M. Malozemoff Voit, J., 1993, J. Phys.: Condens. Matter 5, 8305. (Springer, Berlin), p. 194. Vojta, T., D. Belitz, R. Narayanan, and T. R. Kirkpatrick, Ueda, K., 1977, J. Phys. Soc. Jpn. 43, 1497. 1996, Europhys. Lett. 36, 191. Ueda, Y., K. Kosuge, and S. Kachi, 1980, J. Solid State Chem. Vollhardt, D., 1984, Rev. Mod. Phys. 56, 99. 31, 171. von Barth, U., and L. Hedin, 1972, J. Phys. C 5, 1629. Ueda, Y., K. Kosuge, S. Kachi, and T. Takada, 1979, J. Phys. von der Linden, W., 1992, Phys. Rep. 220, 53. (Paris), Colloq. C2, 275. von Lo¨ hneysen, H., 1996, J. Phys.: Condens. Matter 8, 9689. Uehara, M., T. Nagata, J. Akimitsu, E. Iiseki, N. Motoyama, S. Wada, H., G. Mori, H. Imai, and M. Shiga, 1997, Physica B Uchida, and R. Eccleston, 1997, unpublished. 237-238, 146. Uehara, M., T. Nagata, J. Akimitsu, H. Takahashi, N. Moˆ ri, Wada, N., Y. Nakamura, and K. Kumagai, 1991, Physica C and K. Kinoshita, 1996, J. Phys. Soc. Jpn. 65, 2764. 185-189, 1177. Uemura, Y. J., et al., 1989, Phys. Rev. Lett. 62, 2317. Wagner, J., W. Hanke, and D. J. Scalapino, 1991, Phys. Rev. B Uemura, Y. J., et al., 1991, Phys. Rev. Lett. 66, 2665. 43, 10517. Ulmke, M., R. T. Scalettar, A. Nazarenko, and E. Dagotto, Wallin, M., E. S. Sorensen, S. M. Girvin, and A. P. Young, 1996, Phys. Rev. B 54, 16523. 1994, Phys. Rev. B 49, 12115. Uozumi, T., K. Okada, and A. Kotani, 1993, J. Phys. Soc. Jpn. Wang, Z. Z., T. R. Chien, N. P. Ong, J. M. Tarascon, and E. 62, 2595. Wang, 1991, Phys. Rev. B 43, 3020. Uozumi, T., K. Okada, A. Kotani, Y. Tezuka, and S. Shin, Warren, W. W., Jr., R. E. Walstedt, G. F. Brennert, R. J. Cava, 1996, J. Phys. Soc. Jpn. 65, 1150. R. Tycko, R. F. Bell, and G. Dabbagh, 1989, Phys. Rev. Lett. Uozumi, T., K. Okada, A. Kotani, R. Zimmermann, P. Steiner, 62, 1193. S. Hu¨ fner, Y. Tezuka, and S. Shin, 1997, J. Electron Spec- Watanabe, H., H. Yamamoto, and K. Ito, 1963, J. Phys. Soc. trosc. Relat. Phenom. 83,9. Jpn. 18, 995. Urushibara, A., Y. Moritomo, T. Arima, A. Asamitsu, G. Watanabe, I., K. Kumagai, Y. Nakamura, and H. Nakajima, Kido, and Y. Tokura, 1995, Phys. Rev. B 51, 14103. 1990, J. Phys. Soc. Jpn. 59, 1932. Usuki, T., N. Kawakami, and A. Okiji, 1989, Phys. Lett. A 135, Watanabe, T., N. Moˆ ri, and T. Mitsui, 1976, Solid State Com- 476. mun. 19, 837. Vaknin, D., S. K. Sinha, D. E. Moncton, D. C. Johnston, J. M. Wegner, F., 1976, Z. Phys. B 25, 327. Newsam, C. R. Safinya, and H. E. King, Jr., 1987, Phys. Rev. Wegner, F., 1979, Z. Phys. B 35, 207. Lett. 58, 2802. Wells, B. O., Z.-X. Shen, A. Matsuura, D. M. King, M. A. van der Heide, H., R. Hemmel, C. F. van Bruggen, and C. Kastner, M. Greven, and R. J. Birgeneau, 1995, Phys. Rev. Haas, 1980, J. Solid State Chem. 33, 17. Lett. 74, 964. van der Laan, G., C. Westra, and G. A. Sawatzky, 1981, Phys. Wen, X.-G., and P. A. Lee, 1996, Phys. Rev. Lett. 76, 503. Rev. B 23, 4369. Wentzcovitch, R. M., W. W. Schulz, and P. B. Allen, 1994, van Dongen, P. G. J., 1991, Mod. Phys. Lett. B 5, 861. Phys. Rev. Lett. 72, 3389. van Elp, J., R. H. Potze, H. Eskes, R. Berger, and G. A. Sa- Werfel, F., G. Dra¨ger, and U. Berg, 1981, Cryst. Res. Technol. watzky, 1991, Phys. Rev. B 44, 1530. 16, 119.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998 Imada, Fujimori, and Tokura: Metal-insulator transitions 1263

Werthheim, G. K., V. A. Jaccarino, J. H. Wernick, J. A. Yamaji, K., and Y. Shimoi, 1994, Physica C 222, 349. Seitchik, H. J. WIlliams, and R. C. Sherwood, 1965, Phys. Yamamoto, A., M. Onoda, E. Takayama-Muromachi, F. Lett. 18, 89. Izumi, T. Ishigaki, and H. Asano, 1990, Phys. Rev. B 42, 4228. Wheatley, J. M., T. C. Hsu, and P. W. Anderson, 1988, Phys. Yamane, H., Y. Miyazaki, and H. Hirai, 1990, J. Ceram. Soc. Rev. B 37, 5897. Jpn. 98, 105. White, R. M., and N. F. Mott, 1971, Philos. Mag. 24, 845. Yanase, A., and K. Siratori, 1984, J. Phys. Soc. Jpn. 53, 312. White, S. R., 1992, Phys. Rev. B 46, 5678. Yasuhara, H., and Y. Tokura, 1998, unpublished. White, S. R., 1993, Phys. Rev. B 48, 10345. Yasui, Y., T. Nishikawa, Y. Kobayashi, M. Sato, T. Nishioka, White, S. R., and D. J. Scalapino, 1996, Preprint No. and M. Kontani, 1995, J. Phys. Soc. Jpn. 64, 3890. cond-mat/9605143. Yasuoka, H., T. Imai, and T. Shimizu, 1989, in Strong Corre- White, S. R., D. J. Scalapino, R. L. Sugar, E. Y. Loh Jr., J. E. lation and Superconductivity, edited by H. Fukuyama, S. Gubernatis, and R. T. Scalettar, 1989, Phys. Rev. B 40, 506. Maekawa, and A. P. Malozemoff (Springer, Berlin), 1989, p. Wilson, A. H., 1931a, Proc. R. Soc. London, Ser. A 133, 458. 254. Wilson, A. H., 1931b, Proc. R. Soc. London, Ser. A 134, 277. Yokoya, T., A. Chainani, T. Takahashi, H. Ding, J. C. Campu- Wilson, J. A., 1985, in The Metallic and Nonmetallic States of zano, H. Katayama-Yoshida, M. Kasai, and Y. Tokura, 1996a, Matter, edited by P. P. Edwards and C. N. R. Rao (Taylor and Phys. Rev. B 54, 13311. Francis, London), p. 215. Yokoya, T., A. Chainani, T. Takahashi, H. Katayama-Yoshida, Wilson, J. A., and G. D. Pitt, 1971, Philos. Mag. 23, 1297. M. Kasai, and Y. Tokura, 1996b, Phys. Rev. Lett. 76, 3009. Wilson, K. G., 1975, Rev. Mod. Phys. 47, 773. Yokoyama, H., and H. Shiba, 1987a, J. Phys. Soc. Jpn. 56, Wolfe, R., J. H. Wernick, and S. E. Haszko, 1965, Phys. Lett. 1490. 19, 449. Yokoyama, H., and H. Shiba, 1987b, J. Phys. Soc. Jpn. 56, Wollan, E. O., and W. C. Koehlar, 1955, Phys. Rev. 100, 545. 3582. Woodfield, B. F., M. L. Wilson, and J. M. Byers, 1997, Phys. Yoshioka, D., 1989a, J. Phys. Soc. Jpn. 58, 32. Rev. Lett. 78, 3201. Yoshioka, D., 1989b, J. Phys. Soc. Jpn. 58, 1516. Xu, X. Q., J. L. Peng, Z. Y. Li, H. L. Ju, and R. L. Greene, Yoshizawa, H., H. Kawano, Y. Tomioka, and Y. Tokura, 1995, 1993, Phys. Rev. B 48, 1112. Phys. Rev. B 52, R13145. Yamada, K., et al., 1987, Solid State Commun. 64, 753. Yoshizawa, H., S. Mitsuda, H. Kitazawa, and K. Tatsumata, Yamada, K., et al., 1998, Phys. Rev. B 57, 6165. 1988, J. Phys. Soc. Jpn. 57, 3686. Yamada, K., Y. Endoh, C.-H. Lee, S. Wakimoto, M. Arai, K. Yoshizawa, H., et al., 1998, unpublished. Ubukata, M. Fujita, S. Hosoya, and S. M. Bennington, 1995, Yunoki, S., J. Hu, A. L. Malvezzi, A. Moreo, N. Furukawa, J. Phys. Soc. Jpn. 64, 2742. and E. Dagotto, 1998, Phys. Rev. Lett. 80, 845. Yamada, K., K. Kakurai, Y. Endoh, T. R. Thurston, M. A. Zaanen, J., and O. Gunnarsson, 1989, Phys. Rev. B 40, 7391. Kastner, R. J. Birgeneau, G. Shirane, Y. Hidaka, and T. Mu- Zaanen, J., and P. B. Littlewood, 1994, Phys. Rev. B 50, 7222. rakami, 1989, Phys. Rev. B 40, 4557. Zaanen, J., and A. M. Oles´, 1996, Ann. Phys. (Leipzig) 5, 224. Yamada, K., J. Wada, Y. Endoh, S. Hosoya, S. Noguchi, S. Zaanen, J., and G. A. Sawatzky, 1990, J. Phys. Chem. Solids Kawamata, and K. Okuda, 1995, Physica C 253, 135. 88,8. Yamada, K., S. Wakimoto, G. Shirane, C. H. Lee, M. A. Kast- Zaanen, J., G. A. Sawatzky, and J. W. Allen, 1985, Phys. Rev. ner, S. Hosoya, M. Greven, Y. Endoh, and R. J. Birgeneau, Lett. 55, 418. 1995, Phys. Rev. Lett. 75, 1626. Zener, C., 1951, Phys. Rev. 82, 403. Yamada, K., and K. Yosida, 1986, Prog. Theor. Phys. 76, 621. Zeyer, R., and G. Zwicknagl, 1988, Solid State Commun. 66, Yamada, Y., O. Hino, S. Nohdo, R. Kanao, T. Inami, and S. 617. Katano, 1996, Phys. Rev. Lett. 77, 904. Zhang, F. C., and T. M. Rice, 1988, Phys. Rev. B 37, 3759. Yamada, Y., M. Mori, Y. Noda, and M. Izumi, 1979, Solid Zhang, S. C., 1997, Science 275, 4126. State Commun. 32, 827. Zhang, S., J. Carlson, and J. E. Gubernatis, 1997, Phys. Rev. B Yamaguchi, S., Y. Okimoto, H. Taniguchi, and Y. Tokura, 55, 7464. 1996a, Phys. Rev. B 53, 2926. Zhang, S., N. Kawashima, J. Carlson, and J. E. Gubernatis, Yamaguchi, S., Y. Okimoto, and Y. Tokura, 1996b, Phys. Rev. 1995, Phys. Rev. Lett. 74, 1500. B 54, R11022. Zhang, X. Y., M. J. Rozenberg, and G. Kotliar, 1993, Phys. Yamaguchi, S., Y. Okimoto, and Y. Tokura, 1997, Phys. Rev. Rev. Lett. 70, 1666. B 55, R8666. Zhang, Z., and S. Satpathy, 1991, Phys. Rev. B 44, 13319. Yamaguchi, S., and Y. Tokura, 1998, unpublished. Zotos, X., and P. Prelovsˇek, 1996, Phys. Rev. B 53, 983.

Rev. Mod. Phys., Vol. 70, No. 4, October 1998